University
stringclasses
19 values
Text
stringlengths
458
20.7k
Colorado School of Mines
ER-4609 134 centerline methods. Good control of the phreatic surface in the vicinity of the embankment face is achieved by either arrangement, provided that the downstream shell materials are sufficiently pervious with respect to the core. Prerequisite to the use of cores as a phreatic surface control method is, of course, that suitable low-permeability natural soils be locally available. Cores of some type are usually mandatory where water will stand directly against the upstream face of a pervious embankment. 6.6 Drainage (after Soderberg and Bush, 1980) From the designer's viewpoint, it is desirable to promote drainage of water from the tailings pond in order to keep the phreatic surface as low as possible and help the consolidation and stability of the embankment. For this reason, the relatively pervious tailings dam is the most common design used. It is also the cheapest because it can be built from the coarse fraction of tailings or from the borrow material. The impervious tailings dam is the least common type and is only used where it is necessary to retain polluted water or low-density solids that are slow to consolidate. In either type, the stability of the dam is of paramount importance and necessary provisions must be made to control seepage through and under the embankment and to control surface runoff into the pond. Water control in a tailings pond is probably the most critical single design item affecting slope stability, followed by (1) weak foundations, (2) slopes that are too steep and too high, and (3) low density of the material in the dike and on the beach. Failures are often caused by
Colorado School of Mines
ER-4609 135 overtopping of pond water, by piping of fine materials because of seepage through the embankment or foundation, and by piping along the outside of decants and culverts because of the lack of or inadequate seep rings. 6.6.1 Drains (after Soderberg and Bush, 1980) The position of the phreatic surface in an embankment is very critical for its stability. Of all the factors controlling stability, the control of water within the embankment is by far the most important. For example, a slope composed of cohesionless sand having a total unit weight of 120 pounds per cubic foot and an angle of internal friction of 35° will remain stable at an angle of 35° regardless of the height of the slope, providing the phreatic surface is below the toe of the slope. If the phreatic surface is raised to exit the downstream face of the slope, the steepest stable angle is reduced to approximately 18°. A relatively impervious starter dam very definitely requires drainage on the upstream side, and a pervious starter dam does not necessarily preclude the need for drains. Should the pervious starter dam become ineffective and there is no drain to fall back on, the phreatic surface can rise and seriously reduce stability. Two of the basic types of drains are (1) the blanket drain, which is coarse gravel sandwiched between protective filter material which carries all the water through and beneath the dam, and (2) the perforated pipe drain, which is surrounded by gravel and a protective filter which collects the drainage and carries it through the dam. The choice of
Colorado School of Mines
ER-4609 136 which one to use depends on the availability of drainage material, drainage capacity required, foundation conditions, and construction cost. Variations and combinations of these two methods can be used, but they must be carefully constructed, and the material used for filters must have proper grain-size graduation. 6.6.2 Blanket Drain (after Soderberg and Bush, 1977) Blanket drains would typically be used in a cross­ valley dam with either a pervious or impervious starter dam. Bedrock or a relatively impervious base is close below the natural ground level and the upstream method of dam building utilizing the tailing sands is planned. The purpose of this blanket drain is to intercept the water that moves downward out of the tailings as well as any springs or artesian water that may come up from below. If springs are found in site investigation or artesian water in drill holes, either from the rock or below a stratum of impervious clay, the blanket drains should have capacity to remove all this water plus additional capacity for that which may not have been discovered. It is very important that this water be removed because in mountainous areas it could have a high head and if trapped below the slime layer in a tailings pond it could exert tremendous upward pressure and greatly reduce the factor of safety of the embankment. The drain consists of a layer of clean gravel up to 18 inches thick extending from above the upstream toe to below the downstream toe and wide enough to cover the main valley bottom. This gravel drain is protected by a 9- to 12- inch filter layer of clean sand and gravel both above and below.
Colorado School of Mines
ER-4609 137 An additional drain of unprocessed sand and gravel up to 3 feet deep is also placed upstream to extend the drain area to catch all the seepage (Figure 43) . These drains must have a catchment ditch filled with cobbles to intercept the drainage and prevent erosion on the downstream face. Where the downstream slope of an embankment is composed of fine-grained materials, water should not be allowed to flow out of this slope. Lowering the phreatic surface increases the stability, permitting the use of steeper slopes, and reduces the volume of construction material needed. In cold climates it is especially important that the drain water be directed through a drainage blanket below the compacted soil so that it will not freeze and raise the phreatic surface causing the entire embankment to become saturated behind a frozen blanket of soil on the downstream face of the starter dam. 6.6.3 Pipe Drain (after Soderberg and Bush, 1977) Where drainage pipes are to be used, the pipes should be designed to withstand the maximum anticipated load of the overlying tailings. When perforated pipe is used, it should be perforated on the bottom half only and laid with the perforations down, with a bed of gravel both top and bottom and graded filter surrounding the gravel (Figure 44). The diameter of the perforations should not be larger than one- half of the 85-percent size of the drainage material surrounding the pipe. Pipe drains can be very satisfactory with a good foundation and careful construction, but the blanket or strip drains may be more fail-safe. Various arrangements of pipe drains can be made. A perforated pipe
Colorado School of Mines
ER-4609 140 parallel to the upstream toe of the starter dam with one or more solid pipes through the dam to the downstream toe is the simplest. This same arrangement can be used as a collection for drains up to 600 feet long running parallel to the valley at right angles to the dam axis and spaced at 50- to 100- foot intervals along the valley floor and walls (Figure 45) . Pipes through the starter dam should not be perforated and should have at least three cutoff collars that extend at least 2 feet to prevent "piping". If the foundation beneath a tailings embankment is compressible and differential settlement is possible, pipe drains should be avoided. The stresses may result in opening pipe joints or breaking the pipe, which might allow internal erosion. 6.6.4 Strip Drain (after Soderberg and Bush, 1977) Strip drains are the same as blanket drains in design and construction except that they are narrow strips of drain material laid in the foundation prior to dam construction. They are laid out to carry drainage through the dam and to outlets beyond the downstream toe of the embankments. The drains are laid out in strategic locations to catch the drainage and must be arranged according to the contours of the foundation. Strip drains can be used upstream from the starter dam in the same manner as blanket or pipe drains. If blanket or pipe drains are placed as much as 500 to 600 feet upstream from the toe of the starter dam and spigoting is started with tailings, the slimes begin settling on the top of the drains in the first 100 feet from the upstream toe of the starter dam. As the sand builds up, this fine
Colorado School of Mines
ER-4609 142 material continues to move uphill in an increasingly thicker layer. The layer of fines or slimes is quite impervious even in thin layers, and under consolidation and drainage it can have a permeability as low as 10-6 or 10”7 centimeters per second. This layer of slimes over a drain renders the drain useless except for the first few feet upstream of the starter dam. For this reason when drains are constructed upstream any distance above the starter dam, the tailings should be cycloned and the coarse underflow should be repulped and spigoted on the dam to make a beach of very pervious sand that will cover the drain. Since this coarse sand will have a slope angle of 4+ percent, depending on grind, etc., it will require a lot of material and come up quite high on the starter dam. After the drain is well covered, spigoting of unclassified tailings can be started; the blanket of slime will not cover the drain, keeping it free and operating for the life of the dam. This type of blanket drain would be used only where bedrock is relatively shallow and the natural soil is saturated or will become saturated from the tailings pond. In areas where the bedrock is 100 to 500+ feet deep and the soil is very pervious (10'2 to 10~4 centimeters per second) the blanket, strip, or pipe drains extending upstream from the starter dam would not be used because the seepage through the bottom would go down toward bedrock and not follow the drain. Nearly every mine has a different set of conditions and each tailings area must be designed accordingly. Because of the layering in a spigoted embankment, the permeability in the horizontal direction may be as much as 5 to 10 times that in a vertical direction, especially if the grind is coarse and the pulp density is low. To determine
Colorado School of Mines
ER-4609 143 the seepage from the pool and from the spigoting on the beach, a flow net should be used to estimate the seepage rate to the drains. The quantity of seepage will depend on the permeability values, hydraulic gradient, and area of flow. In some embankments and possibly all of them, the piezometric head from the downstream toe up and under the beach (on a large dam 500 to 600+ feet distance) is determined more by the water flowing on the beach during discharge than by the water escaping from the pond area. The water in the pond is contained in a saucer of slime with the permeability lowest at the center of the pond and increasing toward the beach. Because of this, control of the height and location of the water pool is important in controlling the seepage. An increase of 1 foot in the pond will flood 200 to 600 feet of beach, which has a higher permeability than that area at a lower elevation. To insure that the drain design is adequate, the upper range of the coefficients of permeability of the embankment should be used in these calculations. In the same way the lower range of permeability of the drainage blanket should be used to be sure that the blanket has a greater capacity than the calculated seepage of the embankment. With a high water table, water can enter the drainage system from the foundation strata, thus increasing the required capacity of the drainage system. Layers of impervious materials in the foundation strata will restrict seepage into the foundation. Calculating the thickness and width of blanket and strip drains is probably worth the effort from a cost standpoint because the difference of cost between 1 foot and 2 feet of gravel over a large area could be considerable. The drains should be as large as practicable considering the cost and availability of materials. They should be uniform
Colorado School of Mines
ER-4609 144 and continuous and constructed of the proper gradation of materials, without which they could become useless. Granular materials incorporated in underdrainage systems should be compatible with the properties of the seepage water they are designed to carry. . Drainage materials composed of carbonate rocks are unsuitable if the seepage collected by the system is acidic. Blanket drains and strip drains should be designed to be capable of passing full design flow when the phreatic surface within the drain is at or below the upper surface of the drainage material. 6.7 Filters and Transition Zones (after Soderberg and Bush, 1977) A filter or "protective filter" is any porous material that has openings small enough to prevent movement of soil into the drain and yet is much more pervious than the soil it is protecting. Transition zones are also filters between a very fine and a very coarse material where several different filter materials have to be used. In the construction of a zoned embankment with the permeability increasing in the downstream direction, filters and transition zones should be placed between layers of significantly different gradation to prevent piping and subsurface erosion (Figure 46). Each situation is different, and the filter design should be governed by field conditions, particularly by the gradation of the soil to be protected and the material protecting it. Two different gradations of filters would probably be used where coarse mine rock was to be placed on the downstream toe of a tailings dam to prevent seepage.
Colorado School of Mines
ER-4609 146 Experiments have shown that filters need not screen out all the particles of the soil but only the coarsest 15 percent, or the Dss, of the soil. These coarser particles (Des and larger) will collect over the filter openings and screen out the smaller particles. Therefore, the screen or holes in a perforated pipe must be smaller than the Des size of the soil. By the same reasoning, if soils are used as filters, the effective diameters of the soil voids must be less than Dss of the soil being filtered. Since effective pore diameter is about 1/5 Die, then Die filter < Dss soil. The filter must provide free drainage, and since the permeability coefficient varies as the square of the grain size, the ratio of permeabilities of over 20 to 1 can be secured by Die filter >5Die soil. To satisfy the criteria for filter materials, thefollowing rules should be applied as illustrated in Figure 47. Rules 1, 2, and 3 assure that the protected soil does . not pass through the filter. All filters allow small amounts of fines to pass and collect in the outlet pipes, so provisions should be made for flushing out the pipe if at all possible. Rules 4 and 5 are to assure that the permeability of the filter material is high enough to take all the seepage from the protected soil. Attempts have been made to develop a universal filter that will filter the finest soil and yet have a Dss large enough that it will not pass through the 5/16-inch perforations of commercial drainage pipes. These filters have such a wide range of size that the particles segregate during handling and placement. A filter with a Cu of 20 or less is not as susceptible to segregation (Rule 6). Transition zones and filters are very important in the
Colorado School of Mines
ER-4609 148 design and construction of some tailings ponds, particularly where drains are required either upstream from and through the starter dam or in the starter dam construction itself. They are especially important where a water-type dam is to be constructed. Where remedial measures are necessary for seepage emerging on the downstream face of a tailings dam, a protective surcharge of any available material can be used with protective filters and drains. The thickness of filters required varies with the head of water from a few inches with a low head up to 10 feet with a high head in a water-type dam. The filter thickness also varies depending on whether it is placed with construction equipment or by hand. Steeply inclined filters placed with construction equipment should be wide enough to accommodate this equipment easily, while on level ground these filters should be at least 3 feet thick. Filters placed by hand should be at least 6 inches thick. In summary, filters must be of proper gradation to protect the soil, have a permeability greater than the soil being protected, and have an increasing permeability in the direction of flow. 6.8 Sand yield (after Soderberg and Bush, 1977) The yield of suitable sand obtained in separating cthe coarser fraction from the raw tailings affects the design and construction of the embankments. A cross-valley dam has a big advantage over a flat-country impoundment where three to four sides have to be built using sand. A shortage of sand makes it difficult to keep the embankment crest above the finer tailings in the pond. With fine grind (55 to 60
Colorado School of Mines
ER-4609 149 percent minus 200 mesh and finer) and in flat country, cyclones with the upstream method may be one way to keep the embankment ahead of the tailings. In this case it may be necessary to use mine waste as a supplement to the sand. Operations with : a) 35 or more acres per 1,000 tons of daily capacity; b) with this fine-grind material ; and c) with tailings areas divided into two separate ponds that are used alternatively; can operate with peripheral discharge if the ultimate dams are kept low and the tailings are allowed to drain and consolidate after each 10+/- foot lift. The yield of acceptable sand from cyclones can be calculated from the gradation of the raw tailings and the characteristics of the cyclones. The rate of embankment construction will depend on the amount of available sand, the length of embankment being built, and the weather, or the number of months a year that it is possible to construct embankment. Using cyclones and the downstream method, each foot of rise takes longer and requires more sand than the previous foot. The use of cyclones with the centerline method is nearly as bad. Spigoting and building an upstream embankment with beach sand requires the same amount of sand each lift except for the increased length as the embankment gets higher. In planning any tailings site, the active time for embankment utilization is far below 100 percent. The time required to build embankment and replace spigots or cyclones and the time necessary to raise the entire line to a new berm are times when the pond is not available for discharging tailings unless they can be "dumped” at some other spot in the pond. For this reason, it is better to have two complete and separate dams. This is especially important at the start of a new operation. With two dams
Colorado School of Mines
ER-4609 150 there can be a complete shutdown of an area so that the sand beach can be drained, dike built, and pipes or cyclones replaced. By alternating sites, a regular schedule of maintenance and operation can be set up; also the annual rise of the embankment is reduced, which improves slope stability. Where the winters are severe, dike building can be done only in the 6 to 8 warmer months to prevent the formation of ice lenses in the beach area. Enough sand must be available to build enough dike in the summer to last through the winter months. Where tailings sand is used for mine stope fill, the amount of sand available for embankment construction is further reduced; of course, the total volume to be impounded is also reduced by this amount. When the grind is such that the proportion of minus 200-mesh tailings is more than 55 to 60 percent, the use of cyclones is almost mandatory in order to save the entire volume of sand for dam building. Under certain conditions, a water-type dam should be considered even though the capital cost is high. For these dams the operating cost is very low. Conditions that may warrant water-type dams are high percentage of slimes, harmful chemicals in the tailings, and, for phosphate clay, slimes with no sand in the tailings. 6.9 Spigots (after Soderberg and Bush, 1977) The most common method of placing tailings on a pond in upstream construction is by spigots spaced 16 to 60 feet apart along a header pipe which is installed on the embankment crest (Figure 48). The material immediately adjacent to the spigots is used for dike constructions on
Colorado School of Mines
ER-4609 152 the next lift and should be the best tailings material available. The gravitational separation of the sand and slime as it flows into the pond ranges from very poor to very good depending on the grind, specific gravity, and pulp density. Figure 36 shows the gradation of the mill tails from mine A, the gradation at the dike, and the change in gradation 1,000 feet from the dike of a material that has about 38 percent minus 200 mesh and is spigoted at 30 percent pulp density by weight. Figure 37 shows the gradations for a material that has 58 percent minus 200 mesh and is spigoted at 48 to 50 percent pulp density. Figure 49 shows the uniformity of the permeability change; the vertical permeability is 8 x 10'3 centimeters per second at the dike and decreases to 9 x 10-4 centimeters per second 7 00 feet from the dike. The horizontal permeability had the same relative decrease in a 1,000-foot distance from the spigot point but was approximately five times the vertical. In mines B and C (Figures 50 and 51) , the horizontal and vertical permeability in surface test pits is about the same at the dike as it is 500 feet from the dike and also near the decant area 3,000 feet from the dike. From very limited testing there is also no apparent difference in permeability between the horizontal and vertical samples from surface test pits. The vertical permeability does definitely decrease with depth and is very low just above the natural soil. These two examples might represent the extremes of what can be expected from spigoting. Mine B could certainly improve the separations of the sand and slimes if it were necessary by decreasing the pulp density 10 to 20 percent.
Colorado School of Mines
ER-4609 155 6.10 Cyclones (after Soderberg and Bush, 1977) As the grind in metal mines becomes finer, 50 to 60 percent minus 200 mesh, cyclones can recover a larger percentage of the sand for dike building than can spigoting, and they are being used for this purpose. They are usually mounted on the crest of the embankment but are sometimes mounted on an abutment at one side of the embankment. From the cyclones the fines are piped to the tailings pond, and the coarser sand drops directly onto the embankment or is repulped and discharged through spigots or launders onto the embankment. In some cases, the fines can be piped upstream many hundred of feet so that no slimes are beneath the dike area. In some cases the slimes may be piped to another tailings pond temporarily. 6.11 Compaction (after Vick, 1983) Compaction of cycloned sand is a significant issue in embankment design, and, as explained above, the various cycloning methods vary in their ability to accommodate compaction procedures. Compaction of embankment sands is often desirable to reduce pore pressure buildup during shear and the possibility of flow slides. But compaction becomes of critical design importance in determining the susceptibility of saturated embankment sands to seismic liquefaction. The site seismicity, internal drainage provisions in the embankment design, and compaction requirements are all interrelated. Liquefaction is most likely to occur for loose, saturated sands under high levels of seismic shaking.
Colorado School of Mines
ER-4609 156 The main design measures against liquefaction are good internal drainage to prevent saturation, and densification of the cycloned tailings sands by compaction. In the extreme, these design measures tend to be mutually exclusive : when the sand is unsaturated, liquefaction is unlikely regardless of density, while if the sand is sufficiently dense, liquefaction is improbable even under complete saturation. Since compaction requires mechanical hauling and placement at considerable expense, it is often less costly in areas of lower seismicity to provide for internal drainage rather than high densities. In areas of high seismicity, both compaction to high densities and internal drainage is recommended. Where high seismicity indicates the need for compaction, a remaining design consideration is the relative density requirement. In reviewing various liquefaction- related relative density criteria, it is apparent that compaction to relative densities of 50-60% is appropriate for areas of moderate seismicity with expected accelerations up to 0.10g. Higher relative densities, in the range of 75%, may be necessary for higher seismicity. Usually, reasonably high relative densities can be obtained for clean sand tailings with only modest compactive effort. The major cost associated with compaction results from the need to haul, spread, and place the sands in thin lifts rather than from the compaction operation itself. Use of the hydraulic cell method, however, allows for compaction without mechanical fill handling and at relatively little cost. Compaction of a soil is usually accomplished by spreading the soil in layers of specified thickness and compacting with a mechanical compactor. A number of arbitrary standards for determining the optimum moisture and
Colorado School of Mines
ER-4609 157 maximum densities have been established to simulate different amounts of effort as applied by the full-sized equipment used in soil construction. The simplest and the most widely used are the Standard Proctor and the Modified Proctor. Compaction may be specified by procedure (type of compactor, layer thickness, number of passes, and moisture content to be used) or by end product (minimum inplace density required). The purpose of compaction may be to increase the fill shear strength or to decrease the fill permeability or both. Variables affecting compaction are: The type of compactor : Pneumatic-tired roller, steel wheel, vibratory steel wheel, sheepsfoot, grid roller, vibratory plate compactors, and tract-type tractors. The weight and energy of the compactor. The thickness of layers. The placement water content. Table 17 shows the compactors recommended to be used to attain 95 to 100 percent of the Standard Proctor on various soils, but during construction regular density testing is necessary. For cleanc ohesionless tailings sand, compaction in thin lifts with proper moisture might be attained with a tractor for building dikes, but it should be checked with inplace density tests. Seldom does the sand have less than 8 percent minus 200-mesh material, and the moisture content is in the bulking range where compaction is difficult.
Colorado School of Mines
ER-4609 158 TABLE 17 Compaction equipment and methods (after Soderberg and Bush, 1977) Requirements for compaction to 95 to 100 percent Standard Proctor maximum density Compacted Possible variations in lift Passes or Dimensions and weight of equipment thickness , coverages equipment inch SHEEPSFOOT ROLLERS For fine-grained soils 6 4 to 6 passes F oot For earth dam, highway, and or dirty coarse-grained for fine­ Soil type contact airfield work, drum of 60- soils with more Chan 20 grained area , inch diameter loaded to percent passing the sq in psi I 5 to 3 tons per lineal No. 200 sieve. Not 6 to 8 passes Fine-grained 5-12 250-500 foot of drum generally is . suitable for clean for coarse­ soil, plas­ utilized. For smaller coarse-grained soils. grained ticity projects, 40-inch-diameter Particularly appropri­ index, >30. drum loaded to 0.75 Co ate for compaction of Fine-grained 7-14 200-4 00 1.75 tons per lineal foot impervious zone for soil, plas­ of drum is used. Foot con­ earth dam or linings ticity tact pressure should be where bonding of lifts index, <30. regulated to avoid shearing is important. Coarse-grained 10-14 150-250 the soil on the 3d or 4th soil. Efficient compaction o soils wetter than optimum requires less contact pressure than for the same soils at lower mois­ ture contents RUBBER-TIRE ROLLERS For clean, coarse­ 10 3 to 5 Tire inflation pressures of 60 to Wide variety of rubber-tire grained soils with 4 to coverages. 80 psi for clean granular mate­ compaction equipment is 8 percent passing the rial or base course and subgrade available. For cohesive No• 200 sieve. compac tion. soils, light-wheel loads, For fine-grained soils 6- 8 4 to 6 Wheel load 18.000 to 25,000 such as provided by wobble- or well-graded, dirty coverages. pounds. Tire inflation pres­ wheel equipment, may be sub­ coarse-grained soils sures in excess of 65 psi for stituted for heavy-wheel with more than 8 per­ fine-grained soils of high load if lift thickness is cent passing the plasticity. For uniform clean decreased. For cohesionless No. 200 sieve. sands or silty fine sands, use soils , large-size tires are large-size tires with pressures desirable to avoid shear of 40 to 50 psi. and rutting SMOOTH-WHEEL ROLLERS Appropriate for subgrade 8-12 4 coverages. Tandem-type rollers for base 3-wheel rollers are obtain­ or base course compac­ course or subgrade compaction. able in wide range of sizes. tion of well-graded 10- to 15-ton weight, 300 to 2-wheel tandem rollers are sand-graveI mixtures. 500 pounds per lineal inch of available from 1 to 20 tons . width of rear roller. 3-axle tandem rollers are May be used for fine­ 6- 8 6 coverages. 3-wheel roller for compaction of generally used in the range grained soils other fine-grained soil ; weights from of 10 to 20 tons. Very than in earth dams. 5 to 6 tons for materials of low heavy rollers are used for Not suitable for clean plasticity to 10 tons for mate­ proof rol ling of subgrade or well-graded sands or rials of high plasticity. base course. silty uniform sands. VIBRATING -BASEPUTE COMPACTORS For coarse-grained soils 8-10 3 coverages Single pads or plates should Vibrating pads or plates are with less than about 12 weigh no less than 200 pounds. available, hand -propel led percent passing No. 200 May be used in tandem where or self-propelled, single or sieve. Best suited for working space is available. For in gangs , with width of cov­ materials with 4 to 8 clean coarse-grained soil, erage from 1-1/2 to 15 feet. percent passing vibration frequency should be no Various types of vibratlng- No. 200, placed thor­ less than 1,600 cycles per drum equipment should be oughly wet. minute. considered for compaction in large areas. CRAWLER-TRACTOR Best suited for coarse­ 10-12 3 to 4 No smaller than 08 tractor with (Tractor weights up to 60,000 grained soils with less coverages. blade, 34,500-pound weight, for 1 pounds, than 4 to 8 percent high compaction. passing No. 200 sieve, placed thoroughly wet. POWER TAMPER OR RAMMER For difficult access, 4-6 for silt 2 coverages. 30-pound minimum weight. Con­ Weights Up to 250 pounds. trench backfill. Suit­ siderable range is tolerable , Foot diameter 4 to 10 able for all inorganic 6 for coarse - depending on materials and grained conditions.
Colorado School of Mines
ER-4609 159 The pneumatic tire has proven to be an excellent compactor for cohesionless and low-cohesion soils, including gravels, sands, clayey sands, silty sands, and even sandy clay. It applies a moderate pressure to a relatively wide area so that enough bearing capacity is developed to support the pressure without failure. The light rollers are capable of compacting soils in 4-inch layers to densities approaching the Standard Proctor maximum at optimum moisture with three or four passes. The heavy rollers can obtain densities above Standard Proctor maximum in layers up to 18 inches thick with four to six passes. Because of the small loaded area and high unit pressure, the sheepsfoot roller is adapted best to cohesive soils such as clays. Tailings embankments with more than 20 percent minus 200-mesh material in the dike could be compacted with the sheepsfoot or in the modified sheepsfoot, which has some of the feet removed and flat plates 8 to 10 inches in diameter welded on the remaining feet. This modified version is best for silty soils of low cohesion, which better describes some of the tailings sand. Moisture control is very important to efficiency of compaction of this type of material. For cohesionless borrow and waste materials, such as rockfills and gravels, compaction by track-type tractors and haulage trucks might be adequate if properly controlled. Where additional compaction is required, heavy vibratory steel wheel compactors are very efficient and are not moisture dependent. Compaction is very critical on starter dams whether pervious or impervious and on tailings embankments and dams built of tailings sand to raise density, increase stability, and prevent liquefaction.
Colorado School of Mines
ER-4609 160 A particular compactor has its own optimum water content which may not be the same as that determined in the Standard or Modified Proctor test in the laboratory. The laboratory tests will be within a few percentage points of those applicable to most compactors, and the heavier the compactor, the lower will be the optimum moisture content. Field density tests will help determine the efficiency of a particular method. Use of compaction layers which are too thick for a particular compactor will result in undercompaction at the base of the layers, with increased horizontal permeability of that zone. No one method of compaction can be established as best for the sand zones of tailings embankments because of the wide variations in the gradation of the sand on the beach from one property to another. Each mine must determine the most efficient compactor for the materials it is using. 6.12 Seismic Considerations (after Klohn, 1980) Liquefaction of a soil is a temporary state in which the structure of the soil is disturbed, causing the particles to lose contact, and to transfer the weight of the soil and any superimposed loads onto the water in the voids. This causes the pore pressure to rise and the soil behaves like a dense fluid. It is unable to support loads or to resist significant shear forces, and the resulting pore pressures approach or equal the confining pressures. Liquefaction can be caused by changes in static stress conditions, particularly in very loose soils. However, the more general cause of structural disturbance leading to liquefaction is dynamic loading such as can occur during an
Colorado School of Mines
ER-4609 161 earthquake. A form of liquefaction, in which air acts as a fluid medium, can occur in extremely loose aeolian soils such as loess. However, for soils of the types and densities used in constructing tailings dams, liquefaction is assumed to occur only in saturated zones, i.e. below the phreatic surface. The ease with which liquefaction can be initiated varies. Under the least favorable conditions, a soil can be liquefied by the first pulse of an earthquake, whereas under other circumstances it may not fail even after hundreds of cycles. The factors affecting liquefaction potential are summarized following : Saturation More-or-less complete saturation is essential for liquefaction in sand of the type and density under consideration. Partially saturated sand, such as may exist where water is percolating through the fill above the phreatic surface, will not liquefy because the contained air is highly compressible, thus preventing a significant rise in porewater pressure. Saturation by capillarity is not significant in cycloned tailings sand. Permeability For liquefaction to occur, the soil must be sufficiently low in permeability that no significant migration of water and consequent pressure relief is possible within the time frame of the earthquake. Most cycloned tailings sands
Colorado School of Mines
ER-4609 163 of cycles that may cause liquefaction under a given cyclic shear. It may also affect strain movements once failure has occurred. Initial Shear Stress If the initial shear stress has been acting long enough to allow full drainage, the resistance to liquefaction is increased. However, if drainage has not had time to occur under the initial shear stress, the resistance to liquefaction is decreased. Other Factors There are several other factors which appear to affect the liquefaction potential of a fine sand under earthquake loading. These include such things as: age of deposit, method of deposition, exposure to previous earthquakes, particle shape, overconsolidation, etc. The liquefaction problem is normally assessed by running dynamic shear tests on laboratory samples of the tailings sand. These may be so-called "undisturbed" samples or reconstituted samples. The most commonly used method is the cyclic triaxial test which uses conventional triaxial equipment adapted to allow cycling of the deviator stress. The output includes continuous records of deviator stress, axial strain and porewater pressure. Using the data obtained from the dynamic laboratory tests a dynamic analysis of the tailings dam may be carried out. Such an analysis is involved and complex. Briefly, a seismic analysis involves the following steps :
Colorado School of Mines
ER-4609 165 saturated. In tailings dam designs for areas rated as low seismic risks it is usually more economical to strive for complete drainage rather than for high densities. For areas of high seismicity both high densities and good drainage are considered desirable. 6.13 Rate of Seepage (after Soderberg and Bush, 1977) The rate of seepage through a tailings embankment is governed by the permeability of the materials, the hydraulic gradient, and the elevation difference between the pond and the point of emergence of the seepage. With a coarse grind, low pulp density, and a wide beach the seepage vertically through the beach is tremendous and could equal or exceed that which escapes through the embankment. With a given screen analysis, mineralogy, tonnage, pulp density, area, and construction method (upstream or downstream), there is little that can be done to change the seepage once the operation has started. The length of the flow path cannot be increased after the pond gets to the extreme upstream position, or to the decant, except by adding material to the toe. The difference in elevation has to increase, and no impervious barriers can be placed. Therefore, it is imperative that seepage be anticipated and incorporated into the design. The mathematical expression for the rate of seepage is
Colorado School of Mines
ER-4609 166 nf q = k — h, nd where q = the rate of seepage per unit length perpendicular to the plane of the flow net, k = the coefficient of permeability of the soil. nf = the number of flow paths (determined from the flow net) , nd = the number of equipotential drops (determined from the flow net), and h = the difference in piezometric head between the point of seepage entry and the point of seepage exit. An approximate coefficient of permeability can be determined by laboratory tests if field tests are not possible. 6.14 Pond inflow and control methods (after Portfors, 1980) Water inflow to the pond is made up of tailings slurry transport water, precipitation directly onto the impoundment, runoff from the pond catchment area, seepage recovery, ground water seepage, and other miscellaneous sources, such as pit dewatering, surface diversions, and domestic wastes. If a deficit exists in the system, a supplemental water inflow source may be required. In most tailings ponds, the water required to transport the tailings in a slurry form is the single largest input water source. In a closed circuit operation, reclaim water is withdrawn at a rate proportional to the slurry transport
Colorado School of Mines
ER-4609 167 water input, consequently, the rate of transport water input usually is not critical. Precipitation onto the pond becomes completely available as input water to the pond. On the other hand, runoff from the impoundment catchment and exposed beaches will not equal the total precipitation, but will reflect the basin losses to evapo-transpirâtion, groundwater recharge, diversions, other withdrawals for use, etc. Hydrological studies are required to estimate catchment basin yields in average, wet and dry years. Average annual precipitation and runoff values would be used to estimate available water to balance other losses, and to estimate the average quantity and cost of any make-up water that may be required. Average conditions determine the overall pond management strategy over the life of the mine. Wet year inflow conditions are used in combination with flood discharges to calculate the required dam freeboard or spill capacity. Dry year flows, on the other hand, can be critical to plant operation. If the precipitation and runoff are required as water supply sources, the probability of a dry year, or a series of dry years, must be established to define water deficiencies. Timing of inflows throughout the year will be important as dam construction must be advanced sufficiently to store inflow during the high runoff season. The freewater volume is then depleted during the dry seasons of the year. Available precipitation records are used to establish amounts and timing of precipitation, and stream flow records to estimate runoff. Mines are often located in remote regions where hydrologie data are scarce or entirely absent. The designer then must use information from neighboring areas to estimate flows at the site.
Colorado School of Mines
ER-4609 168 Catchments above tailings impoundments are generally small, and caution must be used when information from large basins is applied on small ungauged basins. For example, in mountainous regions, weather effects on local precipitation can be larger than the regional average. In the case of large upstream catchments, diversion channels probably will be necessary to carry upstream runoff around the pond, thereby limiting the input of surface water. Design discharges adopted for these channels can vary depending upon the consequences of overtopping. Commonly adopted design floods range between the 100 year and 1000 year recurrence interval events. With such a design, only in rare cases will the channels be overtopped, and flood waters enter the pond. Offtakes can also be added to the diversion channels enabling extra water to be added to the pond during dry years. Mine drainage water may be contaminated and consequently cannot be released to the downstream. The tailings pond then becomes a convenient storage location, and the drainage water becomes another input water source. Surface drainage water from around the pit and mill area also may carry a high concentration of suspended sediments and other pollutants. This water can be directed into the tailings pond rather than into separate settling ponds. Controls (after Vick, 1983). The preferred situation is to limit water inflow by proper siting of the impoundment. There are several methods available for handling remaining flood inflows. A primary method of inflow control is storage. Adequate dam freeboard or surcharge is maintained at all times by raising the embankment at a rate such that
Colorado School of Mines
ER-4609 169 sufficient volume is always available for storage of the design inflow. If the design flood inflow has been determined with an appropriate degree of conservatism, it is unlikely to be experienced during the life of the impoundment. In the improbable event that it should occur, the impounded runoff could be stored for eventual evaporation in regions of arid climate. In other areas, the water, if contaminated by mixing with the mill effluent, could be treated and released at a gradual rate. Control of flood inflow by storage avoids the necessity of expensive and sometimes difficult treatment of contaminated flood inflow prior to release into surface watercourses. In some areas, topographic constraints on practical embankment height and impoundment volume, combined with high precipitation or high rates of net mill water discharge, may make storage of flood inflows unfeasible. In such cases, the only option may be to provide treatment of all mill effluent prior to its discharge into the impoundment. Flood inflows entering the impoundment, presumably no longer at risk of contamination by mixing with the treated mill effluent, can be passed through the impoundment and discharged through a spillway designed according to conventional methods. In applicable areas, the thunderstorm rather than general storm PMP may control the design of spillways where peak flow rate rather than total in flow quantity is of primary importance. The use of spillways with raised embankments, however, is awkward at best. With each embankment raise, a new spillway must be constructed at the new crest elevation. This adds appreciably to cost and difficulty of construction. In extreme cases, water handling by spillways may preclude the use of raised embankments, instead requiring water-retention type
Colorado School of Mines
ER-4609 170 structures built initially to full height. Diversion channels are desirable in most cases and essential in some for diversion of normal runoff flows. Although diversion channels can also be used to divert flood flows around the impoundment, major design floods often produce flows that require large channels (channel widths for PM F diversion in excess of 100ft are not uncommon) and heavy riprap for protection of the channel banks against high flow velocities. As a result, diversion channels designed for peak flood flow are sometimes impractical, depending on the channel length required. For raised embankments, however, excavation of even relatively large flood diversion channels may provide a convenient source of starter dike fill material. For open-pit mining operations, creative planning of waste rock dumps and even the pits themselves can provide useful water-control benefits for the tailings impoundment. Process flow considerations in mine facility layout usually dictate that the tailings impoundment and mill be located at lower elevations than the pit and associated waste rock dumps. If the pit is located within the drainage basin of the tailings impoundment, the pit volume itself can be credited with flood water storage for extreme PMF-type floods. If the waste rock dumps can be extended across the tailings impoundment drainage area without excessive haul distance, diversion of extreme floods by the mass of the waste rock can be realized at essentially no cost. This requires that waste dump and tailings impoundment planning be integrated and carried out together. A related method of runoff diversion is to construct non-impounding diversion dikes across the impoundment drainage area. To be effective, such dikes should be
Colorado School of Mines
ER-4609 171 located as close as possible to the maximum upstream extent of the impoundment. This method of surface-water control may be of particular advantage if the impoundment is located over relatively shallow bedrock where construction of diversion channels would require expensive rock excavation. Flow velocities against the diversion dike may be high, however, possibly requiring the use of riprap if the dikes are constructed of native soils. The use of open-pit waste rock for diversion dike construction, usually relatively coarse and erosion resistant, may be attractive. In extreme cases where the tailings impoundment is located in a narrow, constricted valley with a large upstream drainage area, steep valley sidewalls may make it impractical to divert flood runoff around the tailings impoundment by either diversion channels or nonimpounding dikes. Here it may be necessary to construct an entirely separate flood-control dam upstream from the tailings impoundment. The flood-control dam can provide full storage of expected flood runoff from the upstream drainage area, with gradual evacuation of the stored water by a culvert passing through the water-control dam and around the tailings deposit. This arrangement is to be avoided if possible, since fill requirements for the water-control dam may be large, in some cases even exceeding fill required for construction of the tailings embankment itself. Moreover, construction of the flood-control dam cannot be staged; it must be completed prior to tailings impoundment operation in order to serve its intended flood-protection function. Finally, maintenance of buried culverts is problematic, and the limited life of the culvert structure may make it necessary to provide more permanent water-control measures after impoundment abandonment and reclamation.
Colorado School of Mines
ER-4609 172 Figure 52 summarizes the various methods for handling extreme flood inflows from external drainage areas. It is important to note that even with complete diversion of flood runoff from tributary drainage areas, that portion of the extreme flood precipitation which falls directly on the impoundment surface must still be accounted for. 6.15 Effects of floods on impoundments (after Vick,1983) A completely separate consideration is control of floods that may pass at the toe of the embankment. This inflow into the impoundment may result in erosion, undercutting, and eventual failure of the exterior face of the embankment, and it is especially critical for impoundments located in low-lying river floodplain areas or narrow canyons. The same general flood criteria as for impoundment inflow apply, only here the concern is with flow velocities of the water at the embankment toe. Riprap of affected portions of the embankment face provides one solution. The method of choice is to avoid impoundment siting in floodplains, particularly in areas of predominantly soft, sedimentary rocks where riprap may be unavailable locally and difficult or expensive to import. For some Southeastern U.S. phosphate tailings impoundments located in low-lying coastal areas, imported riprap for wave protection during hurricanes constitutes the major expense in impoundment construction.
Colorado School of Mines
ER-4609 174 6.16 Embankment freeboard and wave protection (after Soderberg and Bush, 1977) Tailings embankments built of borrow material need protection from wave action if the water is allowed to come against the borrow dike. Wave action is a very destructive force and can erode and overtop an embankment unless ample freeboard is provided. The height of the waves depends on the wind velocity, the duration of the wind, the fetch (the distance the wind can act on the water) , and the depth of water. On steep upstream slopes, riprap will limit the uprush of the waves to approximately 1.5 times the height of the waves and will prevent erosion by wave action. Tailings embankments should not have free water standing against the dike except where a water-type dam is impounding the tailings, and then they should have riprap on the upstream face. Table 18 gives the approximate wave heights for various wind velocities and fetch, and the freeboard and riprap gradation for the 3:1 slope. For 2:1 slopes, the thickness should be 6 inches greater. A layer of filter gravel should be placed between the riprap and the embankment. Wave action is not a problem if even a small beach is provided between the dike and the water, because the waves dissipate harmlessly in the shallow water on the beach. 6.17 Crest Width (after Soderberg and Bush, 1977) The crest width should be no less than 12 feet for easy equipment operation in a situation where water is against the embankment. The most suitable crest width will depend on the allowable percolation distance through the embankment
Colorado School of Mines
ER-4609 177 CHAPTER 7 SEEPAGE AND SEEPAGE CONTROL MEASURES The information presented here on seepage and seepage control has been extracted from several sources. Section 7.1, 7.4, 7.5, were extracted from Control of Seepage in Tailings Dams, written by Patrick M. Griffin. Section 7.2 and part of section 7.3 were compiled from Earl J. Klohn1 s article Seepage Control for Tailings Dams. Sections 7.3 and 7.8.4 were extracted from the Bureau of Mines information circular 8755, Design Guide for Metal and Nonmetal Tailings Disposal, written by Roy L. Soderburg and Richard A Busch. Sections 7.6, 7.8.1, 7.8.2, 7.8.3, 7.8.5, 7.8.6, and 7.8.7 were compiled from Seepage Control Measures in Tailings Dams, written by Robin Fell. Sections 7.7 and 7.8.8 were compiled from Steven G. Vick's book titled Planning, Design, and Analysis of Tailings Dams. The information in section 7.9 through the end of the chapter was extracted from Mine Waste Management, edited by Ian Hutchison and Richard D. Ellison. 7.1 Introduction to Seepage (after Griffin, 1990) When considering seepage, the designer of a tailings dam is faced with three major concerns: Resistance to piping and internal erosion. Location of the phreatic surface and its effect on
Colorado School of Mines
ER-4609 179 water in the tailings pond can be estimated using approximate flow nets based on the anticipated width of beach and the permeabilities of the materials involved. Estimating the quantity of seepage due to spigotting is much more difficult, however, a reasonable approximation can be made by assuming that the spigotting completely saturates the beach so that in effect the free water in the pond extends to near the top of the spigotted beach. A flow net drawn for this condition should provide an estimate of the combined seepage, due to both the free water in the pond and the effects of spigotting. Estimating the quantity of the construction water that the drains must handle varies from a simple exercise for the case of on-dam cycloning, where all the construction water seeps into the downstream sand dam, to a complex exercise for the case where large volumes of hydraulic fill transport water flow across the dam and then exit either upstream into the tailings pond or downstream behind the seepage recovery dam. For the on-dam cycloning, all the water contained in the cyclone underflow is assumed to reach the underlying drain. This value can be estimated with reasonable accuracy from the density of the underflow and the recorded tonnage of sand placed each day. For the case of hydraulic fill placement, where large volumes of water flow across the sand dam, the seepage loss into the dam may be estimated by assuming downward seepage under a hydraulic gradient of 1 and using the expression : q = k i A where q = rate of vertical seepage A = total area over which the water is flowing
Colorado School of Mines
ER-4609 180 i =hydraulic gradient (1 for this case) k =effective coefficient of permeability = Vk x K h v The fourth potential source of seepage, consolidation water squeezed out of the slimes as they consolidate is more difficult to quantify. In effect, this action adds an additional increment of pore pressure to the normal hydrostatic pore pressure that would exist in the slimes if they were completely consolidated under their own weight. In high rainfall climates precipitation can be a major contribution and may cause the sand dam to be almost fully saturated during long periods of heavy rainfall. Of the above five outlined sources of seepage water, construction water from cyclone underflow or hydraulic fill placement operations is usually several times greater than all other sources combined. In designing the drains, the highest probable seepage flows that can enter the drains should be used and the drains should be assigned their lowest probable permeabilities and gradients. This is essential, as once constructed and buried, the drains must continue to perform satisfactorily throughout the life of the structure and, in many cases, for many years after the mining operation is completed. All drains should be sized to handle flows several times the largest probable flows computed using the above outlined methods. 7.3 Use of flow nets (after Soderberg and Bush, 1977) The uppermost line of seepage that is at atmospheric pressure is known as the phreatic surface and is the
Colorado School of Mines
ER-4609 181 uppermost flow line. An equipotential line is a line of equal head; therefore, water rises to the same level in piezometers installed along a given equipotential line. The equipotential lines in a flow net must intersect the free water surface at equal vertical intervals. Flow lines and equipotential lines must intersect at right angles to form areas that are basically squares when the materials are isotropic. Adjacent equipotentials have equal head losses. The same quantity of seepage flows between adjacent pairs of flow lines. Usually it is best to start with an integral number of equal potential drops by dividing the total head by a whole number and drawing flow lines to conform to these equal potentials. The outer flow path will generally form a distorted square figure, but the shape of these distorted squares (the ratio B/L) must be a constant (Figure 53). In a stratified soil profile where the ratio of horizontal to vertical permeability exceeds the ( K h / K v ) 1 0 , flow in the more permeable layer controls. The flow net may be drawn for the more permeable layer, assuming the less permeable layer is impervious. The head on the interface from this permeable layer is imposed on the less pervious layer for the construction of the flow net within that layer. This situation can occur when a starter dam is impervious to the tailings it is retaining and has insufficient, ineffective, or no drains, so that water at the interface builds up to and goes over the top of the starter dam. The flow net through the starter dam then would be the same as if there was free water against it, as illustrated in Figure 54. In a stratified layer where the ratio of permeability of the layers is less than 10, the flow net is deflected at
Colorado School of Mines
ER-4609 184 the interface in accordance with Figure 54. When materials are anisotropic with respect to permeability, the cross section should be transformed by changing the scale. The horizontal dimension of the section is reduced by the square root of The flow net is then K v K h . drawn as for isotropic materials and can be transposed to a true section. If Kh > Kv the L dimension becomes elongated, and the net is no longer a square. In computing the quantity of seepage, the differential head is not altered for the transformation. Where only the quantity of seepage is to be determined, an approximate flow net suffices. Where pore pressures are to be determined, the flow net must be accurate. Earl J. Klohn, in his article Seepage Control for Tailings Dams states that seepage flows are predicted from the flow net using the following relationship : q = k . h . nf per unit of length nd Where : q = rate of seepage flow k = coefficient of permeability of the soil h = the hydraulic head acting across the structure nf = number of flow paths in the flow net nd = number of equipotential drops in the flow net
Colorado School of Mines
ER-4609 185 7.4 Piping protection (after Griffin, 1990) The major issues which must be addressed to prevent piping within the tailings embankment are control of hydraulic flow gradients, care in the use of erodible materials, and effective use of filters to separate materials with large differences in gradation. Where a beach is used, drainage paths will be lengthened, which will have a positive effect in reducing the seepage flow gradient through the upstream portion of the embankment. Without a beach, it is necessary for designers to provide appropriate control of the gradient in the upstream dam section. If sand tailings are used as an integral part of the dam, thed esigners must take into account the potential erodibilities of these materials. Generally, non-plastic sand and slime tailings can be expected to be extremely erodible. Materials containing some clay minerals, such as phosphate tailings, could have some inherent resistance to erosion and dispersion. Filter systems are used in tailings dams in much the same manner as for conventional embankment structures. Often, the sand fraction of the tailings can be used as a filter or a component of a filter system. Care must be taken, however, that the sand material has sufficient durability and will be chemically stable when exposed to the effluent seepage. Physiochemical decomposition of the tailings sand could create problems over time if not properly accounted for in the design. Durability is commonly evaluated by performing the same types of tests normally required for concrete sand and aggregate (i.e.-pétrographie analysis, specific gravity and absorption, sodium/magnesium soundness, abrasion, freeze-
Colorado School of Mines
ER-4609 186 thaw, etc.), but the acceptance limits are usually more flexible than that required in ASTM Method C-33. For example, chert is deleterious to concrete, but may be perfectly adequate in a filter or drain. Similarly, requirements for resistance to mechanical and chemical breakdown need not always be as stringent as for concrete aggregate. 7.5 Location of the phreatic surface (after Griffin, 1990) The stability of the downstream slope of the tailings dam will be heavily influenced by the depth of the phreatic surface through the embankment. This is not as critical an issue for the upstream slope, because of the presence of solids within the pond. On the downstream side of the embankment, the strength of cohesionless materials is influenced by both the buoyancy effect of water below the phreatic surface, and also the effect of seepage pressures. The technical issues are essentially similar to those for conventional embankment dams. However, with the combination of features such as beach zones and extended drainage paths, it is often possible to reduce the downstream phreatic level to a very low surface within the embankment, and therefore steepen the downstream slope while still maintaining acceptable slope stability. This issue can often have major economic consequences on the overall cost of a tailings dam. For borrow dams, the most desirable materials for the embankment are free draining hard durable rock or gravelly materials. However, less desirable materials, such as mixed soils, and decomposed shales and sedimentary rocks, might be more economical or more readily available locally. In these
Colorado School of Mines
ER-4609 187 cases, it may be desirable to introduce a chimney and/or blanket drain arrangement to help lower the phreatic surface. The added cost of providing the zoning detail in the embankment, including provision for staged enlargement, might well be justified if the provision allows for the use of inexpensive materials and for steepening of the downstream slope. Another approach for dealing with the economies of steepening a downstream slope would include full utilization of the observational approach. The observational approach is well established and well accepted within the geotechnical engineering profession, and has definite applicability to tailings dams. In a specific project, a system can be designed with a reasonably conservative estimation of the phreatic level. Then as the dam is enlarged in stages, instrumentation can be used to determine the actual performance of the structure, including the actual phreatic levels, stresses and displacements. If it can be shown from instrumentation data that the original design assumptions were overly conservative, then there will be ample justification for modifying the design to a more economical, but still reasonably conservative, scheme. For dams constructed with tailings material, it becomes more difficult to introduce zoned elements, such as internal filters and drains. However, it is possible to introduce a blanket drain and appropriate filter along the foundation of the dam which extends to thé eventual downstream embankment. Here again, tailings sands could serve as a filter or a significant component of a filter system. Combined with the observational approach, this method could allow for refinement of the steepness of the downstream slope over time.
Colorado School of Mines
ER-4609 188 Blanket drains and similar underdrain systems have a significant additional advantage where there is a necessity to collect hazardous tailings effluent leachate. It is possible to design an underdrain system which also incorporates groundwater cutoffs and doubles as a wastewater containment and collection system. Another advantage to lowering the phreatic level through the downstream embankment section of the dam is the beneficial effect on long-term durability of the embankment materials. Physiochemical interaction is minimized, and material degradation is significantly retarded when materials do not have tailings fluids flowing through them in a saturated condition. It is often possible to justify use of very low cost "random fill zone" materials in a downstream dam section when the material will be completely protected from saturation by a chimney and underdrain system. 7.6 Contaminants (after Fell, 1990) Many mine and industrial tailings have accompanying water or "liquor" which contains dissolved salts, heavy metals and other residual chemicals from the mineralogical processes. If these escape to the surrounding surface and groundwater in sufficient quantities, they can lead to unacceptable concentrations, making the water unusable for drinking water and affecting aquatic life. Some common problems are : Gold tailings — cyanide Coal — high salts content Copper, lead, zinc — heavy metals, sulphides
Colorado School of Mines
ER-4609 190 contaminants in the seepage water willnot all join the groundwater. Much will be absorbed in the foundation soil and rock. Contaminant load does not equal flow rate x contaminant concentration in the storage having joined the groundwater saturated zone, dispersion takes place, reduction of contaminant concentration may occur in mixing with stream flows it is contaminant concentration in ground and surface water which is generally critical, not the total quantity. Hence, absorption, dispersion and dilution can yield acceptable water quality in streams or well points, even though the original contaminant levels may have been unacceptable. It is important to realize that in many cases, a partially saturated flow condition will exist in the foundation, at least at the start of operations and possibly on a permanent basis if the tailings permeability is low compared to the foundation permeability. Figure 56 shows the stages in the development of seepage. It may take years for the seepage mound to rise to connect to the tailings (stage 3) or it may never happen. Note that flow in the tailings in stages 1 and 2 will be vertical, will not emerge at the toe of the embankment, and will be virtually unaffected by any foundation treatment such as grouting. In stages 2 to 4, flow will be in all directions away from the storage.
Colorado School of Mines
ER-4609 192 In many other cases, the tailings will be stored in a valley in which a dam has been constructed. It is usual to site such dams at the head of a catchment so as to limit the water runoff from the area outside the storage. Catch drains and other measures may be used to reduce this inflow. Figure 57 shows an example of a tailings storage. In this case the final development will consist of several embankments. When estimating seepage from such a storage it is important to remember that seepage will occur under each of the embankments, and depending on the base groundwater levels, into the hillsides adjacent the embankments. Hence in Figure 57 seepage will occur to the west, north and east, but not to the south where natural groundwater levels are higher from the storage. Note that the groundwater does not always mirror the topography and may be affected by local variations in geology — e.g. permeable dikes. Many inexperienced engineers and geologists will either forget completely that seepage will occur in all directions from the storage, or at least apply an excessive amount of the site investigation effort and analysis to the seepage which will flow through and beneath the main embankment. From the examples shown in Figures 55, 56 and 57, it will be apparent that the assessment of seepage flow rates will involve: knowledge of the permeability of the tailings, as these are commonly part of the seepage path. In many cases they may control the seepage rates; In Figure 57 it would be necessary to be able to model the whole of the area between the streams, necessitating knowledge of rock permeabilities well beyond the storage area
Colorado School of Mines
ER-4609 194 mmodeling of the seepage, usually by finite element methods, which may involve several sectional models and/or a plan model. This modeling should account for the development of flow as shown in Figure 55, not just model an assumed steady state coupled flow situation, i.e. the storage and groundwater coupled as in stage 3, Figure 56. A common error is to do this when in fact it does not apply, and worsen the accuracy further by assuming seepage emerges at the toe of the dam and at ground surface, when in fact flows downstream of the embankment are inadequate to raise the phreatic surface to ground level. 7.7 Objectives of seepage control (after Vick, 1983) Some general principles related to seepage control: Not all mill effluents contain toxic constituents. Depending on ore type, mill process, and pH, contaminants may range from toxic heavy metals (that is, cadmium, selenium, arsenic) to such relatively innocuous materials as sulfates or suspended solids. For mill effluent that does contain toxic constituents, it is not necessarily the case that seepage of this effluent will result in pervasive groundwater contamination. Geochemical processes may retard or inhibit movement of some constituents, and these processes are often most
Colorado School of Mines
ER-4609 195 effective in reducing mobility of the most troublesome metallic ions associated with low-pH effluents. If toxic constituents do enter the groundwater regime, the ultimate effects on the groundwater environment must be determined before deciding on a seepage-control strategy intended to minimize these effects. Essential to this determination are hydrogeologic factors, baseline water quality, and intended use of the groundwater resource both present and future. From these principles, it seems reasonable that the type of seepage-control strategy should match the chemical conditions of the effluent and the geochemical and hydrogeologic conditions of the site. Three types of systems used with regard to uranium tailings seepage control can be defined and extended to tailings in general : Type I System. Seepage from the impoundment is essentially uncontrolled. Due either to a lack of troublesome contaminants in the mill effluent or to uptake of these contaminants by geochemical processes, groundwater contamination potential is not serious, irrespective of seepage quantity. Type II System. In this case, impoundment effluent is partially retained, but some seepage loss is anticipated. Contamination potential is greater than for Type I effluents, and a higher level of seepage-contaminâtion analysis is
Colorado School of Mines
ER-4609 198 Liners 7.8.1 Controlled placement of tailings (after Fell, 1990) In many cases the most cost effective way of controlling seepage will be to place the tailings so they blanket the base of the storage. Figures 58 and 5 9 show this effect. The effectiveness of the tailings as a "liner" is dependent on placement methods. If tailings are placed sub- aerially and allowed to desiccate, lower permeabilities will result from the drying. If placed subaqueously lower densities and higher permeabilities are likely to result. A difficulty with this approach is that the coarser fraction of the tailings tend to settle out more quickly than the fine (or slimes) fraction. Hence a "beach” of sandy tailings often occurs near the discharge point. If water is allowed to cover this area subsequently, it can allow high local seepage rates. This can be overcome by shifting the tailings discharge points from one end of the storage to the other, placing slimes under the beach area, and/or by using a liner or seepage collector system under the sandy area. Seepage can also occur down the contact between tailings and embankment if rock riprap is used. The actual permeability of the deposited tailings can be determined by laboratory permeability or consolidation tests, preferably on undisturbed samples from the tailings storage (taking account of variations within the deposited tailings, particularly in regard to distance from the discharge point and the degree of consolidation of the
Colorado School of Mines
ER-4609 201 tailings. Almost invariably, the horizontal permeability will be one or two orders of magnitude higher than the vertical permeability due to a stratification on cycling of deposition (particularly near the discharge point). This often leads to thin sand or silt partings on the top or bottom of each layer (depending if the coarse particles are low or high specific gravity). 7.8.2 Foundation Grouting (after Fell, 1990) It is common to grout the foundation of most large water storage dams, at least those which are founded on rock. The grouting is usually carried out by pumping a slurry of cement into holes drilled into the foundation. Chemical grouts may be used to grout soil (e.g. sand and gravel) or lower permeability rock. In such dams the grouting is carried out to reduce leakage through the dam foundation reduce seepage erosion potential reduce uplift pressures under concrete dams (when used with relief drains) strengthen the dam foundation and reduce settlement (for concrete gravity and arch dams. Grouting has a limited effect on reducing seepage, particularly in rock which is closely fractured. The critical issues are: Portland cement grout is a suspension of silt and sand size particles in water. It cannot penetrate
Colorado School of Mines
ER-4609 202 fractures finer than about 0.15-0.2mm wide. Table 20 shows estimated minimum groutable rock lugeon values. One Lugeon ia a flow of 1 litre/minute/meter of borehole under a pressure of 1000 KPa. The grout penetrates a limited distance into the fractures in the rock. This is dependent on the fracture opening and roughness, grout pressure and time, and grout viscosity. Table 21 shows approximate penetration of grout into fractures based on laboratory tests and comparison with field behavior. Because the grout only penetrates a limited distance, and does not render the rock "impervious", but only reduces the permeability, the seepage is only significantly reduced if the rock is high permeability. A simplified permeability analysis gives the figures in Table . 22 Since it is not uncommon to adopt a closure criteria (i.e. the lugeon value at which no further grouting is carried out) for grouting of say 3 to 5 lugeons, it is apparent that unless the rock is very high in permeability there will be little reduction in seepage yielding little overall reduction in seepage except in high permeability zones in the foundation. Chemical grouts may achieve a lower permeability in the grouted zone than cement grout, but the penetration distances are small.
Colorado School of Mines
ER-4609 204 TABLE 22 Effectiveness of grouting in reducing seepage, (after Fell, 1990) Permeability Grout Seepage with Grouting x 100 % Permeability Rock Seepage without Grouting 0.5 90 0.2 70 0.1 50 0.05 35 0.02 20 0.01 10 On one project, the grouting of a 5km long dam to a depth of about 25m, on average would have reduced the estimated seepage by 1%, nearly all of this in the small portion of the foundation affected by faulting, and having a permeability of the order of 100 lugeons. It will be seen from the above discussion that it is unlikely that grouting of tailings dam foundations can be justified on the grounds of reducing seepage. It may be justified on other grounds, such as reducing potentialerosion in weathered rock, or where the high permeability zones can be identified from geological information, allowing grouting of only those parts of the foundation. 7.8.3 Foundation cutoffs (after Fell, 1990) For tailings storages constructed on soil foundations, particularly sand or sand and gravel, a significant reduction in seepage may be achieved by construction of an earthfill cutoff or a slurry trench cutoff wall as shown in Figures 62 (a) and (b) . There are several types of slurry trench wall. Table 23 summarizes these alternatives, and their
Colorado School of Mines
ER-4609 206 potential application. When a relatively impervious layer is at a shallow depth below the foundation, a core trench can be used to limit seepage. A core trench is illustrated in Figure 61. 7.8.4 Hydraulic barriers (after Soderberg and Bush, 1977) Where the tailings dam is constructed on a thick pervious foundation, and pollution control requirements preclude the escape of water from the tailings pond, seepage losses may be controlled by developing a hydraulic barrier downstream of the tailings dam. The hydraulic barrier (illustrated in Figure 62) can be produced by a line of pumping wells and a line of injection wells downstream of the embankment, the injection wells being located downstream of the pumping wells. Fresh water is supplied to the injection wells, while ground water is extracted from the pumping wells. If the piezometric water levels along the line of the injection wells are maintained at elevations higher than the piezometric water levels along the line of the pumping wells, a hydraulic barrier will be formed that will prevent the flow of seepage from the tailings pond past the line of pumping wells. This should be checked in the field by piezometric measurements. The hydraulic barrier may be usable for up to 100 feet of overburden, but would not be feasible where the depth was much greater. This method does not eliminate contamination of the ground water in the long run, but only as long as the injection and pumping wells are operating.
Colorado School of Mines
ER-4609 210 7.8.6 Toe drains (after Fell, 1990) A drain may be provided at the downstream toe of the embankment with a view to collecting seepage which emerges at that location. These drains can be reasonably successful in intercepting seepage, but only if the seepage naturally emerges in this location. In many cases the flow rates will be such that the phreatic surface stays below the level of the drain. Even when the seepage is sufficient to raise the phreatic surface to flow to the drain, much may still bypass by flowing beneath the drain. Ideally the drain has to penetrate to a low permeable stratus, and this is often not practicable. Such drains may also intercept surface runoff from the downstream face of the dam and groundwater from downstream, and if the seepage is to be returned to the dam or process plant, may exacerbate water management problems if a no release system is being operated. 7.8.7 Seepage collection dams (after Fell, 1990) In many cases, a practical way of collecting seepage from tailings storages will be to construct a seepage collector dam or dams. Figure 57 shows such a system. The seepage collector dams may be located close to the storage, sufficiently close to collect (the bulk of) the seepage but not too far away so as to limit the external catchment— e.g. Dam A on Figure 57. In this case one would be anticipating pumping the water back into the dam or process system.
Colorado School of Mines
ER-4609 211 Alternatively one may deliberately locate the collector dam sufficiently far downstream that the runoff from the catchment to the dam is sufficient to dilute the seepage to acceptable water quality— e.g. Dam B in Figure 57. This has the advantage that the seepage which occurs after shutdown of the tailings storage is also diluted to acceptable quality. Whether such an approach is acceptable will be dependent on the particular circumstances for the tailings storage. For example, it may be unacceptable to have substandard water quality in the stream between Dam B and the tailings storage. Seasonal effects can also be important e.g. If there is a prolonged dry season, water may pond in the stream and concentration of contaminants may occur. In many cases the catch dam with pump-back or dilution may be far more appropriate than expensive measures to control the seepage. Many engineers have an over-optimistic view of the efficacy of these seepage control measures, or an unrealistic view of what costs a mining operation can reasonably bear to construct such measures. 7.8.8 Liners (after Vick, 1983) Liners constitute the final category of seepage-control measure and are usually reserved for conditions where a Type III system is called for because of stringent groundwater protection requirements and relatively high concentrations of toxic constituents in the mill effluent. Liners of any type are inherently high in cost, but their effectiveness is at least somewhat less in doubt than high-cost barrier
Colorado School of Mines
ER-4609 212 systems, principally because a liner is a surface installation that can be constructed under controlled conditions and inspected. This does not guarantee, however, that even a properly constructed liner will function as intended during actual operation or that any lined impoundment will be a "zero discharge" facility. Liners do have a major advantage over seepage barriers or seepage return systems insofar as they are completely independent of subsurface conditions. Whereas the effectiveness and construction feasibility of grout curtains, slurry trenches, and collector ditches or wells depend on the presence of a lower impervious layer as well as the nature of the material to be penetrated, liners suffer no such limitations and can be constructed on any surface sufficiently dry and competent to allow for normal earthwork operation, without regard for the nature of subsurface soil, rock, or groundwater conditions. Liners, however, must be resistant to both chemical attack by the retained effluent and a variety of types of physical disruption. Because of the critical nature of seepage problems in those cases where liners are required, there is considerable debate over the relative merits and effectiveness of the various kinds of liners and liner materials. 7.9 Liner systems and properties (after Hutchison, 1992) Liners and liner system technologies for application to tailings dams have made great strides in recent years, with environmental legislation and restrictions continuing to increase, liner systems installed in closed circuit tailings dams will become more common. This section will identify
Colorado School of Mines
ER-4609 215 TABLE 24 Available materials or procedures for liner system components. (after Hutchison, 1992) A. LOW HYDRAULIC CONDUCTIVITY LINERS A.l Low Hydraulic Conductivity Natural Soil or Rock - Natural soils or rock may be used as a low hydraulic conductivity liner so long as it is possible to demonstrate by field investigations that the material is uniform over the entire area requiring the liner. This demonstration may be particularly difficult for rock that is extensively fractured. A.2 Constructed Low Hydraulic Conductivity Liners - Low hydraulic conductivity liners which are constructed beneath a mine waste management unit may consist of any of the following materials: • Compacted, low hydraulic conductivity soils (e g., clayey-silt to clay depending upon required hydraulic conductivity ). • Soil and bentonite or cement mixtures. • Pre-formed flexible membrane liners, made from a variety of available polymer material, generally called geomembranes; varying in thickness from about 20 to 100 mils. • Field-applied liners, varying from about 80 mils of spray-on asphaltic materials to 6 inches of conventionally placed asphaltic materials. • Composite liners, consisting of combinations of soil and geomembrane low hydraulic conductivity layers. A3 Waste Material - Settled or mechanically placed tailings often have a low hydraulic conductivity and can be used as pan of the long-term liner system, provided the tailings serve one of the low hydraulic conductivity liner functions illustrated in Figure 7.1. B. CUSHION OR LINER PROTECTION MATERIALS B.l Geo textile - Synthetic geotextile materials varying in weight from 4 to 20 ounces per square yard may be used above or below geomembranes to protect against penetrations from rock particles due to loads from construction activities or the weight of the waste material. The suitability of a geotextile material varies with its method (density) of fabrication. B3 Fine-Grained Soil for Geo membra ne Protection - Soils varying from clay to sand can also be used to protect most geomembranes from equipment traffic or static loading of the waste material. Small gravel size material has also been used to protect thick geomembranes. The protective soil must be relatively free of large rock panicles which could cause stress concentrations on the liner. B3 Coarse-Grained Material for Clay Liner Protection - Gravel protection may also be required for a compacted soil liner, if the liner may be subjected to extreme loads such as impacts from large boulders placed by the end dump method. C. HYDRAULIC HEAD CONTROL COMPONENTS O'O'O-O C.l Free-Draining Gravel Layer - Several inches of free-draining gravel (including coarse sand) are usually adequate to rapidly remove small volumes of leakage. However, thicker layers (8 to 18 inches) are usually placed to facilitate construction. The waste material itself may function for this purpose if the material is granular and relatively free draining. Cut Perforated Pipes - Closely spaced perforated pipes can be used to control hydraulic head above a liner. The required spacing is calculated considering the maximum desired head, and the flow rate and hydraulic conductivity of the waste material between the pipes. CJ Geocomposite Systems - Composite systems consisting of synthetic drainage associated with geotextile filters have recently been developed for a wide range of drainage control functions. Presently these systems have limited load-carrying capacity, but are routinely being improved for wider use. D. LEACHATE COLLECTION AND REMOVAL SYSTEMS (LCRS) O O O O O D.l Synthetic Geonet Materials - Geonets arc net-like polymer products designed to allow high rates of transverse flow Typical thicknesses of these materials vary from 0.16 to 0.30 inches. D.2 Free-Draining Gravel Layer (See Item C.l)
Colorado School of Mines
ER-4609 217 pervious layer above the low hydraulic conductivity liner. To be effective, the pervious layer must be free-draining and direct the liquids away from the liner system. Often, the mine waste material itself is porous so that it alone can act as the free-draining over-liner material. For finer-grained wastes such as tailings, the free-draining layer must be constructed from imported materials. In design, control of the hydraulic head can be an important consideration in determining the potential for leakage to occur. Often it is necessary to establish a trade-off between head control and liner hydraulic conductivity, especially at sites where low hydraulic conductivity materials are not readily available. Figure 63 illustrates that cushions or liner protection layers may also be appropriate for this type of system, if required to avoid liner penetrations from underlying conditions or from the free-draining material. 7.9.4 Composite liners with and without overlying head control (after Hutchison, 1992) This type of liner system usually involves the placement of a geomembrane liner directly on top of a low hydraulic conductivity soil or clay layer. To be effective, the two layers must be in close contact so that any leakage through imperfections in the geomembrane liner must also pass through the soil or clay layer as a limited point source. Without this contact, leakage through the geomembrane can spread and hence leak through a large soil or clay area, thus increasing the total leakage through the liner system. Usually in mine waste application, there is sufficient load placed on the upper geomembrane to assure
Colorado School of Mines
ER-4609 218 good contact between the two layers. As for single liner systems, liner protection and/or hydraulic head control layers can be located above the composite liner system (see Figure 63). 7.9.5 Double liners with an intervening leachate collection and removal system (after Hutchison, 1992) This type of system assures a very low hydraulic head on the bottom liner. This is provided by a free-draining LCRS medium above the bottom, low hydraulic conductivity liner, plus a low hydraulic conductivity top liner. Acceptable quantities of leakage through the top liner and into the LCRS should avoid the potential for unacceptable leakage through the bottom liner. Leakage through the bottom liner depends upon the rate at which the LCRS material can remove any leakage which does occur and the hydraulic conductivity of the bottom liner. In many cases, it is possible to remove this leakage so that the head on the bottom liner is just a fraction of an inch in those areas exposed to top liner leakage, and negligible or zero elsewhere. It is important that the LCRS removal rate be adequate to avoid the buildup of a continuous column of water between the top and bottom impervious liners. Otherwise, the hydraulic pressure above the top liner would be transmitted directly through to the bottom liner, and the double-lined system would have essentially no benefit toward reducing the potential for leakage through as a result of the hydraulic head.
Colorado School of Mines
ER-4609 219 7.9.6 Drainage system above double liners, with an intervening collection system (after Hutchison, 1992) This system provides an additional level of assurance that the liner system will not be exposed to excessive hydraulic head. The top, free-draining layer, which can be the waste itself, reduces the hydraulic pressure on the top liner and therefore reduces the potential for leakage to occur into the LCRS between the liners. The reduced leakage into the LCRS, in turn, reduces both the thickness of the film of water which may flow above portions of the bottom liner, and the resulting potential for leakage to occur into the underlying unsaturated zone. 7.9.7 Low conductivity liners (after Hutchison, 1992) Tables 25, 26, and 27 provide summary data for a variety of low hydraulic conductivity liner materials which can be applied to mine waste management units. The liner types are divided into categories of low hydraulic conductivity soils, geomembranes, field-applied liners, and composite systems. 7.9.8 Basic liner properties (after Hutchison, 1992) Table 25 summarizes typical hydraulic conductivity values for a variety of soil and synthetic materials. Table 26 indicates typical physical properties for commonly-used low hydraulic conductivity synthetic liner materials. Table 27 qualitatively indicates problems which can occur for each liner type and indicates the potential degree of difficulty
Colorado School of Mines
ER-4609 222 uonipiitiQ praifofjig « 6 ° o O ° ° o o 0 « uonipuSaQ O O O o O o o O o o O o o o o O o O o o O s Iui^wq uoanraaQ ® © ® © © © o Xupqnsui adoy • • • • • • • • • o o o o o o o lunia*s Xq i$raiiu| Xiqiqtimty o o © ® o uoaipulaa |rjnu*o • • • • e • o o o o • 0 o o o o o o o o o luiptap) may uoninnzy © # # o © o ® @ ® © s « © iiniudmbg uioy nentrouy # # ® © © © o # ® © ® ® © uoniuusjaO oi.ana •UT13«'D o • © o © • o O O ° ° 0 o o o o o o o ° ° qsniyiwquv o • o o o ° o o o o o o o o o uon»put>a wjowmin © © o o @ o o © 0 o o © © o ! Xiipqmui nfcis o O o o o o o «ny o ° o o o o o i t Itnqjuj «tans jtuioqi • o o 0 o o o o o o o o o o O Sunpuo «irqyœay # # © @ © o Ï luiptO uontaroaQm # © @ # © o o Xiiuuojiu^ |>uanH UD OCVYÛ o # o o o o o o o o o o o 0 00 0 O 0 o ipq or yu ua r« oK y t uu om nt up pV u < y" 0 @ © © © © 6 e # © o • iiomdmbg may uontnouy # @ ©© O © o # e © © # « o uontuuopa m joq tuppuj o ° o o o o o O ° ° o 0 o 0 o o o o o ° ° limpuo *eiu/*oay # ©# ® 0 o o ‘0 tuiqnr) uoanWBQ©# @®© o o © • o I © • 1 Î! 1 © © • 1 Î 1 SI 1 ! I 11 o o * 1 i JG 5111 j 9 i1 II »1I 1 !! 1 6 I ° ° • 1 i i i 1 s & B s I Ï u » Qpaatapiqaai) (pumpmtuooQ) Al olA nU nS D vT nI wC OK iAN HOO siQN nn o aN xw ruawaw smm 03nddv<mu US JW SO1L iS KÀ OS D OHWUOdTUd ytivitcudnoc ciluardyh wol rof smelborp laitnetoP 72 ELBAT )2991 ,nosihctuH retfa( srenil DETALER NOITAREPO UUTAIER NOITCURTSNOC METSYS RENIL
Colorado School of Mines
ER-4609 223 that may be realized in attempting to solve each problem. 7.9.9 Potential problems with low hydraulic conductivity liners (after Hutchison, 1992) Table 27 qualitatively illustrates potential problems which may be realized with various low hydraulic conductivity liners. The size of the circle in the matrix indicates the relative potential for the identified problem to adversely affect liner performance. The darkness of shading in each circle indicates the degree to which special efforts may be required to resolve the difficulty. The most consistent potential problem with low hydraulic conductivity soils is associated cracking, from the effects of either desiccation (drying) or freeze/thaw cycles. Desiccation cracking can occur up to several feet deep with a substantial increase in hydraulic conductivity. Freeze/thaw cracking is more difficult (costly) to protect against because a greater cover thickness is required. For this reason, compacted soil liners may not be appropriate when large portions of liner will be leftexposed throughout freezing winter conditions. Additional important potential problems with low hydraulic conductivity soil/rock liners are : QA/QC verification of liner continuity for natural, in situ soil and rock liner conditions. This is particularly important for rock formations that may be fractured due to factors such as stress-releasing tectonic folding, differential weathering, etc.
Colorado School of Mines
ER-4609 224 Soil and rock hydraulic conductivity can be unaffected or greatly increased due to the chemistry of the liquid to be contained. High pH liquids or liquids which contain high levels of dissolved salts which could exchange with ions in the soil, can cause hydraulic conductivity increases in soil liners. The most severe potential impacts are associated with high concentrations of organic solvents. However, this is not generally important, as most mine waste management units are not exposed to organics. The most important potential problems for most of the geomembrane liners are associated with the possibility for penetration of the liner as a result of construction activities or waste material loading. These types of problems are most easily mitigated by the provision of cushioning (or protection layers) below and/or above the synthetic material to reduce the effects of concentrated loading conditions. Polyvinyl chloride (PVC) liners also can be subjected to accelerated deterioration if exposed for long periods to sunlight or to certain organic chemicals. Composite systems, which consist of a combination of compacted soil and a geomembrane, have less potential for either penetration or cracking problems than either material alone because the dissimilarities in the two provide redundancy. The system least likely to suffer from potential penetrations is one in which a geomembrane is sandwiched between two compacted soil layers. This type of system can only be economically feasible, however, when the low hydraulic conductivity soils are readily available and a
Colorado School of Mines
ER-4609 226 7.9.11 Hydraulic head control layers (after Hutchison, 1992) Provisions to control hydraulic head above a low hydraulic conductivity liner are important in minimizing the potential for excessive leakage. This is accomplished by providing a free draining system above the liner, with the primary function of removing liquids without the need to establish a large hydraulic gradient. Table 23 provides descriptions of three basic types of hydraulic head control systems commonly used in mine waste management practice : free-draining gravel, perforated pipes, and geocomposite systems. The rate at which liquids are removed from any of these systems depends on: (1) the system's hydraulic conductivity, (2) the slope of the waste management unit base, and (3) the volume of liquid involved. 7.9.12 Leachate collection and removal systems (LCRS) (after Hutchison, 1992) Instead of controlling the liquids above the low hydraulic conductivity liner systems, the primary function of an LCRS is to detect and collect liquids that may pass through a top, low hydraulic conductivity liner. As shown in Table 23, an LCRS is. placed immediately between two low hydraulic conductivity liners. Therefore, liquids that pass through the top liner can be removed with only a fraction of an inch of hydraulic head buildup on the lower liner. An LCRS usually consists of the following components: The drainage layer itself
Colorado School of Mines
ER-4609 228 of water, especially for ground water at depth. For a given hydraulic head and unsaturated zone characteristic, a geomembrane or composite liner will perform better than a single clay (compacted low hydraulic conductivity soil) liner. Composite liners, consisting of both geomembrane and clay, provide greater protection than either material alone. A liner system performance can be predicted best by its overall design and construction QA/QC procedures, and not simply by its component specifications (e.g., thickness). For example, if a liner will be subjected to constructed or operational loads, geomembrane components may require protective layers below and/or above the liner, to avoid penetrations. In those cases, the properties of the protective layers and not the geomembrane thickness often will be the most important design criteria. Clay liner performance does not increase proportionately to its thickness, but potential leakage does decrease as the thickness is increased. The thickness of the clay liner should be designed to be sufficient to avoid leakage breakthrough which poses a threat to the beneficial uses of ground water during operation of the waste management unit. At sites with limited availability of clay, the minimum practical thickness of compacted soil is warranted. Based on experience, this is
Colorado School of Mines
ER-4609 229 considered to be 1 foot, although with special construction, considerations 8 to 9 feet can be utilized. The potential threat to beneficial uses of water can vary greatly, depending on operational factors such as the duration of waste unit operation and controls employed to reduce the hydraulic head pressure exposed to the liner. Low hydraulic heads can be maintained by providing a control component above the low hydraulic conductivity layer, or a double liner with an intervening LCRS. The potential threat to beneficial uses also will vary depending on liner system design and the condition of the unsaturated zone below the waste management unit. Potential migration periods through the unsaturated zone can be tens or hundreds of years with a properly designed and constructed geomembrane liner and verification by state-of-the-art QA/QC procedures, especially when ground water is deep and the unsaturated zone material has relatively low hydraulic conductivity. The rate of migration and probability of seepage reaching ground water depend upon the hydraulic conductivity of the unsaturated zone and the depth to water. The need to provide redundancy within the liner system or to rely upon natural subsurface conditions will depend on waste and site-specific factors. At some sites, liner system construction
Colorado School of Mines
ER-4609 230 and operating procedures may provide basic containment so that the unsaturated zone can be relied upon top rovide redundancy, with an adequate factor of safety. For example : Double liners may not be necessary at many sites if hydraulic heads are maintained sufficiently low. Small rates of leakage into the LCRS for double liners can be acceptable so long as it is collected and removed to avoid significant head buildup on the lower liner. Uniform design specifications either are likely to be inappropriate, with insufficient containment for sites with unfavorable conditions, or too restrictive, resulting in over-regulation of sites with favorable conditions. This need for flexibility is consistent with the conclusion reached by the ERA in the development of draft criteria for designing Municipal Solid Waste Landfill(MSWLF) containment systems. The ERA studies showed that a site-specific, risk- based approach would be most appropriate. In reaching its conclusion, the ERA specifically recognized the importance of climate and geologic site factors, including the effects of these factors on travel time from the waste management unit to ground water. Table 28 summarizes factors that should be considered for site-specific liner system designs.
Colorado School of Mines
ER-4609 231 TABLE 28 Site-specific factors to consider in liner system design (after Hutchison, 1992) Potential Waste Material Toxicity Chemical Properties of the Waste Physical or Chemical Changes REsulting from Mining or Ore Extraction Net Acid Generation Potential Soluble Constitutents for Anticipated Environmental Conditions Special Treatment or Neutralization Procedures Utilized Total Resulting Mass of Soluble Constituents Which Could be Mobilized Under Site Conditions General Water Resource Values at the Site Adequate Quality for Beneficial Use Sufficient Quantity for Beneficial Use Existing or Identified Beneficial Uses Probable Locations of Future Beneficial Uses Leachate Availability to the Environment Waste Material Characteristics Hydraulic Conductivity Based on Direct Measurement, Laboratory Tests, or Grain Size and Density Moisture Retention Capacity Thickness of the Waste Site Climatic Conditions Provisions at Closure to Restrict Infiltration Site Factors Topography Geology, Including Predictability of Uniformity and/or the Potential for Discontinuities Unsaturated Zone Thickness, Contiinuity, Hydraulic Conductivity and Natural Water Content Potential Migration Time for Seepage to Ground Water Effects of Climatic Conditioins on Long Term Unsaturated Zone Migration Characteristics Constituent Attenuation Potential Waste Unit Management Practices Facility Type Waste Placement Method Protection of Liner Systems From Environmental or Physical Damage Controls on the Hydraulic Head Risk Reduction Practices, Such as Placement of Underdrains, Sub-Aerial Depositions, Limited Time of Operations Non-Liner Barriers, Such as Cutoff Walls Installation of Special Early Warning Monitoring Systems
Colorado School of Mines
ER-4609 232 CHAPTER 8 SLOPE STABILITY In chapter eight, sections 8.1, 8.3, and 8.6 were compiled from the Bureau of Mines information circular 87 55, Design Guide for Metal and Nonmetal Tailings Disposal, written by Roy L. Soderburg and Richard A Busch. Section 8.2 was extracted from Steven G. Vick's book titled Planning, Design, and Analysis of Tailings Dams. Sections 8.4, 8.5, and 8.7 were compiled from the book Geotechnical Engineering and Soil Testing, written by Amir Wadi Al- khafaji and Orlando B. Andersland. Section 8.8 was extracted from an article titled Seismic Assessment of Tailings Dams, written by Thomas G. Harper, Harvey N. McLeod, and Michael P. Davies. 8.1 Introduction to Slope Stability (after Soderberg and Bush, 1977) Since no slope can be regarded as permanently stable, slope stability is a relative matter. However, in soils engineering practice, the term is used in reference to the possibility of a sudden relatively deep-seated slide. Soil and rock materials fail in shear if the applied shearing stresses on any surface exceed the shear strength of the materials along that surface. The resisting forces are the shear strength of the materials, both frictional and cohesive. The cohesive strength is minimal or negligible in most cases. Pore water pressure at the failure surface lowers the resistance to sliding because it reduces the ARTHUR LAKES LIBRARY ^ COLORADO SCHOOL OF MINES GOLDEN, CO 80401 x
Colorado School of Mines
ER-4609 233 effective stress. Artesian water from the substrata into the embankment will also reduce the stability of the embankment owing to hydraulic uplift. The shear strength of the materials can be further reduced by weathering and softening by water. The shear strength may be increased by compaction or by chemical cementing of the waste materials. Cracking of the embankment caused by differential settlement can reduce the shearing resistance along potential failure surfaces. This cracking may lead to slide failures or piping. Dense tills are usually strong, their shear strengths can have both frictional and cohesive components, and they may be relatively impermeable and incompressible. Drilling and sampling are necessary to seek out inconsistencies in the materials. Sands and gravels are relatively incompressible, and their shear strength is primarily frictional with no cohesion. Here again, the density, gradation, and particle shape determine their behavior. Loose, fine sand acts the same as the same gradation material in mine tailings. If it is saturated and below the critical density, it is subject to liquefaction under shock load. Silts develop strength from either friction or cohesion, depending on density, gradation, and moisture content. Clay in the foundation may cause embankment settlement and instability. As the embankment rises, the clay may consolidate and gain shear strength. Uncompacted clays in waste piles saturate and swell, reducing their shear strength to almost zero. The finer portions of tailing from metal mines act as clays. The entire output of some mines is mudstone or clay and requires specially designed dams to contain it. Phosphate slimes are also a special case because of the fineness of the clay and the pulp density of
Colorado School of Mines
ER-4609 234 the material being impounded. 8.2 Phreatic surface determination (after Vick, 1983) Seepage conditions within the embankment exert a controlling influence on its stability, and a primary purpose of seepage evaluation is to assess pore pressures for input to stability analyses. Seepage in tailings embankments is commonly assumed to occur under gravity flow and, for purposes of pore pressure evaluation, is usually determined for steady-state conditions. It is important to note that the Darcy assumptions of steady seepage conditions and gravity flow are useful in the context of stability analysis because theyu sually yield conservative estimates of pore pressures. The most common procedure for determining phreatic surface location is by flow net analyses. 8.3 Rotational Slides (after Soderberg and Bush, 1977) A somewhat idealized concept of a rotational slide is shown in Figure 64. The crosshatched areas of the figure represent a cross section of the sliding mass. The surface on which sliding occurs is curved and may often be approximated in cross section by a circular arc. The sliding tendency is created by the moment of the mass about the center of the arc. This moment is opposed by shearing resistance developed along the sliding surface. When all available resistance is overcome, failure occurs as shown in the bottom panel of Figure 64. Two pictures of a typical rotational slide are shown in Figure 65.
Colorado School of Mines
ER-4609 237 See Table 29 for procedures for stability analyses TABLE 29 Generalized procedures for performing tailings embankment stability analyses. (after Soderberg and Bush, 1977 Step Method (1) Select trial embankment slope configuration Experience and judgment (2) Determine phreatic surface location based Flow nets on internal zoning, material permeabilities, Numerical models and boundary conditions Published solutions (3) Establish whether or not initial excess Compare raising rate to pore pressure pore pressures will result from embank­ dissipation rate for tailings or soft ment raising foundation soils (4) Perform stability computations for Use any of several available computational applicable conditions methods after defining loading ^ cases for analysis, and appropriate strength behavior under drained and undrained conditions (5) Return to Step 1 and revise trial configuration if factors of safety are not adequate Hand calculations for the factor of safety of a given embankment by the "method of slices", utilizing trial and error, is a long and tedious process. 8.4 Stability of homogeneous slopes (after Al-khafaji, 1992) When dealing with slopes in homogeneous soil deposits, it is possible to derive a general expression similar to those developed for infinite slopes. All other cases require use of approximate numerical or graphical techniques. The factor of safety in all approximate methods of analysis for finite slopes is defined in terms of moments about the center of an assumed circular failure arc. This concept is illustrated graphically in Figure 66. The analysis of a finite slope in any soil can be made by first
Colorado School of Mines
ER-4609 239 Thus far, the factor of safety has been calculated for only one failure surface. To determine the minimum factor of safety we must calculate the FS for many trial failure surfaces by assuming other radii and centers. The proper factor of safety corresponds to the lowest value computed for all of the assumed failure surfaces. This concept is depicted in Figure 67. Here, a grid representing the centers of all circles to be investigated is first specified. The radius at each of the nodes is varied so that the lowest factor of safety for circles with radii ranging from bedrock to the top of the slope is normally investigated. The lowest value is then placed at the node j (j=l,2,... ,m) , where m is the number of nodes in the grid. This procedure is repeated for each node, after which contours of equal factors of safety are drawn and the lowest factor of safety is determined. For the case shown in Figure 67, the soil cross section involves stratified soil layers. The individual soil layer i is homogeneous within its thickness Li. The implication is that the weight of each stratus within the assumed failure surface must be determined separately. The factor of safety for a given center j is then given by: n T TiLi (FS)J = H = J=1.....m <3) The stability of homogeneous finite slopes can be determined using stability charts. Typical stability charts for homogeneous clay slopes are shown in Figures 68 and 69. These charts have been extended to include the influence of surcharge loadings, submergence, and tension cracks as shown
Colorado School of Mines
ER-4609 243 in Figure 70. These charts have one thing in common, they apply to finite slopes in homogeneous soils. Nonhomogeneous soils require a more elaborate analysis. 8.5 Total stress analysis (after Al-khafaji, 1992) Not all materials are going to be a clay with T = c = constant along the failure surface. Therefore, we must look at materials with s = c + a tan 0 . Note that it is also possible to use effective stress strength parameters. Stability chart solutions are available for soils with these strength characteristics as shown in Figure 71. This chart is extended to include the influence of surcharge loadings, submergence, and tension cracks in Figure 70. The method of slices deals with slopes in soils with variable shear strength along the failure surface. The method is extremely versatile and is based on dividing a failure mass into several vertical slices. Consider the problem using total stress strength parameters (strength = i = c + a tan 0) . This assumption generally implies that the pore water pressure is not known. Consider the finite slope shown in Figure 72. The factor of safety against stability failure is defined as before using Eg.(2).Assuming the failure mass consists of vertical slices only, the free body diagram for the ith slice can be determined as shown in figure 73. Since the slice is statically indeterminate, assume that the resultant of EL and SL is equal to the resultant of Er and SR and that their lines of action coincide. This assumption is not a bad one, especially if the failure mass is subdivided into several slices.
Colorado School of Mines
ER-4609 252 Substituting Eqs.(ll) and (12) into (2) yields the desired factor of safety n -U1X1-CT2X2 + r Y / W^sinocj 2=1 When hand calculations are necessary, tabular summary sheets are generally used. Otherwise, computer programs are used for analyzing the majority of practical problems. 8.6 Computer input for factor of safety calculations (after Soderberg and Bush, 1977) In most cases it is extremely difficult, if not impossible, to obtain soil samples that are truly representative of the zone being studied. Consequently, the soil properties developed from these samples must be interpreted and applied with great care. Assuming that input values developed are representative of the actual case being studied, the computer factors of safety are general guidelines and are meaningful only if used in conjunction with all of the other design considerations. The engineer must anticipate and design for the worst possible situation; that is, ultimate height, maximum phreatic line, saturated soils, and seismic activity. The safety factor will only be meaningful if such values are used for the computer program. Grain-size distribution, the area of the tailings pond, and the rate of discharge have much to do with the stability
Colorado School of Mines
ER-4609 253 of the embankment. A finer grind in the mill with the coarse fraction being taken out for use as underground fill will make dike building more difficult. Combine these factors with the small impoundment area compared with the tons of waste per day, and the situation becomes worse. The tailings used underground at some properties do reduce the total that must be impounded on the surface by 40 percent or more, but they also remove the coarse sand (the best material for building the dike) and the coarse sand beach, which provides added safety. An example of rapid building is a situation in which a 500-ton-per-day operation is impounded in a 5-acre site with a pond rise of 1 foot in 33 days. Even a 10-acre site with a 1-foot rise in 66 days is much too fast. Such situations could cause a rapid increase of pore pressure because the water does not have time to percolate through the fine material. Even with the best of conditions, a rapid building rate is not good, and every effort should be made to keep the annual rise compatible with the seepage ability of the soil or drains. Piezometers installed in proper places in the embankment will allow monitoring of the pore-water pressure in the dam, which can be related directly to the safety factor as shown in Figure 77, a typical graph showing the variance of safety factor with phreatic water height in the dam. This type of chart can be developed for any dam and used by the operators to predict the safety of the embankments. The phreatic surface is related to the rate that material is placed around the periphery of the dike. Safety factors cannot be accepted at face value if there is the possibility of liquefaction from either earthquake, sonic blasts, or sudden load. Soils in the 80- to 280-mesh
Colorado School of Mines
ER-4609 255 sizes, saturated and above the critical void ratio, are very sensitive to liquefaction. Soils finer than that would be too sluggish in their reaction to shock because of lower permeability, and those coarser would dissipate the water fast enough to make failure from shock unlikely. All soil testing requires stringent testing conditions and experienced personnel. Since Coulomb's theory and stability equations are approximations, these samples have to be studied thoroughly by experienced soils engineers to insure that true values are obtained for any stability analysis. Once these values have been interpreted, it is simple to predict the stability using the computer. Ultimate height, slope, and water movement can be determined by soil engineering analyses. Before any major construction of this type, these analyses should be made either by the operating firm or by a consulting agency. The cost is only a few percent of the total investment. 8.7 Recommended factors of safety (after Al-khafaji, 1992) Table 30 gives factors of safety suggested by various sources for mining operations. All of these factors are based on the assumptions that the most critical failure surface is used in the analysis, that strength parameters are reasonably representative of the actual case, and that sufficient construction control is ensured. There is no substitute for a sound sub-surface investigation and for a credible laboratory program for soil property determination. The lower the uncertainty, the lower the factor of safety required to ensure safety.
Colorado School of Mines
ER-4609 256 TABLE 30 Factors of Safety Suggested for Mining Operations (after Al-khafaji, 1992) United States (Federal Register, 1977) Minimum Safety Factor I End of construction 1.3 II Partial pool with steady seepage saturation 1.5 III Steady seepage from spillway or decant crest 1.5 IV Earthquake (cases II and III with seismic loading) 1.0 Design is based on peak shear strength parameters 1.5° 1.3* Design is based on residual shear strength parameters 1.3° 1.2* Analyses that include the predicted 100-year return period accelerations applied to the potential failure mass 1.2° 1.1* For horizontal sliding on base of dike in seismic areas assuming shear strength of fine refuse in impoundment reduced to zero 1.3° 1.3* °Where there is a risk of danger to persons or property. *Where no risk of danger to persons or property is anticipated. 8.8 Seismic Assessment, (after Harper, 1992) A significant portion of the operating and closed large dams in North America are tailings dams used to impound wastes from mining operations. Often these dams were built by mine personnel, and may not have been subjected to the rigorous quality control measures required for conventional water storage dams. Construction methods and materials vary considerably. Tailings dams’ stability in earthquakes has become an increasing concern as seismic knowledge has advanced. In Chile, failure of the new and old El Cobre tailings dams following the La Ligua earthquake in 1965 killed 200 people. Growing awareness of seismicity and the potential liability associated with tailings dams are raising the requirements for assessing the stability of both closed and operating structures.
Colorado School of Mines
ER-4609 257 The main seismic design factors for tailings dams are the governing site seismicity and the properties of the tailings. Site-specific earthquake-induced motions can be calculated either deterministically (based on the largest earthquake thought possible given the geological evidence within the region) or probabilistically (by estimating ground motions and their annual probability of being exceeded) . Tailings behavior under seismic loading depends on the soil characteristics and the size and duration of the seismic event. The tailings and foundation materials may behave in either a brittle or a non-brittle fashion. Several soil characteristics influence the dynamic behavior of hydraulically placed tailings and foundation materials, including relative density, modulus, fabric of soil, and aging. When planning new impoundments or raising existing ones, designers should look for a deposition method or construction technique that optimizes the impoundment’s seismic resistance. These include using a rockfill toe berm; constructing a dam with cycloned tailings to use the coarser, more liquefaction-resistant fraction in the dam; depositing tailings by methods that ensure that coarse, drained tailings form the dam; or compacting tailings to increase liquefaction resistance. However, selecting the method most appropriate for the anticipated seismic conditions at a given site requires broad experience. Deposition planning that satisfies seismic requirements while not ignoring space restrictions and environmental and economic considerations is very challenging, and must take into account impoundment geometry, seepage control measures, tailings deposition
Colorado School of Mines
ER-4609 258 mode, hydraulic fill composition and tailings line density. 8.8.1 Methods of analysis (after Harper, 1992) Once site seismicity has been evaluated, the next step is to assess the significance of the design seismic loading on the tailings structure. If large-scale brittle behavior is not considered likely, a deformation analysis is recommended. It may be argued that rigorous deformation analyses are more suited to formal water-retention dams, but tailings impoundments are often just as susceptible to catastrophe if significant crest settlement occurs. Many have free water or slimes contained with minimal freeboard and require permanent deformation analyses to ensure that the freeboard is sufficient. If tailings behavior is potentially brittle, some form of seismic stability analysis is required. The level should depend on the consequences of failure in loss of life, property damage or mill downtime. Site seismicity can be evaluated empirically by examining relationships between earthquake magnitude and the furthest liquefied site or through one of a number of analytical procedures requiring an ascending degree of effort : Residual-strength, or steady-state, analyses use the concept of minimum post-earthquake strength with the well-accepted methods of conventional limit equilibrium analyses. This type of analysis is typically conservative, and if an adequate post-earthquake stability is determined using it.
Colorado School of Mines
ER-4609 260 now, this level of analysis is typically reserved for major tailings impoundments whose failure may cause extreme economic loss or potential loss of life; as the body of experience grows effective- stress finite-element analyses will become a commonplace design tool. A companion process to analyzing the stability of tailings structures is the quantification of the risk of failure. In cases where a deterministic assessment indicates that the dam will fail, it is important to quantify the risk. This can be done on a probabilistic basis, in which the probability of seismic events is coupled with the behavior of the dam to calculate the annual probability of failure. One method uses the program PROLIQ (from the University of British Columbia), which calculates the probability of seismically induced liquefaction for level ground sites. Modifications can be developed for sloping ground. Program input involves field data and the results of a regional seismicity evaluation. A slight modification of the PROLIQ model has been used successfully on major tailings projects. Other methods are available but are all essentially similar. Regardless of the methodology, the results of a probabilistic event should be carefully rationalized and used with extreme engineering judgment. 8.8.2 Seismic upgrading (after Harper, 1992) The requirements for seismic upgrading must be assessed according to: (1) impact of the failure on people and the
Colorado School of Mines
ER-4609 261 environment; (2) cost and feasibility of upgrading works; and (3) the risk or probability of failure. To decide whether the risk of failure is acceptable requires determining the risk criteria. For a deterministic assessment, it is commonly accepted that the dam must have a factor of safety of over 1.1 for the maximum credible earthquake (MCE) if its failure could cause loss of life or extensive environmental damage. The MCE criterion normally uses the limit equilibrium method for stability analysis, and it is commonly applied to water- storage dams. A main concern with tailings dams is the potential for liquefaction, and this requires closer integration with the risk of liquefaction. Criteria for a probabilistic assessment has not been rigorously set. An annual probability of less than 0.0001 is currently operative for numerous water-storage dams in the U.S. This could be regarded as a minimum target criterion for tailings dams. The impact of a failure depends on the runout distance for the failed section. Runout assessments have been carried out for numerous failed structures worldwide and form the basis for current estimations. The basic options available for seismic upgrading are : (1) drainage; (2) densification; or (3) construction of buttress/containment structures. Drainage is a very important parameter and can often be improved at a relatively lower cost than other options. It can, however, be very difficult because of the typically fine grain size of the tailings and the fact that loose deposits at the base of the pond could require pumping if gravity drainage is not feasible. Densification methods include dynamic compaction, vibro-wing and vibro-
Colorado School of Mines
ER-4609 264 Pressure cells. 9.2 Piezometers (after Kealy, 1972) Incorrect choice of the piezometer has caused unsuitable monitoring in the past because of mechanical problems or incorrect pressure ranges. A laboratory testing program was initiated by USBM to study the characteristics of various types of piezometers. The knowledge gained should be useful in selecting a suitable piezometer for various water-monitoring applications. In active tailings ponds the slime zone is not readily accessible; therefore, when piezometers are placed in these areas, lines have to be run horizontally to a reading station on the bank or some other solid footing. In these installations only hydraulic, pneumatic, or electric piezometers can readily be used. Standpipes and other Casagrade-type piezometers cannot be used because of the lack of access to the piezometer and air-locking problems. Four pneumatic piezometers, one electric vibrating-wire piezometer, and one electric strain-gage piezometer were obtained. All were tested under laboratory conditions and monitored by instruments more sensitive than the piezometers themselves. Table 31 was compares the six piezometers. Figure 7 8 shows piezometers installed to sample pore water pressures in tailings. Figure 7 9 shows piezometers monitoring changes in water quality in tailings dam seepage.
Colorado School of Mines
ER-4609 265 TABLE 31 Comparison of piezometer devices by weighting system (after Kealy, 1972) — E UH 5 1 E u ill 3 JZ < c 2 S1 ? r Cost of piezometer j 3 ! 5 i 2 * | 6 1 Cose of readout 1 ! 5 ; 3 j A j 6 2 Total cose 2 i1 5 1i3 1 T Ease of reading (hookup and 1 reading) 3 Î 4 ! 1 6 ! 5 2 Repeatability (total average standard deviation) 1 4 5 3 ! 2 Accuracy at maximum reading 1 ! 3 5 ‘ i 2 . 1 Accuracy at maximum reading as percentage of \ ? ! , j , total range tested j ~ ! , 1 i i 6 ! Accuracy at mid-range reading 4 ! 2 j 1 3 ! 5 i Accuracy at mid-range reading as percentage of j , ! j total range tested | 'l j V ^ 3 5 • i 1 1 Ease of calibration ! i 2 3 1 * h Ease of determining head from reading 1 1 3 2 1 Slop lie ity of readout j 1 2 3 ! 1 i i Water sample possible 5 1 ; 5 5 ! 5 j 3 1 Time lag 1 l | 2 I 3 | 2 | 5 Volume water required for head measurement 1 2 j ? ; ?! 6 Se1f-conrained 1 2 I 2 1 Diameter hole required (1 = rain hole) 3 1 4 j 2 j 3 -------------------------------- - 3 1 1 = roost favorable 6 = least favorable cirtcelE ___)ecnatsiseR(__ cirtcelE )eriW gnitarbIV(
Colorado School of Mines
ER-4609 267 9.3 Slope movement (after Soderberg and Bush, 1977) There are inherent hazards in constructing and maintaining mine tailings embankments. These hazards are difficult to recognize and quantify. Inspectors are trained to look for certain "telltale" signs of hazardous conditions, but often factors such as creep or slow local failures are not noticeable without the ability to relate to prior embankment conditions. Numerous systems are available for observation of ground movement and also for mapping areas of interest. Slope indicator devices can monitor small movements of a slope, and standard ground surveying techniques can be applied to mapping. These standard surveying techniques, which enable mapping and locating of specific points on a structure, are time consuming. One of the simplest but most effective ways to measure the movement of embankment is to drive pipe or rebar vertically into the berm of an embankment in a straight line of sight so that any movement can be measured by simply measuring the offset from the line of sight. On cross-valley dams the permanent stations can be placed on the natural ground or rock off the embankment with the recording stations set on an abandoned berm where they can be protected from the spigoting or dike building operation. These should be measured for elevation changes also. In flat country the survey line must be brought up to the embankment from the natural ground at each end of the downstream slope, and the movement markers placed the same as described for the cross-valley dams. This is important because the line of site must be made from points on top of the sand berm, and these points themselves could have
Colorado School of Mines
ER-4609 268 considerable movement which then would not show the true movement of the other points. The permanent points on the natural ground at each end should be placed at a considerable distance from the embankment so as to be unaffected by the foundation deformation. Another device for measuring both horizontal and vertical movement is the slope indicator, which is lowered into a casing with grooves which guide it and measures the movement in two directions at right angles. For this instrument the casing must be set into bedrock, or if the alluvium is deep, the bottom of the casing must be set into the foundation and the casing installed in the embankment as it rises. Another way to install the casing is to drill a hole and install the casing after an embankment is constructed. From the readings of the slope indicator a profile of the hole can be drawn showing where the movement is taking place from top to bottom (Figure 80). 9.3.1 Other detection methods (after Kealy, 1972) As an aid to inspection personnel, a faster and more reliable reconnaissance system must be integrated with a suitable method of cataloging and evaluating the gathered data. In August 1975, CH2M Hill was awarded a contract to identify, develop, and test a rapid system for monitoring coal refuse embankments to aid in inspections. The rapid- monitoring system has been shown to be an effective, economic, and powerful monitoring tool. The system obtains high accuracy by using convergent and vertical photography
Colorado School of Mines
ER-4609 270 from conventional fixed-wing aircraft. Either black and white (B&W) or color-infrared (CIR) film may be used; B&W film is superior from the standpoint of accuracy, while the CIR film offers better evaluation of qualitative stability indicators through photointerpretation techniques. The rapid-monitoring system also provides a method for maintaining a permanent, visual record of site conditions. The rapid-monitoring system of aerial photography proved to be practical for detecting and recording movement of targets on slopes and embankments. The system uses existing technology and conventional aerial photography with fixed-wing photo aircraft* Targets for the rapid-monitoring system are readily available, inexpensive, and easy to install. A destroyed target can be replaced easily, and base coordinates for detection of movement can be determined at the next monitoring. The rapid-monitoring system could be readily implemented for embankment inspection. Economics are enhanced when several areas can be photographed in a single photo mission. Cost is often the most decisive factor in planning the scope of monitoring programs. Some of the factors influencing the cost are site location and size, vertical relief, accessibility, vegetation, weather, and frequency of monitoring. For the landslide and the coal refuse embankments studied, costs of monitoring by the rapid- monitoring system were estimated to be 15 to 40 percent of the costs of monitoring by conventional ground survey methods. Besides being a useful inspection tool for coal refuse embankments, the rapid-monitoring system can also be used for a wide variety of other types of mining problems such as the stability of mine spoil piles, open pit mines, and
Colorado School of Mines
ER-4609 271 tailings dams. The system's accuracy is such that major earth dams and the stability of natural and manmade slopes can also be monitored to detect and record movement. 9.4 Pressure Cells (after Soderberg and Bush, 1977 ) Pressure cells should be placed on the top of the decant lines or other culverts that will be under the embankment to measure the pressure of the soil and water against the plane surface. The reinforced-concrete conduit should be designed for the combined weight of material at maximum height, but the pressure cells placed at strategic spots along the line could verify the actual pressures. This information can be valuable in designing future conduits. 9.5 Records (after Soderberg and Bush, 1977 ) All instruments installed in an embankment should be read and recorded on a regular basis, which at startup should be quite frequently. Later, as patterns are established, the time between readings can be altered to suit the situation. These records are very important and should be reviewed periodically to see if additional records are needed and to note the trends. The piezometric level measurements are probably the most important and have a direct tie-in with the seepage or flow from the drains. Precipitation measurements should be made, and positions of the spigots or cyclones that are operating should be noted. Often the length of time a group of spigots can be left in operation depends on the rise of the piezometric level in
Colorado School of Mines
ER-4609 273 CHAPTER 10 MAINTENANCE AND INSPECTION In chapter ten, sections 10.1 and 10.2 were extracted from the Bureau of Mines information circular 8755, Design Guide for Metal and Nonmetal Tailings Disposal, written by Roy L. Soderburg and Richard A Busch. Sections 10.3 through 10.3.6 were compiled from the document Tentative Design Guide for Mine Waste Embankments in Canada, D.F. Coates being the project officer. Section 10.3.7 was compiled from Jan Eurenius’ article Long Term Studies and Design of Tailings Dams. Section 10.4 through 10.4.2 were extracted from an article titled Reclamation of Mineral Milling Wastes, written by K.C. Dean and R. Havens. Section 10.4.3 was compiled from an article titled Comparative Costs and Methods for Stabilization of Tailings, written by Karl C. Dean and Richard Havens. 10.1 Overview (after Soderberg and Bush, 1977) The active life of mine tailings embankments may be from a few years to as much as 100 years ; during this time many changes can take place that affect the stability of the embankment. This type of construction is radically different from a water-type dam where the construction is done in a relatively short time under close quality-control of material and methods. The physical properties of the tailings used in pond construction may change over the years for many reasons.
Colorado School of Mines
ER-4609 274 Such changes can alter the stability of an embankment, resulting in variations in the factor of safety. One of the most common changes is the increase in tonnage to the mill without a compensating change in tailings area; this will mean an increase in the annual rise of the dam, reducing the factor of safety. A change in grind with an increase in the minus 200-mesh material can cause a higher phreatic line, a possible decrease in the efficiency of the drains, and increased seepage through the starter dam. Any one or combination of these changes can mean a decrease in the factor of safety of the embankment. It is therefore important that a continuous program of inspection and maintenance of the embankment be started at the beginning and maintained throughout the life and even after the abandonment of the embankment. The records of the instrumentation as described previously are one of the most important aids in determining the safety of the tailings dam, and in a high dam are an absolute necessity. High embankments should be thoroughly inspected by a competent engineer at least twice a year during the active life of the pond. A review of the records of the instrumentation in the embankment should be included in this inspection. Daily inspection should be made of the spigots or cyclones, the decant lines, and position of the water pool in relation to the decant or the tailings area boundary. The drain lines should be checked for quantity of water and sediment.
Colorado School of Mines
ER-4609 276 Whether the drains are operating, free of sediment, and flowing at a regular rate. Whether there is seepage on the downstream face of the starter dam indicated by weeds growing along the face, or worse yet, seepage on the downstream face of the sand dam above the starter dam. Whether the decant lines are intact and free of cracks that could allow sand to pipe into the lines and cause a total failure. These can be visually inspected. If the phreatic surface is as planned; or is there excess pore water pressure from within the foundation or perched water tables? Whether there have been variations in the water levels or a sudden rise in the water level, the appearance of any new springs, or new seepage on the face of the embankments, foundation, or abutments. Conditions at the seepage exit points, decant and drain pipe outlets should be reviewed: Is the water clear or does it contain sediments; is there sloughing in the area; is water coming along the outside of these pipes; are there sinkholes in the beach or slime zone which would indicate piping? Whether there has been an increase in embankment movement as indicated by the surface control points or slope indicator.
Colorado School of Mines
ER-4609 277 Any evidence of borrow from the toe or any other area of the embankments that might affect stability. Whether the embankment geometry is no steeper than planned. Whether diversion channels and pipes have withstood spring runoff or storms. Are they adequate and in good repair? All of these points must be watched closely to check stability, but if a tailings dam has been designed for a total height of 500 feet and all seems to be going well at a height of 200 feet, a thorough investigation should be made by drilling holes into the fill material. Undisturbed samples can be checked for inplace density, screen analysis, 0 angle, and cohesion. With this information and the phreatic surface and geometry of the embankment, a static and dynamic stability analysis can be run to get the FS at that time. From this information the FS can be projected for an embankment 500 feet high to determine if the slope can be steeper or must be flatter, or if the construction must be stopped short of the design height. 10.3 Remedial Measures (after Coates, 1972) The extent and nature of remedial measures required to maintain or improve the stability of mine waste embankments will vary with circumstances. Some developments may require
Colorado School of Mines
ER-4609 278 extensive work in addition to that anticipated in design. For others, minor repairs may be adequate. Descriptions of the principal types of measures effective in improving stability and in combating erosion follow. 10.3.1 Slope Flattening (after Coates, 1972) This is illustrated on Figure 81. By removing the weight of material near the crest of the slope, the driving force tending to produce a slide is reduced. If the material removed form the crest is dumped over the toe of the slope, measures should be taken to ensure that it does not impede drainage form the embankment. (A drain could be required if the material has a permeability low in comparison to that of the material at the base.) 10.3.2 Berms (after Coates, 1972) This is a special case of slope flattening. It is also illustrated on Figure 81. Generally, a berm can improve foundation stability but may not be very effective in increasing the factor of safety against slope failure unless it is at least one third to one half the height of the embankment. Its effectiveness will also depend on the shear strength in the berm. Compaction may be necessary. 10.3.3 Height reduction (after Coates, 1972) This could be in the form of excavation to forma berm, as shown on Figure 82, thus flattening the overall slope.
Colorado School of Mines
ER-4609 281 10.3.4 Seepage control (after Coates, 1972) Inverted filters can be used to prevent erosion and sloughing caused by seepage from embankment foundation surfaces, as shown on Figure 82. Such filters can prevent the development of piping. However, if piping has developed to the point where sink-holes are forming on the upstream face of the embankment, it may be necessary to blanket this upstream face with impervious fill or to seal the leaks by grouting. Where sub-surface erosion is occurring into culverts or pipes buried under the embankment, these methods may be the only ones possible. Relief wells can sometimes be used to lower the water table under embankment slopes, as shown on figure 82. Such systems can be expanded as the need arises. They are not very effective where the holes are located outside of the embankment fill. Inclined holes can be drilled under the embankment; however, drilling at angles more than about 30° off vertical usually involves additional costs. Relief well systems are effective when installed under the embankment before the start of waste disposal. 10.3.5 Surface Drainage (after Coates, 1972) Generally, less erosion of waste pile slopes will occur when the upper surface of the embankment is graded down towards the hillside or towards the center, as shown on Figure 83, and the runoff led away through drainage ditches or pipes. A drainage system of this type is preferable to one where drains are located close to the top of the slope, as seepage from such drains can affect the stability of the
Colorado School of Mines
ER-4609 283 slope. Erosion of long slopes can be reduced by breaking the length of the slope by berms and leading the drainage water to drop pipes. Broken rock, coarse gravel or grass can also be used to limit slope erosion. Maintenance of embankment drainage systems shouldbe directed to preventing unnecessary entry of water into the embankments. The drainage system should be inspected at intervals, particularly after heavy rain, with a view to keeping the system free of obstruction. Arrangements should be made for any work required to be carried out such as: clearance of vegetation, sediments and refuse from trash screens and drainage ditches, cleaning pipe drains to clearsediments or salt deposits; cleaning of silt traps, attention to the outlets of any drainage zones, or other seepage outlets. repairs as required, the diversion of any flow or accumulation of water into the permanent drainage system, attention to filters. Tailings settlement ponds may require adjustment of inlet and decant arrangements, the clearance of any blockages, and rectification of any undercutting of the embankment slopes brought about by wave action. For completed tailings embankments it may be necessary
Colorado School of Mines
ER-4609 284 to maintain the decant or overflow systems to prevent the accumulation of rainwater or to control surface run-off; the clearance of silt from these systems is likely to be necessary from time to time. Breaching a tailings embankment is not generally an acceptable means of preventing an accumulation of water; this method should only be used if erosion caused by the outflow will not endanger the embankment, block the drainage system or cause nuisance in other ways. 10.3.6 Toe Embankments (after Coates, 1972) Where waste piles are located on relatively steep hillsides they may start to move downhill, particularly if the materials are fine and become saturated, or they are affected by weathering. Such movement can sometimes be stopped by constructing an embankment of compacted material at the toe of the pile. 10.3.7 Surface Erosion (after Eurenius, 1990) The surface of the dam must be protected against water and wind erosion. The major factors affecting water erosion are amount of rainfall, catchment area, runoff, steepness and length of slopes. The stability of natural soils against water erosion is predicted by studying the influence of these factors upon formations similar to a dam construction. Such studies indicate that gentle slopes are essential for the long-term stability. Flattening of the slopes is considered being effective in preventing surface erosion. For high dams (higher than 15 m) the slopes are
Colorado School of Mines
ER-4609 285 suggested to have an inclination flatter than 2:1 and 3:1 (horizontal to vertical). When non-toxic tailings are used as damfill, the surface is mostly stabilized with an appropriate vegetation. If the tailings contain toxic material, the slopes are covered by suitable sealing material. 10.3.8 Dust Control Dust control has become a subject of importance in the regulatory climate faced today. Just a few years ago, measures taken against water erosion would generally serve the purpose of controlling wind erosion. This is no longer the case. Active tailing ponds must be stabilized to reduce dust and wind erosion. Sucessful techniques are simple, and relitively easy to implement. Tailings areas on the dam face and slope can be stabilized by annual treatment with a water soluble organic polymer. These areas see little traffic. Polymers that have been proven successful include citrus based oils and water soluble organic resin. Typically the Polymer is applied mixed with water and sprayed from a water truck. The polymer or resin bonds the fine grained particles together and coats the surface of the tailing with a glaze. This effectively stops dusting and wind erosion. Active areas such as roads and pipe corridors can be stabilized by similar methods. Polymers and resins can be applied with success, but vehicle traffic will crush the glazed surface making repeated applications necessary. Milo hay and straw has been used to good effect by crimping the straw into the active roads, and then applying polymers and
Colorado School of Mines
ER-4609 286 resins to the surface. Water trucks are then used to hit areas that start dusting when the wind picks up. 10.4 Stabilization Procedures (after Dean, 1972) The principal methods for stabilization of milling wastes include : Physical— the covering of the tailings with soil or other restraining materials. Chemical— the use of a material to interact with fine-sized minerals to form a crust. Vegetative— the growth of plants in the tailings. The vegetative procedure is preferred in that esthetics of the area are improved while obtaining stabilization. Also, if a mineralized waste is to be conserved for possible later retreatment or if the area is to be used later for residential construction or recreational purposes, it is beneficial to stabilize the area with vegetation. Vegetation does not hinder retreatment procedures as much as covering the tailings with other foreign materials. Methods have been developed and applied in many areas of the country using physical, chemical, vegetative, and combined procedures. Several milling companies, either independently or in cooperation with the Bureau of Mines, have applied various stabilization techniques to differing types of wastes and environments. The principal aim of these companies has been to achieve effective low-cost
Colorado School of Mines
ER-4609 287 stabilization requiring minimal maintenance. A summary of various procedures tested is given in the following subsections on physical, chemical, and vegetative stabilization. 10.4.1 Physical Stabilization (after Dean, 1972) Many materials have been tried for physical stabilization of fine tailings to prevent air pollution. Other than water for sprinkling, perhaps the most used material is rock and soil obtained from areas adjacent to the wastes to be covered. The use of soil often has a dual advantage in that effective cover is obtained and a habitat is provided for local vegetation to encroach. Crushed or granulated smelter slag has been used by many companies to stabilize a variety of fine wastes, notably inactive tailings ponds. On active tailings ponds, however, the slag-covered portions are subject to burial from shifting sands. Slag has the drawback, unlike soils or country rock, of not providing a favorable habitat for vegetation. Furthermore, suitable slag, like soil and rock, must be locally available. Other physical methods of stabilization include (1) the use of bark covering and (2) the harrowing of straw into the top few inches of tailings. 10.4.2 Chemical Stabilization (after Dean, 1972) Chemical stabilization involves applying a material which reacts with mineral wastes to form an air and water- resistant crust or layer which will effectively stop dusts
Colorado School of Mines
ER-4609 288 from blowing and inhibit water erosion. Chemicals have the drawback of not being as permanent a stabilizing means as soil covering or vegetation. They, however, can be used on sites unsuited to the growth of vegetation because of harsh climatic conditions or the presence of vegetable poisons in the tailings or in areas that lack access to a soil-covering material. Chemical stabilization is also applicable for erosion control on active tailings ponds. Chemicals can be effectively used on portions of these ponds to restrict air pollution while other portions continue to be active. 10.4.3 Vegetative Stabilization (after Dean, 1972) The successful initiation and perpetuation of vegetation on fine wastes involves ameliorating a number of adverse factors. Mill wastes usually (1) are deficient in plant nutrients, (2) contain excessive salts and heavy metal phytotoxicants, (3) consist of unconsolidated sands that, when wind-blown, destroy young plants by sandblasting and/or burial, and (4) lack normal microbial populations. Other less easily defined problems also complicate vegetative procedures. The sloping sides of waste piles receive greatly varying amounts of solar radiation depending on direction of exposure. Studies have indicated that, contrary to popular belief, photosynthesis of plants is not continuous while the sun is shining; under high-temperature conditions, photosynthesis may almost stop. Furthermore, most accumulations of mill tailings are light in color and may reflect excessive radiation to plant surfaces, thus intensifying physiological stress. For these reasons, vegetation that may be effective on northern and eastern
Colorado School of Mines
ER-4609 291 Ingredients : Lessons learned from experience and the consideration of other factors will permit the selection of plants of specific qualities and proven usefulness. They will be compatible with the environment at large since there appears to be very few that will not grow on a prepared tailings surface. New species will be introduced where it is justified and profitable. Economics : Study of the physical, economic, and social framework will help determine the economic aspects of the reclamation scheme. The best use will be made of the tailings with regard to the creation and maintaining of secondary industries where applicable. 11.2 From Waste to Soil (after Leroy, 1972) In nature, deleterious elements and substances as occur from the presence of a base metal deposit, for example, are not necessarily damaging to the, vegetation because their release into the environment generally takes place at a naturally controlled rate. Conversely, the establishment of a vegetation cover on tailings will result in the deleterious substances, if any, being in fact sequestered or neutralized, precipitated by the developing humus layer and the various natural colloids present. Hence, the vegetation cover becomes a protective, neutralizing, pollution-controlling natural device of unsurpassed effectiveness. A fully vegetated acre will transpire from 5,000 to 10,000 gallons of water daily, and
Colorado School of Mines
ER-4609 292 eliminate all previous erosion brought about with every precipitation. An area of tailings, unprotected and exposed to summer heat, will quickly lose all available moisture in its topmost layer and will readily reach temperatures of 100-120 F. Even a moderate wind, blowing unhindered, will cause sand storm conditions; and at over 15 mph a sand blasting effect will quickly take its toll of any freshly established seedlings if there is an exposed partially reclaimed area. Tailing waste will remain sterile until properly conditioned for plant growth. But since uncontaminated, clean and salt-free sand is the ideal medium for plant growth, conditions very near to optimum will be achieved provided one has : moisture, of which there is generally no lack in tailings; fertilizing chemical elements, of which 16 are necessary for plant growth, from N,P,K, down to the lesser and the micro-nutrients; an adequate bacterial population to promote germination and continued growth. Once the specific needs of the particular waste area to be reclaimed are determined, it will be a mere matter of weeks, with proper conditioning before vegetation blankets the prepared area. The basic combination is one of fast growing grasses to act as nurse crop, and slower, nitrogen-fixing legumes to provide the long-range permanent and maintenance free
Colorado School of Mines
ER-4609 293 vegetation cover. A detail of considerable importance, is the homogeneity of the tailings throughout the mass. It means that it can be shaped to any desired form very easily, and the plants established in it will be able to develop a full, coherent root system, a factor of prime importance in the growth of healthy plants. Therefore, this material/waste, can be made more versatile than the common soil, particularly in poorer podzolic soils with their thin fertile horizon. 11.3 Tailings and Slime Ponds (after BLM, 1992) Tailings and slime ponds consist of impounded mill wastes. Slime ponds are tailings ponds with high percentages of silts and clays, which cause very slow sediment drying conditions. Slime ponds are commonly associated with phosphate and bauxite processing, and reclamation is complicated by the slow dewatering. Tailings impoundments are typically placed behind dams. Dams and the impounded wastes may require sealing on a case- by-case basis to avoid seepage below the dam or contamination of the groundwater. This measure only may be done before emplacement of the wastes. Long-term stability of the structure must be assured in order to guarantee ultimate reclamation success. 11.3.1 Tailings Characterization (after BLM, 1992) The nature of the tailings to be impounded should be determined as early as possible during the development of ' any plan. Tailings exhibiting phytotoxic or other
Colorado School of Mines
ER-4609 294 undesirable physical or chemical properties will require a more complex reclamation plan. Analysis should include a thorough review of groundwater flow patterns in the area and a discussion of potential groundwater impacts. An impermeable liner or clay layer may be required to avoid contamination of groundwater. Where tailings include cyanide, final reclamation may include either extensive groundwater monitoring or pumpback wells and water treatment facilities to assure (ensure) groundwater quality is protected. The presence of cyanide in the tailings will not normally complicate reclamation of the surface. 11.3.2 Dewatering (after BLM, 1992) The first phase of actual reclamation will normally be the dewatering or drying of the impoundment so that equipment can gain access to the surface. This can range from simply letting the tailing material dry naturally to more complicated methods of trenching to allow water to escape from the tailings. This phase of reclamation is often complicated by surface crusting of the tailings. This phase of reclamation can take up to several years. Reclamation of slimes will typically require some form of trenching using either balloon-tired vehicles or cable trenching tools. Slimes reclamation can be greatly accelerated by creating surface drainage for initial stabilization using peripheral and feeder trenches. Feeder trenches which drain into the peripheral trench are typically 251 to 40 ' apart, and up to 2 ' deep. Once the surface of the tailings has dried, heavier equipment can be used. Farm equipment can usually operate when the solid
Colorado School of Mines
ER-4609 295 content exceeds 60% in the top 6 feet. Revegetation of the tailings after trenching can accelerate the drying process through evapo-transpiration. Dust abatement may be required at this stage in order to avoid airborne particulates which may constitute a substantial environmental problem. 11.3.3 Reshaping (after BLM, 1992) Depending upon the nature of the tailings, it may be necessary to modify the overall shape of the top of the impoundment to avoid the concentration of water and to improve visual quality. This can involve the addition of material to develop a "crown” on the impoundment and the construction of artificial drainages. 11.3.4 Surface Treatment (after BLM, 1992) Depending upon the nature of the tailings materials, it may be necessary to construct a cover system to isolate the waste. Where the tailing itself is a suitable growth medium, or can be amended to provide a suitable growth medium, this will not be needed. Cover systems typically include: water exclusion layer, capillary break, and growth medium. In some cases, it may be impractical to revegetate the impoundment. Because dry tailings material is highly susceptible to wind erosion and subsequent dust problems, it is important to cap the tailings material with coarse durable rock.