Sei sulla pagina 1di 11

SME Annual Meeting Feb. 28-Mar.

1, Salt Lake City, Utah

Preprint 00-68

PIT SLOPE DESIGN CHALLENGES IN RESIDUAL SOILS AND WEATHERED ROCK: BACKGROUND AND A CASE STUDY H W. Newcomen and B. T. Burton BGC Engrg Inc. Vancouver, BC, Canada J. Geyer Gold Reserve Corp Spoken, WA, USA ABSTRACT For open pits developed in tropical climates the upper portions of the pit walls are often located in residual soils and weathered rock. Due to the presence of relict structures, and the relatively low strength of the residual soils and weathered rock, design slope angles in these materials have to be developed by blending the results of kinematic assessments of geologic structures with rock mass stability analyses and traditional soil mechanics. Background information on the engineering considerations of residual soils are discussed, and the results of geotechnical assessments for proposed pit design of the Brisas del Cuyuni Project, located in southeastern Venezuela, are presented. RESIDUAL SOILS Introduction Excavations in tropical climates often encounter residual soils and weathered rock. A residual soil can be described as a soil-like material derived from the weathering and decomposition of rock which has not been transported from its original location (Blight, 1997). This general definition is a broad term that includes saprolites, mature soils and laterites. The term residual soil is sometimes used in more specific terms to describe mature soil alone. Saprolites are materials that have soil-like strength or consistency, but retain recognizable relicts of the structure and fabric of the parent rock (Blight, 1997). As an example, a saprolite derived from a lava may contain flow bands, amygdules, and joints. Relict structures often constitute planes of weakness and zones of higher permeability within a soil mass. Mature soil is that part of the soil profile which has undergone physical and chemical weathering to the extent that no evidence of the parent rocks fabric or structures remain. Laterites are highly altered residual soils that have had the silica leached out and have some degree of cementation by sesquioxides (Blight, 1997), giving these soils a granular or nodular appearance. Laterites are typically rich in hematite and boehmite and their high iron content gives them a deep reddish color. The term laterite, however, is used very loosely and is sometimes applied to soils with little influence from sesquioxides. Laterization may occur in ancient transported soils as well as residual soils. Due to the nature of their formation, laterites tend to occur near the surface and extend to limited depths. Laterites are often excellent construction materials and may be a source of aggregate. The progression with chemical weathering of a soil from saprolite to mature soil, to laterite will only occur in a favorable chemical environment and is described in greater detail below. Distribution Residual soils can be found throughout the world, but primarily in the tropics, as shown in Figure 1. Broad classes of residual soils can be seen to extend beyond the tropics where favourable circumstances permit. Conversely, numerous areas exist in the tropics where residual soils are overlain by more recent alluvial and aeolian deposits. Structure, Fabric and Discontinuities The following definitions are proposed for use (Fookes, 1997) when discussing the characteristics and properties of residual soils: 1
Copyright 2000 by SME

SME Annual Meeting Feb. 28-Mar. 1, Salt Lake City, Utah Structure the fabric, texture and discontinuity patterns making up the soil-rock material, mass or unit; Fabric the spatial arrangement of component particles; Discontinuities the nature and distribution of surfaces separating elements of fabric, material or soil-rock mass.

Figure 1. Simplified world distribution of the principal types of residual soils; based on F.A.O. World Soil Map (after Fookes, 1997). The structure of tropical residual soils includes macroscopic features such as relict discontinuities, and microstructure or fabric. Although microstructure is important to understanding the engineering behavior of soils, particularly partially remolded soils such as compacted fills, the behavior of the in-situ soil mass is frequently more influenced by macroscopic features. The structural orientation of macroscopic features such as schistosity, fissures, veins, joints, faults, and voids can have a significant influence on the behavior of the soil mass. Mineralogy The mineralogical composition of residual soils is dependent on the composition of the parent rock and the climatic conditions. The mineralogy of tropical soils has engineering significance in the aggregation and cementation of soils (Burton, 1998). The mineralogy can have a significant affect on index properties such as moisture content, and has been observed to affect field instruments such as 2 nuclear densometers. Swelling or collapsing soils can have a significant effect on slope stability. Figure 2 shows that, in a general sense, clay mineralogy varies in relation to the distance from the equator. In the tropics, feldspar minerals (aluminosilicates) weather initially to kaolinite, hydrated iron and aluminum oxides, such as goethite (Fe2O3H2O) and gibbsite (Al2O33H2O ), also referred to as sesquioxides. Other minerals which are more resistant to weathering, such as quartz (SiO4) and mica (Kal3Si3O10(OH)2 - muscovite) may persist, often as individual sand grains in a clayey matrix. With further weathering the kaolinite content may decrease and the sesquioxides progressively alter to hematite (Fe2O3) and boehmite (Al2O3H2O) (Mitchell and Sitar, 1982). Further chemical action may cement these materials and produce lateritic gravels. Weathering Profile The engineering properties of a soil will vary with depth even if it has developed from a uniform parent
Copyright 2000 by SME

SME Annual Meeting Feb. 28-Mar. 1, Salt Lake City, Utah material. Near the surface, soil layers are affected by humus, seasonal wetting and drying, and bioturbation. At greater depth, the moisture content fluctuates less with the seasons, and less organic matter is present. Also, at greater depth the groundwater movement is often slower, and soil particles and solutes are less likely to be transported. ple, weathered granite from the warm, humid Malaysian peninsular has quite different properties from weathered granite from cooler, semi-arid South Africa (Blight, 1997). However, some general trends exist for residual soils, depending on their depth of occurrence and geologic age. Figure 4 shows how soil properties tend to vary with depth due to the effects of weathering.

Figure 2. The effect of climate on frequency of clay mineral occurrence with climate zones represented in a simplified manner as distance from the equator (from Millot, 1979; see Uehara, 1982).

The transition from saprolite to fresh rock may be sudden or may occur over tens of meters. Crystalline rocks, such as granites, granodiorites, migmatitic gneisses, and some metavolcanics, tend to have a fairly sharp boundary. More fissile rocks, such as schists and phyllites, tend to have gradational transitions over many tens of meters. Welljointed igneous rocks will have an increasing abundance of corestones with depth. A knowledge of typical weathering profiles for different parent rocks can help the slope designer predict how geomechanic properties will vary with depth, allowing a rational approach to be taken to dividing up slopes for design. An example of a typical weathering profiles for metamorphic and igneous rocks are shown in Figure 3. This figure also shows a classification system for weathering grades. Many classifications have been proposed; however, it is out of the scope of this paper to compare them. Selected Residual Soil Properties It is difficult to relate the properties of a residual soil directly to the parent rock because of the complex superposition of effects from climate, topography, geologic age, and structure. For exam3

Figure 3. Weathering classification system proposed by Deere and Patton (1971) with typical weathering profiles for metamorphic and intrusive igneous rocks (after Deere and Patton, 1971).

Figure 4. Changes occurring with depth in a weathering profile (adapted from Tuncer and Lohnes, 1977, and Sueoka, 1988; after Blight, 1997).

Grain Size: Experience obtained in residual soils from Brazil (Mori, 1979) indicates that for those climatic conditions the three main types of bedrock
Copyright 2000 by SME

SME Annual Meeting Feb. 28-Mar. 1, Salt Lake City, Utah commonly found consistently weather to different grain sizes. The corresponding rock types and grain sizes of the residual soil are: basalt clayey gneiss silty (often micaceous) granite (sandy) often requires that the faces be washed or exposed to rainfall to enhance the visibility of the structures. Testing the hydraulic conductivity of saprolite must take into account scale effects. The in-situ permeability is often controlled by relict joints and thus laboratory testing can underestimate the conductivity by orders of magnitude. The success of open borehole testing can be affected by smearing of the sidewalls, and packer testing often has limited success due the difficulties associated with obtaining a proper seal for the packers. Larger scale pump testing can be performed to overcome these difficulties; however, the results obtained will incur significant additional costs. Mapping the base of weathering requires careful consideration. Slope instability has occurred where interpreted bedrock was in fact a very large boulder floating in a weathered rock matrix. In some areas, fresh outcrops often give way to troughs of weathering 15 to 20 m deep over horizontal distances of 100 to 200m, and quite sudden depth increases of 30 to 50m (Fookes, 1997). Visual mapping is not a reliable method, because what appears to be rock from a distance may break down to clay and silt under relatively light pressures. CASE HISTORY- BRISAS DEL CUYUNI PROJECT Regional Setting and Project History Over 100 years ago, the KM 88 Mining District was one of the richest gold-producing areas in the world. This area is located approximately 375km south of Ciudad Guyana (locally known as Puerto Ordaz) near the village of Las Claritas, (Figure 6) in Venezuelas Bolivar State. The area is underlain by the Precambrian Guyana shield which extends into Brazil, Suriname, Guyana and French Guyana. Most of the shield is covered in dense rainforest, and mining has traditionally been carried out in the highly weathered saprolitic bedrock. Considerable mineral wealth has also been discovered more recently in the underlying hard rock. The Brisas property is located in the KM 88 mining district. The property refers to the Brisas alluvial gold concession and the Brisas hard rock concession for the gold, copper and molybdenum contained below the alluvial. The property, through a wholly owned Venezuelan subsidiary, was acquired by Gold Reserve Corporation (GRC) in 1992. In February 1998, a detailed pre-feasibility study was completed by Jacobs Engineering of Denver, CO. In addition, a supplemental study was completed in August 1998 on potential improvements to the milling using the Cominco CESL group milling 4
Copyright 2000 by SME

The resultant grain sizes are generally a function of the mineralogy of the various parent rock types. Plasticity: A correlation between plasticity index and parent rock type is shown in Figure 5. Relationships between the index properties of plastic limit and liquid limit of a soil and shear strength are well accepted for temperate soils. In the authors experience, at the Brisas property and elsewhere, the local variability in plasticity for residual soils is much greater than the variability in shear strength. The relationships developed for temperate soils are considered appropriate as a first approximation of shear strength; however, they should be used with caution.

Figure 5. Plasticity Index Results for Various Residual Soil Parent Rock Types (after Blight, 1997).

Site Characterization Investigating the properties of a saprolite differs from investigations of temperate soils in three general ways: relict structures must be characterized relict structures must be accounted for in permeability testing weathering distributions should be mapped.

Relict structures can be difficult to characterize, often they are not readily apparent in drill core due to smearing from the drilling process. Mapping of trench and test pit walls is generally more successful but often these are also obscured by smeared surfaces resulting from the excavation process. This

SME Annual Meeting Feb. 28-Mar. 1, Salt Lake City, Utah process, to facilitate on-site treatment of copper concentrates. GRC continues to conduct work on updating the mineable reserves, optimizing cutoff grades, and is carrying out additional metallurgical testing and tailings characterization. Slope stability assessments for a deeper and larger open pit are being carried out on an ongoing basis by BGC Engineering of Vancouver, BC (Newcomen, et al, 1997; 1999). Four horizons have been identified at the property by GRC geologists. They consist of: saprolite soils a mixed zone weathered rock fresh rock

The saprolites are formed by the complete weathering of bedrock, and are distinguished by the absence of any significant rock content. The top of the saprolite is generally oxidized, with an orange or reddish-brown color, to a depth of between 15 and 20m. The base of the oxidized saprolite is denoted in the drillhole logs by the marker BOS (Figure 7). The oxidized saprolites generally overly a greenish-grey, sulfide-stable saprolite. The marker BAS is assigned to the base of the saprolite. The depth to the BAS marker is variable across the site, but has been logged to depths of up to 100m. Relict structures, such as joints and bedding planes are common in the saprolite. At Brisas, the saprolites generally consist of silt-size particles. The oxidized saprolite often contains zones of white, chalky low-plasticity kaolin. The mixed zone is a gradational horizon between the overlying saprolite and the underlying bedrock. This zone has layers of saprolite mixed with layers of highly weathered rock. The thickness of this zone is variable, ranging from 5m to greater than 30m. The marker BZM has been assigned to the base of the mixed zone.

Figure 6. Location Map

Engineering Geology

Figure 7. Typical Section Through the Proposed Pit

Copyright 2000 by SME

SME Annual Meeting Feb. 28-Mar. 1, Salt Lake City, Utah Bedrock at the Brisas del Cuyuni site consists of a sequence of metavolcanics composed of andesites, tuffs and volcanic sediments which have been intruded by dykes and stocks of diorite, quartz porphyry, and gabbro. The average depth to the weathered rock is about 50 m across the site. The weathered rock is generally identified by carbonate leaching. Rust staining on the fractures and joints suggests that groundwater flow occurs predominately within this horizon. The fresh rock corresponds to unweathered versions of the andesite, tuffs and volcanics, as described above. The dominant structural features throughout the site are bedding and foliation. Bedding and foliation in the rocks both dip to the west at similar orientations and are difficult to distinguish from on another. No significant faults have been identified within the project area. Up to four additional joint sets have been identified from surface outcrop/trench mapping. Hydrogeology The phreatic surface at the project site fluctuates seasonally between 2 to 5 m below the ground surface, and generally conforms to the existing surficial topography. The principal aquifer identified on the property is the weathered rock. Groundwater flow in this semi-confined aquifer will likely be predominantly controlled by fissures and joints. Recharge is derived from gravity leakage out of the overlying saprolite and regional sources. The fresh rock is believed to act as an impermeable lower boundary. Hydraulic conductivities in the weathered rock are -5 -3 estimated to be between 9x10 and 1x10 cm/s, with -4 -5 calculated transmissivities between 3x10 and 2x10 2 m /s (Hydro-Triad, 1996). Conductivities in the saprolite are estimated to be approximately three orders of magnitude lower. These values fall below the threshold of what would be considered productive for water well installations. However, it is our opinion that the actual values are likely higher than those calculated, as they were determined using packer tests which are often found to be unreliable in ground conditions such as those found at Brisas. Geotechnical Investigations BGC Engineering has been carrying out office and field studies since 1997 to refine pit slope angles for various designs proposed by GRC. A comprehensive approach to slope stability has been taken, incorporating engineering geology and rock mechanics information collected during GRCs exploration program. Mapping: Surface outcrops, trenches and most small pits within the concession have been structurally mapped. All of the exposures in the outcrops consist of saprolite or heavily weathered rock, making identification and mapping of geologic structures a challenge. Several small open pits excavated by previous mining activities are located throughout the property. The pits are often filled with water; however, lower water tables during the dry season have allowed structural mapping of the pits to be carried out on a limited basis. Five test pits were excavated around the property to depths of about 5m during the geotechnical investigations. The test pits were logged for stratigraphy and engineering properties. The undrained strength of the soil was measured in the test pit walls using a Geonor hand vane. Disturbed saprolite samples were collected at regular intervals in the test pits. One undisturbed block sample of saprolite was collected from the test pits. Core Logging: GRCs exploration program includes diamond drilling using single-tube diamond coring techniques. To date, GRC has drilled 763 holes on the property, for a total of over 165,000 m of drilling. The drill holes are typically HQ size initially, reducing to NQ when competent rock is reached. The majority of the drill holes are inclined to the east to intersect the mineralization across its dip. Two vertical drill holes were drilled specifically for the geotechnical program, in an area where the bedding dip was expected to be critical to the pit wall design. Engineering logging has typically been carried out by GRC personnel before the drill core is split. Consequently, accurate estimates of RQD, core recovery and hardness have been determined. The logging is generally carried out in material below the saprolite zone, in the weathered and fresh rock. Additional detailed logging of over 3,500 m of core from eleven selected drillholes was undertaken by BGC to determine the properties of the saprolite and the rock. The following geomechanical parameters were assessed: rock hardness (as per ISRM standards) core recovery RQD fracture frequency joint condition degree of breakage degree of weathering/alteration

Copyright 2000 by SME

SME Annual Meeting Feb. 28-Mar. 1, Salt Lake City, Utah The geomechanics information was recorded in a format amenable to entry into a computer database for statistical analysis and subsequent estimation of Rock Mass Rating (Bieniawski, 1976) and rock mass strength (Hoek, 1994). Point load testing was also undertaken on selected representative core using a Roctest point-load testing machine. Diametral and axial testing along and across the foliation were carried out on split and un-split core, to determine if any strength anisotropy exists in the rock mass. Point load strengths were converted to unconfined compressive strengths for subsequent use in rock mass strength assessments. Laboratory Testing: Extensive triaxial testing was carried out by others (Abel, 1997) prior to BGCs investigations. This information was assessed and used, in conjunction with field vane testing results, to estimate undrained strengths for the saprolite (Figure 8).
Su vs Depth Saprolite Samples
0

liquid and plastic limits moisture content grain size distribution

One undisturbed sample was transported to North America for direct shear testing. The results of drained shear testing on the saprolite indicated a peak strength of =25, c=50 kPa, and a residual strength of =21.5, c=0 kPa. Six core samples with natural fractures from the weathered rock and fresh rock units were collected and transported to North America. Direct shear testing was carried out on two of these samples. The testing indicated a shear strength (peak and residual) of = 35, c'=0 kPa for discontinuities in both the weathered rock and the fresh rock. Geotechnical Units Based on the engineering geology, trench mapping, geomechanics assessments and laboratory testing results, three geotechnical units have been defined according to similarities in engineering properties. The distribution of the various geotechnical units on a typical section through the proposed pit is shown in Figure 7. For geotechnical purposes and slope stability assessments, the saprolite unit has been defined as the material above the BZM marker, which includes both the saprolite and the mixed zone horizons. Although the mixed zone often contains a significant amount of weathered rock, it was considered appropriate to combine the two horizons. The strength of the saprolite has been assigned to this unit, since the weaker materials will dictate the stability. The strength of the saprolite is highly variable, but it generally has the consistency of a firm to very stiff soil. The design strength curve for the saprolite, based on drained direct shear testing and index testing, is shown in Figure 9.

20

40
Design Curve Sulfide Saprolite

60
Oxide Saprolite

80 Depth (m)

100

120

140

160

180

200 0 200 400 600 800 1000 1200 1400 1600 1800 2000 Shear Strength (kPa)
Undrained Strengths Determined by BGC with Hand Vane in Test Pits Interpreted Undrained Shear Strength of Oxide Saprolite from triaxial testing undertaken by Abel (1997) Interpreted Undrained Shear Strength of Sulfide Saprolite from triaxial testing undertaken by Abel (1997)

Figure 8. Undrained Strength of Saprolite

During the field investigations, disturbed samples of oxidized and sulfide stable saprolites were obtained from exploration drill core and from trenches. The saprolite samples were sealed in plastic bags and containers and shipped to a Venezuelan laboratory for standard index and soil classification testing. The index testing consisted of:

The weathered rock geotechnical unit has been defined as the zone between the BZM and the base of weathering (BDM). The weathered rock has an ISRM hardness of R1 to R2, placing it in the category of very weak to weak rock. It has a moderate degree of fracturing and the joints are generally open, highly weathered and have little infilling. Based on point load testing and the rock mass rating conducted on the weathered rock, it has an average unconfined compressive strength of 20 MPa, and a RMR of 49, classifying it as a fair quality rock mass. The rock mass strength curve for the weathered rock is shown in Figure 9.

Copyright 2000 by SME

SME Annual Meeting Feb. 28-Mar. 1, Salt Lake City, Utah towards 288. A bedding joint strikes sub-parallel to the foliation and dips approximately 60 to the northwest. Three other weaker sets of cross joints, dipping east and south were also identified. Structures mapped in Domain 2 have been used to represent the anticipated structure for most of the proposed pit, particularly the central and northeastern portions. The foliations in Domain 2 have a peak dip of about 55, dipping towards 298. Several of the discontinuity sets observed in Domain 1 were absent or very weakly represented in Domain 2. Design sectors for the open pit were derived by combining structural domains and pit wall orientations. A total of sixteen design sectors were developed for the proposed open pit, as shown in Figure 10. Design Method Potential Modes of Instability. Instability in excavated rock slopes is commonly initiated along structural discontinuities in the slope. Potential failure modes can be predicted by comparing the orientation of the discontinuities to the orientation of the proposed excavation. Certain failure modes are kinematically possible when the bench face, interramp or the overall slope undercuts geologic structure (i.e. planar) or the line of intersection of a combination of the discontinuities (i.e. wedge). To assess these types of instabilities, the shear strength of the discontinuities are of primary importance. If a rock mass is highly fractured or of poor quality, such as the saprolite and the weathered rock, failure can occur along a rotational surface or through the rock mass. To assess these types of instability, the shear strength of the saprolite or the rock mass are of primary importance. Approach: Structural mapping information was plotted on lower hemisphere stereonets. The plots have been used to determine kinematically feasible failure mechanisms in each design sector of the proposed pit. Limit equilibrium stability analysis techniques were then used to estimate a Factor of Safety (FS) for each potential mechanism. Potential planar modes of instability were by assuming a simple sliding block of unit width. The weight of the block and the resisting forces along the sliding plane, along with any water pressures were calculated to determine a FS of any potential planar failures. The stability of potential wedges identified was analyzed using the commercially available computer program SWEDGE. The calculated FS values were then used to select allowable slope angles based on kinematics.

Figure 9. Geotechnical Unit Strength Curves

The fresh rock geotechnical unit is generally unweathered, and is defined as the material below the BDM. The fresh rock is classified as a medium strong to very strong rock, with an ISRM hardness of R3 to R5. It generally has few fractures, and the joints are infilled with calcite. Varying degrees of sericitic alteration are present in the fresh rock. Based on the point load testing and the rock mass rating conducted on the fresh rocks, it has an average unconfined compressive strength of between 115 and 183 MPa (from the southeast and northwest sides of the proposed pits, respectively), and a RMR of between 70 and 73. In terms of rock mass quality, the fresh rock is classified as good. The rock mass strength curves for the fresh rocks are shown in Figure 9. Note the difference in the estimated rock mass strength of the fresh rock between the NW and the SE sides of the pit. This difference is due to the anisotropy of the rock mass strength parallel to and across bedding. Structural Domains and Design Sectors Areas with similar geologic structure, based on structural mapping information collected on the Brisas property, were grouped into structural domains. Two structural domains were identified in the pit area. Domain 1 encompasses the southwestern one-third of the currently proposed pit (Figure 10). The dominant structures in this domain are the foliations, which have an average dip of about 42

Copyright 2000 by SME

SME Annual Meeting Feb. 28-Mar. 1, Salt Lake City, Utah

Figure 10. Conceptual Pit Design and Design Sectors

The potential for rotational failures in saprolite and weathered rock was examined using chart solutions (Hoek and Bray, 1981) and the slope stability software SLOPE/W, respectively. Due to the highly variable thickness of the saprolite and the weathered rock throughout the proposed pit area, a generic approach to the analysis was chosen. Curves of slope angle vs. slope height for a specified (allowable) FS of 1.2 were developed to assist in design (Figure 11), using the drained strength of the saprolite and the rock mass strength curve developed for the weathered rock (Figure 9). The average depth of the saprolites and/or weathered rock for each design sector was plotted on the design curves shown in Figure 11 to determine an allowable slope angle. The potential for both rotational/rock mass failure and instability due to kinematic controls was assessed for each design sector. The allowable angle based on kinematic controls was compared to the allowable angle based on rotational/rock mass failure, and the lesser of the two angles was used for design.
Figure 11. Design Curves for Saprolite and Weathered Rock (FS=1.2).

RESULTS AND DESIGN ANGLES Saprolite In terms of kinematic analysis, numerous potential wedge and planar modes of instability were identified within the proposed pit area, for several slope orientations. Most of the potential wedges identified 9
Copyright 2000 by SME

SME Annual Meeting Feb. 28-Mar. 1, Salt Lake City, Utah in the saprolite have a FS <1.0 under completely dry conditions. This indicates that wedges formed by relict structures in the saprolite will be unstable if they are undercut. In terms of rotational failure or shearing through the soil mass, effective stress analyses using chart solutions for two groundwater conditions (dry and partially dewatered) indicate that, for slope heights over about 20m in the saprolite, the allowable slope angle is significantly reduced if dewatering of the slope is not completely achieved. Based on these assessments, with the exception of two structural domains (1B and 2I), the main mode of potential instability in the saprolites, particularly for inter ramp slopes, is anticipated to be rotational failure. Due to the presence of relict structures, it is likely that some element of structural control will exist in these rotational failures. This could result in complex modes of instability occurring; however, these are beyond the current scope of this project. For preliminary planning and economic analysis, 40 is considered to be a reasonable angle for prefeasibility level design in the saprolites. This angle is based on an average depth to the weathered rock of 50m, and the assumption that the saprolites can be fully dewatered. If they cannot be dewatered, slope angles of less than 30 will be required. Weathered Rock Similar discontinuity and pit wall orientations to those used for the saprolite were used to assess the stability of the pit wall in weathered rock, albeit with strength parameters for the discontinuities modified accordingly. A majority of the potential wedges identified in the weathered rock have a FS <1.0, even under completely dry conditions. This indicates that wedges formed by discontinuities in the weathered rock will likely be unstable. The recommended average design angle for the o weathered rock is 45 . The actual angle that can be achieved will also depend on the dominant influence of the overlying and underlying geotechnical units. Achieving these angles will require successful dewatering of the pit wall in the upper slopes. Fresh Rock Due to the limited structural data available for the fresh rock, in which kinematic constraints are expected to control the slope stability, design angles in some design sectors within Domain 2 cannot be accurately defined. The high competency of the fresh rock, and the apparent absence of any unfavorably oriented discontinuities, indicate that interramp pit wall angles in excess of 55 may be possible. However, 55 is currently considered a reasonable upper bound for feasibility level planning. Justification of steeper angles for a pit of the currently proposed depth (>350m) will require additional geotechnical information at greater depth, and more sophisticated analyses techniques where stresses and slope deformations are assessed (Newcomen, et al, 1999). SUMMARY AND CONCLUSIONS In summary, the key features of residual soils and weathered rock, and their impact on pit wall stability are: Weathering profiles are influenced by climate conditions, parent rock type, structure, topography, and geologic age; Weathering profiles have a large influence on the strength and permeability of a rock; Permeability of intact soil may be considerably less than the permeability of the in situ soil mass; Relict structures often act as planes of weakness; Pit slope designers are familiar with grouping pit slope sections into design sectors as part of a rational approach to design. Pit design in areas of residual soils should incorporate the potential variability in geomechanical properties due to weathering into the design approach; and Geotechnical engineers must collaborate closely with mine planners to produce a workable and practical mine plan where great variations in geomechanical properties are observed at a property.

The anticipated modes of instability in the weathered rock in the southeast and southwest walls of the proposed pit (i.e. Domain 1) are predominantly kinematic (i.e. planar and wedge) failures. Based on the current structural geologic information from Domain 2, which indicate that fewer kinematic controls exist, the anticipated mode of potential instability along the northeast and northwest walls is rock mass failure. The rock mass failure analysis conducted indicated that the potential for rock mass or rotational failure in the weathered rock can be significantly reduced if the weathered rock can be partially dewatered.

At the Brisas del Cuyuni Project, the results of engineering geology and kinematic assessments of 10
Copyright 2000 by SME

SME Annual Meeting Feb. 28-Mar. 1, Salt Lake City, Utah the structural geology have been blended with rock mass stability analyses and soil mechanics analyses to develop design angles for proposed pit wall slopes in saprolites and weathered rock. The steepness of a significant portion of the lower pit wall, which is located in relatively strong, fresh rock, does not currently have any kinematic or rock mass constraints. However, additional structural geologic information is required at depth to confirm that kinematic controls are not a concern. Implementation of the design angles presented in this paper will require a concerted effort to dewater the saprolites and weathered rock. REFERENCES 1. Abel, J., 1997. Brisas del Cuyuni mine Recommended open pit slope angle design, report to Gold Reserve de Venezuela, March, 143 p. 2. Bieniawski, Z.T., 1976. Rock mass classification in rock engineering, Proceedings of the Symposium on Exploration for Rock Engineering, March, pp. 97-106. 3. Blight, G. E., editor, 1997. Mechanics of Residual Soils. Balkema, 237 p. 4. Burton, B.T., Earthworks with wet, fine grained tropical residual soils., Unpublished Master of Engineering Report, University of Alberta. 5. Deere, D.U., and Patton, F.D., 1971. Slope stability in residual soils, Proceedings of the 4th Panamerican Conference on Soil Mechanics and Foundation Engineering, San Juan, Puerto Rico, Vol. 1, pp. 87-170. 6. Fookes, P.G., editor, 1997. Tropical residual soils - A geological society engineering group working party revised report, The Geological Society, London, 184 p. 7. Hoek, E., 1994. Strength of rock and rock masses, extract from Support of Underground Excavations in Hard Rock, by E. Hoek, P. K. Kaiser and W.F. Bawden, A.A. Balkema Publishers, the Netherlands. 8. Hoek, E. and Bray, J., 1981. Rock slope engineering, The Institution of Mining and Metallurgy, London, Revised Third Edition, pp. 226-241. 9. Hydro-Triad., 1996. Brisas del Cuyuni, Hydrogeologic Report, December. 10. Millot, G., 1979. Clay, Scientific American, Vol. 240, pp. 109-118. 11. Mori, R.T., de Bare, F.R., and Pan, Y.F., 1979. Properties of some typical compacted th saprolites, Proceedings of the 6 Panamerican Conference on Soil Mechanics and Foundation Engineering, Lima, Peru, Vol. 2., pp. 583-591. 12. Newcomen, H.W., and Tape, R.T., 1997. Feasibility level open pit slope design for Brisas del Cuyuni Project, Las Claritas, Venezuela, September, (Volumes I and II), BGC Engineering Inc. 13. Newcomen, H.W., and Burton, .T., 1999. Brisas del Cuyuni Project, review of 250 million tonne pre-feasibility open pit, April, BGC Engineering Inc. 14. Newcomen, H.W., and Porter, M.J., 1999. Modern tools for assessing the stability of footwall slopes, Proceedings of Canadian Institute of Mining Workshop on Rock Mechanics and Ground Control in the Soft Rock and Coal Industries, Edmonton, Alberta, 26p.. 15. Sueoka, T., 1988. Identification and classification of granitic residual soils using chemical nd weathering index, Proceedings of the 2 International Conference on Geomechanics in Tropical Soils, Singapore, Vol. 1, pp. 55-62. 16. Tuncer, E.R., and Lohnes, R.A., 1977. An engineering classification for certain basaltderived lateritic soils, Engineering Geology, Vol. II-4, pp. 319-339. 17. Uehara, G., 1982. Soil science for the tropics, Engineering and Construction in Tropical and Residual Soils, A.S.C.E. Geotechnical Division Special Conference, Honolulu, Hawaii, pp. 30-57.

11

Copyright 2000 by SME

Potrebbero piacerti anche