You are on page 1of 171
CEP 702 FEB 1990 Retaining Wall Design Notes CEP 702 FEB 1990 Retaining Wall Design Notes Published by: Group Office PO Box 12343 Wellington. Tais document and its contents is the property of Opus International Consultants Limited, / ny unauthorised employment or reproduction, in full or part is forbidden © Opus International Consultants Limited 1997 PREFACE The aim of these notes on Retaining Wall design is to provide the designers within WORKS, guidance on design methods, bringing together the commonly used charts. The revisions in this edition generally incorporate up to date design methods, and include major revisions to the section on Earthquake Earth Pressures and Design. A brief new section on the design of sheet retaining structures is also included. The section on Earthquake Earth Pressures was prepared by Dr John Wood of Phillips and Wood Limited, Lower Hutt and has been incorporated into this document. This revised document was prepared by the Special Project Office of Works Consultancy Services, Wellington, with useful comments from other divisions of Works Consultancy Services and the Geotechnical Control Office in Hong Kong. Special Projects Office January 1990 CONTENT: RETAINING WALL DESIGN NOTES CONTENTS ‘S LIST OF FIGURES LIST OF ‘SYMBOLS SECTION 55 re we RE iG 14 TABLES. 1 - INTRODUCTION Scope Definitions and Symbols Design Principles 1.3.1 Free Standing Retaining Walls 1.3.2 Other Retaining Structures Load Cases 1.4.1 Basic Loadings 1.4.2 Other Considerations SECTION 2 - SOTL PROPERTIES Rie beloukunL General Selection and Use of Backfill Density Effective Stress and Pore Pressures Shearing Strength Base Friction Wall Friction Modulus of Elasticity and Poisson's Ratio Coefficient of Subgrade Reaction Swelling and Softening of Clays Permeability Liquefaction Page 10 15 15 16 Page SECTION 3 ~ STATIC EARTH PRESSURE 3.1 States of Stress 32 3.2 Amount and Type of Wall Movement 32 3.3 Limiting Equilibrium Conditions 4 3.3.1” The Rankine Earth Pressure Theory The Coulomb Earth Pressure Theory The Trial Wedge Method 3 Earth Pressure in Soils with Cohesion 3 Passive Earth Pressures 3 Geometrical Shape of the Retaining Structure 3 Limited Backfi11 3.4 Elastic Equilibrium Conditions 39 3.4.1 At-Rest Pressures 3.4.2 Over-Consolidation Pressures 3.4.3 Elastic Theory Methods 3.4.4 Compaction Induced Earth Pressures SECTION 4 - EARTHQUAKE EARTH PRESSURES AND DESIGN 4.1 Background 42 4.1.1 Wall Categories 4.1.2 Soil - Structure Interaction 4.1.3 Simplifications for Design 4.1.4 Plastic Theory and Failure Modes 4.1.5 Resonance Effects 4.2 Design Seismic Coefficients 45 4.3 Load Combinations 46 4.4 Factors of Safety 46 4.5 Dynamic Forces and. Pressure Distributions 46 4.5.1 Rigid Wall 4.5.2 Stiff Wall 4.5.3 Flexible Wall 4.5.4 Displaceable Wall 4.5.5 Forced Wall 4.5.6 Water Pressures 4.6 Applications to Various Types of Walls 53 4.6.1 Free Standing Walls Founded on Soils 4.6.2 Free Standing Walls Founded on Rock or Piles 4.6.3 Tied Back Walls 4.6.4 Basement Walls 4.6.5 Bridge Abutments SECTION 5 - EFFECT OF SURCHARGES 5.1 Uniform Surcharges 5.2 Line Loads 5.3 Point Loads 5.4 Non-uniform Surcharges SECTION 6 - EFFECTS OF WATER 6.1 General 6.2 Static Water Level 6.3 Seepage Pressure 6.4 Dynamic Water Pressure 6.5 Drainage Provisions 6.6 Filter Requirements 6.7 Geotextiles SECTION 7 - STABILITY OF RETAINING WALLS 7.1 General 7.2 Sliding Stability 7.2.1 Base Without a Key 7.2.2 Base With a Key 7.3 Overturning Stability 7.3.1 Gravity and Counterfort Walls 7.3.2 Sheet Walls 7.4 Foundation Bearing Pressures 7.4.1 Vertical Central Loads 2 Inclined Loads, Tilted Base and Sloping Ground 3 Miscellaneous Foundation Conditions 4 Foundations on Rock 7.4. 7.4, TA 7.5 Slope Failure in Surrounding Soi] SECTION 8 - STRUCTURAL DESIGN 8.1 General 8.1.1 Codes 8.1.2 Limit State Design 8.1.3 Cover to Reinforcement 8.1.4 Selection of Wall Type 8.2 Toe Design Page 59 59 60 60 61 61 62 62 63 64 64 66 67 68 69 74 75 76 Page 8.3 Stem Design 76 8.3.1 Stem Loading 8.3.2 Lower Section of Counterfort Wall Stem 8.3.3 Horizontal Moments in Counterfort Wall Stem 8.4 Heel Slab Design 7 8.4.1 Loading 8.4.2 Heel Slabs for Counterfort Walls 8.5 Counterfort Design 78 8.6 Key Design 78 8.7 Control of Cracking 79 SECTION 9 - SPECIAL PROVISIONS FOR CRIB WALLS General 80 Design Loading 80 Foundation Depth 80 Drainage 80 Multiple Depth Walls a1 Walls Curved in Plan 81 SECTION 10 - SHEET RETAINING STRUCTURES 10.1 General 82 10.2 Strutted Excavations 82 10.3 Anchored Flexible Walls 84 10.3.1 Walls Anchored Near the Top 10.3.2 Multiple Anchored Walls 10.3.3 Effects of Anchor Inclination 10.4 Cantilevered Walls 85 References 86 Figures 10 MW 12 13 14 15 16 17 18 19 20 ai 22 23 LIST_OF FIGURES Loading on typical retaining wall Rankine earth pressure, cohesionless soil, constant backfill slope Rankine active earth pressure coefficients, cohesionless soil, uni- form sloping backfill Rankine earth pressure, soil with cohesion, horizontal ground surface Coulomb earth pressure, cohesionless soil, constant backfill slope Coulomb active earth pressure coefficients, cohesionless soil, uniform sloping backfill o = 25° “ o . . o . @ = 30° . « " " ® " = 35° " " " " " " = 40° Coulomb failure plane for active pressure, cohesionless soil with uniform sloping backfi11 Trial wedge method, cohesionless soil, irregular ground surface Trial wedge method, cohesionless soil, Culmann's construction Trial wedge method, soil with cohesion, irregular ground surface Trial wedge method, layered soil and pore water pressures (A) Approximate method for direction of Rankine earth pressure. (B) Geometric conditions for Rankine and Coulomb methods. Point of appiication of active pressure Point of application of resultant pressure and pressure distribution Earth pressure coefficients, sloping ground Earth pressure coefficients, sloping wall Earth pressure due to compaction Seismic zone factor Stability analysis of retaining wall under earthquake loading Earthquake increment of pressure on rigid retaining wall 24 25 26 27 28 29 30 31 32 33 34 35 Equivalent wall height for estimating earthquake earth pressures for wall with sloping backfill Earthquake increment of earth pressure - stiff and flexible walls Mononobe constant Mononobe vertical Mononobe vertical (A) o= (8) o= Mononobe vertical (c) o (0) = Mononabe vertical (E) o= Mononobe vertical Mononobe vertical (A) o= (B) o= Mononobe vertical (C) = (0) o= Mononobe vertical (E) o= Mononobe - Okabe earthquake earth pressure - cohesionless soil, backfill slope - Okabe active earthquake earth pressure increments for walls : Horizontal Backfill : 6 = @ : $ = 20° - 45° - Okabe active earthquake earth pressure increments for walls : Sloping Backfill : 6 = @ 20° 25° ~ Okabe active earthquake earth pressure increments for walls : Sloping Backfill : 6 = > 30° 35° - Okabe active earthquake earth pressure increments for walls : Sloping Backfill : 6 = @ 40° - Okabe active earthquake earth pressure increments for walls : Horizontal Backfill : & = 26/3, @ = 20° - 45° - Okabe active earthquake earth pressure increments for walls : Sloping Backfill : 6 = 29/3 20° 25° - Okabe active earthquake earth pressure increments for wall : Sloping Backfill : § = 29/3 30° : 35° - Okabe active earthquake earth pressure increments for walls : Sloping Backfill : & = 26/3 40° = Okabe active earthquake earth pressure increments for fon-vertical walls : Wall Slope = -14° : Horizontal Backfill : 8 = 26/3 19 = 20° - 45° 36 37 38 39 40 41 42 43 4 45 46 47 48 49 50 51 52 Mononobe - Okabe active earthquake earth pressure increments for non-vertical walls : Wall Slope = -14°; Sloping Backfill; 6 = 26/3 (A) @ = 20° (8) @ = 25° Mononobe - Okabe active earthquake earth pressure increments for fon-vertical walls : Wall Slope = -14° : Sloping Backfill : & = 26/3 (C) @ = 30° (0) 9 = 35° Mononobe - Okabe active earthquake earth pressure increments for non-vertical walls; Wall Slope = -14°; Sloping Backfill; & = 26/3 (Cee Trial wedge method for earthquake force on wall Displaceable wall - Wall displacements from Sliding Block Theory Mononobe - Okabe passive earthquake earth pressure for vertical walls, horizontal backfill and zero wall friction (A) Earthquake passive earth pressure increment. (B) Total passive pressure. Forced wall - earth pressure distributions (A) Rotated wall (8) Translated wall Static and Earthquake Pressures on Walls with Submerged Backfill Failure surfaces for tied back wall under earthquake loading Earthquake and gravity forces on bridge abutments Uniform surcharge load cases Trial wedge active pressure including 1ine load Lateral pressure distribution on wall due to point and line loads Static water pressure, and seepage pressure for drained walls Drainage system for low walls and crib wall drainage Drainage details for backfilled walls Permeability of drainage materials 53 54 55 56 87 58 59 60 61 62 63 64 65 66 67 Stability criteria for retaining walls Sheet wall stability - cohesionless soils Bearing capacity factors (A) Loading arrangements for Brinch Hansen's bearing capacity formulae (B) Transformation of eccentrically loaded foundations Brinch Hansen bearing capacity formula - shape factors Brinch Hansen bearing capacity formula - depth factors Brinch Hansen bearing capacity formula - load inclination factors Design loading on heel slab Cut-off positions of main tensile steel in counterfort Crib wall design curves, normal loading, @ = 30°, 40° Crib wall design curves, earthquake loading, 6 = 30°, 40° Pressures on typical crib wall Braced Excavation Pressure Distributions Factor of Safety with Respect to Base Heave Large Excavations Settlement Guide 10. LIST OF TABLES Representative Values for Unit Weight of Soils Representative Values for the Angle of Shearing Resistance Typical Friction Angles and Adhesion Values for Bases Without Keys Indicative Proportions of Maximum Wall Friction Developed Modulus of Elasticity for Selected Soils (Undrained Compression) Typical Values for Poisson's Ratio Coefficient of Subgrade Reaction (Vertical) Permeabilities of Soils Movement of Wall Necessary to Produce Active Pressures Allowable Bearing Pressure on Jointed Rock g HyHy etc 10 SYMBOLS effective area of base base width effective base width seismic response function (for a 150 year return period) horizontal earthquake coefficient (period T) horizontal earthquake coefficient (period T 0) cohesion of soil in terms of total stress adhesion at base cohesion of soil in terms of effective stress foundation depth particle size at which 20% of the material is finer forward displacement of centre of mass due to an earthquake foundation depth correction factors modulus of elasticity of soil eccentricity of load on base factor of safety gravitational acceleration vertical height of plane on which earth pressure is calcu- lated (from underside of base or bottom of key to ground surface) horizontal component of loading on the base vertical height of wall depth below which active pressure exceeds compaction induced earth pressure equivalent height of fi11 to uniform surcharge loads piezometric head hydraulic gradient foundation load inclination factors 1 coefficient of earth pressure at rest coefficient of active earth pressure coefficient of active earthquake earth pressure (Kq+AKAg) coefficient of active earthquake earth pressure increment coefficient of passive earth pressure coefficient of subgrade reaction coefficient of permeability length of base effective length of base length of retaining wall stem length of failure surface normal reaction on a soil failure surface standard penetration test blow count for 300 mm penetration threshold acceleration factor stability factor bearing capacity factors slope stability number resultant force due to lateral pressure force due to active lateral earth pressure force due to active lateral earthquake earth pressure (ParaPae) force due to earthquake pressure component from inertia forces in soil force due to earthquake pressure component from forcing of the wall against the soil force on forced wall due to translation of wall force on forced wall due to rotation of wall horizontal component of force due to lateral earth pressure inertia load acting on abutment mass load from superstructure Pe Py p(z) Ap(z) Q OL % q 9a qq quit qu R RARAsRp»Ry sete RQD $ s Su 12 force due to at-rest earth pressure force due to passive earth pressure force due to net passive pressure force due to lateral earth pressure due to line or point surcharge load (per unit length of wall) force on wall due to gravity pressure component vertical component of force due to lateral earth pressure force due to hydrostatic water pressure increment in force due to active earth pressure due to an earthquake increment of earthquake force effective surface line load due to compaction plant active earth pressure consolidation pressure water pressure earth pressure at depth z below ground surface earthquake pressure increment at depth z below surface total load Vine load point load pressure on base or surcharge load intensity allowable soil bearing pressure flow rate through drain ultimate soil bearing pressure unconfined compressive strength risk factor (seismic design) resultant forces rock quality designation total shearing resistance at underside of base shearing strength of soil undrained shear strength of soil SesSqo8¥ ScL> ScB» SqLs SqB> SyL> SyB T u 20 2 aA AE y! Yd Ysat Yw 13 foundation shape correction factors foundation shape correction factors for inclined loading tangential force along a failure surface resultant force due to pore water pressures pore water pressure vertical component of resultant of loading on the base peak ground velocity (seismic) weight of soil wedge used in calculation of earth pressures weight of backfill over heel of wall weight of wall total weight of wall, soil above toe and soil above heel horizontal translational displacement of wall displacement of superstructure earthquake zone factor vertical depth of tension crack in cohesive soil critical depth for compaction pressure distribution depth below ground surface inclination of failure plane from the horizontal for active state (degrees) inclination of failure plane from the horizontal for active state under earthquake conditions inclination of failure plane from the horizontal for passive state (degrees) Slope of back of the wall (degrees, anti-clockwise positi- ve) bulk unit weight of soil (kN/m3) submerged unit weight of soil = Ysat - Yw dry unit weight of soil unit weight of saturated soil unit weight of water a ® ho 14 increment angle of wall friction (anti-clockwise positive) angle of base friction angle = tan-! c(0) rotation of forced wall rotation of base of wall Poisson's ratio settlement angle measured clockwise from vertical to direction of Pa total normal stress horizontal earth pressure effective normal stress effective vertical overburden pressure shear stress angle of shearing resistance in terms of total stress angle of shearing resistance in terms of effective stress angle of shearing resistance under undrained conditions angle of inclination of loading on base inclination of ground slope Section 1 INTRODUCTION La 1.2 1.3 1.3.1 SECTION 1 ‘SCOPE These notes are intended as a guide for use in the estimation of earth pressure forces and the design and construction of retaining walls and similar earth retaining structures. These notes are not intended to be used as a detailed text encompassing all aspects of retaining wall design. The intention is to provide brief notes and recommended methods covering most aspects of design. If a more detailed knowledge of a particular subject is required, the references given should Prove helpful. Reference is also made to standard texts for detailed methods such as the construction of flow nets for pore water pressure determination, and reinforced concrete design methods. Aspects such as the use of classical earth pressure equations, the effect of earthquakes on earth pressures, and allowable bearing pressures under inclined loads are covered in detail. Engineering judgement must always be used when applying the theories and methods given in these notes and strict notice must be taken of the limitations of the various assumptions. Special retaining systems such as reinforced earth and soil Nailing are beyond the scope of these notes, and reference should be made to specialist literature. DEFINITIONS AND SYMBOLS Throughout these notes, static earth pressure means the pressure exerted by the earth due to gravity forces. Earthquake earth pressure means the combined static and dynamic earth pressure which acts during or because of an earthquake. A list of symbols used, with their meanings, is included in the front of these notes. DESIGN PRINCIPLES Free Standing Retaining Walls In the design of free standing retaining walls, the following aspects need to be investigated: (a) the stability of the soil around the wall; (b) the stability of the retaining wall itself; and (c) the structural strength of the wall. For these walls it is usual to assume that some outward movement of the wall takes place so that the lateral earth pressure from 1.3.2 1.4 1.4.1 16 the retained soil is a minimum (active condition) for both static and earthquake loadings. However, the designer should check that the required movement can take place and that it does not affect the serviceability or appearance of the wall or cause damage to nearby structures or services. If the deformation that is required to reduce the earth pressure to the active case is not available due to the rigid nature of the structure or foundation, either the wall must be designed to withstand a higher pressure or some change made to the structure or foundation. If cohesive backfill is used, the large displacements necessary for the active condition means that the lateral earth pressure will almost always be higher than the active value. For the determination of earth pressures it is usual to consider only a unit length of the cross-section of the wall and retained soil. A unit length is also used in the structural design of cantilever walls and other walls with a uniform cross-section. Other Retaining Structures Where an earth retaining wall is part of a more extensive structure (eg, a basement wall in a building or an abutment wall of a portal structure) or is connected to another structure (eg, a bridge abutment connected to the superstructure) the wall is usually subject to static earth pressures which are greater than active since the structure does not allow full "yielding" of the soil. In these cases, the main structure generally provides the stability for the wall which then only needs to have adequate structural strength. The earth pressure on this type of structure under earthquake conditions depends on the movements of the structure and the forces exerted on the wall by the rest of the structure as well as the inertia forces from the soil. LOAD CASES Basic Loadings Two basic earth pressure loadings are considered for design. These are: (a) Normal loading —= Static effective earth pressure + water pressure + effective pressure due to live loads or surcharge. (b) Earthquake loading = Earthquake earth pressure + water pressure + surcharge (but not live Voads). However, earth retaining structures should be designed for not Jess than the pressure due to a fluid with a unit weight of 4 kN/m?. 1.4.2 wv For normal loading, static earth pressure and pressure due to surcharge should be derived in terms of effective stress unless short term load conditions exist, see section 2.5. For short term loading conditions earth pressures should be derived in terms of total stress. For many walls of lesser importance, earthquake loading need not be applied, see section 4. Other Considerations Consideration should also be given to the possible occurrence of other design cases or variations within the two design cases given above, caused by the construction sequence or future development of surrounding areas. For instance, additional Surcharges should be considered in calculating active pressures and allowance made for any possible future removal of ground in front of the wall. Usually the passive resistance of the material in front of the wall is ignored in the design of gravity or cantilever retaining walls. Section 2 SOIL PROPERTIES. 21 2.2 18 SECTION 2 - SOIL PROPERTIES GENERAL In advance of the design, tests should preferably be carried out on the proposed backfill material and natural ground behind and under an earth retaining structure. It is good practice to make further soil tests on the material exposed after excavation. For all walls higher than 6m, especially those with sloping backfill, the soil properties of the natural ground and backfill should be estimated from tests on samples of the materials involved. For less important walls, an estimation of the soil properties may be made from previous tests on similar materials. However, a careful visual examination of the material, particu- larly that at the proposed foundation level, should be made with the help of identification tests to ensure that the assumed material type is correct. SELECTION AND USE OF BACKFILL The ideal backfill is a free draining granular material of high shearing strength. However, the final choice of material should be based on the costs and availability balanced against the desired properties. In general the use of cohesive backfills is not recommended. Clays are subject to seasonal variations, swelling (see section 2.10), and deterioration which all lead to an increase in pressure on a wall. They are difficult to consolidate and long term settlement problems are considerably greater than with Cohesionless materials. For cohesive backfills, special attention must be paid to the provision of drainage to prevent the build-up of water pressure. The wall deflection required to produce the active state in Cohesive materials may be up to 10 times greater than that for cohesionless materials. This, together with the fact that the former generally have lower values of shearing strength, means that the amount of shearing strength mobilised for any given wall movement is considerably lower for cohesive materials than for cohesionless materials. The corresponding active earth pressure for a particular wall movement will therefore be higher if cohesive soil is used for backfill. In cases of a high seismic coefficient and for a steeply sloping backfill, the active earth pressure will be substantially reduced if the failure plane occurs in a material with a high angle of shearing resistance. In some circumstances it may be economical to replace the existing soil by a material with a high angle of shearing resistance so that the above situation occurs. However, also see section 3.3.6. 2.3 2.4 19 It is essential to specify and supervise the placing of backfill to ensure that its properties ($, c and y) agree with the design assumptions both for lateral earth pressure and dead weight calculations. DENSITY The density of soil depends on the specific gravity of the solid particles and the proportions of solids, air and water in the soil. The typical specific gravity of soil particles is about 2.65 for sand or rock and 2.70 for clays. However this may vary from area to area. The proportion of the total volume that is made up of this solid material is dependent on the degree of compaction or consolidation. An estimate of the density of backfill material to be used behind a retaining structure may be obtained from laboratory compaction tests on samples of the material. The density chosen must correspond to the compaction and moisture conditions that will apply in the actual situation. The density of natural soil should be obtained from undisturbed samples kept at the field moisture content, and volume. In earth pressure calculations, density must be in force units, ie, Unit weight (N/m? or kN/m?). For low, relatively unimportant walls, the unit weight of the soil behind the wall may be estimated from the typical values given in Table 1. In general the saturated unit weight should be used in calculations involving clay. EFFECTIVE STRESS AND PORE PRESSURE An effective stress may be considered to be the stress trans- mitted through the points of contact between the solid particles of the soil. It is this stress that determines the shearing resistance of the soil. The effective stress, o', at any point in a saturated soil mass may be obtained by subtracting the pressure transmitted by water in the voids, u (pore water pressure), from the total stress, 0, thus: An increased pore water pressure gives a reduced effective stress and therefore a reduced soil shearing resistance. This Jeads to an increased force against a wall in the active case. Conversely, an increase in the negative pore pressure (ie, a pore suction) gives an increased shearing resistance and reduces the force against a wall in the active case. 20 TABLE 1: REPRESENTATIVE VALUES FOR UNIT WEIGHT OF SOILS (Basic Data from Terzaghi and Peck (1967) and US Department of the Navy (1982)) Unit Weight Material Dry, Ya | Ysat (kN/m?) | (kN/m?) Clean gravel or rock ~ loose 16 - 17 | 19 - 20 dense, poorly graded 18 - 20 | 20 - 22 dense, well graded 20 - 21 Well graded, clean sands, gravelly sands - loose 14 - 16 19 dense 17 - 20 20 Poorly graded clean sand, sand-gravel mix - loose 16-17 20 dense 7-19 a1 Clayey sand - loose, poorly graded 4-17 dense, poorly graded 16.- 18 Fine silty sands and silt ~ loose 14 - 16 dense 17 - 19 Sand-silt clay mixed with slightly plastic fines VW - 20 Clayey gravel, poorly graded gravel-sand clay 18 - 20 Silty gravel, poorly graded gravel-sand silt 19 - 21 Glacial till - very mixed grained | 20 - 21 23 Glacial clay - soft 16 - 19 stiff 20 - 21 Organic clay ~ soft slightly organic 15 - 16 soft very organic 13-14 Pumice 10 - 12 2.5 a1 Positive pore water pressure results from a number of factors, the most important being static water pressure, seepage of groundwater or rainfall and seepage from other sources, such as burst or leaking water supply mains. In some soils, shock or vibration can cause transient increases in pore pressure. In low permeability soils, changes in pore water pressure can result when changes in total stress due to ground loading, dewatering or excavation are more rapid than the pore water can flow. These pore pressures dissipate with time, but may need to be considered in design. Pore water pressures due to static water pressure and seepage of water are covered in chapter 6. Negative pore pressures are present in many partially saturated soils as a result of capillary tension. Capillary tension, and hence soil suction, may be destroyed by surface infiltration or seepage, and in general its beneficial effect on the shear resistance of the soil should not be used in retaining wall design. SHEAR STRENGTH In all earth pressure problems the magnitude of earth pressure on a particular structure is a function of the shear strength of the soil. The shear strength is not a unique property of the material but depends upon the conditions to which the soil is subjected when it is sheared. Shear strength 1s also a function of effective stress and water content, and is dependent on volume change of the soil. This gives rise to two separate conditions which determine the type of analysis to be carried out. These are: (a) "Short term" (or undrained) which applies when the water content of the soil cannot change rapidly and hence at the end of construction excess pore water pressures have not dissipated. This commonly occurs in saturated soils with low permeability such as saturated clays. The shear strength of such a soil does not change when it is sheared quickly and the undrained shear strength may be used to calculate earth pressures. Analysis of "short term" stability may be carried out in terms of total stress, o, and strength parameter, Sy (undrained shear strength) with ¢y = 0. (b) “Long term" (or drained) conditions, apply when the water content of the soil can change quite rapidly with a conse- quent change in pore pressure and hence with a change in shear stress. This generally occurs in cohesionless soils with a high permeability or when after a long period of time excess pore water pressures have dissipated in soils with Tower permeability. In this case it is necessary for earth pressures to be calculated from shear strengths expressed in terms of effective stresses o', and strength parameters c' and 9’. 22 If analysis is carried out in terms of effective stresses the effect of any field pore water pressure must be included in the analysis. The shear strength of a soil is proportional to the normal or confining stress acting on the failure plane. The maximum shear stress that a sample of soil can sustain under different normal stresses should be obtained from laboratory tests or site investigation, see WORKS publication on Site Investigation (WORKS, 1982). Results of laboratory testing may be plotted to form a relationship between shear stress at failure and normal stress on the soil. This relationship forms an envelope which is commonly termed the strength envelope. The envelope is generally curved, particularly in the low stress range. But portions of the curve can be approximated by a straight line relationship as follows: S=c+otan@ (in terms of total stress) or s = c' +o! tan 9' (in terms of effective stress) Where c and are termed the strength parameters in terms of total stress and c' and $' are termed the effective strength parameters. For important walls it is desirable that the design strength parameters are determined for the range of stress, moisture content and density that is appropriate to the field situation. Guidance on determination of strength parameters is given in WORKS (1982) and Lambe and Whitman (1969). For less important walls the values given in Table 2 may be used. 23 TABLE 2: REPRESENTATIVE VALUES FOR THE ANGLE OF SHEARING RESISTANCE {Values obtained mainly from Terzaghi and Peck (1967)] (c = 0 in all the cases except clay where c = qu/2) (degrees) Material (degrees) | (saturated) Sandy gravel or rock filling 35-45 Sand - loose, round grains, uniform 28 dense, round grains, uniform 34 loose, angular grains, well graded 33 dense, angular grains, well graded 45 Silt and silty sand - loose 27-30 20-22 dense 30-35 25-30 Clayey sand 20-25 14-20 Clay, normally loaded or slightly preconsolidated| 22-30 0 2.6 BASE FRICTION Typical values of friction angle (8) and adhesion (cp) for calculating the shearing resistance between a concrete base and the foundation material are given in Table 3. These values may be used for low walls in the absence of specific test data. If a base key is used the failure plane will generally be through the foundation soil and therefore the shearing resistance is that of the soil (Sp = 6' and ch = c'). 24 TABLE 3: TYPICAL FRICTION ANGLES AND ADHESION VALUES FOR BASES WITHOUT KEYS [Values taken from US Department of the Navy (1982)] Friction Adhesion Interface Materials Angle (8) (kPa) Degrees Mass concrete on the following foundation material - Clean sound rock 35 Clean gravel, gravel-sand mixtures, coarse sand 29-31 Clean fine to medium sand, silty medium to coarse sand, silty or clayey gravel 24-29 Clean fine sand, silty or clayey fine to medium sand 19-24 Fine sandy silt, non-plastic silt 17-19 Very stiff and hard residual or preconsolidated clay 22-26 Medium stiff and stiff clay and silty clay 17-19 Formed concrete on the following foundation material - Clean gravel, gravel-sand mixtures, wel] graded rockfil) with spalls 22-26 Clean sand, silty sand-gravel mixture, single size hard rockfill 17-22 Silty sand, gravel or sand mixed with silt or clay 7 Fine sandy silt, non-plastic silt 4 Soft clay and clayey silt 10 to 35 Stiff and hard clay and clayey silt 35 to 60 27 WALL FRICTION The magnitude and direction of the wall friction developed depends on the relative movement between the wall and the soil. In the active case, the maximum value of wall friction develops only when the soil wedge moves significantly downwards relative to the rear face of the wall. In some cases, wall friction cannot develop. These include cases where the wall moves down with the soil, such as a gravity wall on a yielding foundation or a sheet pile wall with inclined anchors, and cases where the 25 failure surface forms away from the wall, such as in cantilever and counterfort walls, see section 3.3.6. The maximum values of wall friction may be taken as follows: Timber, steel, precast concrete, Smax. = ¢ Cast in-situ concrete, Snax. = ae" In general, the effect of wall friction is to reduce active pressure. The effect is small and often disregarded. An exception to this is when large vertical loads are applied to the top of the wall, an inclined anchor is stressed to an appreciable load or the wall is founded on compressible soil. In these cases the wall has to move down, and because the friction acts on the soil wedges in a downward direction, this increases the earth pressures acting on the wall. However, this can be ignored because limiting conditions are considered in the calculation of overall stability (Padfield and Mair, 1984). The effect of wall friction on passive pressures is large (see section 3). However, considerable structural movements may be necessary to mobilise maximum wall friction, for which the soil in the passive zone needs to move upwards relative to the structure. Generally, maximum wall friction is only mobilised where the wall tends to move downwards, for example, if a wall is founded on compressible soil, or for sheet piled walls with inclined tension members. Some guidance on the proportion of maximum wall friction which may develop in various cases is given in Table 4. 26 TABLE 4: INDICATIVE PROPORTIONS OF MAXIMUM WALL FRICTION DEVELOPED (Granular Soils - Passive Case) (Rowe & Peaker, 1965) Proportion of Maximum Wall Friction Developed Structure Type Loose Dense Gravity or free standing walls with horizontal movement. Sheet pile walls 0 0.5 bearing on hard stratum Sheet walls with freedom to move downwards under active forces or inclined anchor 1.0 1.0 loads Walls where passive soil may settle 0 0 under external loads Anchorage blocks, etc. which have freedom to move upwards on mobilisation of passive 0 0 pressure. Where a wall will be subjected to significant vibration, wall friction should not be included, if the soil is cohesionless. 2.8 MODULUS OF ELASTICITY AND POISSON'S RATIO The relationships between stress and strain in soils are important in the settlement of soil-supported foundations. They also determine the change in earth pressure due to the small movements of retaining walls or other earth supports. These relationships are complex since they depend on stress, strain, time, initial degree of saturation and various other factors. However, it is often convenient to express them in terms of modulus of elasticity and Poisson's ratio, since for small stress differences the soil behaviour closely approximates that for a perfectly elastic, homogeneous material. The modulus of elasticity of the soil, Es, is important problems where displacements are to be calculated. The value is usually determined from triaxial compression tests, but plate bearing tests may be used. Seismic methods may be used to check a larger mass of material. These methods give higher values of Es than those obtained from static testing, especially in 27 Jointed rock and results are not directly applicable to problems of static load problems. For all soils the elastic modulus increases with increasing consolidation pressure, pc. For loose sand, Es approximately equals 100 pc. A range of values for the modulus of elasticity in compression for common selected soils is given in Table 5. Poisson's ratio, v, is very important in stress oriented Problems (eg, stresses on retaining walls for no wall movement) Since it controis the relationship between orthogonal stresses. It may be determined from triaxial tests; however like the elastic modulus, it is dependent on the confining pressure and rate of loading amongst other factors. For granular or normally consolidated materials, v may be estimated from the relation- ships for at-rest pressure coefficients, see section 3.4. Representative values are given in Table 6. TABLE 5: MODULUS OF ELASTICITY FOR SELECTED SOILS (UNDRAINED COMPRESSION) [Values Taken from Bowles (1982)] Soil Es (WPa) Very soft clay 2-15 Soft clay 5 - 25 Nedium clay 15 - 50 Hard clay 50 - 100 Sandy clay 25 - 250 Silty sand 7 - 20 Silt 2 - 20 Loose sand 10 - 25 Dense sand 50 - 80 Loose sand and gravel 50 - 145 Dense sand and gravel 95 - 190 Loess 15 - 60 Sandstone 6,900 - 20,600 Limestone 13,800 - 41,300 Basalt 48,200 - 89,500 29 28 TABLE 6: TYPICAL VALUES FOR POISSON'S RATIO [Values Taken from Bowles (1982)] Soft v Clay, saturated 0.4-0.5 Clay, unsaturated 0.1-0.3 Sandy clay 0.2-0.3 silt 0.3-0.35 Sand dense 0.2-0.4 loose to medium dense (void ratio 0.4 - 0.7) coarse 0.15 fine 0.25 Rock 0.1-0.4 Loess 0.1 - 0.3 COEFFICIENT OF SUBGRADE REACTION In the design of footings and wall foundations, the simplified concept of subgrade reaction 1s often used to determine founda- tion settlement. This concept is based on the assumption that the settlement, p, of any element of a loaded area is entirely independent of the load on the adjoining elements. It is further assumed that the ratio ks 3 between the foundation pressure q on the element and the corresponding settlement p is a constant, Ks. The foundation pressure, q, is called the subgrade reaction and the coefficient Ks is known as the coefficient of subgrade reaction. Representative values of Ks for foundation design are given in Table 7. Allowance should be made for errors in this approximation by applying a suitable factor of safety, see Terzaghi and Peck (1967). 2.10 29 TABLE 7: COEFFICIENT OF SUBGRADE REACTION (VERTICAL) kK, Soil Type (kPa/mm) Dense gravel and gravelly soils (no clay fines) >80 Dense sand and sandy soils including clayey sand, clayey gravel 55-80 Silts, clays of low compressibility | 25-55 Clays of high compressibility 15-25 Note: For clays Ks may be assumed to vary linearly with the unconfined compressive strength qu, from 8 kPa/mm for qy of 100 kPa to 90 kPa/mn for qu of 380 kPa. SWELLING AND SOFTENING OF CLAYS Some clays, particularly those with high plasticity (plasticity index exceeding 20) tend to expand in the presence of water and if restrained by a structure can develop very high earth pressures exceeding 500 kPa. These pressures are not related to soil strength, but to the mineralogy and moisture content of the clay. Swelling pressures can be estimated from laboratory swell tests, but at .present such predictions may not be reliable. These pressures usually only develop in the zone of weathering at a depth of up to 1 to 1.5 m below ground level. The above pressures should be considered if cohesive soil is to be used behind 'non-yielding' walls, but need not be allowed for in the case of free standing walls where a small yield can be tolerated. When a natural deposit of clay or silt is disturbed by an excavation for a retaining wall the change in stress conditions and water content may lead to a change in shearing strength with time. With stiff fissured clays it has been shown that progressive softening can reduce the shearing strength to a small fraction of its original value. This is usually due to water percolating into the fissures which are open at the time of excavation for the wall resulting in a reduction in the value of cohesion c'. Earth pressures should therefore be calculated using a ‘softened’ strength to allow for this deterioration. For details. see Chandler and Skempton (1974), and Cullen and Donald (1971). 30 In fissured clays and clay fill the rate of softening is reduced by adequate drainage and when the wall is prevented from yielding progressively. However the latter requirement will mean that lateral earth pressures higher than active will result. au PERMEABILITY The permeabilities of soils in broad terms are given in Table 8. The effects of seepage pressures and permeability of the backfill material is detailed in section 6. TABLE 8: PERMEABILITIES OF SOILS [Values Taken from Terzaghi and Peck (1967)] Coefficient of Permeability Soil Type (m/s) Clean gravel 10-7 - 1 Clean sands, clean sand and gravel 10-5 = 107? mixtures etc Very fine sands, organic and inorganic silts, mixture of sandy silt and clay, 10-* - 10-5 glacial till, stratified clay deposits, Homogeneous clays below zone of weathering 10-14 - 10-* 2.12 LIQUEFACTION Liquefaction is the process which causes saturated cohesionless soils to lose strength or stiffness during earthquake ground motion. The process is associated with densification of soil grains, with a corresponding build-up in pore water pressure and hence a reduction in effective stress and shear strength. Liquefaction of saturated backfill material and/or foundation soils has been responsible for a large number of documented wall failures (e.g. Quay wall failure in Puerto Montt during 1960 Chilean earthquake). The liquefaction potential of a wall site and the backfill material used behind the wall must be considered. In the extreme case liquefaction may lead to an increase in lateral soi] pressures acting on the wall or a decrease in those resisting failure. 31 The potential for liquefaction in the backfill material is reduced by providing drainage and using a free draining gravel sized material which is well compacted in place. Where site conditions indicate that liquefaction is possible, steps should be taken to prevent this occurring since it is not likely to be feasible to design a retaining wall for this condition. Liquefaction potential may be reduced on site by either providing drainage or through densification of the deposit. Removal and replacement of localised deposits may also be considered. For walls of low importance the following method of assessing liquefaction potential is recommended. Saturated sandy soil layers which are within 9 m of the ground surface, have a standard penetration test N-value less than 10, have a coefficient of uniformity less than 6 and also have a Dgq-value between 0.04 mm and 0.5 mm, have a high potential for liquefaction during earthquakes. Saturated sandy soil layers which have a Dg9-value between 0.004 mm and 0.04 mm or between 0.5 mm and 1.2mm have a potential for liquefaction during earthquakes. Very soft/loose and sensitive silts can also liquefy. Soils outside this range of sizes or layers deeper than 15 m are less likely to liquefy. A more rigorous analysis is recommended for important retaining walls, see National Research Council (1985). Section 3 STATIC EARTH PRESSURE 3.1 3.2 32 SECTION 3 - STATIC EARTH PRESSURE STATES OF STRESS The stresses at any point within a soil mass may be represented on the Mohr coordinate system in terms of shear stress, t, and effective normal stress, o', see Scott (1963), Lambe and Whitman (1978) or Henry (1986), for the plotting of stresses and use of the system). “With this system, the shearing strength of the soil at various effective normal stresses gives an envelope of the possible combinations of shear and normal stress. When the maximum shearing strength is fully mobilised along a surface within a soil mass, a failure condition known as a state of plastic (or limiting) equilibrium is reached. Where the combinations of shear and normal stress within a soil mass all lie below the limiting envelope, the soil can be considered to be in a state of elastic equilibrium, see Terzaghi and Peck (1967). A special condition of elastic equilibrium is the 'at-rest' state, where the soil is prevented from expanding or compressing laterally with changes in the vertical stress. Any lateral strain in the soil alters its horizontal stress condition. Depending on the strain involved, the final horizontal stress can lie anywhere between two limiting (failure) conditions, known as the active and passive failure states. AMOUNT AND TYPE OF WALL MOVEMENT The earth pressure which acts on an earth retaining structure is strongly dependent on the lateral deformations which occur in the soil. Hence, unless the deformation conditions can be estimated with reasonable accuracy, rational prediction of the magnitude and distribution of earth pressure in the structure is not possible. For no movement of a retaining wall system, at-rest earth pressures (or pressures due to compaction) act on the wall. When a wall moves outward, the shearing strength of the retained Soil resists the corresponding outward movement of the soi] and reduces the earth pressures on the wall. The earth pressure calculated for the active state is the absolute minimum value. When the wall movement is towards the retained soil the shearing strength of the soil resists the corresponding soil movement and ‘increases the earth pressure on the wall. The earth pressure (or resistance) calculated for the passive state is the maximum value that can be developed. 33 TABLE 9: MOVEMENT OF WALL NECESSARY TO PRODUCE ACTIVE PRESSURES Soil Wall Yield Cohesionless, dense 0.001 H Cohesionless, loose 0.001 - 0.002 H Clay, firm 0.01 - 0.02 H Clay, soft 0.02 - 0.05 H where H is the height of the wall The amount of movement required to produce the active or passive states in the soil is dependent mainly on the type of backfill material. Table 9 gives the outward movement of a wall which is necessary to produce an active state of stress in the retained soil. The movements required to produce full passive resistance are considerably larger, especially in cohesionless material with movements of 0.05H to 0.1H being indicated (Wu, 1975), where H is the height of the wall. These requirements apply whether the movement is a lateral translation of the whole wall or a rotation about the base. The pressure distributions for full active and passive states are basically triangular for constantly sloping ground, see section 3.3. If a wall rotates about its top in the direction away from the soil, the soil between the wall and the surface of sliding does not all pass into the active state. The soil near the top of the wall stays near the at-rest state. This condition arises in cuts that are braced as excavation proceeds downwards from the top. The distribution of pressure may be represented by a trapezium with dimensions which vary according to the soil type. The amount of wall movement which will take place depends mainly upon the foundation conditions and the flexibility of the wall. The designer must ensure that the calculated earth pressures correspond to the available wall movement. A free standing wall need only be designed for active earth pressure as far as stability is concerned since, if it starts to slide or overturn under higher pressures, the movement will be sufficient to reduce the pressures to active. However, if it is on a strong foundation or otherwise fixed so that adequate stability ts provided, the stem may be subject to pressures near those for the at-rest state. The following pressure coefficients should be used for rigid foundation conditions unless a more exact analysis of movements is mad (a) Counterfort or gravity type walls founded on rock or piles Ko 3.3 3.3.1 34 (b) Cantilever walls less than 5 m high founded on rock or piles 0.5(Ko#Ka) (c) Any wall on soil foundations or cantilever walls higher than 5 m Ka In situations where soil-structure interaction is significant (eg bridge abutments) a rigorous analysis of earth pressure should be carried out by an experienced engineer. Broms and Ingieson (1971), and Mathewson et al (1980), describe some solutions for static earth pressures at bridge abutments. The tilting movement that will result when earth pressures act on a retaining wall may be estimated by simulating the founda- tion soil as a series of springs with an appropriate coefficient of subgrade reaction, Ks, see section 2.9. The base rotation, @ (in radians), is then given by: @p = 12Ve/KLB3 fore s Where Vis the vertical component of resultant of loading on the base eis the eccentricity of the load on the base L,B are the length and breadth of the base respectively LINITING EQUILIBRIUM CONDITIONS The Rankine Earth Pressure Theory Rankine's equations give the earth pressure on a smooth vertical plane, which is sometimes called the virtual back of the wall. The earth pressure on the vertical plane acts in a direction parallel to the ground surface and is directly proportional to the vertical distance below the ground surface (ie, a triangular Pressure distribution with the resultant acting at 1/3 H). Rankine's method is theoretically only applicable to retaining walls when the wall does not interfere with the formation of any part of the failure wedges that form on either side of the vertical plane as shown in Figure 2, or where an imposed boundary produces the conditions of stress that would exist in the uninterrupted sot! wedges. The method assumes that the earth pressure acts parallel to the backfill slope, w. Rankine's equations for earth pressures for cohesionless soils are given in Figures 2 and 3 and for cohesive soils in Figure 4. Where there is submerged backfill behind the wall or the possibility of build up of groundwater level, then the hydrostatic water pressures should be added to the earth pressures derived from Rankine's equations, see section 6.2 for details. Care should be taken in assessing the earth pressure coefficient in cohesive soils, see section 3.3.4, 3.3.2 35 Rankine‘s active earth pressure coefficients are presented graphically in Figure 3. Passive earth pressure calculations using Rankine are not recommended, since the direction of wall friction will be incorrect and an underestimation of passive resistance will result. Coulomb Earth Pressure Theory Coulomb theory assumes that a wedge of soil bounded by a planar failure surface slides on the back of the wall. Hence shearing resistance is mobilised on both the back of the wall and the failure surface. The resultant pressure can be calculated directly for a range of wall frictions, slopes of wall and backfill slopes. Where the wall friction is at angles other than the backfill slope angle the equations given in Figure 5 are an approximation due to the curved nature of the actual failure surface and the fact that static equilibrium is not always satisfied. The error is only slightly on the unsafe side for the active case, but is more unsafe for the passive case when § > 9/3. Hence the charts for Coulomb active pressure coefficients given in Figures 6 to 9 may be used at all times, but the passive pressures from Figure 5 should only be used when & < $/3 or when wall friction is ignored (conservative). In the active case the soil can slip down along the back of the wall, causing the resultant earth pressure to be inclined at a Positive angle, 6, to the normal of the wall, see Figure 5. It is recommended that an angle of wall friction, 8, of 42/3 9 be used in the equation for active pressure for concrete walls which have been cast against formwork. However, wall friction may not always result in an increase in wall stability, see section 2.7 for a detailed discussion of wall friction. The inclination of the failure surface assumed by the Coulomb theory is given by the charts in Figure 10. Coulomb wedge theory may be used for either the long or short term conditions of drained or undrained soil respectively (see section 2.5). For drained soils the effective strength para- meter $', should be substituted for the total strength parameter 4, in the Coulomb equations given in Figure 5. For undrained soils analysis in terms of total stress, the use of saturated unit weight Ysat allows for any pore water pressures. Hence pore water pressure need not be considered separately. Linear interpolation may be used to find the earth pressure coefficient or failure plane angle for intermediate values of $. 3.3.3 3.3.4 36 The Trial Wedge Method Where the ground surface is irregular or where it is constantly sloping in cohesive soil, a graphical procedure using the assumption of planar failure surfaces is the simplest approach. This procedure is known as the trial wedge method, see Figures 11 to 14. The backfill is divided into wedges by selecting planes through the heel of the wall. The forces acting on each of these wedges are combined in a force polygon so that the magnitude of the resultant earth pressure can be obtained. A force polygon is constructed even though the forces acting on the wedge are often not in moment equilibrium. | This method is therefore an approximation with the same assumptions as the equations for Coulomb's conditions and, for a ground surface with a constant Slope, will give the same result. If the conditions are the same as those for Rankine's equations, the trial wedge earth Pressures will correspond to these also. The limitations on wall friction and passive pressures mentioned in the use of the Rankine and Coulomb equations also apply to the trial wedge method. The adhesion of the soil to the back of the wall in Cohesive soils is neglected since it increases the tension crack depth and hence reduces the active pressure. For the active case the maximum value of the earth pressure calculated for the various wedges is required. This is obtained by interpolating between the calculated values. For the passive case the required minimum value is similarly obtained. The direction of the resultant earth pressure in the force Polygons should be obtained from the considerations in sections 3.3.1 to 3.3.3. For the cases where this force is assumed to act parallel to the ground surface, a substitute constant slope should be used where necessary, as shown in diagram (A) of Figure 15 for soil both with and without cohesion. For cohesionless material, Culmann's graphical construction shown on Figure 12 provides a compact method of plotting the resultant earth pressures for the various wedges and obtaining the maximum value with the corresponding failure plane. For an irregular ground surface the pressure distribution {s not triangular. However, if the ground does not depart Significantly from a plane surface, a linear pressure distribution may be assumed, and the construction given in Figure 16 used. A more accurate method is given in Figure 17. The latter should be used when there are abrupt changes in the ground surface or there are non-uniform surcharges. Earth Pressure in Soils with Cohesion In cohesive soils tension cracks can exist to a depth zp (see Figure 4) where,

You might also like