You are on page 1of 308
SUBJECTS TOPICS AUDIENCE BACKGROUND: OBJECTIVE APPROACH RESULTS REP Oe Sw WN Mes Ge ‘Overhead structures and foundations / Overhead transmission Soils ‘Transmission towers Testing ‘Transmission tines Foundations Design ‘Transmission managers and engineers Manual on Estimating Soil Properties for Foundation Design This manual provides foundation engineers with a comprehensive reference on estimating engineering soil parameters from field or laboratory test data. Empirical correlations are used extensively to evaluate soil parameters. The manual describes the most impor- tant of these correlations completely and systematically with an emphasis on the correlations of relatively common tests, including those that are seeing increased usage in practice. The analysis of all geotechnical problems, such as transmission structure foundation design, requires the adoption of a soil behavioral model that must include all relevant soil properties. These soil properties are not known in advance and require the design engineer to either measure or ‘estimate properties using correlations. However, the source, extent, and limitations of correlations are most often obscured in the presentation of the relationships. When plotted, most correlations are presented as a sim- ple line, but in reality they may be based on a veritable shotgun blast of data points. ‘To present a readily usable, comprehensive set of correlations for estimat- ing soil properties with each correlation presented in the context of its historical evolution and statistical variability; to update existing correlations with new data when possible. The researchers established a context for basic soil characterization, in- cluding simple soil descriptions, classification, unit weight, relative density, and consistency. Next, they developed correlations for in situ state of stress, strength, elastic behavior, time-dependent deformability, and permeability— both for common tests and for newer tests coming into increasing use. This work is a collection of correlations that organize a huge body of dis- persed knowledge into a coherent framework. Comprehensive correlations are given for basic soil characterizations, in situ stress state tests, strength tests, tests of elastic and time-dependent deformability, permeability tess, and liquefaction resistance tests. Each correlation is constructed from its beginnings in the literature. Some correlations are original amalgams of EPRI EL-68008 lecvic Power Research Insite EPRI PERSPECTIVE PROJECT several different presentations, and several correlations are consider- ably enhanced by the addition of new data. Further, many new correla- tions were developed when sufficient data were available. All of the presentations give the foundation designer an immediate feel for the variability of each relationship. This manual is intended to make the job of the transmission structure foundation designer easier. A second application is to aid in the devel- ‘opment of local scil property correlations specific to particular utility service areas. This use of the soil properties manual will tie in directly with the use of the TLWorkstation™ foundation task modules, CUFAD and MFAD (EPRI report EL-6420, volumes 16 and 17), and the recently released CUFAD+ EPRIGEMS module (report EL-6583-CCML). Finally, the manual can serve to alert the design engineer, who previously had only standard penetration test data on which to base soil characteriza- tions, that several other in situ tests are vastly superior predictors of soil properties. The engineer is thus presented with the data to make a cost-benefits analysis of the worth of better data on which to base de- sign. For other EPRI work on soil properties and foundation design see EPRI reports EL-2870 and EL-6420, volume 2. RP 1493.6 EPRI Project Manager: Vito Longo Electrical Systems Division Contractor: Cornell University For further information on EPRI research programs, call EPRI Technical Information Specialists (415) 855-2411 Manual on Estimating Soil Properties for Foundation Design EL-6800 Research Project 1493-6 Final Report, August 1990 Prepared by CORNELL UNIVERSITY Geotechnical Engineering Group Holster Hel! Ithaca, New York 14853-3501 Authors FH. Kulhawy BW. Mayne Principal Investigator FH, Kulhawy Prepared for Electric Power Research Institute 3412 Hillview Avenue Palo Alto, California 94304 EPRI Project Manager V. J. Longo Overhead Transmission Lines Program Electrical Systems Division ORDERING INFORMATION Requests for copies o this report should be directed to Research Reports Center (PRC), Box $0490, Palo Alto, CA 94303, (415) 965-4081, There is no charge for reports Fequested by EPRI member ules and afflates, US. utlty associations, US. government agencies (federal, state, and loca), media, and foreign organizations witn which EPRI has an information exchange agreement. On requast, RAC will send a catalog of EPRI reports. Evecare Power Research insture are EPR! ave vgiteed sorvce math of lect Pone: Reseach nse. ne (Sopyint © 1990 Electie Power Recsten Ini ne A ight reser Nonice [Ps repo nar erepred by te ocanzaon) raed below 88 an account of otk sponsored bythe cis owe: Research insite. ne (EPR), Noter EPR. memons ot PR, he exgancalonl names fon, nev ey Berson acting on behalf sy of hem: (a) makes ary werany, expat er pe. wh respect te oe O oy Information, soparaus, method. or process closed ini repo that Such vee may st singe pratt (ened “ts: of essumes any kbies wi respec tothe uae oo for daages resting Hom he oe {ny information, apoaraus, meod, or aoaess scoeed nhs repo" Prepared by (Carel Unitersty "aca, New York ABSTRACT This manual focuses on the needs of engineers involved in the geotechnical design of foundations for transmission Line structures, It also will serve as a useful reference for other geotechnical problems. In all foundation design, it is neces- sary to know the pertinent parameters controlling the soil behavior. When it is not feasible to measure the necessary soil parameters directly, estimates will have to be made from other available data, such as the results of laboratory index tests and in-situ tests, Numerous correlations between these types of tests and the necessary soil parameters exist in the Literature, but they have not been synthe- sized previously into readily usable form in a collective work. This manual sunma- rizes the most pertinent of these available correlations for estimating soil paran- eters. In many cases, the existing correlations have been updated with new data, and new correlations have been developed where sufficient data have been avail- able. For each soil parameter, representative correlations commonly are presented in chronological order to illustrate the evolutionary development of the particular correlation. The emphasis is on relatively common laboratory and in-situ tests and correlations, including those tests that are seeing increased use in practice. ACKNOWLEDGMENTS ‘The authors appreciate the assistance of numerous people during the course of this study and want to acknowledge their contributions. C. H, Trautmann and H. E. Stewart of Cornell University provided many useful comments and suggestions as this report was compiled, and they provided detailed review comments on both the first and final drafts of this report. K. J. Stewart prepared the text, and A. Avcisoy drafted the figures Several colleagues graciously responded to a request for review and evaluation of aes the first draft of this report. These included: J. I. Adams, Consultant Campanella, University of British Columbia; A.M. DiGioia, Jr. and L. F. Rojas- Gonzalez, GAI Consultants; M. D. Grigoriu and S, Vidié, Cornell University; E. B Lavless, III, Potomac Electric Power Company; V. J. Longo, EPRI; J. K. Mitchell, University of California at Berkeley; H. $. Radhakrishna, Ontario Hydro; T. E. Rodgers, Jr., Virginia Power; J. W. Rustvold, Bonneville Power Administration; J. H. Schmertmann, Schmertmann ané Grapps; J. J. Wolf, Western Area Power Administra- tion; and C. P, Wroth, University of Oxford. The detailed reviewers’ conments were very helpful in preparing the final text of this report. ‘CONTENTS Section Page 1 INTRODUCTION AND BACKGROUND ry Soil Correlations - Lea Sofl and Test Variability 12 Soil Modeling 13 Scope of Manual 15 References 17 2 BASIC SOTL CHARACTERIZATION aad Simple Descriptions 21 Particle Size and Distribution 21 Index Parameters for Cohesive Soils 2-2 Index Parameters for Cohesionless Soils 2-3 characterization by Simple Field Tests 2-6 Color and Odor 2-6 Classification 2-8 General Classification and Identification Systems 2-8 Cone Penetration Test (GPT) Classifications 2-8 Piezocone Penetration Test (GPTU) Classifications 2-14 Dilatometer Test (DMT) Classifications 2-16 Unit Weight 217 Relative Density of Cohesionless Soils from In-Situ Test Correlations 2417 Standard Penetration Test (SPT) Correlations 27 Cone Penetration Test (CPT) Correlations 2-26 Dilatometer Test (DMT) Correlations 2-36 Consistency of Cohesive Soils from In-Situ Test Correlations 2-358 Standard Penetration Test (SPT) Correlations 2-35 Cone Penetration Test (CPT) Correlations 2-37 Relationship Between SPT N and CPT qc Values 2-36 References 2.39 Section 3 IN-SITU STRESS STATE Basic Definitions Reconstruction of Stress History Effective Preconsolidation Stress in Cohesive Soils Correlations with Index Paraneters Conments on Field Test Correlations Correlations with VST Strength Correlations with SPT N Value Correlations with CPT qc Value Correlations with CPTU Results Correlations with PMT Results Correlations with DMT Results Effective Preconsolidation Stress in Cohesionless Soils Overconsolidation Ratio for Cohesive Soils Correlations with Index Parameters Correlations with Laboratory Strength Correlations with VST Strength Correlations with SPT N Value Correlations with CPT and CPTU Results Correlations with DMT Results Overconsolidation Ratio in Cohesionless Soils Effective Horizontal Stress in Cohesive Soils Correlations with Index Parameters Direct Correlations with SBPMT and DMT Results Indirect Correlations with SPT, CPT, CPTU, and DMT Results Effective Horizontal Stress Direct Correlations with SBPMT and DMT Results Indirect Correlations wich SPT and CPT Results Gonbined DMI/CPT Approach for Ky of Sands Empirical Approach References ia Cohesionless Soils 4 STRENGTH. Basic Definitions Effective Stress Analysis Total Stress Analysis Relevance of Laboratory Strength Tests to Field Conditions Effective Stress Friction angle of Cohesionless Soils - General Evaluation Basis Page 34 31 3-6 37 3-8 3-8 3-8 3-10 3-10 313 3-13 3-14 3-14 3-15 3-16 3-17 3-17 3-18 3-18 3-19 3-20 3-21 3-24 3-26 3-29 3.29 3-30 3-32 3-33 3-35 el wl a1 beh 45 47 Section Typical Values Correlations with Index Parameters Influence of Strength Envelope Curvature Influence of Test Boundary Conditions Effective Stress Friction Angle of Cohesionless Soils Correlated with In-Situ Tests Correlations with SPT N Value correlations with GPT ge Value Correlations with PMT Results Correlations with DMT Results Effective Stress Friction Angle of Cohesive Soils Correlations with Critical Void Ratio Friction Angle Influence of Test Boundary Conditions Correlations with Residual Friction Angle Undrained Shear Strength of Cohesive Soils - General Evaluation Basis Correlations with Index Parameters for Undisturbed Clays Correlations with Index Parameters for Remolded Clays General Behavior Under Triaxial Compression Loading Influence of Overconsolidation Influence of Test Boundary Conditions Influence of Strain Rate During Testing Sumary of Factors Influencing the Undrained Strength Ratio Undrained Shear Strength of Cohesive Soils Correlated with In-Situ Tests Correlations with VST Results correlations with SPT N Value correlations with CPT qe Value Correlations with CPTU Results Correlations with PNT Results Correlations with DMT Results References 5 ELASTIC DEFORMABILITY Basic Definitions Poisson's Ratio Undrained Modulus of Cohesive Soils Typical Values Correlations with sy 412 4-13 413 ae 4-18 4-20 4-21 421 4-22 4.26 4-27 4-28 4-30 4-32 437 4.39 448 4-49 4-50 4-53 4-55 4-58 4-59 4-59 4-61 5-1 sa 5-3 56 5-6 Section Page Correlations with SPT, OPT, and PMT Results 5-10 Back-Figured from Full-Scale Load Tests 5-11 Estimation from Dynamic Keasurenents 5-12 Modulus for Cohesionless Soils 5-13 ‘Typical Values 5-14 Correlations with Strength 5-15 Correlations with SPT N Yalue 5-17 Young's Modulus 5-17 Pressuremeter Modulus 5-18 Dilatoneter Modulus 5-18 Correlations with CPT qc Value 5-19 Back-Figured from Full-Scale Load Tests 5-20 Estimation from Dynamic Measurements 5-20 Subgrade Reaction 5-23 References 5-24 6 ‘TIME-DEPENDENT DEFORMABILITY 6-1 Basic Definitions 6-1 Compression and Unload-Reload Indices for Cohesive Soils 6-3 Typical Values 63 Correlations with CPT qc Value 6-7 Constrained Modulus for Cohesive Soils 6-7 ‘Typical Values 6-7 Correlations with SPT N Value 6-7 Correlations with CPT Results 6-9 Correlations with DMT Results 6-9 Compression Index for Cohesionless Soils 6-10 Constrained Modulus for Cohesionless Soils 6-11 Typical Values on Correlations with SPT, CPT, and DMT Results 6-13 Coefficient of Consolidation 6-13 ‘Typical Values 6-13, Correlations with CP? ané DMT Results 6-13 Coefficient of Secondary Compression 6-15 References 6-17 7 PERMEABILITY 7 ‘Typical Values Te References 1b Section 8 LIQUEFACTION RESISTANCE cyclic Stress Ratio Correlations with SPT, CPT, and DMT Results References Appené A. STANDARD PENETRATION TEST Procedure Advantages and Disadvantages Sources of Error, Reliability, and Cost References B CONE PENETRATION TEST Procedure Advantages and Disadvantages Sources of Error, Reliability, and cost References © PRESSUREMETER TEST Procedure Advantages and Disadvantages Sources of Error, Reliability, and Cost References D —_DILATOMETER TEST Procedure Advantages and Disadvantages Sources of Error, Reliability, and cost References = VANE SHEAR TEST Procedure Advantages and Disadvantages Sources of Error, Reliability, and Cost References F COMPARISON OF IN-SITU TEST METHODS References CRITICAL STATE SOTL MECHANICS (CSSM) CONCEPT References Appendix Page GPT CALIBRATION CHAMBER DATA FOR SANDS HL Data Summary HL Chamber Boundary Influence Ha References 7 I UNIT CONVERSIONS ra J SUMMARY CORRELATION TABLES ga wad 2-13 2-14 2-19 2-20 ILLUSTRATIONS Connon Laboratory Strength Tests and Field Tests Importance of Proper Data Presentation Liquidity Index Variations Particle Roundness Definitions Generalized Curves for Estimating egax and eain Early Soil Classification by CPT Soil Classification by Mechanical Friction CPT Soil Classification by Fugro Electric Friction CPT Simplified Soil Classification by Fugro Electric Friction CPT Most Recent Soil Classification by Fugro Electric Friction CPT Soil Classification Based on ap and Bq Soil Classification Based on CPTU Data Unequal End Areas of Electric Friction Cone Determination of Soil Deseription and Unit Weight by DMT Effect of Overburden Stress and Dy on SPT N Value Relative Density-N-Stress Relationship Relative Density-N-Stress Relationship for Several Sands Donut and Safety Hanmer Comparisons Energy Ratio Variations Comparison of SPT Overburden Corrections Comparison of Recommended Cy Factors and Available Data from OC Sands Particle Size Effect on Blow Count for Sands Figure 2-21 2-22 2-23 2-26 3-2 3.3 364 Aging Effect on Blow Count for Sands Relative Density from CPT for Uncenented and Unaged Quartz Sands Relative Density from CPT, Including Soil Coupressibility Correlation Between D, and Dimensionless qe (Uncorrected for Boundary Effects) Calibration Chamber Data on NC Sande Calibration Chamber Data on 0G Sands Summary of Calibration Chamber Studies Correlation Between DMT Horizontal Stress Index and Relative Density for Normally Consolidated, Uncemented Sand Variation of qc/N with Grain Size for Electric and Mechanical Friction Cones Recommended Variation of qe/Ml with Grain Size for Fugro Electric Friction Cones Variation of qo/N with Fines Content Recommended Variation of qc/M¥ with Fines Content Stresses in Soil Stress Paths for Simple Stress listories Horizontal Stress Coefficient for NC Soils from Laboratory Tests Unload Coefficient for 0C Soils Reload Coefficient for 0C Soils Generalized dp - Liquidity Index - Sensitivity Relationships 4p Correlated with VST 5, Field Vane Coefficient versus PI ap Correlated with SPT N jp Correlated with CPT q Gp Correlated with CPTU op jp Correlated with CPTU dur Bp Correlated with CPTU aye, Correlated with SBPMT py, 2-25 2-28 2-28 2-37 2-38 2-38 3-5 35 37 39 3-10 3-11 Figure 3-19 3-20 3-21 3-22 3-23 3-26 3-25 3-26 3-40 wl 42 Correlated with SBPMT sy and Ty Bp Correlated with DMT Pp OCR Correlated with VST sy OCR Correlated with SPT W OR Correlated with CPTU ar OCR Correlated with DMT Kp OCR - Xp Relationships for Clays of Varied Geologic Origin Kone Correlated with Atterberg Limits Xo Correlated with PT and OoR Apparent Lack of Trend Between Koge ané PI for 135 Clay Soils Ko Correlated with OoR Ko Correlated with Undrained Strength Ratio Ky from SBEMT Correlated with OCR Ko Correlated with Kp Ko ftom SBPMT Correlated vith Kp Ko Correlated with SPIN Xo Correlated with OPTU gr Xo Correlated with CPTU au Comparison of Ky Values for London Clay at Brent Cross Kp Correlated with Kp in Sands Cone Factor versus Ky as a Function of Bee and D/B Estimation of Ky in Coastal Plain Sand from CPT Simplified qc - Ky - dee Relationships Ky Correlated with g¢ and Kp Tentative Correlation Between ajo, dc, and D; for NC and OC Sands Tested in Calibration Chambers Comparison of Ko Values at Stockholm Site General Coulonb-Mohr Failure Effective Stress Coulomb-Hohr Failure 3-34 3-35 4-2 4-15 4-16 47 4-18 as 4-20 a2. 4-22 4-23 4-04 4-25 4-26 4-27 4-28 4-29 4-30 Serengch Envelopes for a Range of Soil Types Friction Angle Definitions Total Stress Coulomb-Mohr Failure Relevance of Laboratory Strength Tests to Field Conditions Jee versus Relative Density Gee versus Relative Density and Unit Weight Nonlinear Failure Envelope Representation Dilatancy Angle Relationships Influence of Intermediate Principal Stress on Friction Angle N versus $cc N versus 3¢¢ and Overburden Stress de Versus Fee and Vertical Stress for NC, Uncemented, Quartz Sands Bee from CPT Data Simplified ae - Ko - dee Relationships Trend of 3¢¢ with Normalized ae PAT Data Representations Friction Angle Evaluation fron PAT Results Jey for NC Clays versus PI Jee Variation as a Function of Consolidation Stress for NC Clays Fpse versus Sp¢ for NO Clays Jee versus Bee for NC Clays ‘$e from Ring Shear Tests and Field Studies }r for Amay Soils sy(VST)/8yo versus PI for NC Clays sy/Pvo for NC Clay versus Liquidity Index General Relationship Between Sensitivity, Liquidity Index, and Effective Stress Reuolded Undrained Shear Strength versus LT Undisturbed Undrained Shear Strength versus LI 48 49 49 4-15 416 417 47 4-26 426 4-26 4-28 4-29 4-31 4-32 4.33 4-34 4-35 4-36 4-37 4h 442 4-43 4h 445, Comparison of Undrained Strength Ratio for NC Clays After Anisotropic and Isetropic Consolidation Undrained Strength Ratio versus Zee for NC Clays su/Bro versus OCR Normalized s4/Byo versus OCR Observed Trends Between Ap and OCR Relationship Between Ag and PI for NC Clays Undrained Strength Ratio as a Function of Test Type Comparison of Undrained Strength Ratios from PSC and CKgUC Tests Comparison of Undrained Strength Ratios from PSE and CKGUE Tests Effect of Loading Direction on ey Undrained Strength Ratio from DSS versus te Comparison of Undrained Strength Ratios from DSS and CKgUC Tests Comparison of Undrained Strength Ratios from DSS, CKGUC and CKGUE, and PSC and PSE Tests Comparison of Undrained Strength Ratios from CKUE and CIUC Tests Undrained Strength Ratios in Extension and Compression versus Plasticity Index Serain Rate Influence on sy Normalized Undrained Strength Ratios for Major Laboratory Shear Test Types Field VST Correction Factor Vane Undrained Strength Ratio versus Plasticity Index for NC, Young and Aged Clays Selected Relationships Between N and sy Apparent Decrease of N with Increasing Sensitivity Relationship Between sy and SPT N Value Reported Range of Ny Factors from CPT Data Effect of Pore Water Stress on Cone Tip Resistance Sy as a Function of Au in the CPTU BMT Results in Bartoon Clay Page 438 4-36 438 4-38 4.39 4-40 4-41 4-46 4-47 4-47 4-49 451 4-52 4-52 4-54 455 Figure 4-57 5-5 5-6 5-7 5-16 5-17 5-18 ss 6-1 6-2 6-3 6-4 6-5 6-6 Sy as a Function of Kp from the DMT Modulus Definitions Drained Poisson's Ratio Parameters for Granular Soils Drained Poisson's Ratio versus PI for Several 10 Soils Drained Poisson's Ratio versus OGR and Stress Level for Sydney Kaolin Normalized Undrained Modulus versus Stress Level and OCR Generalized Undrained Modulus Ratio versus OGR and PI Cam Clay Prediction of Undrained tial Tangent Modulus Ratio PNT Modulus of Clay versus N Value Undrained Modulus for Deep Foundations in Compression Undrained Modulus for (a) Drilled shafts in Compression and Uplift and (b) Spread Foundations in Uplife Shear Modulus versus Shear Strain for Sands Dynamic Shear Modulus versus N for Cohesive Soils Couparative Plot of Drained Modulus Correlations for Sand PAT Modulus of Sand versus N Value Trend Between Dilatometer Modulus and N in Piedmont Sandy Silts Variation of a with Dy for Sands in Calibration Chambers CPT @ Correlation for Ticino Sand Normalized Drained Modulus for (a) Drilled Shafts in Uplift and (b) Spread Foundations in Uplige Variation of Shear Modulus of Dry Sands with Void Ratio and Confining Stress Consolidation Behavior TMme-Sectlement Behavior Representative Cc Relationships for Cohesive Soils Coupression and Unload-Reload Indices versus PI Sensitivity-Conpression Index Relationships Compression Ratio versus Water Content 5-8 5-8 5-9 5 5-13 S14 5-23 6-2 6-3 6-4 6-5 6-6 6-6 Figure 6-1 6-12 6-13 Bl B2 B-3 General Relationship Between Modulus Nuber and Porosity for NC Soils Modulus Number for NO Clay SPT Constrained Modulus Coefficient f versus PI Constrained Modulus versus qr from CPTU for Clays Constrained Modulus from DMT Parameters Effect of Grain Size on Sand Coupressibility Effect of Dy on Sand Compressibility Modulus Nusber for NC Silts and Sands ey versus wy, Pore Water Stress Decay versus Piezocone Time Factor Pore Water Stress Decay versus Dilatoneter Time Factor Coefficient of Secondary Compression versus Water Content for NC Clays Coefficient of Permeability versus Particle Size Coefficient of Permeability versus Particle Size and Relative Density Vertical Coefficient of Permeability for Clay Permeability Anisotropy for Various Natural Clays Liquefaction Resistance Correlated with Modified SPT N Value Liquefaction Resistance Correlated Indirectly with CPT Results Liquefaction Resistance Correlated Dixectly with OPT Results Liquefaction Resistance Correlated with DMT Kp Standaré Split-Spoon Sampler SPT Safety Hanmer SPT Donut Hanmer Equipment Used to Perform the SPT Mechanical Cone Penetronetors Typical Designs of Electric Cone Penetroneters Electric GPT Data Acquisition System 6-10 6-10 6-12 Comparison of Begemann Mechanical and Fugro Electric Cones Correlation of q¢ Between Electric and Mechanical Cones Conmon Piezocone Geometries Measured Pore Water Stresses in CPTU Tests Menard Pressureneter Equipnent Typical Pressuremeter Test Curve Self-Boring Pressuremeters Examples of "Lift Off" Pressure Dilatoneter Test Equipment Vane Geonetries and Sizes Connon Vane Borers Qualitative Relationship Between Relative Test Cost and Accuracy Typical Soil Stress-Strain Behavior SSM Notation Calibration Chamber Data for Various Sands CPT Calibration Chamber Correction Factor 3-6 Table La 4-3 4h 45 a7 TABLES cacegories of Analytical Methods for Soil Modeling Soil Particle Size Identification Relative Density of Cohesionless Soils Approximate Plasticity and Dry Strength of Sofl by Simple Tests approxinate Cohesive Soil Strength by Sinple Tests Approsinate Cohesionless Soil Relative Density by Simple Tests Unified Sotl Classification Syste burmister Soil Identification Systen typteal Soil Unit Weights Relative Density of Sand versus 8 ser Correction Factors for Field Procedures SPT Correction Factors for Sand Variables Relative Density of Sand versus ge consistency of Clay versus N consistency Tadex of Clay versus N and de Representative Values of Bc Relative Values of Effective Stress Friction Angles for Cohesionless Soils N versus $c Relationships de versus tc Relative Values of Effective Stress Friction Angles for Normally Consolidated Cohesive Soils Classification of Sensitivity ‘Typical Ranges in Ag for Isotropically Consolidated Clays 2h 2-6 27 2-7 Table 48 5-1 5-2 5-3 5-4 5-5 5-6 5-7 61 Evaluation of Modified Cam Clay Exponent A Correction Factors for sy Compared with sy from CIUC Test Results Approximate s, versus N Relationship Typical Ranges of Drained Poisson's Ratio Typical Ranges of Undrained Modulus for Clay Typical Undrained Hyperbolic Modulus Paraneters Exponent M for Shear Modulus Typical Ranges of Drained Modulus for Sand Typical Drained Hyperbolic Modulus Parameters Values of Rigidity Index for Selected Cohesionless Soils Dogree of Compressibility Coupressibility Data for Six Sands in Calibration Chamber Tests Compilation of Cge/Cg for Natural Soils Coefficient of Permesbility Major Sources of Error in The Standard Penetration Test Major Sources of Error in The Core Penetration Test Major PMT and SBPMT Variables Major DMT Variables Major Sources of Error in the Vare Shear Test Assessment of In-Situ Tests Usefulness of In-Situ Tests in Common Soil Conditions Historical Use, Mobilization and Access Requirements, and Costs of In-Situ Tests Estimates of In-Situ Test Variability Calibration Chanber Data Base for Sands Boundary Conditions in Flexible-Wall Calibration Chambers Correlations for Cohesive Soils Correlations for Cohesionless Soils voit, 4-48, 4-50 4-56 57 5-13 5-15 5-16 5-16 6-4 F2 F3 PS. EvGLISH ae reo — mh a ma asm - 7. oc - cae che cm - cy Cock - cp: CR ci SWMBOLS LETTERS - UPPER CASE dilatometer test reading failure pore water stress paraneter surface area of cone sleeve cone area over which wpe acts cone area over which ug acts American Society for Testing and Materials foundation diameter or width; cone diameter; dilatometer test reading calibration chamber éianeter piezocone paraneter experinental constant in modified Cam clay sodel N correction for aging N correction for borehole dianeter N correction for energy ratio N correction for overburden stress N correction for overconsolidation N correction for particle size N correction for rod length N correction for sampling method compression index Qc correction for overburden stress swelling index uniformity coefficient = Dgo/D0 unload-reload index cr cre cae cre ox gue cK QE cov crt opru cr cst css Ds pss coefficient of secondary compression ~ Cg in terms of void ratio Cq in terms of vertical strain consolidated anisotropic undrained triaxial compression clay fraction consistency index consolidated isotropic drained triaxial compression consolidated isotropic undrained triaxial compression consolidated isotropic undrained triaxial extension Xo consolidated undrained triaxial compression Ko consolidated undrained triaxial extension coefficient of variation (standard deviation/nesn) cone penetration test piezocone test compression ratio = Go/(1 + 9) critical state Line critical state soil mechanics Gepth; vane diameter; cone diameter particle size at 5 percent finer particle size at 10 percent finer particle size at 50 percent finer particle size at 60 percent finer naximum particle size ninimum particle size relative density = (mex - €)/(emex - nin) dilatoneter test éirect shear direct simple shear Young’s modulus BOPT Ailacometer modulus pressuremeter modulus dzained modulus dcained secant modulus Young's sodulus of foundation initial tangent modulus secant modulus tangent modulus undrained modulus wndrained initial tangent modulus undrained secant modulus undrained tangent modulus electric GPT energy ratio percent passing No. 200 sieve Poisson's ratio parameter cone friction ratio = fe/a¢ (alternate for Rp) shear modulus = £/2(1 +») initial tangent ¢ aynanic shear modulus reload-unload specific gravity of solids; secant 6 unload-reload G Poisson's ratio parancter vane height; height of drainage path heavily overconsolidated density index; DMT material index relative dilatancy index nonent of inertia of foundat Neo Nk P OD 6 New Re oc. ocr ocr; OCR tims rigidity index reduced rigidicy index coefficient of horizontal soil stress = 3/éy; vane constant coefficient of minimum active soil stress DMT horizontal stress index coefficient of maximum passive soil stress in-situ or at-rest K = Ino/%vo normally consolidated Ko Ko during primary unloading calibration chamber correction factor subgrade reaction modulus; ratio of sy in extension to compression Liquidity index = (vy - vp)/(L - ¥p) Lightly overconsolidated critical state failure parameter; constrained modulus; earthquake magnitude; exponent in Gnax relationship drained constrained modulus érained secant constrained modulus érained tangent constrained modulus mechanical CPT standard penetration test value corrected N for field procedures cone bearing factor PAT bearing factor Ngo corrected to reference stress of one atnosphere piezocone factor normally consolidated overconsolidated overconsolidation ratio = Ip/Fyo isotropic OCR = Buax/Po Limiting OCR at which passive failure occurs Bsc PSE Qe Ww ocr To 1. vst maximum OCR plasticity index = wy - vp pressuremeter test plane strain compression plane strain extension sot] mineralogy and compressibility coefficient for strength dilatancy soil mineralogy and compressibility coefficient for cone dimensionless cone tip resistance = (a¢/Pa)/(yo/Pa) "> Qc correction for overconsolidation (equal to Cocr) particle roundnes: Fitting coefficient; ra equivalent radius failure ratio; cone friction ratio = fs/ac dynamic stiffness coefficient; particle sphericity sensitivity self-boring pressureneter test standard deviation stress level (fraction of strength mobilized) standard penetration test time factor; torque triaxial compression triaxial extension unconfined compression test unconsolidated-undrained triaxial cospression test volume vane shear test ENGLISH LETTERS - LOWER CASE ocr ‘RATE - cone area ratio; modified Cam clay parameter for anisotropic compression ~ sy correction for overconsolidation - sy correction for strain rate wav anst oh sy correction for test mode naximum horizontal acceleration at ground surface intermediate effective principal stress factor cohesion intercept effective stress cohesion intercept alternative form of sy coefficient of consolidation horizontal ¢y internal cone diameter; modified Cam clay parameter for plane strain compression void ratio (usually in-situ) maximum void ratio minimum void ratio initial void ratio unit side resistence; SPT nodulus coefficient cone side resistance normalized cone side resistance corrected cone side resist: gravitational acceleration coefficient of pernesbilicy horizontal k; alternative form for kg subgrade reaction modulus vertical k modulus number; OOR exponent reload coefficient coefficient of volumetric compressibility hyperbolic modulus exponent; cone exponent measurements or data points porosity = e/(1 +e); number of coefficient of subgrade reaction alternative form of By; applied stress Su(vst) t Ue effective mean normal stress = (51 + 32 + 35)/3; applied stress DMT expansion stress dilatoneter ¢ reading at a particular tine PAT Limit stress ataospheric pressure or stress (See Appendix I for munerical values.) alternative form of 3p BMT expansion stress PMT yield stress F at failure maximum 6 DMT contact stress; PMT total horizontal stress initial effective p shear stress DMT tip resistance corrected Ge cone tip resistance dc standardized to reference stress of one atnosphere unconfined compression strength = 2 sy critical state spacing ratio; sample correlation coefficient coefficient of determination slope of PAT plot of pe VS. €e undrained shear strength renolded sy Sy from VST time pore water stress behind cone tip measured pore water stress during cone penetration hydrostatic stress pore water stress behind cone sleeve ue “L vp pore water stress on cone tip/face in-situ, natural water content Liquid Limit plastic Limit depth DMT gage pressure deviation GREEK LETTERS - UPPER CASE av au bupe bug aug 01,3 volume change excess pore water stress excess pore water stress measured behind the cone tip ‘hu at time zero excess pore water stress measured on the cone tip or face uajor, minor principal stress increment corrections to Jey change in residual friction angie critical state paraneter GREEK LETTERS - LOWER CASE evr Pe Bo 7 7 va sat Trotal Ww OPT parameter relating modulus to qc; Ko unload coefficient empirical vane factor DMT Ko coefficient DMT OCR coefficient unit weight; shear strain effective unit weight ayy saturated 7 total 7 of water stress rotation angle; displacenent va vas ea Pauax. Penin °1,2,3 21,2,3 Cab,c Bae et strain strain rate axial strain cavity strain radial strain vertical or volumetric strain modulus nunber; critical state parameter for isotropic swelling index critical state parameter for isotropic compression index micron (10-6 meters); VST correction factor Poisson’s ratio drained v initial tangent drained v undrained v = 0.5 density in-situ dry density maximum pg minimum pg total normal stress effective normal stress major, intermediate, minor principal ¢ major, intermediate, minor principal 3 confining stress a, b, and c minor principal effective confining stress initial horizontal ¢ initial horizontal 7 isotropic overburden 7 current vertical } mean principal @ maximum vertical preconsolidation stress vertical ¢ ed vertical alternative form of 3p vertical (or overburden) ¢ vertical (or overburden) 7 shear strength; shear stress average cyclic stress shear stress in DSS friction angle effective stress friction angle fully softened, constant volume, or critical state 3 direct shear 3 peak $ plane strain compression $ residual relative secant 3 triaxial triaxial dilation 3 friction angle = Bee = 25°)/(45" - 25*) coupression § extension § angle Section 1 INTRODUCTION AND BACKGROUND ‘This manual has been prepared to assist foundation engineers in the selection of soil parameters, primarily for the geotechnical foundation design of transmission Line structures. It also will serve as a useful reference for other geotechnical problems. Soil is a complex engineering material, and its properties are not unique or constant, Instead, they vary with many environmental factors (e.g., time, stress history, water table fluctuation, etc.), as discussed in most geotech- nical reference books Because of the complexity of soil behavior, empirical correlations are used exten- sively in evaluating soil parameters. In this manual, an attempt has been made to summarize the most pertinent of these empirical laboratory and in-situ test corre- lations in an organized manner. The emphasis is on relatively common tests and several newer tests that are seeing increased use in practice. Within this section, the necessary background is presented to understand and appre- ciate the nature of soil correlations and modeling, and the scope of this manual is outlined. SOIL CORRELATIONS ‘The analysis of all geotechnical problems requires the adoption of a soil bebi vioral model, couplete with all relevant soil properties. These so{l properties are not known beforehand, and therefore the design engineer mist either measure the properties under controlled conditions in the laboratory or field or estimate the properties from other test data, These estimates are made most often from labora- tory index tests and in-situ test results, which are correlated to the soil proper- ties either by calibration studies or by back-caleulation from full-scale load test data obtained in the field. Comprehensive characterization of the soil at a particular site would require an elaborate and costly testing program, well beyond the scope of most project budg- ets. Instead, the design engineer wust rely upon more limited sofl information, and that is when correlations become most useful. However, caution mist always be exercised when using broad, generalized correlations of index parameters or in-situ test results with soil properties. The source, extent, and Limitations of each correlation should be examined carefully before use to ensure that extrapolation is not being done beyond the original boundary conditions. “Local” calibrations, where available, are to be preferred over the broad, generalized correlations Im addition, many of the common correlations in the literature have been developed from test data on relatively insensitive clays of low to moderate plasticity and on unaged quartz sands reconstituted in the laboratory. Extrapolation of these corre- lations to “special” soils, such as very soft clays, organic clays, sensitive clays, fissured clays, cemented soils, calcareous sands, micaceous sands, collap- sible soils, and frozen soils, should be done with particular care because the cor- relations do not apply strictly to these soil deposits. Careful exaination of the soil samples and reference to available geologic and soil survey maps should be made to detect the possible presence of these soils. The same special care should be exercised in remote areas and where no prior experience has been gained. Tf any "special" soils are present, or if no experience has been documented in a given area, a qualified geotechnical expert should be consulted for guidance. SOTL AND TEST VARIABILITY Soil is a complex engineering material which has been formed by a combination of various geologic, environmental, and chenical processes. Many of these processes are continuing and may be modifying the soil in-situ. Because of these natural processes, all soil properties in-sicu will vary vertically and horizontally. Even under the most controlled laboratory test conditions, soil properties will exhibit variability. This variability becomes more pronounced in the field where the natu- ral geologic environment is introduced. When empirical correlations are used, additional uncertainty is introduced, These levels of uncertainty must be con- sidered when assessing the reliability of a particular foundation design. Variability also may be introduced by the type of laboratory or in-situ test used. Each available test will provide a different test result because of differing boun- dary conditions and loading mechanisms. Figure 1-1 illustrates these variables for Some of the connon laboratory strength tests and field tests. For the laboratory strength tests, corrections are necessary to interrelate the particular test results because of the different boundary conditions. For the field tests, differ- ent in-situ responses are being measured in the different tests, as described in Appendices A through E, Each test has its own variability, and the relative merits 12 Laboratory Strength Tests eSSees sss cc = ese Field Tests SPT CPT PMT. OMT ST. SYMBOLS: TC - triaxial compression SPT - standard penetration test TE - triaxial extension CET - cone penetration test DS - direct shear PMT - pressuremeter test DSS - direct simple shear DMT - dilatometer test PSC - plane strain compression VST - vane shear test PSE - plane strain extension Figure 1-1. Common Laboratory Strength Tests and Field Tests of each test should be considered within the overall project context. provides 2 general comparison of these field test methods SOTL MODELING ‘Appendix F Wroth and Houlsby (1) have stated succinctly that correlations ideally should be (a) based on a physical appreciation of why the properties can be expected to be related, (b) set against a background of theory, and (c) expressed in terms of Gimensionless variables to allow sealing. These thoughts should always be kept in mind when using any type of correlation. It also must be remenbered how complex soil behavior really is. Ladd, described this complexity as follows. "A generalized model of the stress-strain behavior of soils should ideally account for nonlinearity, yielding, variable dilatancy (volune changes caused by shear stress), and ani- sotropy (both inherent and stress system induced), plus the behavioral dependence on stress path, stress system (orien- tation of 3 and relative magnitude of 07), and stress his- tory (both initial and changes due te consolidation)" 13 et al. (2) Table 1-1 sumarizes the major categories of analytical models that currently are available for representing the behavior of soils. These models range from rather complex (1) to advanced (II) to simple (ITI) descriptions of soil. Constitutive models for soil behavior require input in the form of soil properties and in-situ Parameters. In most cases for transmission line structure foundations, Category III models may be most appropriate at the present time. Jamiolkowski, et al. (3) also discuss the available laboratory and field tests in use for characterizing soil. Their discussion focuses on a wide range of soil behavior issues and might suggest that soll modeling is a most difficult task However, new efforts in research and development have resulted in considerable Progress in understanding soil behavior. The calibration and modification of soil models have been made possible by the back-analysis of performance data from full- scale field structures, such as deep foundations, enbankments, tunnels, offshore platforms, and high-rise buildings. As additional field performance data become available, newer ané more reliable correlations undoubtedly vill be developed. This progress in research ideally will allow foundation design to evolve from Cate- gory IIT in Table 1-1 to Categories II and then I, at which time all of the neces- sary soil behavior issues will be addressed. Table 1-1 CATEGORIES OF ANALYTICAL METHODS FOR SOIL MODELING Category Main Features of Models Determination of Soil Parameters 1 Very advanced models using non- Only from sophisticated laboratory Linear elastic-plastic time-depen- tests, with the exception of vari. éent lavs which possibly incorpo- ables which must be obtained from rate anisotropic behavior in-situ tests 0 Advanced models using constitu- Laboratory tests which are only a tive incremental elastic-plastic little more sophisticated than con- laws and nonlinear elastic rela- ventional tests; in-situ tests also tionships appropriate TIT Simple continuum, such es isotro- Conventional laboratory and in-situ pic elastic continuum, including tests layering and empirical models Source: Adapted from Janiolkowski, et al. G), p. 58 At the present time, there is one modeling concept of soil behavior which is of sone practical use for estimating soil properties. This concept is known as Criti- cal State Soil Mechanics (GSSM) and is described in Appendix G. With this concept, a general predictor for soil behavior has emerged. Strictly speaking, CSSM is applicable only to renolded, insensitive soils without aging, cementing, and other environmental influences. However, the resulting model predicts well the behavior of normally consolidated, insensitive soils, also without aging, cementing, and other environmental influences. In other soils, the model effectively provides a lower bound on the predicted property, such as the undrained shear strength. For these reasons, property prediction by CSSM has been included in this manual as a valuable reference on probable lower bound behavior of natural soils. SCOPE OF MANUAL In the following sections, commonly used correlations have been compiled that are helpful for estimating soil properties. Within a particular topic, these correla tions are selected and presented in an approxizate evolutionary order to represent the development of the relationship as newer research findings became available In certain instances, it was necessary to develop new correlations to supplement existing ones. Where new correlations have been developed, the complete data set and regression analysis results are presented to provide 2 measure of the validity of the relationship, The regression equation is presented first, normally using an ascuned intercept of zero for simplicity. The number of data points in the corre- lation is denoted by n, and the standard deviation (S.D.) is given to allow assess- ment of the dispersion around the regression line. Also given is the coefficient of determination, r2, which is the ratio of the explained variation to the total variation. For r2= 1, a perfect correlation exists; for r? = 0, no correlation exists; and for r2 = 0.75, 75 percent of the observed variation in y may be attrib- uted to x. In almost all cases presented, the value of r? for a zero intercept was only 1 or 2 percent less than the x? for a regression line with an intercept. The sample correlation coefficient, r, is the statistic for testing the significance of a simple two-variable linear relationship (i.e., how well the data fit a linear relationship). For x = 0, no linearity exists while, for r= +1, direct linearity By presenting the complete data set, the regression equation, and some pertinent statistics (n, $.D., x2), the user will be able to assess the quality of the rela- tionship and use the results accordingly. This format also will allow direct incorporation of the results into evolving reliability-based design procedures. Moroney (4) states rather directly in Figure 1-2 the importance of presenting the Ls IT IS DISHONEST TO PRESENT THIS FoR ~~ THIS Figure 1-2, Importance of Proper Data Presentation Source: Moroney (4), p. 29 data properly. Since this manual is directed toward the practicing engineer, its focus has been Limited to the more conmon tests available on 2 comercial basis and to those tests that are seeing increased use in practice. Included are the common laboratory index and performance tests and the field standard penetration test (SPT), cone penetration test (CPT), pressuremeter test (PNT), and vane shear test (VST). The newer tests included are the dilatometer test (DMT), piezocone or cone penetration test with pore water stress measurement (CPTU), and the self-boring pressuremeter test (SBPMT). Intentionally not included are the wide variety of simple hand devices which are intended primarily for field inspection purposes, such as the pocket penetrometer, torvane, geostick, dynamic cone, etc. These are not design or performance devices and should not be used as such. Also not included are scaled tests such as the plate load test or centrifuge test, which may be used to model full-scale foundation performance on a smaller scale. Section 2 addresses basic so{l characterization to define the soil material, while Section 3 focuses on evaluating the in-sita sofl stresses. The evaluation of soil ons 5 and 6 address elastic and tine- dependent soil deformability, respectively. Section 7 covers soil permeability, strength is covered in Section 4, while S while Section & briefly addresses the special topic of liquefaction resistance. Appendices A through F provide information on the various in-situ tests used in the correlations, primarily for those readers sho are not familiar with the tests. This information was extracted largely from EPRI Reports EL-2870 (5) and EL-5507, Vol. 2 (6) Appendix G gives a brief sumary of the Critical State Soil Mechanics concept, and ‘These reports should be consulted for further details on the tests. Appendix H summarizes available CPT calibration chamber data used to develop a nun- ber of correlations in this manual Within this manual, an effort has been made to present the relationships in dimen- sionless form for ease in sealing to whatever units are desired by the user. Therefore, stresses have been made dimensionless by the atmospheric pressure or stress, Pa which is equal to 1.058 tsf, 14.7 psi, 101.3 kN/m?, etc. A simple, approximate conversion for preliminary work is that 1 atm ~ 1 tsf = 1 kg/em? = 100 k/a2, These approximate conversions have been used liberally with previously published work where the 1 or 2 percent variation would not be significant. All unit weights have been made dimensionless by the unit weight of fresh water, yw, which is equal to 62.4 pef or 9.80 kN/m3. Where lengths are included, dual units are given. A detailed unit conversions guide is given as Appendix I. Lastly, Appendix J presents summary tables to assist the user in locating specific recomended correlations in this manual. These tables are not intended to be a substitute for the text, which puts the correlations in proper perspective. Instead, they are intended to be a quick reference guide for the experienced user. [REFERENCES 1. Wroth, ¢. P. and Houlsby, G. T., "Soil Mechanics - Property Characterization ad Analysis Procedures", Proceedings, llth International Conference on Soil Mechanics and Foundation Engineering, Vol. 1, San Francisco, 1985, pp. 1-55. 2, Ladd, C. C., Foott, R., Ishihara, K., Schlosser, F., and Poulos, H. 6., *Stress-Deformation and Strength Characteristics", Proceedings, 9th Interna- tional Conference on Soil Mechanics and Foundation Engineering, Vol. 2, Tokyo, 1977, pp. 421-494. 3. Jamiolkowski, M., Ladd, C. C., Germaine, J. T., and Lancellotta, R., "New Developments in Field and Laboratory Testing of Soils", Proceedings, 11th International Conference on Soil Mechanics and Foundation Engineering, Vol. 1, San Francisco, 1985, pp. 57-154 4, Moroney, M. J., Facts From Figures, 3ré Ed., Penguin Books, Balt: 472 p. more, 1956, 5. Kulhawy, F, H., Trautman, C. H., Beech, J. F., O'Rourke, T. D., McGuire, ¥., Wood, W. A., and Capano, C., "Transmission Line Structure Foundations for Uplife-Conpression Loading”, Report E1-2870, Electric Power Research Insti- tute, Palo Alto, 1983, 412 p. Orchant, C. J., Kulhawy, F. H., and Trautmann, C. H., "Reliability-Based Foun- dation Design for Transmission Line Structures: Critical Evaluation of In-Situ ‘Test Methods", Report EL-5507, Vol. 2, Electric Power Research Institute, Palo Alto, 1988, 216 p. 1-8 Section 2 BASIC SOIL CHARACTERIZATION One of the first steps in any geotechnical design problem is to develop an under- standing and knowledge of the soil materials 2 in > 305 Cobble 12 in to 3 in 305 to 76 Coarse gravel 3 in to 3/4 in 76 to 19 Fine gravel 3/4 in to Wo, 4 sieve 19 to 4.75 Coarse sand No. 4 to No. 10 sieve 4.75 to 2.0 Medium sand No. 10 to No. 40 sieve 2.0 to 0.42 Fine sand No. 40 to No. 200 sieve 0.42 to 0.075 Fine-Grained Silt and/or clay < No. 200 sieve < 0.075 Note: Particles finer than fine sand can not be discerned with the naked eye at a distance of 8 in. (203 mn) range of particle sizes, while soils with a low value of C, are uniformly graded and contain particles of similar sizes The relative consistency of cohesive soils is described by several useful index Patameters which are expressed as water contents at particular soil states. These consistency states are known as Atterberg limits, determined by ASTM D4318 (3) The most common index paraueters are: wp = in-situ natural water content, wy = liquid limit, wp = plastic Limit, PI = wy - wp = plasticity index, and LI = (iq - wp)/(wy - wp) = liquidity index. Soils with a Liquid limit (wy) greater than 50 percent are termed “highly plastic". A plasticity index (PI) greater than 25 to 30 may mean troublesome soils vith low strength, high compressibility, high shrink- swell potential, ete. The liquidity index (LI) is an excellent indicator of geo- logic history and relative soil properties, as shown schematically in Figure 2-1. 2-2 Me "wp Decreasing woter content Sensitive _NC_ LOC HOG LOC __HOC __.- increasing OCR, Ko —————= Increasing strength, modulus = Decreasing compressibility art Q <0 atte pcereasing Lr NC = normally consolidated Loc = Lightly overconsolidated HOC = heavily overconsolidated OCR = overconsolidation ratio = 2p/évo Bp = maximum vertical effective stress in soil during its geologic history Byo = Vertical effective stress in-situ Tho = horizontal effective stress in-situ Ky = in-situ coefficient of horizontal soil stress = Gho/8vo Figure 2-1. Liquidity Index Variations Index Paraneters for Cohesionless Soils Cohesionless soils also can be represented by simple index paraueters, generally expressed in terms of either "unit weight" or "density". Unit weight (7) is defined as the soil weight per unit volume and is given by the units kW/m? or Ib-force/ft3, Density (p) is defined as the soil mass per unit volume, with units of ke/m? or 1b-mass/fe' SI usage, conventional engineering practice has favored unit weight, which will be Although density actuslly is the preferred term in modern used in this manual. The ratio (7/p) is the grevitational acceleration (g), which is equal to 9.807 m/sec? or 32.17 ft/sec? For cohesionless soils, the relative density (D;) expresses the degree of compact- ness with respect to both the loosest and densest states achieved by standard labo- ratory procedures [ASTM D4253 (4) and D4254 (5)]. Most comonly, the relative den- sity is expressed in ters of void ratio: by = eS ae max ~ min ce) in which © = in-situ void ratio, egax ~ watimm void ratio (loosest), and enin ~ minimun void ratio (densest). Alternatively, D; can be expressed as: Panax(Pa ~ Pdnin) PalPamax ~ Péain) st Dy = im which pq = in-situ dry density, pagay ~ maximm dry density, and pain = minimus dry density. In this equation, unit veight can be used alternatively in place of density. In some instances, the degree of relative compactness is described in terms of the density index (Ip): Pa - Pdmin | Te eee i ? Pamax ~ Paain —- Relative density is a useful parameter for describing the relative behavior of cohesionless soils. Standard terminology is given in Table 2-2. Column (a) tends to be used more commonly in the U.S. Increasing Dy generally means increasing strength and decreasing compressibility. If Dy is negative, a collapsible soil structure may be present, such as can occur with honeycombed soils and very loose cemented or calcareous sands vith e > egax. The applicability of Dy is limited to cohesionless soils having less than 15 percent fines. In practice, it has been misapplied occasionally to soils having greater than 15 percent fines, with ques- tionable results. Since it is very difficult to obtain truly undisturbed samples of clean sands, the direct measurement of Dy also is difficult. In addition, the Table 2-2 RELATIVE DENSITY OF COHESIONLESS SOILS Relative Density Pe) @ y ‘Very loose 0 to 15 0 to 20 Loose 15 t0 35-20 to 40 Medium 35 t0 65 40 to 60 Dense 65 to 85 «60 to 80 Very dense &5 to 100 80 to 100 @- Source: Lambe and Whitman (6), p- 31. b - Source: Meyerhof (7), p. 17 in-situ void ratio (e) is compared to epax and egin, both of which are subject to considerable error in their determination in the laboratory. For these reasons, Dy should be considered only as an index parameter. For a variety of natural and artificially-prepared mixtures of sands, ¢max and min depend primarily on the particle roundness (R) and the uniformity coefficient (Gy). The roundness is defined as the ratio of the minimum radius of the particle edges to the inscribed radius of the entire particle. Although R is difficult to measure, it can be estimated from the apparent angularity of the grains, as shown in Figure 2-2. Conbined with a particle size analysis, the enax and Cnin values can be estimated from Figure 2-3. This figure is valid for clean sands with normal to moderately-skewed particle size distributions. R 30 High Impossible to crush with fingers Source: Sowers (3), p. 83. Table 2-4 APPROKIMATE COHESIVE SOIL STRENGTH BY SIMPLE TESTS oa Strength (ks£) (e/m2) Field Test Very soft 0 to 1/2, 0 to 25 —_—Squeezes between Fingers when fist is closed Soft 1/2 t01 25 to 50_—_—Easily molded by fingers Firm 1t02 50 to 100 Molded by strong pressure of fingers sestt 203 100 to 150 ented by strong pressure of fingers Very stiff 3 to 4 150 to 200 __Dented only slightly by finger pressure Hara >4 > 200 Dented only slightly by penetl point Note: qq = unconfined coupressive strength = 2 54 84 = undrained shear strength Source: Sowers (3), p. 80. Table 2-5 APPROXIMATE COHESIONLESS SOIL RELATIVE DENSITY BY SIMPLE TESTS Density Dr (#) Field Test Loose 0 to 50 Easily penetrated with 0.5 in. (12 mm) reinforcing rod pushed by hand Firma 50 to 70 Basily penetrated with 0.5 in. (12 mm) reinforcing rod driven with 5 Ib (2.3 kg) hammer Dense 70 to 90 Penetrated a foot with 0.5 in. (12 mm) reinforcing rod driven with 5 1b (2.3 kg) hamuer Very dense 90 to, 100 -—-Penetrated only a few inches with 0.5 in. (12 mm) reinforcing rod ériven with 5 lb (2.3 kg) hammer Note: generally refers to shallow depths in uncemented quarts end feldspar sands Source: Sowers (9) +p. 81 CLASSIFICATION General Classification and Identification Systems Classification systems are useful for grouping together soils of similar particle size and plasticity characteristics. By this grouping into pre-established cate- gories, consistent terminology can be employed to represent a soil fitting within the bounds of a particular category. The most widely used of these systems is the Unified Soil Classification System [ASTM 22487 (10) and D2488 (11)], given in Table 2-6. To use this system properly, both particle size and Atterberg limits data are needed. With the particle size and Atterberg limits data, the soil is classified using the pre-established group symbols in Table 2-6. Plastic soils utilize the plasticity chart shown as well. Note that if any soils plot above the "U" line in the plasticity chart, the data should be questioned and verified. Further details are given in the ASTM Standards Other well-known special purpose classification systems have been developed by the U. S. Department of Agriculture (USDA) for agricultural purposes, the Federal Avia- tion Administration (FAA) for airport pavements, and the American Association of State Highway and Transportation Officials (AASHTO) for highway pavenents. These systems normally are not used in foundation engineering. As an alternative, Burmister (12, 13) developed a soil identification system for both field and laboratory use. As compared with the classification systems which use pre-established soil group categories, Burmister's approach uses rapid and sim- ple visual-manual procedures to approximate the particle size and gradation and overall plasticity index. Essential features of the resulting soil identification are given in Table 2-7. With this system, approximate percentages of the principal and minor components are estimated using the notation in Table 2-7a. Particle size and gradation terms are defined in Tables 2-7 and c. For the fines (percent < No. 200 sieve), the overall plasticity is estimated and then described using the nota- tion in Table 2-76. Example identificaticns also are given with this table. Once the straightforvard visual-manual procedures are mastered, some 15 to 30 samples per hour can be identified in terms of their approximate particle size distribution and plasticity index Gone Penetration Test (CPT) Classifications The CPT has been used widely for many years as a site investigation device Although no soil sample is recovered, the cone tip resistance (qq), cone side resistance (f,), and friction ratio (Rg = FR = £4/dc) have been employed to 28 Table 2-5 UNIFIED SOIL CLASSIFICATION SYSTEM Beata Conta Asta Gow Sts ae Gap Ra iy Lion Te* oon! : = Seren d0arcanedonne Mow'tans0% ctcore Unter it ings? ete SOGSS __ GN tet ge— eee eee Sueeman ts ce3e Feo reece? = See Tapes Frey at Sey gna eee Frescamsy ck Sh (66 carey eo ae ‘cane Giz6eatscess" 0 Wg Segre, Enestin’ Clcsmaeisase 3 Fory ome sr emnie peony armour Sere Mere rant2s tres? Foes ans ar Gko'GH 80 Gey nl Freres Se Seeeten me P37 a os on aoe RT Geen gr SErorore pues tena, gas cn ano rece ee Seescee amen oe N e re Sette ernon Erorn an ete ose ee a Carer o ous lear irom Fy apa a Ad OH et 7 Beso rar pg be mm) oF 7 ae corre = 30% pam te SN PO me Forney 5h coquray ome ns nero pe ae Taig sno nana seo wee fF corns = ured wih saer TBS Seapow evermore ee renee iit sty se c.se_ cna sige: | Rimeworeue sme ear win So es tes mnie at gyn SEs oo SS egress loa nen ccespeseaceeeecsenaete Savant mtoodeate SR Sree 8% yee hoo oS, nor Pe eae meee al quak surg met a a 8 Batre 15 208 oe No. 00, a Beet weap ee en ot Sak orl tgrated SPSC wrguoe sro won ony ae baer SS sory gens nh Tie comare = 20% ots No. 26, oe ‘St oon goo sean Sy cons o Seay prone 60 a g 40 B 20) a 7 4 al fo, 4 | 0 1620 405060 80 Too Liguid Limit, w__ Source: American Society for Testing and Materials (10), pp. 289, 292 29 Table 2-7 BURMISTER SOIL IDENTIFICATION SYSTEM a (@) Terns Describing Composition of Cohesionless Soils nS Identification Proportion Component Written Symbol. «= Hitten Symbol by Weight ee Principal GRAVEL. 6 . : > 50 SAND s c : > smut y a 2 = Minor Gravel « and a 35 to 50 Sand s some s 20 to 35 sile y litele 1 10 to 20 trace © 1 to 10 eee SS Ee (b) Terms Deseribing Gradation of Cohesionless Soils — Designation Witten Symbol “Defining Proportions nO coarse medium to fine emf _—all fractions > 108, coarse to mediun ca < 108 fine medium to fine af < 108 coarse coarse e < 10% medium and fine medium = < 108 coarse and fine fine £ < 108 coarse and medium ee eee NOTE: For proportions in (a) and (b), use + for upper Limit and - for lower limit, Table 2-7 (cont'd) BURMISTER SOIL IDENTIFICATION SYSTEM (ce) Particle Size Definitions Soil Fraction Sieve Number and Size Gravel coarse 3 into lin (76 mm to 25 mm) medium 1 in to 3/8 in (25 wm to 9.5 mm) fine 3/8 in to No. 10 (9.5 mm co 2.0 mm) Sand coarse No. 10 to No. 30 (2.0 wm to 0.6 mm) medium No. 30 to No. 60 (0.6 mm to 0.25 mm) fine No. 60 to No. 200 (0.25 mm to 0.075 mn) sile - < No. 200 (< 0.075 mm) (a) Terms Describing Cohesive Soils Based on Overall Plasticity overall Plasticity Principal Component. Minor Component, Plasticity Written Symbol Index Written Symbol. «== Written Symbol. Non-plastic — - ° stLt ¢ sile 5 slight si Lto5 — Glayey SILT. Gy Glayey Silt ys Low L Stol0 SILT&CUAY ¥&C Silt&Clay sec Medium " Wto 20 GLA ASTIT «Gay Clay&silt ces High R 20 to 40 Silty CLAY = gy= Sty Clay $y Very High ve > 40 cuay c clay c BUMPLES: Fall 7 coarse? medium to Fine” SAND, some” uediun fine Gravel, trace? sile Abbreviated - cfnf” SAND, s“-mf Gravel, tt-Sile Shorthand - ctné" S, s*-més, thy Pull = CLAY & SILT, little* coarse” medium to fine* Sand, Medium Plasticity Abbreviated - CLAY & SILT, 1*-c"mf* S, M-PL Shorthand - C & ¥, It-c'mf*s, MPL NOTE: Principal component (> 508) always listed first. If no principal component, List sand first. Source: Burmister (12, 13). a classify the soil in-situ. Since soil classification by the CPT is an empirical ‘epproach, it has been an evolutionary process which has required periodic updates es new and larger data bases have been collected and evaluated, Two representative examples of the earlier interpretations of CPT data are shown in Figure 2-4. Fur- ther research led to empirical classification charts for the mechanical Begemann friction-cone, as shown in Figure 2-5. Similar developments led to classification charts for electric friction cones, as shown in Figure 2-6 in original form and in Figure 2-7 in simplified form. Recently, it has been realized that the correlations should be made dimensionless by appropriate scaling factors (Wroth, 18), Numerous field studies have shown that the cone side resistance increases proportionally with confining stress. For the tip resistance, the proportionality varies with soil type (e.g., Jamiolkowski, et al., 19). Therefore, at the present time, the most rational approach to soil clas- sification by the CPT is by using dimensionless parameters, as given in Figure 2-8 Soil classification using Figure 2-8 requires an iterative approach, since qo is divided by a power function of the vertical effective stress, (Byo)", and the expo- nent (n) depends upon the soil type. This exponent (n) increases from about 0.5 for sands to approximately 1 for clays. 4n initial estimate of soil type may be obtained from Figure 2-7. A first estimate of n for the iterative solution then 00k Mechanical Friction Electric Friction =| cone Cone ua © 300) ue 5 wg 3 200 22 & é 28 = 20 8 . 33 € 100) eae & 50 3° eo Cone Side Resistance, fs/pq Cone Side Resistance, fs/Pg |. Mechanical Friction Cone Electric Friction Cone Figure 2-4. Early Soil Classification by CPT Source: Laboratorium voor Grondnechanica (14), p. 29. 212 1000, North 3 ae j 400k 8 Florida] ole Soils S € g obs} ee /—___, too] 8 3 cLavey e 2 /Sanoy AND 5 SaNDS 2 BES sano (Ae / Ss 8 ao % aoa 2 é gq é > > - & ® 1 A 2 5 8 sehsitive Zig 2 8 1 8 8 & 4 4 2 1 5 ° 2 4 6 Friction Ratio, f,/a¢ (%) Friction Ratio, f,/a¢ (%) Figure 2-5. Soil Classification by ‘Figure 2-6. Soil Classification by Fugro Mechanics] Friction CPT Electric Friction CPT Source: Sehmertmann (15), Pp. Source: Douglas and Olsen (16), p. 222 400) 400 < ‘Sonds e Z 2 © Psy e So 100 7 sands / &, 100} o JY - Fronbs 2 S af 7” 7 sis.” eoyey SF leon to g 4, Zond silts /sits ond,/ “e 40f Silly sond 2 7883 099f ¢ 2 4 7 & 2 ca 7 cave 2 E ig 7 7 i. o 8 s IQ ar Peat os & 2 8 4 2 4 6 Friction Ratio, f5/a¢ (%) Figure 2-7. Simplified Soil Classification by Fugro Electric Friction CPT Source: Robertson and Campanella (17), p. 721, Cone Side Resistance, fy" + fan*fs/Fy0 Figure 2-8. Most Recent Soil. Classification by Fugro Electric Friction CPT Source: Olsen and Farr (20), p. 858 can be made from Figure 2-8. As described in Appendix 3, different results commonly are obtained using different cones. ‘Therefore, adjustments to the following figures may be warranted as a func- tion of cone type and shape, as given in Appendix B. Piezocone Penetration Test (GPU) Classifications With the recent developnent of the piezocone, which measures the total penetration pore water stress (ug) in addition to qo and fg, the ability of the cone penetrome- ter to delineate soil stratigraphy and provide an accurate classification of soil type is enhanced greatly. In loose, contractive sands, the value of ug closely follows the hydrostatic stress (vp). In dense, dilatant sands, ug may be less than Uo. In clays, cone penetration generates excess pore water stresses which are recorded by the pore water transducer. Two of the recent soil classification sys- tems based on CPTU measurements are given in Figures 2-9 and 2-10. Other classifi- (23). In the first of these figures, the parameter By is used, which is defined as: cation charts are given by Robertson, et Ya = Yo 9” ar = eve Cw in which up ~ measured total pore wats stress (usually behind the tip), Us = hydrostatic pore water stress, ap = corrected cone tip resistance, and oy ~ total overburden stress One important finding which has evolved from the development of piezocones is that the cone tip and side resistances must be corrected for pore water stress effects acting on unequal areas of the cone geometry. The corrected tip resistance is given by: Qn de + CL ~ aupe (2-5) in which qe = measured cone tip resistance, a = net area ratio for the particular cone (See Figure 2-11.), and wr = pore water stress behind the tip. Similarly, the correction for cone side resistance is given by: fem fe + (us Ag? - Ube As)/As (2-6) in which us ~ pore water stress behind the sleeve, Ag = surface area of the sleeve, 3S 8 100} 8 Corrected Cone Tip Resistonce, a, /Pq 8 =: Sond | ast) medium 5 = Ba: or50,_.-100" 150-200 Sr-Pv0 30, a. 50. PSSST iy sand . hi Coy ih ue htt 2 et tem ‘2. very soft Figure 2-9. Soil Classification Based Figure 2-10. Soil Classification Based Source: on ar and 8g on CPTU Data Senneset and Janbu (21), p. 48. Source: Jones and Rust (22), p. 612 Us Hd Hit Us 4 As? Surf reo, se face a a an {hpiealy 15,000 m2) Cone sleeve measuring side resistonce Figure 2-11. Unequal End Areas of Electric Friction Cone 2-15 f, - measured cone side resistance, and As) and Ag are the net internal areas of the sleeve, as given in Figure 2-11. Dilatometer Test (DMT) Classifications The flat éilatometer test (DMT) also is capable of providing an estimate of the soil type and consistency. The original development of the DMT (Marchetti, 24) included 2 classification based on the material index, Ip, defined as: PL - Po . b*5 ova (2-7) in which po = contact stress, pj = stress to expand meabrane 1 um into soil, and up ~ ambient equilibrium pore vater stress (often assuned to be hydrostatic, although not necessarily so). A more recent interpretation is shown in Figure 2-12, which is based on Ip and the dilatometer modulus, Ep, defined as: 2000-——+ or m3 str "|" sano 38 rey Sy sity | a 1000,- “A 4 oo 2 Soop al we € cuay os of a sy LL A Notes we 200P Res a- Number in parenthesis 3 Do © is normalized unit 3 “oD 2 seq ‘eight (y/>w) S100, » eo b= If PI>50, (y/yq)is > overestimated by 5 » wy ‘about 0.1 2 Peas : 7 2 we at NS 4 gh —sasyeetsizis 3s J VERY SOFT CLay/ PEAT sto ieee L a Material Index, Ip Figure 2-12, Determination of Soil Deseription and Unit Weight by DMT Source: Schmertoann (25), p. 98 Ep ~ 34.7(P1 - Po) (2-8) This correlation also provides an estimate of the sofl unit weight ‘UNIT WEIGHT As previously defined, the soil unit weight (7) is determined as the weight of soil per unit volume. The relationship between dry (7g) and total (y¢otal) unit weight Yeotal ~ (1 + ¥n)71a (2-9) in which vq = natural water content (as a decimal). Table 2-8 presents typical soil unit weights RELATIVE DENSITY OF COHESTONLESS SOILS FROM IN-SITU TEST CORRELATIONS The standard penetration test (SPT) N value and the CPT cone tip resistance (qc) have been used extensively to estimate the relative density of cohesionless soils in-situ, Although they are used commonly in practice, different approaches have been adopted by different authors. Sone of these differences in methodology result from improvements in the understanding of penetration tests and the relevant fac- tors affecting the test values. Also, the estination of the relative density using the SPT and CPT results is an evolutionary process during hich never ané larger Gata bases are compiled to allow for more statistically significant trends to be established, Furtheraore, some earlier studies vere based on penetration tests conducted in one type of soil. Testing of more soils of differing geologic ori- gins, stress histories, and mineralogies allows for refinenents and adjustnents to existing correlations. Standard Penet sion Test (SPT) Correlations Early work on this subject simply correlated the SPT N value directly with relative density, as shown in Table 2-9. Later laboratory research demonstrated that the SPT N value also was influenced significantly by the overburden stress. Figure 2-13 shows these results, which were based on calibration chanber tests. For prac- tical use in estimating Dz from N and Jy, these results were presented in alterna- tive forms such as that shown in Figure 2-14. Additional research showed that these relationships are even nore complex and dependent upon other factors, including vertical stress, stress history, and sand 247 “GE ‘ve ‘dd ‘(5% yBnoy aang PUD ALS 38) cU/AM OR" ~ ¢8/2 C96°O = cUO/u T = €23/at Weg = BL 10204 002 FT ox't ae oL-0 ony : so 8 (2218 He > wos 09 06) Keo opuL8z9 ore eT TO SSO GE : a 5 5 0 soz mT n't 0 09°09 ooret - yor t0"0 Corrs He > 405 3900) ke19 LePFoL Le we wT LT ORO aD Oe : too'0 0 s0°0 (o2t# He > a05 92 08) OS'% — oo'z kez opt ETO OLD BOOT 9 Sz 200° t00°0 ose az wt wee set Oro OT ~ yoo. ose 92 OFT TZ 9670 Sz"D ORT ©— OL 3 OT EoD'D 100 OE Se oT fez eeT ero e'9 or eB ST tO" S000 nt ee eet tet zet ovo oe ~ oe wee MET Taz get ozo 0H go.0 soo ve ez wt moe GET OED OS oT EOD 00" Oe az oct oat geet ort 0° 02 ZT ZO. $000 s0°0 ae ST Gat eet oor overgr ee EOS ore rt eT tet ovo Ut wo oso mo wo ot ne! xen sxe cu TeX TG /09q@ Tg. Ty HG MAREE spoveanaes M/A kag “OFA PTON— UETOTII909 “Te BETS WTO aon aTun Poa ToNTG — ‘Aaqw203 100 oaeuyxorddy SIHOTAM LIND ‘110s “TwoTaAL 8-2 erarL Table 2-9 RELATIVE DENSITY OF SAND VERSUS N N Value Relative (blows/ft or 305 om) Density Dp (8) 0 to % very loose 0 to 15, 4 0 10 loose 15 to 35 10 vo 30 medium 35 co 65 30 to 50 dense 65 to 85 > 50 very dense a5 to 100 ‘terzaghi and Peck (I), p. 341 and Lambe and Whitman (6), p. 31 SPT N Value (blows/ft or 305mm) o 20 40 60 a — ‘i { Togo, (9 ! 0 % 1 te l 20} 60 20} Vertical Stress, d9/D9 SPT N Volue (blows/ft or 305mm) rr a Relative Density, D, (%) Figure 2-13. Effect of Overburden Stress Figure 2-14. Relative Density-N-Stress ‘and Dz on SPT N Value Relationship Source: Gibbs and Holtz (28), p. 37 Source: Holtz and Gibbs (29), p. 4él type (primarily compressibility influences), as 2 minimum, Figure 2-15 illustrates sone of these complexities. The studies presented in Figure 2-15 led to a corre- lation for estimating Dy from SPT N values that includes the effect of overburden ao too; T r = == Platte river soné (4) é —— Standord concrete sand (8) = 5 1HU111 TO pshReid Bedford medel sond (C} ond Ottawa send (0) Q BOP BE 40psifeid Bedford medel sand {C} ond Ottawa sond (0) a ANY a0 psied Bestora medel sane (€) 5 — Gibbs ond Holtz 1957) tor'cv-aey coarse (€) ond fine sane SI == Gibbs and Holtz (1357) fer submerged sands (E} S gol ~-~ Bazoraa (1987) g (psi #6.89 N/m?) @ 40- s z= 20 b a 6 ol 0 20 30 40 50 60 70 80 90 100 Relative Density, 0, (%) Figure 2-15. Relative Density-N-Stress Relationships for Several Sands Source: Marcuson and Bieganousky (30), p. 1301. stress (yo), particle size distribution (G,), and stress history (OCR ~ 3p/Byo), as given below: Dy(K) = 12.2 + 0.75[222N + 2311 - 711 OCR - 779(yo/Pa) - 50 G2}9-3 (2-10) Regression analyses of the data gave r? = 0.77. The data all were unaged with OCR equal to 1 or 3 ‘An important factor affecting the SPT N value is the energy efficiency of the drop hammer onto the drill rods. The theoretical free-fall energy for the SPT is 140 1b (0.623 kN) times 30 in (0.76 m) or 4200 in-Ib (0.475 KN-m). Typically, the average energy ratio (ER) is about 55 to 60 percent in the U.S.A., although this value can vary from 30 to 90 percent for particular drillers and SPT equipment in practice. Skempton (31) reviewed SPT calibration data from Japan, China, the U.K., and the U.S.A. and suggested correction factors based on standard practice in these coun- tries. Some of the variables affecting the energy efficiency include the type of hammer, age of the rope, borehole size, and use of liners in the split spoon sam- pler. For example, the donut hammer is less efficient than the safety hammer, as shown by the energy ratio examples in Figure 2-16, Correcting the hammers to a 2-20 SPT N Value (blows/fi or 305mm) 020 40 60 oO 20 40 60 34 (Measured) % Corrected) Sr esate 20 20 aL 7 Donut 2 -Hommer at SS Monammertes aL f/ Comected 25 < 56 a) to ERSS = 2 ‘ 30 63 30 percent _ 2 1b 63 10 * = = [35 ‘ 3s = _ \ 12| 63a lao |2} > 40, sau s6 , —. | nergy '4F ratio (ER)? "4 Figure 2-16. Donut and Safety Hammer Comparisons Source: Robertson, et al. (32), p. 1454. constant energy ratio eliminates the differences. The energy efficiency also depends upon the size of cathesd and number of turns of the rope, as indicated in Figure 2-17. Standard U.S. practice is two turns of rope on 2 large cathead. ‘The SPT N value, corrected for Field procedures, is given below: Ngo = CER Cp Gs OR N @-y im which Ngg = N value corrected for field procedures to an average energy ratio of 60 percent, N= measured SPIN value, and Cea, Cg, Cs, and Cp are correction fac- tors for energy ratio, borehole diameter, sampling method, and rod length, respec- tively, as given in Table 2-10. Since the SPT N value also varies with stress level, overburden stress correction factors are used to provide a consistent point of reference. This correction takes the form: p60 ~ x Neo (2-12) in which (N)¢0 = Ngo value corrected to a reference stress of one atmosphere and Gy = correction factor for overburden stress = a-Rope thrown off catheod yo Small 41.2 os! ~ A & | stenet 10 a (> 30 #0) co 6 to 10 m (20 to 30 fe) 0.95 4 to 6 m (13 to 20 £t) 0.85 3 to 4 m (10 to 13 ft) 0.75 Source; Based on Skeapton GI) Perhaps the simplest expression for Cy is given below (Liao and Whitman, 33): Ox = (Pa/%v0)5 (2-13) A comparison of different Cy recommendations is given in Figure 2-18. Basically, all methods give similar corrections for 3,9 > 0.5 pg within the range of expected accuracy for the SPT. The correction factors proposed by Skempton are based large- ly on Laboratory test data, while the others have been derived from field data. Although Equation 2-13 is simple, high values of Cy develop at very low values of ayo. Alternatively, Skempton (31) suggested the following for fine sands: Oy = 2/1 + ay0/Pa) (2-14) ‘This equation gives 2 maximum Cy of 2 at the ground surface, Figure 2-19 shows that both equations are adequate for Jyo > 0.5 pa and also appear applicable for use in overconsolidated sands Once the SPT N value has been corrected for field procedures and overburden effects to give (1l,)g0, it can be used £0 evaluate the relative density asa function of the soil characteristics. Figure 2-20 shows (N,)¢0/Dy? as a function of the soil particle size (Ds9). The Laboratory data in this figure were obtained from studies SPT Overburden Correction Factor, Cy oe os 10 1S 20 25 os 8 1.0] 8 Peck,et a, 1974(34) Liao ene Whitman, 386 (33) — Skempion, 1986 (31) Gere sone — Skempton, 1986.3) Fine onde —-— Skempton, 1986 (30: ie eonds Seed, 1979 (35); Sse? Vertical Stress, 5 Seg, 1979 (35) Op ro% 30 Figure 2-18. Comparison of SP? overburden Corrections SPT Overburden Correction Factor, Cy Osieee onsets seeonmmas + Fy /Py Vertical Stress, * Reid Bedford Model Sond (ocr= 3) 6 Figure 2-19. Comparison of Recommended Oy Factors and Available Data from 0C Sands 100, Peck ond Bozoroa goto | (Nc, — v ae) ‘Laboratory. rv ST noses) oom 3 (ea ea { ost? Zz 40 u 2 6 r 2o| WES saa © Skempon interoreton 8a Recevauten ving Serolt mehod fr evete eft Bie ecscuoton van] Lie inon mcnce fr carson elect ol Clem 02 35 1 2 3 Oso (mm) Figure 2-20. Particle Size Effect on Blow Count for Sands at the Waterways Experiment Station (WES) on three sands (30, 36, 37). Most of the data were for unaged, normally consolidated (NC) sands (OCR ~ 1), although a small series of tests was conducted on overconsolidated sands with OCR - 3. Skempton’s interpretation (31) of these data is shown, but it is believed that the averaged 224 curves and smoothed data he used led to an underestimation of (W})60/Dz2. Re-eval- uation of the original data (36, 37) leads to the higher values shown, using either Skenpton’s Linearized overburden effect (31) or Liso and Whitman's (33) nonlinear overburden effect. These results can be approximated as follows: (8) 60/Dz? = 60 + 25 log Dso. (2-15) which is applicable for NC, unaged sands. The 0¢ data give higher values than Equation 2-15, and aged sands also give higher values. The data from Niigata, Japan were tabulated by Skempton (31), but they were re-evaluated individually. The Peck and Bazaraa (38) curve represents coarse sands (no exact particle size given) from field test evaluations. These data represent aged sands that Likely were overconsolidated. Figure 2-21 illustrates the data as a function of age of the deposits. The WES Laboratory data are plotted at an age of several days. The Niigata, Ogishina, and Kawagishi data summarized by Skeapton (31) represent NC recent fills that vere assigned approximate ages of 30 to 40 years. ‘The time is not known for the OC, aged, Peck and Bazaraa data, so it is estimated at 100 to 10,000 years. The other four sites (A, B, C, D) are given by Barton, et al. (39). They represent 0C, aged, Hine and fine to medium sands of four geologic periods, as noted. 200; o A, B,C,0 ore OC, aged, Srontnar fine ‘ond fine to medium sands Cross | (Grates) 150) Ken 8 (ine estcers) Eben loo Peck ond Soteraa- (Essene! [ fosge75mnt f LE ‘Aging Effect sol karen od cy212+Q05I0gt1/1001 INcvuneges 16 Ogishima-NC, aged (Ny)60/0? “e_ Kowogishi-NC, oged tee fe oe f 10 10 10) 10% 0? Age or Time Since Deposition, t (years) Figure 2-21. aging Effect on Blow Count for Sands 2-25 Based on Figures 2-20 and 2-21, it is clear that particle size, aging, and overcon- solidation significantly influence the (N;)59/D,2 ratio. These effects can be quantified as follows: @V60 poe ae tC Ca Coor, 2-16) im which Cp, Cy, and Cogg are the correction factors given in Table 2-11. Gp is based on Figure 2-20. Cy is based on a conservative interpretation of the impre- cise data in Figure 2-21. Cogg is based on direct evaluation of the WES data and Interpretation of the Niigata data. It also is consistent with the studies pre- sented by Tokimatsu (40) Finally, the complete expression for relative density (Dy) in terms of SPT N value, including 211 corrections and modifying terms, is CER OB Cs Cp Gy N Dy? Sp Cx Cocr (with Dy in decimal fora) 17) in which N ~ measured N value and the corrections are as follows: energy ratio (Gp), borehole diameter (Cg), sampling method (Gg), rod length (Cp), overburden stress (Cy), particle size (Gp), aging (Gq), and overconsolidation (Coca). Cone Penetration Test (CPT) Correlations Early work on this subject was similar to the SPT, and therefore the CPT qo value Table 2-11 SPT CORRECTION FACTORS FOR SAND VARIABLES Correction Effect Parameter Tem Value Particle size Dso of sand op 60 + 25 10g Dso (D5 in mm) aging Time (e) Ca. 1.2 + 0.05 log (¢/100) Overconsolidation OR = 3p /2yo ocr oc 18 simply was correlated directly to relative density, as shown in Table 2-12. As with the N values, recent research has shown that the relationships are more com- plex. Figure 2-22 shows the generalized relationship for Ticino sand, which is of medium compressibility. The vertical effective stress can be used with this figure 4€ the sand is unaged and normally consolidated. The horizontal effective stress should be used if the sand is aged or overconsolidated. Figure 2-23 illustrates that the generalized CPT correlations vary for soils of ifferent compressibilities. Curve 3 corresponds to data on Monterey sand, which is of low compressibility. Monterey sand is characterized by subrounded to sub- angular grains, waich are composed mainly of quartz and sone feldspar, with zero percent fines. Gurve 2 is for Ticino sand, a granular soil of moderate compressi- bility with subangular grains composed of quartz and 5 percent mica, with less than 1 percent fines. Curve 3 is for the high compressibility Hilton Mines sand, con- sisting of angular iron mine tailings of quartz, feldspar, and mica composition, with 3 percent fines. To compare cone tip resistances obtained at different depths, it is necessary to reference the values to a standardized reference stress level, usually taken as 2yo/Pa = 1 atmosphere, The standardized cone tip resistance (qq) then becones: Qn = Cq de (2-18) Table 2-12 RELATIVE DENSITY OF SAND VERSUS qo Cone Tip Relative Resistance, de/Pa Density De (*) < 20 Very loose < 20 20 to 40 Loose 20 to 40 40 to 120 Mediu 40 to 60 120 co 200 Dense 60 to 80 > 200 Very dense > 80 Source: Meyerhof (1), p. 17. 2-27 Cone Tip Resistance, qe/Pq ° 200 400 i g |e Fost = 1 2 ope 7 B / sa Bust g : 2 ee ere 2ol 8 cs | Del%)=40 30 60, 70 80 30) Figure 2-22. Relative Density from CPT for Uncemented and Unaged Quartz Sands Source: Robertson and Campanella (17), p. 723. Cone Tip Resistance, ae /Py 9 100 200 300 400 _s00 © Hilton mines send €, @ Ticino sand ne @ Monterey sond a 4 Compressibilty Vertical Effect o 0,=40% Figure 2-23. Relative Density frox CPT, Including Soil Conpressibility Source: Robertson and Campanella (12), p. 722 in which qo = measured cone tip resistance, and Cy ~ overburden stress correction factor. For all practical purposes, Cy is nearly identical to Cy, proposed for the SPT and given as: 228 1g = ON = (Pa/Byo °F (2-19) Much research on the CPT has been conducted in calibration chambers, which are described briefly in Appendix H. These studies allow the use of controlled sand properties and in-situ stresses, which is not possible in the field. One sumary of Dy data from calibration chanber tests on five different normally consolidated sands is shown in Figure 2-24. This figure illustrates the range in actual data taken under controlled laboratory conditions after uniform soil placement. The generalized figures shown earlier in this section do not show the data range and perhaps suggest @ high confidence level. Figure 2-24 shows what the actual ranges are for only five sands under controlled laboratory conditions; field cases are Likely to exhibit more variability. Calibration chamber data are useful, but the tests are performed with flexible walls of limited dimensions. Therefore, the boundary effects result in lower qc values than obtained for "field conditions", corresponding to an infinite half- To correlate the field and chamber qo values, Jamiolkowski, et al. (19) recommended dividing the field value of ae by Ky, as given below: space. oor Ss ,.| o:tarse li )-1 X20 (ole LG " = 2 60 49 é ae g fez, Sond 2 LY SS 3 /38/ © Ottowe Be 3/° © Edgar es © Hotkeund wo Jet & hiton mines \ te \ 10 30 105 500 1000 Ge! Po Figure 2-24. Cone Tip Resistance, (Gp/e— Correlation Between Dy and Dimensionless qc (Uncorrected for Boundary Effects) Source: Jamiolkovski, et al. (19), p. 120 2-29 Kg = 1+ r - 30)/300 (2-20) before entering Figure 2-24, The equivalent chanber values then can be used to evaluate Dy. This process requires iteration, because the value of Dy is not known. ‘As an alternative approach, the results of 24 sete of calibration chamber tests on sands were compiled, in which the values of qc were corrected for the effects of boundary conditions. These sands vere predominantly fine and medium sands. A sun- nary of this compilation is given in Appendix H. For the majority of chambers with flexible walls, the boundary correction required an increase in qc to reflect "field" values. The results of this study are given in Figures 2-25 and 2-26 for unaged, uncemented sands. In all cases, a linear relationship was obtained for the square of the rel- ative densiey (0,2) versus the dinenstonless cone tip resistance, given as show by Qcp. Figure 2-25 shows the normally consolidated sands, separated into lor, nediua, and high compressibility. Low compressibility (Figure 2-25a) generally corresponds to quartz sands with Hittle, if any, fines. Medium compress{b{lity (Figure 2-25b) suggests quartz vith sous feldspar, with perhaps severel percent fines. High compressibility (Figure 2-25c) indicates more fines, mica, and other compressible minerals. Most natural sands likely will be more tovard the mediun to high range of compressibility. As shom in these figures, the correlation is good below the Limit of possible particle crushing. This init was established by sta- tistical analysis of the data, optimizing the x? value as a function of different Limiting Qgp values from 250 up to the entire data set. The Limiting Qcp values shown provide the maximum r2 for the data and define the boundary of possible par- ticle crushing. Data points beyond the Limiting Qcp values are not included in the statistics. Figure 2-26 shows comparable calibration chamber data on overconsolidated sands, separated into low (< 3), medium (3 to 8), and high (8 to 15) OCR ranges. These data also were optimized using x2 for different Qcp Limiting values, resulting in the regression lines and possible particle crushing limits shown. ‘A summary of these relationships is given in Figure 2-27 for all of the corrected calibration chamber data. This figure clearly shows the influence of compressibil- ity and OCR on the relationship between Dy? and the dimensionless cone tip resis- tance, These relationships can be quantified approximately as follows: Lo os 4 | ost 4 oat a 4 o2t 4 Compressibility Pr NC Sands 7 100 200300400 500 600 (q,/P9) = co (o/P> Lo tt r a ost 7 o L Q fePossibie particle | é crushing O6- i 4 oF oF pzn.2e "Qe 292 or (n=145, r2=0.885, | L $020.10) 02+ & (b} Medium 4 Compressibility r NC Sands 4 oo 1 1 1 1 100-200-306 40 500 600 (96/60) 9 aaa (So/P)* Figure 2-25. Calibration Chaaber Data on NC Sands 10 ago ) ®, e 3 oa ole? 4 2 Pog 580 e & 008 8 06: o BF 7 4 ae le H+ Possibie particle crushing | o ° 02.860 . co oat OR "280 | ‘00 (n#59, «220,769, i 8.0.=0.14) 4 oat fy (c) High 4 Compressibility NC Sands ° 0-100-200-3500 a00 500 €00 (e/Pa) jm pe co a Figure 2-25. Calibration Chamber Data on NC Sands (continued) 10 os| os 4 2 Low OCR OF Possible particle | grushing range 04 High OCR 2065/0, 4 o2| Ock__symtel_ oro 2 so} OF Bnei open 35030 G7 Ola p> MediS-8) ha 403. 56 0.649 O10.] high (8) fies 483 50 0808 OIE Q 0 "100-200-300 400-500-600 (Gea) a (Go/Pgh? °° Figure 2-26. Calibration Chanber Data on OC Sanés NG-high comp. 280 59 0796 O14 NC-med comp, 292 145 0.885 0.10 NC-low comp, 382 190 0.711 O14 Nc-everage 305 404 0776 013 Low OCR (3) 390 34 O71 O18 Med. OCR (3-8) 403 56 0.849 0.10 High OCR OB) 443 50 0.858 0.12 02! ol ‘0100-200 300400500600 (4¢/P9) Gp/Pgi® Figure 2-27. Summary of Calibration Chanber Studies cp ~ 505 Rc Goce ome) Cq(qe/Pa) 2. ane - Dra 1305 ;0n/Gnce| (2-21b) (Ge/Pa) ee (2-216) 305 Qe 06R9-18 (ayy /p,)0-5 in which Qc ~ compressibility factor (0.91 for high, 1.0 for medium, and 1.09 for low) and Qoga = overconsolidation factor (~ 020-18), comparable to Gocg for the standard penetration test. ‘The Qocg factor was evaluated using the mean OCR values for the low, medium, and high OCR data equal to 2.3, 5.1, and 10.1, respectively The majority of natural sands are likely to be of medium to high compres and low to medium ock. Tt should be noted that Equation 2-17 for the SPT is similar in form to Equation 2-21 for the CPT, although some differences are evident, Perhaps the most important difference is that the SPT relationship includes aging, while the CPT relationship is only for unaged sands. If the same functional relationship for aging holds for both the SPT and CPT, then Cy (as given in Table 2-11) would be introduced into the denominator of Equation 2-21, This addition is speculation at this time. However, the C, changes qualitatively explain the effects of aging ina reasonable manner. Dilatoneter Test (DMT) Correlations ‘The DMT is a relatively new test for which broad correlations have not yet been developed for relative density (Dp), However, it has been used to estimate Dy in normally consolidated, uncemented sands. This correlation is shown in Figure 2-28 for Dy as a function of the DMT horizontal stress index (Kp), described in Appendix D and defined as: RD = (oe - t9)/2vo0 (2-22) in which po ~ initial contact stress, up = hydrostatic stress, and Gyo ~ effective vertical stress. This correlation is based on few data and should be considered preliminary at thie time, © Chomber tests AO Field sites k,=0.40 / : / Horizontal Stress Index, Kp 2040-60-80 100 Relative Density, D, (%) Figure 2-28. Correlation Between DMT Horizontal Stress Index and Relative Density for Normally Consolidated, Uncemented Sand Source: Robertson and Campanella (41), p. 39. 2-34 CONSISTENCY OF COHESIVE SOILS FROM IN-SITU TEST CORRELATIONS The standard penetration test (SPT) N value and the cone penetration test (CPT) qe value also have been used to estimate the consistency of cohesive soils in-situ, However, little published work has been presented on these correlations, and there- fore all should be considered approximate at best. Standard Penets wn ‘Test (SPT) Correlations The consistency of cohesive soils has been correlated with the N value, as shown in Table 2-13. In general, these values are to be considered only approximate guide- lines, since clay sensitivity can greatly affect the N value (Schnertmann, 42) Although the correlations with N value in clay commonly are considered to be less reliable than those in sand, increasing N values do, in general, reflect increasing stiffness and therefore decreasing liquidity index. To express this general corre- lation, the consistency index (CI) has been defined as follows cr ae 1- Lt (2-23) eee ° which effectively is a mirror image of the liquidity index, Table 2-14 is Table 2-13 CONSISTENCY OF CLAY VERSUS N N value (iows/Ee or 305 ma) _Gonsisteney oto? Very soft 204 sore 4 t08 Mediun 8 to 15 sees 15 to 30 Very seiee > 30 fara Source: Terzaghi and Peck (22), p. 347 2-35, Table 2-14 CONSISTENCY INDEX OF CLAY VERSUS N and qc N Value Cone Tip (biows/£t or 305 mm) Resistance, qc/P, Consistency Consistency Index <2 <5 Very soft <0.5 2008 5 to 15 Soft to mediua 0.5 t0 0.75 8 to 15 15 to 30 seis 0.75 t0 1.0 15 to 30 30 to 60 Very seife 1.0 t0 1.5 > 30 > 60 Hard >1s Source: Szechy and Varga (3), ps 105, representative of the CI correlations. Cone Penetration Test (CPT) Correlations ‘The consistency of cohesive soils also has been related to the cone tip resis- tance. Again, as with the N values, the correlations in clay are less reliable. A typical correlation is given also in Table 2-16 RELATIONSHIP BETWEEN SPT N AND CPT qo VALUES Because of the numerous relationships developed for either SPT or CPT data, it is advantageous to have a procedure to interrelate N and qe. Both are penetration resistances (although the SPT is dynamic and the CPT is quasi-static), end they are the most common forms of in-situ testing used worldwide today. ‘A number of investigators have proposed single numerical values of qo/ll. However, recent studies have shown that q¢/N generally correlates with grain size, as shown im Figure 2-29. Unfortunately, most of these data do not include N or qe value corrections as noted previously. Newer data (4é - 50) have been combined with the previous results in Figure 2-29 to result in Figure 2-30. This new relationsiip confirms the general trend of the Gata, and it extends the relationship to mean grain sizes up to 10 mm. The new REFERENCES 1 2 13 cra American Society for Testing and Materials, "Standard Method for Particle-Size Analysis of Soils [D422-63(1972)]", Annual Book of Standards, Vol. 4.08, ASTM, Philadelphia, 1989, pp. 86-92. American Society for Testing and Materials, "Standard Practice for Wet Prepa- ration of Soil Samples for Particle-Size Analysis and Determination of Soil Constants (D2217-85)", Annual Book of Standards, Vol. 4.08, ASTM, Philadel- phia, 1989, pp. 270-272 american Society for Testing and Materials, "Standard Test Method for Liquid Limit, Plastic Limit, and Plasticity Index of Soils (D4318-84)", Annual Book of Standards, Vol. 4.08, ASTM, Philadelphia, 1989, pp. 579-589. Anerican Society for Testing and Materials, "Standard Test Methods for Maximum Index Density of Soils Using a Vibratory Table (D4253-83)", Annual Book of Standards, Vol. 4,08, ASTM, Philadelphia, 1989, pp. 560-571 American Society for Testing and Materials, "Standard Test Methods for Minimum Index Density of Soils and Calculation of Relative Density (D4254-83)", Annual Book of Standards, Vol. 4.08, ASTM, Philadelphia, 1989, pp. 572-578. Lambe, T. W. and Whitman, R. V., Sofl Mechanics, John Wiley and Sons, New York, 1969, 553 p. Meyerhof, G. G., "Penetration Tests and Bearing Capacity of Cohesionless Soils", Journal of the Soil Mechanics and Foundations Division, ASCE, Vol. 82, No. sii, Jan, 1956, pp. 1-19. Youd, T. L., "Factors Controlling Maximum and Minimum Densities of Sands", Evaluation of Relative Density and Its Role in Geotechnical Projects Involving Gohesionless Soils (STP 523), ASTM, Philadelphia, 1973, pp. 98-122. Sowers, G. F., Introductory Soil Mechanics and Foundations: Geotechnical Engi- neering, 4th Ed., Macmillan Publishing Co., New York, 1979, 621 p. american Society for Testing and Materials, "Standard Test Method for Classi- Eication of Soils for Engineering Purposes’ (02487-85)", Annual Book of Standards, Vol. 4.08, ASTM, Philadelphia, 1989, pp. 288-297 American Society for Testing and Materials, "Standard Practice for Description and Identification of Soils (Visual-Manual Procedure) (D2488-84)", Annual Book of Standards, Vol. 4.08, ASTM, Philadelphia, 1989, pp. 298-307. Burmister, D. M., "Physical, Stress-Strain, and Strength Responses of Granular Soils", Symposium on Field Testing of Soils (STP 322), ASTM, Philadelphia, 1962, pp. 67-97. Burmister, D. M., "Suggested Methods of Test for Identification of Soils", Special Procedures for Testing Soil and Rock for Engineering Purposes (STP 478), ASTM, Philadelphia, 1970, pp. 311-332, Laboratorium voor Grondnechanica, “Cone Penetration Testing", Civiele and Bouvkundige Techniek, No. 3, May 1982, pp. 16-36 as. 16. u 1s. 1s 20. 21 22. 23. 2. 25. 26 27. 28 29 Schnertmann, J. H., "Guidelines for Cone Penetration Test Performance and Design", Report FiA-TS-78-209, U.S. Department of Transportation, Washington, 1978, 145 p. Douglas, B. J. and Olsen, R. S., "Soil Classification Using Electric Cone Penetrometer", Cone Penetration Testing and Experience, Eds. ¢. M. Norris and R. D. Holtz, ASCE, New York, 1981, pp. 209-227. Robertson, P. K. and Campanella, R. ., "Interpretation of Cone Penetration Tests. Part I: Sand", Canadian Geotechnical Journal, Vol. 20, No. 4. Nov. 1983, pp. 718-733 Wroth, C. P., "Penetration Testing - A More Rigorous Approach to Interpreta~ tion", Proceedings, lst Internatioral Symposium on Penetration Testing (18orT-1), Vol. 1, Orlando, 1988, pp. 303-311. Jamiolkowski, M., Ladd, C. C., Geruaine, J. T., and Lancellotta, R., "New Developments in Field and Laboratory Testing of Soils", Proceedings, 11th International Conference on Soil Mechanics and Foundation Engineering, Vol. 1, San Francisco, 1985, pp, 57-153, Olsen, R. S. and Farr, J. V., "Site Characterization Using the Cone Penetro- meter Test", Use of In-Situ Tests in Geotechnical Engineering (GSP No. 6), Ed 5. P. Clemence, ASCE, New York, 198, pp. 854-868. Senneset, K. and Janbu, N., "Shear Strength Parameters Obtained from Static OPE", Strength Testing of Marine Sediments (STP 883), Eds. R. C. Chaney and K. R. Demars, ASIM, Philadelphia, 1985, pp. 41-4 Jones, G. and Rust, E., "Piezoneter Penetration Testing", Proceedings, 2nd European Symposium on Penetration Testing, Vol. 2, Ansterdam, 1982, pp. 607- 613. Robertson, P. K., Campanella, R. C., Gillespie, D., and Grieg, J., “Use of Piezometer Cone Data”, Use of In-Situ Tests in Geotechnical Engincering (GS? No. 6, Bd. S. P. Glemence, ASCE, New York, 1986, pp. 1263-1280. Marchetti, S., "In-Situ Tests by Flat Dilatometer", Journal of the Geotechni. eal Engineering Division, ASCE, Vol. 106, No. G3, Mar. 1980, pp. 299-321. Schmertmann, J., "Suggested Method for Performing the Flat Dilatoneter Test", Geotechnical Testing Journal, ASTM, Vol. 9, No. 2, June 1986, pp. 93-101. Hough, B. K., Basic Soils Engineering, 2nd Ed., Ronald Press, New York, 1969, 634 p. Terzaghi, K. and Peck, R. B., Soil Nechanics in Engineering Practice, 2nd Ed., John Wiley and Sons, New York, 1967, 729 p Gibbs, H. J. and Holtz, W. G., "Research on Determining the Density of Sands by Spoon Penetration Testing", Proceedings, 4th International Conference on Soil Mechanics and Foundation Engineering, Vol. 1, London, 1957, pp. 35-39. Holtz, W. G. and Gibbs, H. J., Discussion of “SPT and Relative Density in Coarse Sand", Journal of the Geotechnical Engineering Division, ASCE, Vol. 105, No. GT3, Mar. 1979, pp. 439-461, 2-40 Cloy Slayer sits: Sandy sit oy sity ley Gs! Sity sond Sond Dota from 18 sites 0.001 0005001 005 01 os! Mean Particle Size, Os (mm) Figure 2-29. Variation of qo/N with Grain Size for Electric and Mechanical Friction Cones Source: Robertson and Campanella (17), p. 730 Robertson ond Campcnello, 1983 (17) Zervogionnis and Kalteriotis, 1988 (45) Chin, ef al, 1988 (46 Jamiotkowski, et al., 1985 (44) Andrus and Youd, 1987 (50) Kosim, et ol., 1986 (47) Seed ond deAlba, 1986 (48) Muromachi, 198! (49) eemrax0d Robertson and Componella, (Figure 2-29) \ ok (aq /99)/N*5 44 D502 {ns197, 220,702, S. 0.001 0005 0.01 0.05 OF ost 5 10 Mean Particle Size, Dso (mm) Figure 2-30. Recommended Variation of q¢/X with Grain Size for Fugro Electric Friction Cones recommended relationship is given by the solid line and regression equation on the figure In other studies, the ratio of qo/N has teen correlated to the percentage of fines (clay and silt sizes). For example, Jamfolkowski, et al. (44) indicate the trend presented in Figure 2-31 for Italian soils, In addition to these data, other available data were summarized (46 47, 49) to substantiate 2 general trend between the q¢/N ratio and fines content, as shown in Figure 2-32, Use of Figures 2-20 and 2-32 will provide the best estimate relationship between qq and N, with decreasing with increasing fines content. the ratio 6 a7. ‘No. experimental e points g 3 . °. ° 20 40 60 80100 Percent Passing No. 200 Sieve Figure 2-31. Variation of qo/N with Fines Content Source: Jamiolkowski, et al. (44), p. 1895. 19) © Jomiotkowski, et al., 1985 (44) 4 Kasim, et all, 1986 (47) 8 © Muromachi, 1981 (43) 4 E Chin, et ol. 1988 (46) Percent Passing No. 200 Sieve, F Figure 2-32. Recommended Variation of qc/N with Fines Content 2-28 30 31. 32. 33 34. 35 36 37 38 39. 40. a. 42. 43. Mareuson, W. F., III and Bieganousky, W. A., "SPT and Relative Density in Coarse Sands", Journal of the Geotechnical Engineering Division, ASCE, Vol. 103, No. GTIL, Nov. 1977, pp. 1295-1309. Skempton, A. W., "Standard Penetration Test Procedures and the Effects in Sands of Overburden Pressure, Relative Dereity, Particle Size, Aging, and Overconsolidation", Geotechnique, Vol. 36, No. 3, Sept. 1986, pp. 425-447 Robertson, P. K., Campanella, R. G., and Wightman, A., "SPT-CPT Correlations", Journal of the Geotechnical Engineering Division, ASCE, Vol. 109, No. 11, Nov. 1983, pp. 1649-1459 Liao, S. S. and Whitman, R, V., “Overburden Correction Factors for SPT in Sand", Journal of Geotechnical Engineering, ASCE, Vol. 112, No. 3, Mar. 1986, pp. 373-377. Peck, R. B., Hansen, W. E., and Thornburn, T. H., Foundation Engineering, 2nd Ed., John Wiley and’Sons, New York, 1974,'514 p. Seed, H. B., "Soil Liquefaction and Cyclic Mobility Evaluation for Level Ground During Earthquakes", Journal of the Geotechnical Engineering Division, ASCE, Vol. 105, No. 2, Feb. 1979, pp. 201-255. Bieganousky, W. A. and Marcuson, W. F., III, "Liquefaction Potential of Dans and Foundations: Laboratory Standard Penetration Tests on Reid Bedford Model and Ottawa Sands", Research Report $-76-2 (Report 1), U.S. Army Engineer Waterways Experiment Station, Vicksburg, 1976, 156 p. Bieganousky, W. A. and Marcuson, W. F., III, "Liquefaction Potential of Dans and Foundations: Laboratory Standard Penetration Tests on Platte River Sand and Standard Concrete Sand", Research Report $-76-2 (Report 2), U.S. Army Engineer Watervays Experiment Station, Vicksburg, 1977, 87 p. Feck, R. B. and Bazaraa, A. R. S., Discussion of "Settlement of Spread Foot- ings on Sand", Journal of the Soil Mechanics and Foundations Division, ASCE, Vol. 95, No. S43, May 1969, pp. 905-909 Barton, M. E., Cooper, M. R., and Palmer, S. N., "Diagenetic Alteration and Micro-Seructural Characteristics of Sands", Penetration Testing in the U.K. Thomas Telford, London, 1988, pp. 57-60 Tokimatsu, K., "Penetration Tests for Dynamic Problems", Proceedings, Ist International Symposium on Penetration Testing (ISOPT-1}, Vol. 1, Orlando, 1988, pp. 117-136. Robertson, P. K. and Campanella, R. C., "Estimating Liquefaction Potential of Sands Using the Flat Plate Dilatoneter", Geotechnical Testing Journal, ASTM, Vol. 9, No. 1, Mar. 1986, pp. 38-40 Schzertmann, J. H., "Measurement of In-Situ Shear Strength", Proceedings, ASCE Conference on In-Situ Measurement of Soil Properties, Vol. 2, Raleigh, 1975, pp. 57-138. (closure: pp. 175-179). Szechy, K. and Varga, L., Foundation Engineering - Soil Exploration and Spread Foundations, Akademiai Kiado, Budapest, 1978, 508 p. 4, 4s. 46. 47. 48, 49 50 Jamiolkowski, M., Baldi, G., Bellotti, R., Ghionna, V., and Pasqualini, E., "Penetration Resistance and Liquefaction of Sands", Proceedings, llth Incer- national Conference on Soil Mechanics and Foundation Engineering, Vol. 4, San Francisco, 1985, pp. 1891-1896. Zervogiannis, C. $. and Kalteziotis, N. A., "Experiences and Relationships from Penetration Testing in Greece", Proceedings, 1st International Symposium on Penetration Testing (ISOPT-1), Vol. 2, Orlando, 1988, pp. 1063-1071. Chin, C. T., Duann, $. W., and Kao, T. C., "SP-CPT Correlations for Granular Soils", Proceedings, 1st International Symposium on Penetration Testing (1s0PT"1), Vol. 1, Orlando, 1988, pp. 335-339. Kasim, A. G., Chu, M. Y., and Jensen, 6. N., "Field Correlation of Cone and Standard Penetration Tests", Journal of Geotechnical Engineering, ASCE, Vol 112, No. 3, Mar. 1986, pp. 368-372 Seed, H. B. and de Alba, P., "Use of SPT and CPT Tests for Evaluating the Liquefaction Resistance of Sands", Jse of In-Situ Tests in Geotechnical Engi- neering (GS? 6), Ed. S. P. Clemence, ASCE, New York, 1986, pp. 281-302 Muromachi, T., "Cone Penetration Testing in Japan", Cone Penetration Testing and Experience, Eds. G. M. Norris and R. D. Holtz, ASCE, New York, 1981, pp 49-75 Andrus, R. D. and Youd, T. L., "Subsurface Investigation of a Liquefaction- Induced Lateral Spread, Thousand Springs Valley, Idaho", Miscellaneous Paper GL-87-8, U.S. Army Engineer Waterways Experiment Station, Vicksburg, 1987, 131 > Section 3 IN-SITU STRESS STATE In most geotechnical engineering problems, a knowledge of the in-situ state of stress is necessary for two reasons. First, these stresses represent the original conditions onto which any engineered construction imposes stress increments. These initial through final stress conditions are used to evaluate the overall engineer- ing performance of the constructed facility. Second, nearly all engineering prop- erties of soil are a function of the soil stresses, either directly or indirectly. Therefore, the stresses are needed to evaluate the soil properties In this section, procedures are presented to evaluate the in-situ stresses in both cohesive and cohesionless soils. Vertical stresses are covered first, followed by horizontal stresses. In each case, correlations are presented with soil index parameters and in-situ test results, where available. BASIC DEFINITIONS The in-situ state of stress in soil is defined in terms of the current values of effective vertical stress (ayo) and effective horizontal stress (Gyo). For hori- zontal, level ground, the in-situ stress state is shown in Figure 3-1 The current vertical stress is determined in a straightforvard manner, being equal to the effective overburden stress in which dy) = 72. However, the horizontal stress is more difficult to evaluate. The stress ratio is Ko, the at-rest coeffi- cient of horizontal soil stress, vhich is defined as dho/Byo. AS a lower bound, Ky could equal Kg, the coefficient of minimum active soil stress. The upper bound for Ko is Kp, the coefficient of maximun passive soil stress. For horizontal, level Effective unit weight i Figure 3-1. Stresses in Soil 341 ground and an effective stress cohesion () = 0, these limit states are given by Rankine theory as below: 1+ sia Gpse fp = pe ~ LTE Pose en 1” sim Gose in which pec ~ effective stress friction angle for plane strain compression conéi- tions. Using these Limits for a cohesionless ofl with Fpse - 40", for example, Ko could range from 0.2 to 4.6. Many factors affect the in-situ state of stress in soil, including overconsolida- tion, aging, chemical bonding, etc. Overconsolidation is probably most influen- tial for the majority of soils, because it is caused by glaciation, erosion, desic- cation, excavation, ground water fluctuations, and possibly other factors. In this regard, the effective vertical preconsolidation stress (denoted 7p, Bmax» OF Be) is an important measure of the soil stress history. This maximum past stress affects the compressibility, strength, censistency, and overall state of stress It is often convenient to represent the stress history in terms of a dimensionless parameter defined as the everconsolidation ratio (OCR): OCR = Fp/Byo (3-2) ‘The magnitude of 3p and OCR can be evaluated directly from the results of one-di- mensional consolidation tests conducted on undisturbed cohesive soil samples. Cor- relations with other tests and soil types are presented in this section. The magnitude of Ko may be measured directly either in the laboratory using special vesting equipment, or in the field using devices such as the pressureneter or total stress cells. However, these direct methods may be subject to unavoidable distur- bance effects during sampling and in-situ testing. Alternatively, several empiri- eal approaches can be used to evaluate the in-situ value of Ko, including: (1) reconstruction of stress history, (2) correlations with soil index parameters, and (3) correlations with in-situ test results. All three approaches are described in this section. RECONSTRUCTION OF STRESS HISTORY Reconstruction of the soil stress history involves tracing the stress paths of the soil as in Figure 3-2, from virgin loading, to primary unloading, to primary Primary untoading Figure 3-2. Stress Paths for Simple Stress Histories reloading, and then cyclical load-unload looping from water table fluctuations, etc. (above point E in figure). Virgin loading represents normally consolidated (NC) soils with ocR = soils with ook > 1. All other stress paths represent overconsolidated (0C) Based on a study of 171 different laboratory-tested soils, Mayne and Kulhawy (1) shoved that a general equation can be used to model stress paths OB-BD-DE, as given below: ck, ocr ie tal OCR ax ane Ko = Kone [ v1 G3) in which Kone = Ko during virgin (normally consolidated) loading, a ~ at-rest unload coefficient, m, = reload coefficient, OOR ~ current overconsolidation ratio, and 0¢ ~ maximum past OCR (e.g., point D for a soil currently at point £). % 8 For virgin loading, the simplified Jaky equation (2) provides reasonable estimates for Kone, at given below: Kone “1+ sin Bee on in which Jee = effective stress friction angle for triaxial compression. Figure 3-3 shows this equation to be a reasonable estimate for a wide range of soils, In this figure, Kone vas determined fron oedometer or triaxial tests. During rebound or unloading, the general relationship for Ky is often expressed as: Ko = Kone O0R* 3-5) eo Roca ea st (n2124, 120.707, S.0.=0.06) 4 aN . Doge ee Kone N, SS sin dy_) #0. os 04] 02} Horizontal Stress Coefficient for NC S 04 06 sin Bre igute 3-3. Rorleontal Seraso Cocffctant for NC Solis frou Laboratory Teste Source: Mayne and Kulhawy (1), p. 862. As suggested by Schmidt (3), the exponent @ may be expressed as a function of Fre! @~ sin dec 3-6) Alternatively, the exponent may be expressed as: @-1- Kone on which also appears reasonable, as shown in Figure 3-4. For reloading, the stress path from D to E in Figure 3-2 may be approximated as a straight line with slope m = Who/tvo. Review of laboratory data from 35 soils (Figure 3-5) indicates that the reload coefficient can be estimated adequately from: By = 0.75(1 - sin dee) @-2) Linear regressions on these data for m, give 0.76 Kone (x2 = 0.583 and $.D. = 0.06) and 0.77(1 - sin dec) with x2 = 0.534 and S.D. = 0.06. Regression 91.00 (I-Kone) 2 log K, (unloading) 9 log OCR ° O2 04 06 08 10 Horizontal Stress Coefficient, Kone Figure 3-4. Unload Coefficient for 0¢ Soils Source: Mayne and Kulhavy (1), p. 864. 19 Clays “ 15 Clays O8F 16 Sands 7 C 13 Sands LY 7 ol “1 oa o2| oat 1m,#0.75 Kygc¥0.! T my20.75(I-sin Fy_)£0.! ° ol 0 OR Oa OSB CCC Horizontal Stress Coefficient, Kone sin Pte Figure 3-5. Reload Coefficient for 0C Soils Source: Mayne (4), p. 269. Combining the above relationships gives Ko = > sin ee) [8 + Baa» ge) OcRpay(t * #1" Bee) D1 3-9) in which OCRggx is the OOR at point D in Figure 3-2. For primary unloading, OcR = OCRnax and therefore Ko = (1+ sin Bee) ocr?” te» Kay (3-10) For virgin loading, OCR = 1 and therefore: Ko = 1 - sin $te = Kone (3-11) Most natural soils have undergone a stress history of loading-unloading-reloading, and therefore Ky is likely to be within points ¢ and E in Figure 3-2. Therefore, Ky at point E és an appropr: te lover bound for the in-situ Xj. All that is needed £2 Fer OR, and OCRgqx, which can be evaluated by direct laboratory measurements, geologic generalization of the soil stress history, or experinental test prograns Sn-situ to establish the values, It should be noted that, if an NC assumption is made (Equation 3-11), it will underestimate K, in the majority of soil deposits. One last point to mention regarding Figure 3-2 is that the soil can reach passive failure during primary unloading if the vertical effective stress is reduced suffi- ciently. This limit state can be developed from Equations 3-1 and 3-10 and is given by: OR init = [CL + sin Fpgc)/(L ~ sin Bpge)(L ~ sin dee)) 1/8" Fee) G2) ‘As shown in Section 4, psc ~ 1.1 Bcc. If this Lime state is reached, soft fail- ture occurs, and the stresses change. Tt ie uncertain vhat this stress state may be, although Kp might approach 1 EFFECTIVE PRECONSOLIDATION STRESS IN COHESIVE SOILS Cohesive soils consolidate and stiffen during overconsolidation and effectively retain a "memory" of the largest preconsolidation stress (3p) to which they have been subjected (e.g., point B in Figure 3-2). This process was illustrated quali- tatively in Figure 2-1 as a function of the water content and Atterberg limits. ‘Therefore, these index paraneters represent a starting point for estimating Bp. Correlations with in-situ test results follow the index parameter correlations. Details on the in-situ test strength parameters are given in Section 4. Correlations with Index Parameters The effective preconsolidation stress (p) has been correlated with the Liquidity index by several authors. A recent analysis of laboratory consolidation test data by Stas and Kulhawy (5) suggested the following ap/pq = 201-11 = 1.62 L1) 2-13) in which pg = atmospheric stress in the desired stress units and LI ~ liquidity index. This equation is based on 150 data points for clays with a sensitivity between 1 and 10. This relationship has 2 standard deviation of 0.33 and x? equal to 0.740. Other generalized relationships are shown in Figure 3-6, which gives the precon- solidation stress as a function of Liquidity index (LI) and sensitivity (S_) For comparison purposes to evaluate the soil stress history, the effective vertical stress (@yo) is needed. This stress can be evaluated directly as in Figure 3-1, or it can be estimated from the Liquidity index. Based on the modified Cam clay model and empirical observations, Wood (Z) developed the following approximation for @yo? Byo/Pa = 0.063+102(1-LT) 14) Liquidity Index, LI ° 02 005 01 0S I Suiowsso Preconsolidation Stress, & /Pg Figure 3-6. Generalized ap - Liquidity Index - Sensitivity Relationships Source: NAVFAC (6), p. 7.1-142. 37 Although this equation strictly applies only to insensitive soils at the critical state, it is a useful approximation for uncemented, low sensitivity soils. Comments on Field Test Correlations It should be noted that the following figures correlating 3p with field test meas- urements are presented all in similar form, on log-log plots because of the range in the parameters involved. These figures were developed from the sources noted, and the symbols used correspond to the clay types referenced in the source papers For each figure, the number of intact ard fissured clays is noted, the fissured clays are located separately because their behavior is different, and a linear regression equation is presented for the intact clays only. ‘The regression was done assuming a linear, arithmetic relationship through the origin, The statistics given with each regression include the number of data points (n), coefficient of determination (x2), and the standard deviation of 3p (S.D.) for a given field test measurement. The given relationships should be used only as predictors for Correlations with VST Strength The field vane shear test (VST) has been used for many years as an estimator of Bp. In 1957, Hansbo (8) developed the following equation for Swedish clays: 1p = aust Su(VST) (3-15) in which aygp is an empirical factor approximately equal to 222/w,, with wy = liquid Limit (in percent) A more recent compilation of worldwide clays, shown in Figure 3-7, indicated the general nature of Equation 3-15. This study further showed that aysr could be correlated weakly with the plasticity index (PI), as shown in Figure 3-8 Correlations ch SPT N Value The standard penetration test (SPT) N value may be used to provide a first-order estimate of dp for cohesive soils. Figure 3-9 shows the available data for 51 clays. The regression shows a fair correlation vith a relatively large standard deviation. It should be noted that the reported N values have not been corrected for the fac- tors which significantly affect the SPT N value. Until the N values are corrected to 2 consistent standard, the SPT is likely to be of limited use in evaluating ap 38 50 fissured“ = 20 é £3.54 ay (UST) a (n=205, r?=0.832, 1 tok ; . Sea P10 Snobs eet g sk DO: a 5 2 Zoi 4 5 osk * 4 é og 96 toe! cays o2| / r I fissured eloy - fm baal aC Field Vane Strength, sy (VST)/p. Figure 3-7. dp Correlated with VST sy Based on Mayne and Mitchell (9), p. 154, and others (10) 5 8 ayst* % * 20 40 65 80 Too Plasticity Index, PI (%) Figure 3-8. Field Vane Coefficient versus PI Source: Mayne and Mitchell (9), p. 154. 39 8 = 8 58 fation Stress, &/Pq Preconsoli ° i 49 intoct clays 2 fissured clays L Lesa Losl 2 5 10 20 80 100 200 SPT N Volue (biows /ft or 305mm) °° ° Figure 3-9. 3p Correlated with SPT N Based on Mayne and Kenper (11), p. 144, and others (12). Correlations with CPT qq Value The cone penetration test (CPT) tip resistance, qc, has been used effectively to Profile the preconsolidation stress in clays. Figure 3-10 presents the available ata from 49 clays. This correlation is somevhat better than with the N value, and the standard deviation is smaller. This correlation also shows nore clearly that the fissured clays behave differently from the intact clays. However, it is impor- tant to note that the data in Figure 3-10 are not corrected for pore water stress effects Correlations with OPTU Results The piezocone (CPTU) provides additional data during penetration and generally is considered to be a more sensitive type of cone penetration test. Tavenas and Leroueil (20) demonstrated that the preconsolidation stress (3) was well-corre- lated with the net corrected cone tip resistance (ap - ovo) for eleven Canadian clays. A larger sample of piezocone dats is shown in Figure 3-11. The regression in this case gave an even higher r2 with lower standard deviation. In addition to measurenents of cone tip resistance, piezocones provide the g fesree ~~ Pigg 8) | Bo L . ' 4 20F & 20.2946 ae? ok (ort, 0.888, CEOS 4 ) sk 4 2b al Preconsolidation Stress, o)/Pq 39intoct cloys id lOfissured clays | Obed bl os | 2 5 10 20 50 100 200 Cone Tip Resistance, a./Pq Figure 3-10. 3p Correlated with CPT ac Source: Based on Mayne (13), p. 786, and others (14 - 19) so € 2 1 =o 2 sE (n=74, 20.904, a S.0.=1.02 pg) 5 2 eas & Eos = 26 intact cloys i oa| 9 issued clays 1 02 05 1 2. 5 10 20 50 100 Net Cone Tip Resistance, (¢7-o,9)/Pq Figure 3-11, &p Correlated with CPTU qr Source: Based on Mayne and Holtz (21), p. 25, and others (14 aL magnitude of pore water stress (Au) caused by penetration. A relationship between ap and dur from CPTU tests with tip or face pore water stress measurements is shown in Figure 3-12. For pore water stress measurements behind the tip, the relation- ship is given in Figure 3-13. The resuits are similar for the intact clays. How- ever, for piezocones in heavily overconsolidated fissured clays, pore water stresses measured behind the tip are near zero and sometimes are even negative. On the cone tip, positive pore water stresses are observed for all clays at all OCR values, regardless of whether fissuring is present. From cavity expansion theory, the general relationship between dp and the excess Pore water stress measured at the tip during plezocone penetration can be given by the following (23): Bp/au = 3/(4 In T,) (3-16) In which M ~ critical state parameter (Appendix 6) and I, = rigidity index (6/sy). For measurements behind the tip, the coefficient 3 becomes equal to 4. This equa tion gives values consistent with those in Figures 3-12 and 3-13 for the intact 50 T T T T Pah e? T fissured ——e} Sg / 3 20h a 4 < ig / 1 tok be S/ 4 2 520.47 Buy é a ee 3 & sociated 5 2 4 3 a3 2 ob 4 8 ost Tip/toceZeor 4 & L 37 intact cloys S fissured clays otbo et 0208125 1020-30100 Excess Pore Water Stress from Piezocone, Auy/Pg Figure 3-12. dp Cerrelated with CPTU dug Source: Based on Mayne and Holtz (21), p. 23, and others (14, 3-12 50, area ge ogee 8 | fissured [=e 0) Wa 4 we 6 8 ost Behind 1 . pine feo" & ve o2k 7 intact clays | 5 fissured clays on basil La 1 oz 05 1 2 5 10 20 50 Excess Pore Water Stress from Piezocone, Ausy/Pg Figure 3-13. ap Correlated with CPTU aime Source: Based on Mayne and Holtz (21), p. 24, and others (14, 15, 17, 19, 22). clays Correlations with PMT Results Several correlations have been attempted with the pressureneter test (FMT) to esti- mate the value of 3p. Early work with the Menard pressuremeter indicated that the PMT creep pressure was approximately equal to dp for Chicago area lake clays (24) Later work showed that the limit stress from the self-boring pressuremeter test (SBPNT) could be correlated with 3p, as shown in Figure 3-14. Other studies have shown the correlations given in Figure 3-15, including the undrained shear strength (sy) and the rigidity index (Ty) Correlat: The initial contact stress (po) from the dilatoneter test (DMI) is a measure of the induced total pore water stress caused by insertion of the DAT blade. Analogous to the previous relationship between 3p and su for piezocone tests, a similar rela- tionship applies for the DMT between 3p and (pe - Up), as shown in Figure 3-16 313 og tasured—ey . Eso ff 7 aC {n=89, 120.908, \__ 1” gh s:0-0.28p0 aK j 8 e ‘2 i se ° By = ere 4 8 ost 4 é . 35 Sc ot woe ce | on 1 1 1 4 7 1 a a a SBPMT Limit Stress, p_/p, Figure 3-14, ap Correlated with SBPMT py Source: Data from Mayne and Kulhavy (25), and others (12, 16, 19, 26). EFFECTIVE PRECONSOLIDATION STRESS IN COHESTONLESS SOILS Cohesionless soils also consolidate and stiffen during overconsolidation and retain a “nemory" of the preconsolidation stress. However, cohesionless soils are diffi- cult to sample and test in the laboratory in the undisturbed state, and therefore little correlation information is available to estimate the preconsolidation stress in these soils. More work has focused on evaluating OCR and Ky directly, as described later. OVERCONSOLIDATION RATIO FOR COHESIVE SOILS Im Lieu of describing soil stress history by the preconsolidation stress (3p), the in-situ everconsolidation ratio (OCR) may be estimated directly using normalized parameters developed from laboratory or field test measurements. These correla- tions strictly apply only to insensitive clays. Furthermore, the sane comments nade previously on field test correlations with respect to 3p also apply to the OCR ° fissured A 20 a fawn a/’ BY y=0.76 (sy/oq) In Ty, € 19 (n=105, 2#0.895, (7 = $.0.#0.85 56) & g 4 é oa 5 Der 2 ? cee i e so £ os é f 02 [ 6 fissured clays on (Oz yeios eae ee Sima com=eeso, 38 intact cloys (sy/pq) In I, from SBPMT Figure 3-15. p Correlated with SBPMT sy and I; Source: Based on Mayne and Bachus (23), p. 293, and others (12, 16, 19, 26). correlations. Correlations with Index Parameters Equation 3-13 can be re-cast in terms of OCR as follows: OcR = (Pa/Byo) 10L-11 - 1.62 LI) @-17) As noted previously, this relationship is based on statistical anslysis of labora- tory consolidation test data on clays with sensitivity from 1 to 10. Based on the modified Cam clay model and empirical observations, Wood (7) developed Equation 3-14 to correlate dyo with LI. He also developed the following: Log OCR = [2 - 2 LI - log (15.87 Byo/Pal/A G18) 50) ToT toon —— ca YY 0.51 (Pore) Po 6 T o 2 4 L “ (n=76, 150.896, E 2 sk 4 a § 2b 4 2 ib 4 8 ose 4 i 24 intact cloys 02 7 fissured cloys ote eben tel Ld Effective DMT Contect Stress, (p,-u.)/p, Figure 3-16. 3p Correlated with DMT po Source: Based on Mayne (27), p. 148, and others (26) in which A = critical state parameter (Appendix G). Using a typical value of A = 0.8, ané combining Equations 3-14 and 3-18, results in the following: mu yolt - 2-5 LT - 1.25 10g (yo/Pa) (3-19) Although this equation strictly applies only to insensitive soils at the critical state, it is a useful approximation for uncemented, low sensitivity soils, as noted previously Correlations with Laboratory Strength Laboratory undrained shear strength (sy) data may be used to estimate the in-situ OCR of clays. Using empirical observations from isotropically and anisotropically consolidated triaxial compression tests, Mayne (28) observed the following for OCR: OCReTUC = [(Sy/Fvo)/0-75 sin Bee ]t-43 (3-20) Ocecaue = [(#u/Pv9)/0.67 sin beg)! 3 (21) 3-16 ‘These results are consistent with the modified Cam clay model, which would predict the following: OCR = 21 (sy/By9)/0-5 M)-1/A (3-22) Correlations with VST Strength ‘The undrained strength from the field vane shear test (VST) may be related to the in-situ OGR according to: OCR = aysr (Su/evo) VST (3-23) in which aygp has been shown in Figure 3-8 to be related weakly to plasticity index (PI). Figure 3-17 shows a direct relationship between OCR and sy/ayo for 96 clays. Corel ons with SPT N Value Attempts have been made to correlate the SPT.N value with OCR. Figure 3-18 is typ- feal of these correlations, using uncorrected N values. This relationship is only a first-order estimator 50 OCR#3.22 (54/Fghvsr (n=209, r=0.806, 8 3 Overconsolidation Ratio, OCR 96 intact clays 10 Normalized Field Vane Strength, (5/0) vst Figure 3-17. OGR Correlated with VST sy Source: Based on Mayne and Mitchell (9), p. 152 317 59; 5 8 20] 3 OCR+058 Npg/F yp OE tnalt2, e2=0.661, & $.0.=3.82) Zs > 8 arr ph fF 3 Fo 4 2 o 200 48 intact clays ~] 2 FE oo 3 fissured clays V 2 5 10 20 30 100 SPT N Volue,N pg/yo (blows/ft or 305mm) Figure 3-18. OCR Correlated with SPT N Source: Based on Mayne and Kemper (11), p. 143, and others (12). Correlations with CPT and CPTU Results A number of authors (e.g., 13, 29) have denonstrated that OCR correlates with the OPT qe value through the normalized cone tip resistance, (q¢ - cyo)/Svo- However, Ge also should be corrected for pore water stresses acting on unequal areas of the cone. Figure 3-19 shows the variation of OGR with the corrected cone tip resis- tance, gz, as obtained from piezocones. Other piezocone studies (31) suggested a general trend with By (Equation 2-4) and OCR that was strongly dependent on the rigidity index. However, By is so site- dependent that the relationship was of little predictive use. More recent work (32) considered a combined critical state/cavity expansion model to correlate OCR with piezocone results. However, at the present time, the relationship given in Figure 3-19 probably is most appropriate to use. Gorrelations with DS Results Im the initial introduction of the dilatometer test, Marchetti (33) proposed the correlation in Figure 3-20 between OcR and the DMT parameter Kp, given by: 8 T TF 7 s 5 S OCR=0.32 (47-49) Gyo ni (o=161, 220.762, © g $.0.#0.76) é ™ a Sa B 3 6 52 intact cloys L L 1S 20 25 Corrected Cone Tip Resistance, (47-0) /Gyg Figure 3-19. OGR Correlated with CPTU qr Source: Data from Mayne (30), and others (18, 19, 22) OCR = (0.5 Kp)}-56 (3-24) in which Kp = horizontal stress index = (Po - o)/@yo, Po = initial contact stress, ty = hydrostatic pore water stress, and Gy, ~ effective vertical stress. Subse- quent research with the DMT in other countries suggests a more general expression: OCR = (Bo Kp)}-56 (3-25) in which the parameter fo depends upon the degree of fissuring, sensitivity, and geologic origin, as shown in Figure 3-21. OVERCONSOLIDATION RATIO IN COHESIONLESS SOILS It is difficult to estimate the in-situ OOR of natural sand deposits. The best approach is through a detailed geologic study to evaluate the stress history of the formation. Indirectly, oedometer tests on interbedded clay strata or seams may give clues to the in-situ OCR of the surrounding sands. With the DMT, a value of OCR in sands can be back-calculated from the estimated Ko as (Bullock, 36): 100) Fissured clays aeeash \ Insensitive cloys (89*0.50) origina! Morcheti] 50; 8 8 8 6 Overconsolidation Ratio, OCR 3 Sensitive cloys Overconsolidation Ratio, OCR (Bo=0.35) H Lavi i a a) 2. S10 20 Horizontal Stress Index, Ky Horizontal Stress Index, Kp Figure 3-20. OCR Correlated with DMT Figure 3-21. OCR - Kp Relationships for Kp Clays of Varied Geologic Origin Source: Marchetti (33), p. 315. Source: Based on Marchetti (33), Powell and Uglow (34), and Lacasse and Lanne (35). OCR = [Ko/(1 - sin $¢¢)) 729/84 Fee) (3-26) which is a form of Equation 3-10 that has been rearranged and modified to fit the results of laboratory calibration chanber tests on sands. EFFECTIVE HORIZONTAL STRESS IN COHESIVE SOILS As noted previously, soils retain a "memory" of preconsolidation. With vertical stresses, this menory is reflected by the preconsolidation stress (3p) which, in 0C soils, is greater than the effective overburden stress (8y). In the horizontal Sirection, the process is somevhat different, because the soil can not unload as freely as it can in the vertical direction. The result is that the retained memory of the maximum horizontal effective stress is less clear. If the soil is young and has experienced only a relatively simple stress history, then the procedures described earlier under "Reconstruction of Stress History" can be used to evaluate the horizontal effective stress in terms of Ko, defined as Jpo/Byo. For older 3-20 soils or sofle with more complex stress history, the reconstruction process can be nore difficult. By default in these cases, it my be necessary to assume only pri- nary unloading, as shown in Figure 3-2, This assumption will result in an upper bound on Ko, which must be used with some considered engineering judgment, taking into account the loading level and differences between the virgin loading and pri- mary unloading values of Ky. Alternatively, Ko may be estimated from index peraneters or correlations with in-situ measurements. Ideally, these approaches reflect the soil in-situ and therefore should be good indicators of the current Kj. However, all correlations contain uncertainties and must be considered within the context of the stress history of the soil. The predicted Ko should be consistent with this information Gorrelations with Index Parameters A number of studies have attempted to correlate Ky with the Atterberg Limits. Figure 3-22 shows one of these relationships for NC clay. As shown, organic clays should be excluded from the general trend. For 0¢ soils, an early study demonstra- ted the behavior shown in Figure 3-23, with the overconsolidation ratio (OCR) domi- nating the resulting Ky value. These two figures suggest a high degree of correla- tion with the Atterberg limits. However, more comprehensive data compilations show the lack of correlation given in Figure 3-24, which has an r? equal to 0.147 One simple alternative estimator is to assume overconsolidation by simple unload- ing, which was described previously as: Lo m7 1 ‘ Kone 4 05) ° al ‘© Inorganic clays © Organic clays ° ° ° 40 80 ° 40 80 120 160 Plasticity Index, PI (%) Liquid Limit, w, (%) Figure 3-22. Kone Correlated with Atterberg Limits Source: Larsson (37), p. 21. 21 20 40 60 Plasticity Index, PI (%) Figure 3-23. Ko Correlated with PI and OCR Source: Brooker and Ireland (38), p. 14. 0.27-— —r © Fre < 20" A 20°< die < 30" cei Vv 30°< gee < 40°] Oo | mw b> 40" b a 7 “fad Ake An Kone 0.6 ¥ eae Sage &. aks v ya os Be RA * oy J Vea? s v oat ave - v S| 135 clays z CY ost ys ° 20 30 60 80 Plasticity Index, PI (%) Figure 3-24. apparent Lack of Trend Between Kone and PI for 135 Clay Soils Ko = (1 = sin Bee) oon®® Fee G-10) Figure 3-25 illustrates this approach for 48 clay soils. Also, the following approximation was given earlier: 322 2.5 T - 48 cloys Brett 95g L Qe 20h 4 ° ® s / ol 1 1. 1 1 v 2 S 10 20 50 Overconsolidation Ratio, OCR Figure 3-25. Ko Correlated with OCR Source: Data from Mayne and Kulhavy (1). OCR = (pg/Byq) 10TH - 2-62 LE) 17) By combining these two equations, Ko can be estimated simply from a knowledge of Beer Bvo» and LI One further simplification is to note that de = 20° is a reasonable fit of the data in Figure 3-23. Using this value, Equation 3-10 reduces to: Ko = 0.5 0cr0-5 (3-27) Then, combining this result with Equation 3-17 yielés: Ko = 0.51 (Pa/B yo) 1000.56 - 0.81 L1)} (3-28) which is a simple, first-order estimator requiring only dy and LI. If information is available for the undrained shear strength (s,), then the corre- lation shown in Figure 3-26 can provide an estimate for Ko. Direct Correlations with SBPMT and DMT Results ‘The self-boring pressuremeter test (SBPMT) has shown promise as one of the few devices capable of providing a direct measurement of the in-situ horizontal stress. There is no need for correlations because the stress is measured directly, taking into account equipment calibrations. Figure 3-27 shows results summarized for 56 clays in the literature, in which both Kg and OCR values were given, As can ‘be seen, the trends are consistent with those shown previously (Figure 3-25) for laboratory data, It should be noted thet the fissured and intact clays behave sim- ‘larly when tested with the SBPMT because this test involves an expanding device which compresses the soil and fissures to mimic an intact soil. —1—t —+ (suBlog}® isp S05! (wae) pa (n=49, r2=0.930, v $070.09). Sp 2 13 clays CKGUC tests rr ee 3 5 7 (5y/Boe (su /Gve)ne Source: Modified after Mayne (39) 3-26 go Ko#0.47 OCR ry —(nv59, 120.763, : i} 8.0.=0.83) e os fB Bo 4 K, from SBPMT | sé intact cloys oo (2 fissured clays ° 02 1 Joset 1 1 iaeeeee| Ss 10 20 50 100 OCR from Oedometer Figure 3-27. Ko from SBPMT Correlated with OCR Source: Based on Mayne and Kulhawy (25 and 40). ‘The original intent of the dilatometer test (DMT) was to model the soil modulus for the laterally loaded pile problem, which requires an assessment of the horizontal stress, However, all in-situ testing devices cause some disturbance upon insertion into the ground. Therefore, Marchetti (33) found it necessary to develop 2 corre- lation between a best estimate Ky and the DMT horizontal stress index (Kp), as shown in Figure 3-28. The original Marchetti equation was based primarily upon data from insensitive Italian clays and uncemented normally consolidated sands and was given as: Ko = (Kp/1.5)9-47 - 0.6 (3-29) Powell and Uglow (34) tested heavily overconsolidated and fissured clays from the United Kingdom with the DMT and found that, although the in-situ Ky trended with Kp, the relationship was offset from the original one established for Italian clays, Similarly, Lacasse and Lunne (35) used the DMT at several Norvegian sites 3-25 | T T T Fissured clays (Br08) Ne 2b 4 Insensitive cloys (Be=1.5) original IL Marchetti] 4 ost 2 4 Sensitive clays (222.0) °. L n L L = 2 5 LOB 20) Horizontal Stress Index, Ky Figure 3-28. Ke Correlated with Kp Source: Based on Marchetti (33), Powell and Uglow (34), and Lacasse and Lunne (35) and suggested further modifications to the original Marchetti correlation. Both data sets also are shown in Figure 3-28. Considering these other data, a general equation for Ky is: Ko = (Rp/x)°-47 - 0.6 (3-30) in which fj depends upon soil type and geologic origin. Where possible, local calibration of the DMT should be made relative to Ky measure- ments obtained with SBPMT or push-in space cells. For preliminary estimating pur- poses, the values of f, in Figure 3-28 may be used. Figure 3-29 shows a direct comparison of K from the SBPAT with Kp from the DHT. ‘As can be seen, the SBPMT K, for stiffer clays is higher than the original Ky pre- diction by Marchetti (33). Indirect Correlations with SPT, CPT, CPTU, and DMT Results Indirect Correlations with SPT, CPT, CPTU, and DMT Results The standard penetration test (SPT), cone penetration test (CPT), and piezocone test (CPTU) all are measurements of vertical penetration, and therefore they do not address Ky directly. However, vertical penetration is coupled with the horizontal 3-26 Marchetti (1980) Ky from SBPMT 7 intact cloys 5 fissured clays L is 20 Horizontal Stress Index, Kp Figure 3-29. Kp from SBPMT Correlated with Kp Source: Data from Mayne and Kulhawy (25 and 40). stresses because they control the vertical "stiffness" of the soll and the shearing resistance of the advancing in-situ device. Alternatively, the DMT provides meas- urements of horizontal total stress. These measurements are taken immediately after penetration of the blade into the clay and, as such, reflect large increases in total horizontal and pore water stresses over the geostatic state of stress. Consequently, the SPT, CPT, CPTU, and DMT provide indirect measurements of Ko. Figure 3-30 shows the trend of Ky obtained from laboratory tests and DMT, PNT, and SBPNT measurements with the normalized SPT N value. From regression analyses of these data, K can be given by the following: Ky = 0.073 ¥pa/Bvo G-31) Figure 3-31 shows the trend of Ky from SBPMT measurements with the normalized cone tip resistance. From these data, Ky can be given by the following: Kg = 0.10(ar - 6ve)/Bv0 (3-32) Ko also can be estimated from the piezocone pore water stress, as shovn in Figure 3-32, These data show that Ky can be given by: Ko ~ 0.24 bue/Pyo (3-33) 3-27 Ko Sy a 13 intact clays 2 a o A 1 si 01020 3040800 Nea / G0 Figure 3-30. Kp Correlated with SPT N Source: Kulhawy, et al. (42), p. 129, 5 . K 20.10 (ay-ey9)/B9 (n=67, 2=0.816, 4b s.0.20.54) . 1 o & . = 5 1 8 § =a 1 é ib 4 . 12 intact clays 5 fissured clays 10 20. 30 40 50 60 =) from CPT Figure 3-31. Ko Correlated with CPTU ap Source: Kulhawy, et al. (41), p. 128, and others (40). 3-28 et b ® 7 K,20.24 Au/é, © eee (n=73, 2-0.827, eo $.0.=0.52) nw at i 12 intact clays 28 5 fissured cloys ol L 1 1. 0 5 10 15 20 A Ge ) from CPTU Ko from SBPMT Au, ot tip Figure 3-32. Ko Correlated with CPTU Au Source: Data from Mayne and Kulhawy (15 and 40) an example of Ky profiling by several in-situ tests in London clay is presented in Figure 3-33. Measured values from the SBPMT and estimated values using the origi- nal Marchetti (33) DMT correlation are given, along with correlations developed from the SPT, CPT, and the liquidity index. Although there is obvious scatter, all of the results are consistent with each other EFFECTIVE HORIZONTAL STRESS IN COHESIONLESS SOILS Cohesionless soils also retain a "memory" of preconsolidation. However, as noted previously, the stress history in cohesionless soils is more difficult to determine because of sampling problems. Therefore, the focus has been almost exclusively on in-situ teste. Dizect Correlations with SBPMT and DMT Results The self-boring pressuremeter test (SBEMT) has shown promise as one of the few devices capable of providing a direct measurement of the in-situ horizontal stress. As such, there is no need for correlations because the stress ideally is measured directly. However, the SBPMT has not been used widely in cohesionless, soils because of the relatively high cost, low productivity, and difficulties in advancing the device in the field. H 2 3 4 London Clay at Brent Cross, Hendon 2F (Powell and Uglow, 1986) gL © SBPMT results A DMT results Depth, m Figure 3-33. Comparison of Ky Values for London Clay at Brent Cross Source: Kulhawy, et al. (41), p. 129. One intent of the dilatometer test (DMT) vas to provide a measurenent of the hori- zontal soil stress, as noted previously. Unfortunately, all in-situ testing devices cause sone disturbance upon insertion into the ground. Therefore, Mar- chetti (33) found it necessary to develop a correlation between a best estimate Ky and the DMT horizontal stress index (Kp), as shown in Figures 3-28 and 3-29. How- ever, Schmertnann (42) showed by calibration chamber tests that the original rela- tionship should also be dependent upon the effective stress friction angle (ec), as given in Figure 3-34. Other correlations with CPT results are given below. Indirect Correlations with SPT and GPT Results No correlations have been developed to date between Ky and the standard penetration test (SPT) N value. However, it was shown in Section 2 that the N value could be correlated with the cone penetration test (CPP) qc value, Therefore, the qo corre- lations below could be used approximately with N values converted to *equivalent*™ Qc values For the GPT ge value, Durgunoglu and Mitchell (44) developed a theoty to relate the cone factor, Ko, dzc, and depth (D) to dlaneter (8) ratio. ‘This theory has been used to develop Figure 3-35, from which an estinace of Ky can be made. This figure rust be used cautiously because small changes in the cone factor ot Fee can result a¢/78 Cone Factor, Nya &,q Morchetti Horizontal Stress Index, Ky Figure 3-34. Ko Correlated with Kp in Sands Source: Marchetti (43), p. 2668. F cveeso Ip 2 o7ar19 © 078+ 100 2 be. o—o — : 7 eee sod ee fret a - 4250 pe B=cone diameter lo! A ot Ko Figure 3-35. Cone Factor versus Ko as a Function of $e and D/B in large Ko changes. However, careful use of this figure with a good knowledge of og the soil stress history can result in reasonable Ky predictions. An example us: this approach is given in Figure 3-36. Marchetti (43) also used this theory and developed « more simplified relationship, as shown in Figure 3-37. In this figure, B = 35.7 mm for a standard cone was introduced. Note that these curves also are quite flat, and that small changes in the input parameters can give large Ky changes. Conbined DMT/CPT Approach for Ko of Sands : In 4 novel approach, the combined results of DMT and CPT calibration chamber tests on laboratory-prepared sand (Figure 3-38) indicated a best fit expression for Ky in terms of both the horizontal stress index (Kp) and normalized cone tip resistance (qe/8vo)+ as given below: Ko = 0.359 + 0.071 Kp - 0.00093 (ae/Byo) 34) This equation was modified to account for field CPT and DMT measurements obtained in a natural sand deposit. The differences between the laboratory and field rela- tionships may be a result of aging effects. This phenomenon of aging is quite important, but it is not very well understood at present, as noted in Section 2. cet (using Fig. 3-38 with 6°35) Figure 3-36. Estimation of K, in Coastal Plain Sand from CPT Source: Kulhawy, et al. (41), p. 130 3-32 7 Toy sf Assumed Cone 8/4205 A OW 7 226" 28 30° 3634" 36° 35° 46° HLL L L 1020 “30165200 "300 “1002050 Cone Tip Resistonce, 4 /jq °. 02 Figure 3-37. Simplified qe - Ko - dtc Relationships Source: Marchetti (43), p. 2668. Calibration t 2 5 10 20 Horizontal Stress Index, Kp Figure 3-38. Ko Correlated with q¢ and Kp Source: Marchetti (42), p. 2672 Empirical Approach Data from CPT studies using electric cones in calibration chanber tests (e.g., Appendix H) indicate that the initial effective horizontal stress (po) is more influential on the magnitude of qq than the verti cal stress. Furthermore, the relationship between Jo and qe appears to be independent of OCR. The advantages of using laboratory chamber tests include known stress state, stress history, and in-place density prior to penetration. A tentative evaluation of the calibration chamber data is shown in Figure 3-39, indicating a general trend between Iho, dc, and Dp. The value of Oyo is the imposed effective horizontal stress prior to cone penetration. With this figure, measured values of qc and Dy are used to obtain dno, as given below: Pro _(de/Pa)?-2> a 7 5 exp OI) (3-35) Once Spo is known, Ky can be computed from Bho/Byo- Application of this empirical approach for estimating in-situ Ky from CPT data in an overconsolidated sand near Stockholm is shown in Figure 3-40, The stress his- tory of this sand has been documented well in the literature, and Equation 3-34 was 5 20 sands summarized in appendix H. Ne OC _Dr(%) & © DF 201040 Sp 2h @ mt 401060 oa 5 & 601080 Z ¥ ® 8010 100 /o o 1 Soo 0° Pe? ° Effective Horizontal Stress, 10 20 30 100 200 300 1000 Cone Tip Resistonce, a./Pg Figure 3-39. Tentative Correlation Between Ih, de, and D; for NC and OC Sands Tested in Calibration Chambers Source: Kulhawy, et al. (41), p. 132 334 limit for \ Ft4or 4 CPT estimate Stockholm sand | Depth, m Coteuoted from stress Nstorpencg & SPT estimate Ke20.39 OCHO? @ PMT measurement Figure 3-40, Comparison of Ky Values at Stockholm Site Source: Kulhawy, et al. (41), p. 132 uused to evaluate the in-situ Ky. As can be seen, the agreements are quite good. The CPT approach may be extended to SPT results through an approximate correlation between the cone tip resistance and N value. The ratio of qq to N has been corre- lated to mean particle size (expressed as Dsq), as shown in Figure 2-30. For the Stockholm site, the value of Ds averages about 0.9 + 0.1 mm, suggesting 2 qc/N ratio of about 6.5. This conversion has been used to estimate a profile of Ky from SPT data using the CPT empirical procedure. Figure 3-40 shows reasonable agreement between the profiles of Ko estimated from CPT and SPT resistances and values deter- mined from the known stress history and PMT data. [REFERENCES 1. Mayne, P. W. and Kulhawy, F. H., "Ko - OR Relationships in Soil", Journal of the Geotechnical Engineering Division, ASCE, Vol. 108, No. OT6, June 1982, pp. 351-872. 2. Jaky, J., "The Coefficient of Earth Pressure at Rest”, Journal of the Society of Hungarian Architects and Engineers, Budapest, Oct. 1964, pp. 355-358. 3. Schmidt, B., Discussion of "Earth Pressures at Rest Related to Stress His- tory", Canadian Geotechnical Journal, Vol. 3, No. 4, Nov, 1966, pp. 239-242. 4, Mayne, P. W., Discussion of "Cy/Co Concept and Ky During Secondary Compres- sion", Journal of Geotechnical Engineering, ASCE, Vol. 115, No. 2, Feb. 1988, pp. 267-270. 3-35, 10. u a2. 13 4. 15 16. vu 1B 1s. Stas, C. V. and Kulhawy, F. H., "Critical Bvaluation of Design Methods for Foundations Under Axial Uplift’ and Compression Loading", Report EL-3771, Elec- tric Power Research Institute, Palo Alto, Nov. 1984, 198 p. NAVFAC, Soil Mechanics (DM 7.1), Naval Facilities Engineering Command, Alexan- dria, 1982, 355 P. Wood, D. M., "Index Properties and Critical State Soil Mechanics", Proceed. ings, Symposium on Recent Developments in Laboratory and Field Tests and Analysis of Geotechnical Problems, Bangkok, Dec. 1983, pp. 301-309. Hansbo, S., "A New Approach to the Determination of Shear Strength of Clay by the Fail Cone Test", Report 14, Swedish Geotechnical Institute, Stockholm, 1957. Mayne, P. W. and Mitchell, J. K., "Profiling of Overconsolidation Ratio in Clays by Field Vane", Canadian Geotechnical Journal, Vol. 25, No. 1, Feb. 1988, pp. 150-157. Marsland, A., "Design Parameters for Stiff Clays", Proceedings, European Con- ference on Soil Mechanics and Foundation Engineering, Vol. 5, Brighton, 1979, pp. 159-162 Mayne, P. W. and Kemper, J. B., "Profiling OCR in Stiff Glays by CPT and SPT", Geotechnical Testing Journal, ASTM, Vol. 11, No. 2, June 1988, pp. 139-147, Garrett, C. and Barnes, $. J., "The Design and Performance of the Dunton Green Retaining Wall", Geotechnique, Vol. 34, No. 4, Dec. 1984, pp. 533-548. Mayne, P. W., "CPT Indexing of In-Situ OCR in Clays", Use of In-Situ Tests in Geotechnical Engineering (GSP 6), Ed. S. P. Clemence, ASCE, New York, 1986, Pp. 780-789. Campanella, R. G. and Robertson, P. K., "Current Status of the Piezocone Test", Proceedings, 1st International Symposium on Penetration Testing (ISOPT- 1), Vol. 1, Orlando, 1988, pp. 93-116. Lunne, T., Eidsmoen, T., Powell, J., and Quarterman, R., "Piezocone Testing in Overconsolidated Clays", Proceedings, 39th Canadian Geotechnical Conference, Octawa, 1986, pp. 209-218. Mahar, L. and O'Neill, M. W., "Geotechnical Characteristics of Desiccated Clay", Journal of Geotechnical Engineering, ASCE, Vol. 109, No. 1, Jan. 1983, pp. 56-71 Mayne, P. W. and Frost, D. D., "Geotechnical Report, White House Communica- tions Agency, Anacostia, Washington, D.C.", Report W6-5523, Law Engineering, McLean, 1986, 95 p. Powell, J. J. M., Quarterman, R., and Lunne, T., "Interpretation and Use of the Piezocone Test in U.K. Clays", Penetration Testing in the U.K., Thomas Telford, London, 1988, pp. 47-52. Rad, N. S. and Lunne, T., "Direct Correlation Between Piezocone Test Results and Undrained Shear Strength of Clay", Proceedings, lst International Sympo- sium on Penetration Testing (ISOPT-1), Vol. 2, Orlando, 1988, pp. 911-917. 3-36 20. au. 22 23. 24 25 26. 27 28 29. 30 31 32. 33 34. Tavenas, F. and Leroueil, S., "State-of-the-Art on Laboratory and In-Situ Stress-Strain-Time Behavior of Soft Clays", Proceedings, International Sympo- sium on Geotechnical Engineering of Soft Soils, Mexico City, 1987, pp. 1-46. Mayne, P. W. and Holtz, R. D., "Profiling Stress History from Piezocone Sound- ings", Soils and Foundations, Vol. 28, No. 1, Mar. 1988, pp. 16-28. LaRochelle, P., 2ebdi, M., Leroueil, S., Tavenas, F., and Virely, D., *Piezo- cone Tests in Sensitive Clays of Eastern Canada", Proceedings, 1st Interna- tional Symposium on Penetration Testing (ISOPT-1), Vol. 2, Orlando, 1988, pp. 831-841 Mayne, P. W. and Bachus, R. C., "Penetration Pore Pressures in Clay from CPTU, DMT, and SBP", Proceedings, 12th International Conference on Soil Mechanics and Foundation Engineering, Vol. 1, Rio de Janeiro, 1989, pp. 291-294. Lukas, R. G, and de Bussy, B., "Pressuremeter and Laboratory Test Correlations for Clays", Journal of the Geotechnical Engineering Division, ASCE, Vol. 102, No. GT9, Sept. 1976, pp. 945-962. Mayne, P. W. and Kulhazy, F. H., Discussion of "Independence of Geostatic Stress from Overconsolidation in Sone Beaufort Sea Clays", Canadian Geotechni- eal Journal, Vol. 25, No. 3, Aug. 1988, pp. 617-621. Powell, J. J. M. and Uglow, I. M., "Marchetti Dilatometer Testing in U.K. Soils", Proceedings, 1st International Symposium on Penetration Testing (1SOPT1), Vol, 1, Orlando, 1988, pp. 555-562. Mayne, P. W., "Determining Preconsolidation Stress and Penetration Pore Pres- sures from DMT Contact Pressures", Geotechnical Testing Journal, ASTM, Vol. 10, No. 3, Sept. 1987, pp. 146-150 Mayne, B. W., "Determining OCR in Clays from Laboratory Strength", Journal of Geotechnical Engineering, ASCE, Vol. 114, No. 1, Jan. 1988, pp. 76-92. Weoth, C. P., "Penetration Testing - A Rigerous Approach to Interpretation", Proceedings, Ist International Symposium on Penetration Testing (ISOPT-1), Vol. 1, Orlando, 1988, pp. 303-311 Mayne, B. W., "Cavity Expansion/Critical-State Theory for Piezocone Penetra- tion in Glays", submitted for review in 1990. Robertson, P. K., Campanella, R. ., Gillespie, D., and Greigs, J., "Use of Piezometer Gone Data", Use of In-Situ Tests in Geotechnical Engineering (GSP 6), Ed. S. P. Clemence, ASCE, New York, 1986, pp. 1263-1280. Mayne, P. W. and Bachus, R. C., "Profiling OCR in Clays by Piezocone Sound- ings", Proceedings, 1st International Symposium on Penetration Testing (ISOPT- 1), Vol. 2, Orlando, 1988, pp. 857-864 Powell, J. J. M. and Uglow, I. M., "Dilatoneter Testing in Stiff Overconsoli- dated Clays", Proceedings, 39th Canadian Geotechnical Conference, Ottawa, 1986, pp. 317-326. 35 36 37. 38. 29. 40. 4a 42. 43, Lacasse, S. and Lunne, T., "Calibration of Dilatometer Correlations", Proceed- ings, Ist International Symposium on Penetration Testing (ISOPT-1), Vol. 1, Orlando, 1988, pp. 539-548. Bullock, P., "The Dilatometer: Current Test Procedures and Data Interpreta- tion", Civil Engineering Report, University of Florida, Gainesville, 1983, 308 Pe Larsson, R., "Basic Behavior of Scandinavian Soft Clays", Report 4, Swedish Geotechnical Institute, Linkoping, 1977, 125 p. Brooker, E. W. and Ireland, H. 0., "Earth Pressures at Rest Related to Stress History", Canadian Geotechnical Journal, Vol. 2, No. 1, Feb. 1965, pp. 1-15. Mayne, P. W., "Ko-cy/@yo Trends for Overconsolidated Clays", Journal of Ge technical Engineering, ASCE, Vol. 110, No. 10, Oct. 1984, pp. 1511-1516, Mayne, P. W. and Kulhawy, F. H., "Direct and Indirect Determinations of In- Situ Ky in Glays", Research Record sxx, Transportation Research Board, Wash- ington, 1990, in press. Kulhawy, F. H., Jackson, C. $., and Mayne, P. W., "First-Order Estimation of Kg in Sands and Clays", Foundation Engineering: Current Principles and Prac- tices, Ed. F. H. Kulhawy, ASCE, New York, 1989, pp. 121-134. Schmertmann, J. H., DMT Digest No. 1, GPE Inc., Gainesville, 1983, 3 p Marchetti, S., "On the Field Determination of Ky in Sand", Proceedings, 11th International Conference on Soil Mechanics and Foundation Engineering, Vol. 5, San Francisco, 1985, pp. 2667-2672. Durgunoglu, H. T. and Mitchell, J. K., "Static Penetration Resistance of Soils", Proceedings, ASCE Specialty Conference on In-Situ Measurement of Soil Properties, Vol. 1, Raleigh, 1975, pp. 151-189. Section & STRENGTH A knowledge of the strength of soils is necessary for most geotechnical analyses. From a foundation engineering standpoint, the vength is necessary primarily to evaluate the capacity. However, soil strength varies with many parameters, and therefore it is not uniquely defined. In this section, basic definitions are pre- sented first to establish the general background, notation, and relevance of the strength tests to field conditions. Then methods for estimating the effective stress friction angle are presented, first for cohesionless soils and second for cohesive soils. For each soil type, typical values, influencing factors, and in-situ test correlations are presented. Finally, methods for estimating the undrained shear strength are presented, including typical values, influencing factors, and in-situ test correlations. BASIC DEFINITIONS The strength of soils commonly is expressed by the Coulomb-Mohr failure criterion, as illustrated in Figure 4-1. For this criterion, failure is given by: reetetang 1) in which + = shear stress at failure (i.e., shear strength), ¢ = cohesion inter- cept, ¢ = normal stress, and § = friction angle Effective Stress Analysis Although Equation 4-1 is the general form of the criterion, it is rarely appro- priate to use the complete equation. Instead, the criterion is used in two alter- native forms. First, when effective stress analyses of cohesionless or cohesive soils are conducted, Equation 4-1 is expressed as: r=dtand (4-2) in vhich 2 = effective nomal stress and J = effective stress friction angle, at shown in Figure 4-2. 7 ¢ g $ a a Zc 2 5 5 oy 3 a Normal Stress, o Effective Normal Stress, & Figure 4-1, General Goulomb-Mohr ‘Figure 4-2. Effective Stress Couloab- Failure Mohr Failure No effective stress cohesion intercept (8) is shown because it occurs only in spe- cial cases, such as with truly cemented sofis, partially saturated soils, and heav- Ky overconsolidated clays, in which @ is interpreted as gradually decaying with time on an engineering time-scale. For these special cases, could be included in Equation 4-2. However, it is prudent to seek expert geotechnical advice before considering use of @ for design. Many times, effective stress laboratory test data are interpreted incorrectly to show a moderately high @ and an unrealistically low because the true failure envelope curvature is not being addressed. Figure 4-3 shows actual curved fail- ure envelopes, with € = 0, for a wide range of soils from clay to rockfill. Linear interpretation of any of these data over a limited stress range would suggest a & and J, but these values would not be the true soil strength parameters. The friction angle of soils also varies wich many other factors, as will be described throughout this section, For a given soil at a constant normal effective stress (3), the friction angle varies with density state and strain, as shown in Figure 4-4. Expressing J in corms of the effective major and minor principal stresses (21 and 33, respectively) gives: Gyee-1 C1-V¢e ~ Gest i” Baa ee) sin $ in which the subscript £ represents failure conditions Different peak friction angles (Bp) develop as a function of soil density state. At one limit is the very dense cohesionless soil or the heavily overconsolidated = 60] 2 % 40) 2. a 2 20) & eewrn ondsttbed Lond" consolicoied from slurry ° 20406080100 Effective Normal Siress, &/pg Figure 4-3. Strength Envelopes for a Range of Soil Types Source: Bishop (1), p. 104. Effective Principal Sress Roto, 6/25 Dilative (very dense, heavily OC} A 14 ‘Mat very ¥ large Contractive strains) (very loose, NC) Strain, € Figure 4-4, Friction Angle Definitions cohesive soil which exhibits strongly dilative behavior during shear. For these soils, the peak friction angle is high, and it develops at very small strains, typically on the order of a few percent. At the other limit is the very loose cohesionless soil or the normally consolidated, insensitive, uncemented, cohesive soil, which exhibits contractive behavior during shear. For these soils, the peak friction angle is lower, and it develops at larger strains, typically upwards of 10 The difference between these linits occurs because of the volume Digterent behavior is to 20 percent change behavior during shear (dilative to contractive) noted for sensitive, cemented, and other structured cohesive soils, which normally peak at small strains, much like the intermediate curve in Figure 4-4. As a dilative soil is strained past ite peak, it strain-softens to a limiting state known as the fully-softened or eritical void ratio state (Joy). The contractive soil strain-haréens to reach the critical void rat jo state, which also corresponds to its peak friction angle. The critical state (Joy) typically occurs at strains upwards of 10 to 20 percent. Therefore, regardless of the initial density state, Fey is unique for a given soil at a constant normal effective stress. With subsequent large straining in cohesive soils, typically in excess of 100 per- cont, Jey is gradually reduced to an ultinate limit known as the residual state (Bp. The resulting Jy is commonly several degrees lover than Jey. For cohesion- less soils, Jy is essentially equal to Jqy. The residual state would be considered in foundation engineering only for very large strain problens, such as siting in soils containing pre-existing shear failures. Common examples would be landslide debris or slopes in stiff-fissured clays. Total Stress Analysis The second use of Equation 4-1 is defined as the total stress (or ¢ ~ 0) analysis of cohesive soils, given by: reer cus sy 4-4) in which all four terms can be used interchangeably to represent the undrained shear strength of the soil. This relationship is shown in Figure 4-5. Also in this figure, qu is defined as the unconfined compressive strength = 2 sy. Im many older references, the term “cohesion” was used to designate sy. In recent 4:0 Shear Stress, r 9 a 2 Total Normal Stress, Figure 4-5. Total Stress Coulonb-Mohr Failure references, sy is referred to as the undrained shear strength or undrained shearing resistance. The older definition has led to much confusion and misinterpretation with the effective stress cohesion intercept (2) Total stress analysis normally is adopted for simplicity. In reality, the failure of all sotis (sands, silts, and clays) occurs on the effective stress envelope shown in Figure 4-2, In low permeability sofle such as clays, loading generates changes in pore water stresses (tu). These pore water stresses change the offec- tive stresses, which in turn influence the state of stress relative to the effec- tive stress envelope. Since the total stress loading path and the magnitude of the changes in pore water stresses may not be known with confidence, a total stress analysis provides a simple analysis alternative. However, it must be remenbered that sy includes J and au, and it varies with stress level in-situ. Therefore, sy must be determined carefully to represent the in-situ conditions at a particular depth, as described in detail lacer in this section. Relevance of Laboratory Strength Tests to Field Conditions The strength of soils can be measured by a number of different laboratory strength tests, as noted previously in Figure 1-1, Each of these tests will give different results because each subjects the soil to different boundary conditions and loading stress paths. In the field, different elements of soil also will be subjected to different boun- dary conditions and loading stress paths. Figure 4-6 shows a mmber of common field loading cases and the test types pertinent for each case. For an embankment loading, the bearing capacity is represented most correctly by a combination of compression (PSC or TC), direct simple shear (DSS), and extension (PSE or TE) tests along the potent: 1 shear surface noted. For ease in computation, an average of these three test types normally is used. With 2 loaded wall, the direct simple shear and extension test types are averaged. With a vertical cut, the compression test is most relevant When addressing foundations, different strengths are appropriate for different field loading and behavior modes. These modes are described in detail by Kulhavy, et al. (2). For a drilled shaft in compression, the tip resistance can be evalu- ated from an average of the triaxial compression, direct simple shear, and triaxial extension tests. The side resistance is modeled by the direct simple shear test up to first yield or slippage along the interface, after vhich direct shear is more appropriate. The results of these two tests are similar, so they commonly are used 45 a) Embankment b) Loaded Wall a a Ln a €) Vertical Cut Compression Direct Simple Extension Test (PSC Shear Test Test (PSE or TC) (oss) or TE) PSC or Te d) Orilled Shaft Uplift 2 Loterai/ Moment /OSS/DS- ae rel e) Spread Foundation upiit uplift pitt Comp. w Sheor w. Cone Punching TJLboss SL grre —— oss (OE a es Te/0Ss/TE pss Note: Plane strain tests (PSC/PSE) used for long fectures Trioxial tests (TC/TE) used for near symmetrical feotures _ Direct shear (DS) normally substituted for DDS to evaluate ¢ Figure 4-6. Relevance of Laboratory Strength Tests to Field Conditions where they are best-suited, the DS being used for sands with the DSS being used for clays. For a shaft in uplift, the side resistance is the same as in compression For lateral or moment loading, triaxial extension is more appropriate. For spread foundations in compression, the same bearing capacity approach is used In uplift, the behavior can range from the normal situation of a vertical shear surface to a vertical shear with cone breakout to a punching limit controlled by bearing capacity. As noted in the figure, the shear case is given by the DSS. The cone case is an average of TC and DSS. ‘The punching is evaluated using an average 46 of TC, DSS, and TE. The various tests pertinent for a particular field condition are likely to be an excessive requirement for conmon and routine design cases. Therefore, it is more convenient to establish a standard “test of reference” which would be appropriate for many design cases, and which would be simple and expedient from a commercial testing standpoint, The recommended test (e.g., Wroth, 3) is the isotropically consolidated, triaxial compression test for undrained loading (CIUG) and for drained Loading (CIDC). Using the results of this test as a standard reference, the results of all other tests can be compared simply and conveniently. Te should be noted that most soils in-situ actually will be consolidated aniso- tropically. This difference in consolidation stresses has no appreciable influence fon the soil friction angle ($). However, {t does influence the evaluation of the undrained shear strength, as will be shown Later EFFECTIVE STRESS FRICTION ANGLE OF COHESIONLESS SOILS - GENERAL EVALUATION BASIS Correlations for estimating the effective stress friction angle for cohesionless soils have been presented by mumerous authors. Representative relationships are given below. ‘Typieal Values arly work on this topie suggested staplified tabulated values for the effective stress friction angle, such as those given in Table 4-1. Although never stated explicitly, it is probable that these values refer to peak values measured in triaxial compression tests Gre). Tabulated values such as these only establish the general order of magnitude for Fre. They should not be used for design. Correlations with Index Parameters Subsequent approaches have correlated the value of Ste with one or more soil index paraneters, such as sofl type, relative density, and unit weight or void ratio Figures 4-7 and 4-8 show two comon relationships for estimating 3¢c from soil index paraneters. Figure 4-7 refers specifically to Jee from triaxial compression vests on soils composed of hard minerals, at stress levels typical of those used in footing design. Figure 4-6 is a more general relationship based on the groups in the Unified Soil Classification System and presumably also refers to Bec. Although these figures address more of the variables, they still are simplifications of actual behavior and tend to be somewhat conservative Table 4-1 REPRESENTATIVE VALUES OF ee Bee (degrees) Soil Material “Teose—~—~SC«é ese Sand, round grains, uniform 27.5 34 Sand, angular grains, well-graded 33 45 Sandy gravels 35 50 Silty sand 27 to 33 30 to 34 Inorganic silt 27 to 30 30 to 35 Source? Terzaghi and Peck (@), p 107- Friction Angle, dye 0-20 40 60 «80100 Relative Density, 0, (%) Figure 4-7. $e versus Relative Density Source: Schmertmann (5), p. 41. Influence of Strength Envelope Curvature Table 4-1 and Figures 4-7 and 4-8 imply although data such as that in Figure 4-3 show that the failure envelopes normally \t the soil failure envelope is linear, are nonlinear. This nonlinearity is well-established in the literature (e.g., 1, Z, 8) and is attributed to soil dilatancy. This dilatancy increases with increas- ing relative density and decreases with increasing stress level. ray as" Friction Angle, 4 aot asF Unified soit elossification 30°F For cohesioniess soils without plostic fines as" 12 we 20. 22 24 Dry Unit Weight, y /y,, 1210 0808 04 02 Void Ratio, e for G.=2.68 Figure 4-8. Jee versus Relative Density and Unit Weight Source: NAVFAC (6), p. 7.1-149 The most convenient vay to include the strength envelope curvature is to use secant peak friction angles which vary with stress level, as {llustrated in Figure 4-9. By taking successive secants through the origin at varying normal stresses, the values of secant With normal stress can be obtained. Loose sotls approximate Joy and exhibit an essentially linear envelope. It should be noted at this point that the soil behavior illustrated in Figures 4-3, 4-4, and 4-9 is general and that the same patteras will develop regardless of the laboratory test type. From this point forvard, it will be presumed that the fric- tion angle given represents 2 peak value obtained as a secant to the failure Friction Angle, Psecont Ken | Loot Shear Stress, T Effective Normol Stress, 7 Figure 4-9, Nonlinear Failure Envelope Representation 49 envelope. For clarity, the subscripts to be used will refer only to the test type, such as ye for peak secant friction angle in triaxial compression. No test desig- nation is needed for the critical void ratio friction angle (Jey) because this value is unique and independent of test type (e.g., 8, 9). The same is true for the residual friction angle (,). Recent work by Bolton (6) has unified much prior research in a convenient way, uti- Lizing critical state concepts and a data base primarily of clean sands. This vork demonstrated that the dilatancy component of the friction angle can be estimated as follows: —"" 5) Gre - dev) = 3 Ipp for triaxial compression (4-6) in which Igp is a relative dilatancy index, given by: Ipp ~ De[Q - 1n(100 Be/pa)] - R Cipp > 0) (7) In this equation, Dy ~ relative density, Q = soil mineralogy and compressibility coefficient (10 for quartz and feldspar, 8 for limestone, 7 for anthracite, 5.5 for chalk), Pg = mean principal effective stress at failure (21 +3) + 3)¢/3]. Pa = atnospheric stress in the same units as pe, and R= fitting coefficient (equal to 1 for the evaluated test conditions and data). Figure 4-10 illustrates this rela- tionship for eight different quartz and feldspar sands, The equation noted on the figure would be typical of trisxial compression tests on silica-type sands. The relative dilatancy index (Igp) should be Limited to 4 unless detailed laboratory test data indicate otherwise. Equation 4-7 unfortunately relates to 1 ean principal effective stress at fatl- ture, 4 parameter which includes the initial stress state, stress path to failure, sest conditions, and foundation type. For preliminary estimating purposes, Be cen be assuned to approxinate two tines yo, which should lead to a computed @ - ev) within 1 to 2 degrees of the actual value for most cases. For final design, the value of Pg corresponding to the specific foundation conditions should be used. To estimate the value of Gey, Koerner’s vork (10) on single mineral soils can be considered, which led to the following: bev in which: atancy Component, Py_~ Fey 36° + a6, 242 = 364 {0, L10-Inl1005 4 /p)-I} © 0,08 (6 sonds) © 00.5 eae ol ° On 1 10 100 1000 Mean Principal Effective Stress ot Foallure, B/D, Figure 4-10. Dilatancy Angle Relationships Source: Bolton (8), p- 73 Ab] + Ab2 + B83 + Ady + Ob, (8) correction for particle shape 41 ~ -6° for high sphericity and subrounded shape $1 = +2" for low sphericity and angular shape correction for particle size (effective size, 419) gg = -11* for dyo > 2.0 mm (gravel) Ag = -9° for 2.0 > dyo > 0.6 (coarse sand) gg = -4° for 0.6 > dio > 0.2 (medium sand) 42-0 for 0.2 > dy > 0.06 (fine sand) correction for gradation (uniformity coefficient, C,) 863 = -2° for Cy > 2.0 (well-graded) 43 ~ -1° for Cy = 2.0 (medium graded) 463 = 0 for Cy < 2.0 (poorly graded) correction for relative density (Dr) Ay = -1" for 0 < Dy < 0.5 (loose) 44-0 for 0.5 < Dy < 0.75 (intermediate) 44 = 44° for 0.75 < Dy < 1.60 (dense) correction for type of mineral ¢5- 0 for quartz $5 = 44° for feldepar, caleite, chlorite Ags = +6° for muscovite mica Current understanding (e.g., 8) is that Joy is essentially independent of relative density, and therefore the relative density correction (Ay) should be set equal to zero. Relative density primarily influences the dilatancy component. Equation 4-8 also must be kept within the context of Bolton's work (8) on natural soils, vhich showed that Jey = 33° for representative quartz sands and Gey = 40" for representa- tive feldspar sands. However, nost natural deposits of sand include silt. There fore, Bolton concluded that Joy for most natural sand deposits rarely will be much above 30° to 33°, and may be as lov as 27° vhen the silt content is high. Influence of Test Boundary Conditions For simplicity, most analyses assume that the peak, secant, effective stress fric- tion angle is independent of direction of loading, and therefore the intermediate effective principal stress (2) is disregarded. However, this influence can be important in some loading cases. To evaluate this effect, the intermediate effec- tive principal stress factor (b) can be defined as: 22 - 83 b= (4-9) Normalized test data on five sands are shown in Figure 4-11 to illustrate the importance of b. The mean and range are shovn for both the loose and dense sands. For plane strain compression (b = 0.3 to 0.4), the increase ranges from 7 to 18 percent with an average on the order of 12 percent. For triaxial extension (b= 1), the increase ranges from 0 to 23 percent, again with an average on the order of 12 percent. A similar increase should be expected when comparing plane strain extension to compression. Other studies (e.g., 9) have shown that the plane strain compression (PSC) and direct shear (D5) tests can be interrelated as follows: tan Bde tan Fpse 608 Fev (10) For typical ranges of Joy, the PSC values fron this equation will be sone 2 to 7 degrees higher than the dixect shear values, corresponding to increases from & to 19 percent, Comparison of the direct shear values from Equation 4-10 with the triaxial compres- sion values from Equations 4-5 and 4-6 indicates that the triaxial compression val- ues may be larger or smaller than the direct shear values, depending on the values az Trioxiol Trioxia! compression [Plone strain extension compression rm o|S Teac BIS ‘Dense sands. ei she ‘ 2 ou g fo & os 0 02 08 10 Figure G-11. Influence of Intermediate Principal Stress on Friction Angle Source: Data from Ladd, et al. (11), p. 431. of dey, relative density, and stress level. Table 4-2 summarizes the relationships for friction angle as a function of test type. As can be seen from this table and Figure 4-6, use of the triaxial compres- sion friction angle (je) alone will alnost slays be a conservative assumption. EFFECTIVE STRESS FRICTION ANGLE OF COHESIONLESS SOILS CORRELATED WITH IN-SITU TESTS At the present tine, correlations of the effective stress friction angle have been nade with the standard penetration test (SPT), cone penetration test (PT), pres- suremeter test (PMT), and dilatometer test (OMI). The CPT correlations are perhaps the best-developed, followed by the SPT. The INT correlations are newer and less developed, while the DMT correlations are of Limited use at this time. In all cases, it is presuned that the correlations use the triaxial compression friction angle ($¢¢) corresponding to the appropriate stress and/or relative density condi- tlons. Correlations with SPT N Value Correlations of the effective stress friction angle with the SPT N value have been made for many years. Early work on this subject attenpted to relate N to Fee a3 ‘Table 4-2 RELATIVE VALUES OF EFFECTIVE STRESS FRICTION ANGLES FOR COHESTONLESS SOILS est Type Friction Angle (degrees) Triaxial compression (TC) 1.0 be ‘Triaxial extension (TE) 1.12 be Plane strain compression (PSC) 1.12 Fx Plane strain extension (PSE) 1.12 (for PSC/TC) x 1.12 (for TE/TC) = 125 de Direct shear (05) vant [tan dpne £05 Boy) a tant [tan (1.12 $4) cos dev) @irectly, as shown in Table 4-3. ‘The Peck, et al, (12) approach appears to be more common, perhaps because it is more conservative. These values also are shown in Figure 4-12. As discussed in Section 2, che N value actually depends upon stress level. Figure 4.13 As representative of the correlations between N and Zq, as a function of stress level. This correlation can be approximated as follows: Gee = tan l[N/(12.2 + 20.3 Byo/Pa) 0-34 (11) These results tend to be somewhat conservative and should not be used at very shal- low depths, less than 1 to 2 m (3.3 to 6.6 £t). Improved correlations with the other variables described in Section 2 have not been developed to date Correlations with GPT q¢ Value Similarly, correlations of 3 with cone tip resistance, gg, have been developed Eazly work attempted to correlate qe to Fy: dizectly, as shown in Table 4-4. As described in Section 2, ae is affected by the vertical stress. Therefore, Fy: should be correlated to both q¢ and Gy, as shown in Figure 4-14, This correlation Table 4-3 N VERSUS J gp RELATIONSHIPS reproxinate Fy (degrees) N Value @lows/ft or 305 om) @ @) Ome very loose < 28 < 30 4 wm 10 loose 28 t030 30 to 35 2 to 30 medium 30 to 36 © «35 to 40 3 to 50 dense 36 to Wl 40 to 45 >» very dense > 41 > 45 a> Source: Peck, Hanson, and Thoraburn (12), p. 310. b= Source: Meyerhof (13), p. 17. ¥ eo Ss 8 & & SPT N Value (blows/ft or 305mm) 8 2a Bae Be Friction Angle, De Figure 4-12. N versus dip Source: Peck, Hanson, and Thorburn (12), p. 310. ean be approximated as follows: Sue = tant (0.1 + 0.38 10g (4¢/%v0)) (4-12) Adjustments to this figure and equation for soils of different compressibility and stress history should be made as described in Section 2. SPT N Volue, Blows/ft or 305mm 20 4060 7 ges ie ' g ost Be 3 ao" ss \30" A \ Figure 4-13. N versus $4. and Overburden Stress Source: Schmertmann (14), p. 63. Table 4-4 de WSUS Fy Normalized Gone Tip Relative Approximate ep Resistance, qo/Pa Density (degrees) <® very loose <3 2 t0 40 Loose 30 t0 35 w £0 120 medium 35 t0 40 220 to 200 dense 40 v0 45 > m0 very dense >45 Source: Meyerhof 3), p. 1 Villet and Mitchell (16) presented a nore general approach to evaluating Fee from CPT data which includes qq, stress level, shape effects, and sofl stress history. Their results are shown in Figure 4-15 and are suitable for lov compressibility sands. Cone Tip Resistance, a_/P, 200 400 Vertical Effective Stress, do/Pa al Figure 4-14. gg versus Bee and Vertical Stress for NC, Uncenented, Quartz Sands Source: Robertson and Campanella (15), p. 726. 6 10! Re Cone to send pe & ” x = 2 8 sor asa a 3° 35 3 a Friction Angle, $4. Friction Angle, $4. Figure G15. ge from OPT Data Source: Villet and Mitchell (16), p. 193. 4-17 Using the standard cone dianeter (B) of 35.7 mn, Marchetti (17) revorked the data in Figure 4-15 to result in the more simplified Figure 4-16. Consistent with the development in Section 2 vhich related relative density to the normalized cone tip resistance, a similar correlation has been developed from 20 data sets obtained in calibration chanbers and is shown in Figire 4-17. Mineralogy, particle shape, con pressibility, and percent fines largely account for the observed range of dec at any normalized qe value. Correlations with PMT Results The results obtained from pressureneter tests also can be correlated with the effective stress friction angle, using procedures developed by either Schmertmann (14) or Hughes, et al. (18), The Hughes, et al. approach is presented below. in a pressuremeter test, the basic data cbtained are the expansion stress (pe) and the volune changes (AV) in the pressuremeter of known volume (V). The resulting data can be plotted as shown in Figure 4-18a, using the cavity strain (¢¢) which is Gefined as the change in membrane radius divided by the initial radius and is given by: fe = 1 ey 5 4-13) in which ey = AV/V = volumetric strain. These data then are re-plotted as in T ToT [Assumed Cone 8/6,+0.5 RELL satel n 1020 80 100 200 500 1000 2000 Cone Tip Resistonce, a,/Gq Figure 4-16, Steplified qo - Ky - Fre Relationships Source: Marchetei (17), p. 2668. ais 7 T T - ‘@ Fy I7E+ NO log Lla/P)/(Fyo/25 I G SOT ine eo ‘1 do east 4 < = ao 4 2 aslo 4 v 8 eae og Solid Symbots~Oc Sands _] 8 ce Other Symbols~NC Sands. &% (os given in. Appendix H) 1 1 L 1020 30 100 200 00 1000 Cone Tip Resistance, (a¢/Pg) /(Gyq/Pq)°> Figure 4-17, Trend of 8g, with Normalized qo 2 27 S |f (a) Typicat Prot =B gf (b) Normalized Pict e ie Be 2° eo Eisai € & 2| moe ge ; & ol Blt. . 1. 4 Oo 4 8 @ 6 gos0e 1 2 4 6 10 16 Covity Strain, €,(%) Covity Strain, €¢ (%) Figure 4-18, subtracting the init: Figure 4-18. PMT Data Representations Source: Mair and Wood (19), p. 76. al pore water stress at the pressuremeter level. The resulting log-log plot is essentially linear with a slope, s. By considering cylindrical cavity expansion theory, s can be given by: s = sin Bey (1 + sin y)/(1 + sin ey) 1) in which Bey = critical void ratio friction angle and y = dilation angle (Jee - dev, a5 described previously). Equation 4-14 can be rearranged to give sin p= 5 (1 + sin doy)/sin Joy (4-15) Therefore, by re-plotting the PMT data to give s and estimating Jey as described Previously, the friction angle rc) ean be obtained. Figure 4-19 provides graphical procedure to evaluate Bye, using Bolton's (8) approximation that Bee - Boy + 0.8 ¥ (4-16) Correlations with DMT Results Recently, @ correlation also has been presented between the effective stress fric- tion angle and the thrust pressure (tip resistance) on the dilatometer during pene- tration. Using the Durgunoglu and Mitchell (20) theory, Schmertmann (21) showed that the dilatometer tip resistance (qp), obtained from thrust measurements during Penetration of the blade, could be related to the cone tip resistance (qc) and the effective stress friction angle (psc) under plane strain compression. This 580" Friction Angle, $4, & 3 03 0.4 05 06 8 log (pup) Blog «, Figure 4-19. Friction angle Evaluation from PMT Results Source: Mair and Wood (19), p. 78 4-20 relationship is given below: pse ~ 25(2.3 - ap/ae) 17) To evaluate Spec from the DMT results, an iterative process is necessary. An ini- tial estimate is made of Fpse for triaxial compression conditions, from which an equivalent qe is determined from Figure 4-15 or 4-16. Using this qc and the qp measurement, pso is computed from Equation 4-17, This plane strain 3ps¢ then is converted to an equivalent triaxial Jeo using the relationships shown in Figure Acl1 or Table 4-2. The final $ee is compared with the initial assumption. If they agree, then dpgc {# correct. Otherwise, iteration mist be done until the initial estimate and final value converge. At the present tine, the DMT versus Zpse corre- lation should be considered only as a first order approximation until sufficient field confirmations become available. EFFECTIVE STRESS FRICTION ANGLE OF COHESIVE SOILS Correlations for estimating the effective stress friction angle for cohesive soils have focused on only two areas: (1) the friction angle for normally consolidated (Wic) and remolded clays, which will approximate Gey, and (2) the residual friction angle Gr). No generally accepted procedure has been presented for estimating the peak friction angle of overconsolidated (OC) clays as a function of overconsolida- tion ratio (OGR) and other controlling factors, although the behavior should be qualitatively similar to that for cohesionless soils. Similarly, no generally accepted correlations have been presented with in-situ test results. Correlations with Critical Void Ratio Friction angle As described at the beginning of this section, the peak friction angle for insensi- tive, uncemented NG cohesive soils basically is equal to the critical void ratio friction angle Gey). For sensitive, cemented, or other structured NC cohesive soils, Jey will represent a lover bound for the peak friction angle. For OC soils, renolding will destroy the stress history and therefore result in "newly-created NC soil”, with the friction angle being given by Zoy. Other complex factors such as leaching, sensitivity, stress state, ete. influence this simple explanation to some degree. However, first-order correlations can be made using this simple approach. Many authors have shown that Jey can be correlated with simple index paraneters such as the plasticity index. One such relationship {s presented in Figure 4-20, which shows that Gey decreases with increasing plasticity index and increasing clay © Undisturbed soll oad g 2 Remoided soi 40° sin $e, = 0.8 - 94 In PL po dey= 0.8-0.0981 foe, § ost 20 oz tie soo Monmaionte = 22Z 6810 152030 46-86 G85 100180200 Plosticity Index, PI (%) Figure 6-20. Jey for NC Clays versus PI Source: Mitchell (22), p. 284, mineral activity (ksolinite + illite + montmorillonite). This general trend has ‘been corroborated by others (e.g., 11, 23). However, it should be noted that the error band with this correlation is fairly large Influence of Test Boundary Conditions Laboratory testing conditions can influence the friction angle of NC clays. The data in Figure 4-20 were obtained largely from isotropically consolidated, undrained triaxial compression (CIUC) tests with pore water stress measurements In-situ, the initial stresses would correspond to anisotropic consolidation (CAUC), most comonly restricted to Ky consolidation (CKUC), Fortunately, comparative studies such as that shown in Figure 4-21 have cexonstrated that eq essentially ts the sane, regardless of initial consolidation state. Although the regression shows a sual variation from equality, this variation is small and can be ignored. However, the same can not be said for other testing conditions. For plane strain compression, Wroth (3) suggested analytically that dpsc would be approximately 9/8 tines Fee. Figure 4-22 {Illustrates that this relationship is satisfactory, al- though the regression gives a slightly lover value. This value is similar to that for sands. Figure 4-23 compares the friction angles for NC clays in extension and compression. As can be seen, dre always is equal to or greater than dec and, on the average, Sce/Btc ~ 1.22. This average is the sane for both anisotropic and isotropic test conditions, even though their statistics differ a small anount Additional Limited data (26) show that the pattern should be similar for plane strain conditions as well. 50" 407 30° 20" $e (CKUC and CAUC) Figure 4-21. Jee Variation as a Function of Consolidation Stress for NC Clays Se lOKQuC/CAUC)=0.96 F,lctUe) (nea, r#0.923, $.D.=2. “\e % 47 intact cloys ed $y_(CIUC) Source: Data from Mayne (24) and Nakase and Kanei (25) 50" 40" 30" Pose 20 f 12 intoct loys io 20 50° %e Figure 6-22. Jpse versus dee for NO Clays Source: Data from Mayne and Holtz (26) T° © CKU/CAU (n=37, 1: © CTU (n=33, 120,843, S.0+5.0%) 730, 8.1 50° Fre? 1.22 $e (n=70, r220.790, S.. 40°| 30°} 2081 55 intact clays 50° Fe Figure 4-23. dee versus Jee for NC Clays Source: Data from Mayne and Holtz (26) and Nakase and Kamei (25). Table 4-5 sumarizes the relative values of the friction angle for the different testing conditions. Although no detailed comparisons have been presented for the @ixect shear test, the results should exhibit patterns similar to those presented earlier for cohesionless soils. Therefore, the sane relationship is proposed for cohesive soils. It should be noted that use of $r¢ alone will almost always be a conservative assumption Correlations with Residual Friction Angle 4s described earlier in this section, the residual friction angle (,) develops when a cohesive soil undergoes very large strains, and the soil structure is totally renolded and re-oriented into a minimum strength orientation. Currently, iy 2 Table 4-5 RELATIVE VALUES OF EFFECTIVE STRESS FRICTION ANGLE FOR NORMALLY CONSOLIDATED COHESIVE SOILS Test Type Friction Angle (degrees) Triaxial compression! (16) 1.0 Ste Triaxial extension (TE) 1.22 Gee Plane strain compression (PSC) 1.10 dee Plane strain extension (PSE) 1.10 (for PSC/TC) x 1.22 (for T = 1.36 bce Direct shear? (Ds) tanh [ean ¢psc €08 Fevl or tan} (ean(1.10 Sec) c08 Jey) 1 - GTUC, cKUC, or cauc 2 - Speculative, based on results from sand it is understood that the strains necessary to accomplish this remolding may exceed 100 percent. Earlier studies of this subject may not have subjected the soil to the necessary strains, and therefore residual angles quoted in earlier sources may be somewhat on the high side. Extensive research (e.g., 27, 28) hat shown that the clay fraction (percent finer than two microns) and mineralogy perhaps are uost important in evaluating dy. If the soil clay fraction is less than about 15 percent, the soil behaves much like @ cohesionless soil, with 3; typically greater than 25° and not mich different from Jey. If the clay fraction is greater than 50 percent, $y is appreciably lover than Jey and is governed entirely by sliding of the clay minerals. For the most conon to 10° for illite, and then to 5° for montmorillonite. Soils with clay fractions between 15 and 50 clay minerals, Fy ranges approximately from 15* for kaolinite percent exhibit transitional behavior, as shown in Figure 4-24 ‘the value of dy also is stress-dependent because of curvature of the failure enve- Lope (22, 27, 28). Values given in Figure 4-24 are appropriate for an effective noraal stress equal to about one atmosphere. Figure 4-25a iIlustrates che typical changes in Gy vhich occur with changes in effective normal stress and plasticity 4-25 Residual Friction Angle, dy 40° r 30°S\ _ O —_-Range for 8 field sites ~ xo PI/CF =0.5 10 0.9 Son \ nds EN 20" Mat LON, 2 Yow So ae — oS 10° = tee Values of F, ot &/p9=! Bentonite—-a e—koolin Ce eee ° 20 40 5 30 100 Clay Fraction, CF (%) Figure 4-24. $y from Ring Shear Tests and Field Studies Source: Skempton (28), p. 14. 307 + r + ad Gog 02 g os t = oh 2 4 ore 3 (a) . 1 L L a) Plasticity Index, PL(%) 4 T , 2 Eg NN. 2 0 °. é ~~ 3 To) ° ] 6 oso is ao as Effective Normol Stress, ¢/po Figure 4-25. J, for Amuay Soils Source: Based on Lambe (29), p. 144. 4-26 index for the soils at the Amuay landslide sites. These curves essentially are parallel, indicating that the change in $y as 4 function of stress change is inde- pendent of the plasticity index. Re-plotting these changes in friction angle (Aér) results Figure 4-25. Other data (e.g., 27) are consistent with these sy values ‘The final values of 3 therefore should be evaluated from Figure 4-24, modified for effective normal stress level as given in Figure 4-256 UNDRAINED SHEAR STRENGTH OF COHESIVE SOILS - GENERAL EVALUATION BASIS ‘The undrained shear strength (sy) may very well be the most widely used paraneter for describing the consistency of cohesive soils. However, sy is not a fundamental material property. Instead, it is a measured xesponse of soil during undrained loading which assumes zero volume change. As such, sy is affected by the mode of testing, boundary conditions, rate of loading, confining stress level, initial stress state, and other variables. Consequently, although not fully appreciated by many users, sy is and should be different for cifferent test types (See Figure 1-1 for test types.). As described earlier in this section, it is appropriate to use a standard "test of reference", which is the isotropically consolidated, triaxial compression test for undrained loading (CIUC). With the CIUC test as a standard reference, the results of all other tests can be compared simply and conveniently. I¢ should be noted that simpler forms of triaxial test are available, such as the unconsolidated, undrained (UU) triaxial and unconfined (U) compression tests, With the UW test, a total confining stress is applied, but no soil consolidation is allowed under this confining stress. With the U test, the soil is unconfined with a zero confining Many detailed studies (e.g., 11, 23) have shown that the UU and U tests often are in gross error because of sampling disturbance effects and omission of a reconsoli- dation phase. Based on studies such as these, the CIUC test also is considered to be the minimum quality Laboratory test for evaluating the undrained shear strength of cohesive soils. Other simple tests such as the torvane and pocket penetrometer have an error potential that is comparable to that of the UU and U tests. There- fore, these tests should only be considered general indicators of relative beha- vior. They should never be used directly for design Since sy is stress-dependent, its value commonly is normalized by the vertical effective overburden stress (3yo) at the depth where s, is measured. This 4-27 undrained strength ratio, sy/dyo, has been expressed in many alternate forms in the Uterature, including 64/9, 4/9, cy/By, 6/p, ete. All are equal to s,/Bo, which will be used in the remainder of this section. Gorrelations with Index Paraneters for Undisturbed Clays Gorrelations with Index Parameters for Undisturbed Clays Early work by Skempton (30) suggested the general correlation in Figure 4-26 for sy determined from the field vane shear test (VST) as a function of the plasticity index. All of the data are for normally consolidated (NC) clays. A linear fit of these data results in: 8y(VST)/Byq = 0.11 + 0.0037 PI (4-18) In general, this relationship has been corroborated by others (e.g., 31), but there usually is more spread in the data than that shown in Figure 4-26. Recent vork by Chandler (32) suggests that this approximation may also be valid for OC clays, using the modification below with the preconsolidation stress ()): su(VST)/8p = 0.11 + 0.0037 Pr 19) He notes that the accuracy of this method vill be on the order of + 25 percent, but he cautions against its use in fissured, organic, sensitive, or other unusual clays. However, in a surprisingly large number of case histories, direct use of sy from the field VST in stability analyses of nuerous embankments, excavations, and foot- ings in clay has led to failures. Back-nalysis of these failures has led to 0 20 40°60 80 100 120 Plasticity Index, PI (%) Figure 4-26. sy(VST)/ayo versus PI for NC Clays Source: Skempton (30), p. 306. 4-28 empirical correction factors for the field VST. These factors will be described later in the section on sy correlations with the VST. Subsequent studies (e.g., 33) showed that sensitive clays with high liquidity index did not £ie the trend in Figure 4-26 very well. For these sensitive clays, the undrained strength ratio could be correlated better with the liquidity index, as shown in Figure 4-27, These data were obtained from triaxial compression tests on Nc clays ‘The undrained strength ratio for triaxial compression also can be determined from Critical State Soil Mechanics (CSSM) using the modified Cam clay model (e.g-, 34) For NC clay, this relationship is given by: su/84 = 0.129 + 0.00435 PT (4-20) in which 2; = effective overburden stress after isotropic consolidation. Other useful approximations include the following for low OCR clays with low to moderate PI (Jamiolkowski, et al., 35): Su/Bp ~ 0.23 + 0.06 (4-21) in which 3p = preconsolidation stress. Alternatively, Mesri (36) suggested the following: (s,/p)e1ue for NC Clay Liquidity Index, LT Figure 4-27. sy/dyo for NC Clay versus Liquidity Index Source: Bjerrum ané Simons (33), p. 722. 4-29 su/Bp = 0.22 4-22) In both cases, the sy corresponds approximately to direct simple shear (DSS) condi- Correlations with Index Parameters for Renolded Clays The sensitivity (Sz) is defined as sy in the undisturbed state divided by sy when remolded (both tested normally in unconfined compression at the sane natural water content), and therefore it is a measure of strength loss upon disturbance. Table 4-6 gives the typical terminology used to describe sensitivity, while Figure 4-28 illustrates a generalized relationship for sensitivity as a function of liquidity index and effective stress. The undrained renolded strength represents the lover bound on sy and, when Sp approaches one, sy = Sur- Figure 4-29 indicates that syy correlates reasonably well with the liquidity index. Data on undisturbed natural clays of low sensitivity are presented in Figure 4-30 and indicate good agreement with Figure 4-29, suggesting that sur is a fair predictor of sy for many clays of low sensitivity ‘The undrained shear strength for triaxial compression also can be predicted from the modified Cam clay model as follows (Wroth and Wood, 38) InS= (1-11) Ink (4-23) Table 4-6 CLASSIFICATION OF SENSITIVITY Clay Deseription Se Clay Description Se Insensitive #1 Slightly quick 8 to 16 Slightly sensitive Leo2 Medium quick 16 to 32 Medium sensitive 2to4 Very quick 32 to 64 Very sensitive 408 Extra quick > 64 Source: Mitchell (22), p. 208. Liquidity Index, LI on 1 10 100 Effective Vertical Stress, Fo/Pq Figure 4-28. General Relationship Between Sensitivity, Liquidity Index, and Effective Stress Source: Mitchell (22), p. 229. 4 oer 3 4 Approximote limits sb oo of data i Sef di ib dl ° : oul ead Jul tl oosot 600 Sor 6 io Remolded Undrained Shear Strength, 5y/Pg Figure 4-29. Remolded Undrained Shear Strength versus LI Source: Mitchell (22), p. 228 431 12 Y 74ZJ 2 ae z Range of dte ef a ee yj laylrgst7e 8h) / Lif Wi, Z -o4 om a4 si 0 Undrained Sheor Strength, $,/pg Figure 4-30. Undisturbed Undrained Shear Strength versus LI Source: Wood (37), p. 7 in which $ = 5y/(sy at WL) and R= (sy at ¥p)/(sy at wy). Typically, R= 100 and (Sy at wp) = 0.017 pg, yielding: Su/Pa ~ 1.7 e-4-6 LT (4-28) This equation is plotted as the straight line in Figure 4-30 and shows good agree- ment with the data in the range 0.1 < sy/pa < 3. General Behavior Under Triaxial Compression Loading ‘The undrained strength ratio in triaxial compression can be expressed in terms of more fundamental soil parameters by analysis of the Coulomb-Nohr failure envelope geometry. For Ko consolidation, the uncrained strength ratio is given as: [Ko + Ag(L = Ko)] sin dee 1+ (2dg - 1) sin Ste (Su/2vo eave ~ (4-25) in which Ko = coefficient of horizontal soil stress and Ag - Skempton’s pore water stress parameter, defined as: au - A053 Ae Tass (4-26) 4-32

You might also like