FINITE ELEMENT ALGORITHMS FOR ELASTOPLASTICITY AND CONSOLIDATION · FINITE ELEMENT ALGORITHMS FOR...

285
i FINITE ELEMENT ALGORITHMS FOR ELASTOPLASTICITY AND CONSOLIDATION by Andrew John Abbo B.E, B.Math A Thesis submitted for the Degree of Doctor of Philosophy at the University of Newcastle. February 1997 (3rdEdition, October 2005)

Transcript of FINITE ELEMENT ALGORITHMS FOR ELASTOPLASTICITY AND CONSOLIDATION · FINITE ELEMENT ALGORITHMS FOR...

i

FINITE ELEMENT ALGORITHMS

FOR ELASTOPLASTICITY AND

CONSOLIDATION

by

Andrew John Abbo

B.E, B.Math

A Thesis submitted for the Degree of

Doctor of Philosophy

at the University of Newcastle.

February 1997

(3rdEdition, October 2005)

ii

“I hereby certify that the work embodied in this Thesis is the result of original research

and has not been submitted for a higher degree to any other University or Institute”

(signed)

iii

ACKNOWLEDGEMENTS

The author gratefully acknowledges the financial assistance received through the

receipt of an Australia Postgraduate Award during his candidature. He also is

thankful for the ‘top up’ scholarship provided by the Department of Civil , Surveying

and Environmental Engineering at the University of Newcastle.

The author is indebted to Dr. Scott Sloan for his interest, guidance and provision of

financial assistance during this research. Dr. Sloan’s commitmentand assistancewere

limitless and this is greatly appreciated.

Thanks are also extended to Mr. Peter Kleeman and Dr. Mark Allman for their

valuable time spent proof reading this Thesis.

Finally, thankyou to my wife for her support and encouragement throughout the

period of my studies.

Preface to Third Edition

The third edition of this theis incorprates the minor changes listed below.

¯Minor changes to pagination due to bug in latest version of the publishing

software used in the preparation of the thesis.

¯Correction of some equations in chapter 2.

iv

ABSTRACT

Finite element analysis of nonlinear problems invariably uses piecewise linearisation

to generate approximate solutions. In geomechanics, this linearisation may appear

as:

¯ Discrete strain increments for the integration of nonlinear constitutive laws.

¯ Discrete load increments in nonlinear analyses.

¯ Discrete time steps in the analysis of consolidation.

The size and distribution of these increments (or steps) has a direct bearing on the

accuracy of the resulting solution.

This Thesis describes several new algorithms for controlling the error caused by the

use of discrete increments in nonlinear finite element analysis. The new schemes are

unified by the fact that they all treat the governing relations as a system of ordinary

differential equations. These equations are solved by adaptive integration with

respect to real or pseudo time, and automatically adjust the size of each step by

computing a local error measure. By holding this local error below a specified

threshold, the schemes aim to constrain the global linearisation error to lie near a

known tolerance.

Adaptive substepping schemes for controlling the linearisation error in the solution

of elastoplastic constitutive laws were first formulated by Sloan (1987). These

methods are explicit and automatically subincrement the imposed strain increment

at each stress point. Several important improvements to thesemethods aredeveloped

in this Thesis. The performance of the enhanced explicit schemes is compared to

several implicit schemes for a variety of boundary value problems. These examples

illustrate that adaptive explicit methods are very competitive with implicit methods,

and have the advantage of being simpler to implement for complex constitutive laws.

The remainder of the Thesis is concerned with the development of new adaptive

integration schemes for the solution of elastoplastic and coupled consolidation

v

problems. Thesemethods are applied at the global level and, for a givenmesh, govern

the overall accuracy of the solution. While the elastoplastic and consolidation

schemes both have essentially the same structure, they differ in the method used to

estimate the local error. The algorithm for integrating the global elastoplastic

equations uses an explicit forward Euler/modified Euler pair to provide the error

estimate and incorporates a correction to reduce drift from equilibrium. In contrast,

the consolidation algorithmuses an implicit pair of equations to ensure unconditional

stability. Numerical examples are presented which demonstrate the performance of

both types of schemes. The results suggest that the algorithms are not only efficient,

but also very robust. The latter attribute is very important in geomechanics

computations which often employ complex constitutive relations. While this Thesis

is concerned primarily with the behaviour of nonlinear solids, themethods developed

are quite general and can be extended to deal with many types of nonlinear problems

in structural mechanics.

vi

CONTENTS

ACKNOWLEDGEMENTS iii. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

ABSTRACT iv. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

CONTENTS vi. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

PREFACE ix. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

NOTATION xi. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1.0 INTRODUCTION AND HISTORICAL REVIEW 1. . . . . . . . . . . . . . . . . . .1.1 INTRODUCTION 2. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1.2 HISTORICAL REVIEW 6. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1.2.1 Plasticity 7. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1.2.2 Consolidation 10. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

2.0 GOVERNING EQUATIONS OF ELASTOPLASTICITY 15. . . . . . . . . . . . .2.1 INTRODUCTION 16. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

2.2 GOVERNING STRESS-STRAIN RELATIONS 16. . . . . . . . . . . . . . . . . . . . . . . . .

2.3 GOVERNING LOAD-DEFLECTION EQUATIONS 20. . . . . . . . . . . . . . . . . . . .

2.4 YIELD CRITERIA 25. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

2.4.1 Rounded Mohr-Coulomb Yield Function 27. . . . . . . . . . . . . . . . . . . . . . . . .

2.4.2 Rounded Hyperbolic Mohr-Coulomb Yield Function 31. . . . . . . . . . . . . . .

2.5 YIELD SURFACE GRADIENTS 34. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

2.5.1 Rounded Mohr-Coulomb Gradients 35. . . . . . . . . . . . . . . . . . . . . . . . . . . . .

2.5.2 Rounded Hyperbolic Mohr-Coulomb Gradients 37. . . . . . . . . . . . . . . . . . .

2.6 GRADIENT DERIVATIVES 39. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

2.7 NUMERICAL IMPLEMENTATION 41. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

APPENDICES 43. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .2A SUBROUTINE “YIELD” 44. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

2B SUBROUTINE “GRAD” 48. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

3.0 INTEGRATION OF STRESS-STRAIN RELATIONS 59. . . . . . . . . . . . . . . .3.1 INTRODUCTION 60. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

3.2 STRESS-STRAIN INTEGRATION 61. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

3.3 EXPLICIT INTEGRATION SCHEMES 65. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

3.3.1 Yield Surface Intersection 65. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

3.3.2 Negative Plastic Multiplier 68. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

3.3.3 Correction of Stresses to Yield Surface 73. . . . . . . . . . . . . . . . . . . . . . . . . . .

3.3.4 Modified Euler Scheme with Substepping 76. . . . . . . . . . . . . . . . . . . . . . . .

3.3.5 Single Step Modified Euler Scheme 85. . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

vii

3.3.6 Dormand-Prince Scheme with Substepping 85. . . . . . . . . . . . . . . . . . . . . . .

3.4 IMPLICIT INTEGRATION SCHEMES 89. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

3.4.1 Single Step Backward Euler Scheme 89. . . . . . . . . . . . . . . . . . . . . . . . . . . . .

3.4.2 Backward Euler Return Scheme 92. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

3.5 COMPARISON OF INTEGRATION SCHEMES 96. . . . . . . . . . . . . . . . . . . . . . . .

3.5.1 Rigid Strip Footing on Tresca Layer 99. . . . . . . . . . . . . . . . . . . . . . . . . . . . .

3.5.2 Rigid Strip Footing on Associated Mohr-Coulomb Layer 102. . . . . . . . . . . .

3.5.3 Rigid Strip Footing on Nonassociated Mohr-Coulomb Layer 106. . . . . . . .

3.6 CONCLUSIONS 107. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

4.0 INTEGRATION OF LOAD-DISPLACEMENT RELATIONS 109. . . . . . . . .4.1 INTRODUCTION 110. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

4.2 BACKGROUND 111. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

4.3 EXPLICIT INCREMENTAL METHODS 113. . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

4.4 MODIFIED EULER SCHEMEWITH SUBSTEPPING 115. . . . . . . . . . . . . . . . . .

4.4.1 Correcting for Drift from Equilibrium 119. . . . . . . . . . . . . . . . . . . . . . . . . . . .

4.4.2 Prescribed Force Loadings 121. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

4.4.3 Efficient Formation of the Global Stiffness Matrix 122. . . . . . . . . . . . . . . . .

4.4.4 Implementation 123. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

4.5 APPLICATIONS 126. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

4.5.1 Thick Cylinder 128. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

4.5.2 Rigid Strip Footing 134. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

4.5.3 Flexible Strip Footing 139. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

4.5.4 Rough Trapdoor 141. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

4.6 CONCLUSIONS 146. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

5.0 INTEGRATION OF CONSOLIDATION RELATIONS 147. . . . . . . . . . . . . .5.1 INTRODUCTION 148. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

5.2 BACKGROUND 148. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

5.3 FORMULATION OF GOVERNING BIOT CONSOLIDATION EQUATIONS151. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

5.4 SOLUTION OF ELASTIC CONSOLIDATION EQUATIONS 160. . . . . . . . . . . . .

5.4.1 Single-Step Schemes 161. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

5.4.2 Two-Step Schemes 162. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

5.4.3 Two-Stage Single-Step Schemes 164. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

5.5 AUTOMATIC TIME STEPPING SCHEME FOR ELASTIC CONSOLIDATION165. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

5.5.1 Theory 166. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

5.5.2 Scaling of Linear Equations 173. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

5.5.3 Implementation 174. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

5.6 AUTOMATIC TIME STEPPING SCHEME FOR ELASTOPLASTICCONSOLIDATION 178. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

5.6.1 Theory 179. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

viii

5.6.2 Implementation 184. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

6.0 CONSOLIDATION APPLICATIONS 191. . . . . . . . . . . . . . . . . . . . . . . . . . . . .6.1 INTRODUCTION 192. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

6.2 ELASTIC CONSOLIDATION 195. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

6.2.1 One Dimensional Compression of a Finite Layer 195. . . . . . . . . . . . . . . . . .

6.2.2 Finite Layer Compressed Between Two Rigid Plates 206. . . . . . . . . . . . . . .

6.2.3 Flexible Strip Footing on Finite Layer 213. . . . . . . . . . . . . . . . . . . . . . . . . . . .

6.3 ELASTOPLASTIC CONSOLIDATION 222. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

6.3.1 Drained and Undrained Analysis of Thick Cylinder 224. . . . . . . . . . . . . . . .

6.3.2 Undrained Analysis of Strip Footing 232. . . . . . . . . . . . . . . . . . . . . . . . . . . . .

6.3.3 Strip Footing with Associated Flow Rule 235. . . . . . . . . . . . . . . . . . . . . . . . .

6.3.4 Strip Footing with Nonassociated Flow Rule 243. . . . . . . . . . . . . . . . . . . . . .

6.4 CONCLUSIONS 248. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

7.0 CONCLUDING REMARKS 251. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .7.1 SUMMARY 252. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

7.2 ROUNDED APPROXIMATION TO THE MOHR-COULOMB YIELDCRITERION 252. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

7.3 INTEGRATION OF ELASTOPLASTIC CONSTITUTIVE LAWS 253. . . . . . . . .

7.4 SOLUTION OF ELASTOPLASTIC LOAD-DISPLACEMENT RELATIONS 254

7.5 SOLUTION OF THE GOVERNING EQUATIONS IN CONSOLIDATION 256.

REFERENCES 259. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

ix

PREFACE

The research work presented in this thesis was conducted in the Department of Civil,

Surveying and Environmental Engineering at the University of Newcastle from

February 1992 to February 1997. This work was performed under the supervision of

Dr. Scott Sloan. During the term of the candidature, a number of papers and reports

were published. These are listed below:

1. Abbo, A.J. and Sloan, S.W., ‘Accelerated initial stiffness schemes for

elastoplasticity’, Proceedings of the 5th International Conference on

Computational Plasticity, Barcelona, Spain, Invited paper (1997).

2. Abbo, A.J. and Sloan, S.W., ‘Load path control of iterative schemes’,

Proceedings of the 5th International Conference on Computational Plasticity,

Barcelona, Spain, Accepted for publication (1997).

3. Abbo, A.J. and Sloan, S.W., ‘An automatic load stepping algorithm with error

control’, International Journal for Numerical Methods in Engineering, 39,

1737-1759 (1996).

4. Abbo, A.J. and Sloan, S.W., ‘Automatic time step control in finite element

analysis of consolidation’, in Proceedings of the 7th Australia New Zealand

Conference in Geomechanics, Adelaide, Australia (1996).

5. Abbo, A.J. and Sloan, S.W., ‘A smooth hyperbolic approximation to the

Mohr-Coulomb yield criterion’, Computers and Structures, 54, 427-441 (1995).

6. Abbo, A.J. and Sloan, S.W., ‘An algorithm for controlling load path error in

non-linear finite element analysis’, Proceedings of the 8th International

Conference onComputerMethods andAdvances inGeomechanics,Morgantown,

USA, 1945-1950 (1994).

7. Abbo, A.J. and Sloan, S.W., ‘A comparison of integration schemes for

elastoplastic constitutive laws’,ResearchReport 091.02.1994,Department ofCivil

Engineering and Surveying, University of Newcastle, Australia (1994).

x

8. Abbo, A.J. and Sloan, S.W., ‘Backward Euler and subincrementation schemes

in computational plasticity’, Proceedings of the 2nd Asian-Pacific Conference on

Computational Mechanics, Sydney, Australia, 319-324 (1993).

9. Sloan, S.W. and Abbo, A.J., ‘Automatic load path control in non-linear finite

element analysis’, in Proceedings of the 2nd Asian-Pacific Conference on

Computational Mechanics, Sydney, Australia, 1295-1300 (1993).

Preface to Third Edition

The third edition of this theis incorprates the minor changes listed below.

¯Minor changes to pagination due to bug in latest version of the publishing

software used in the preparation of the thesis.

¯Correction of some equations in chapter 2.

Two addtional papers have subsequently been published based upon the content in

the final chapters of the thesis.

1. Sloan, S.W. and Abbo, A.J., ‘Biot consolidation analysis with automatic time

stepping and error control. Part 1: Theory and implementation’. International

Journal for Numerical and Analytical Methods in Geomechanics, 23, 467---492

(1999).

2. Sloan, S.W. and Abbo, A.J., ‘Biot consolidation analysis with automatic time

stepping and error control. Part 2: Applications’. International Journal for

Numerical and Analytical Methods in Geomechanics, 23, 467---492 (1999).

xi

NOTATION

All variables used in this Thesis are defined as they are introduced into the text. For

convenience, frequently used variables are described below. The general convention

adopted is that vector and matrix variables are shown in bold print while scalar

variables are shown in italic. In addition, lower case bold print is used to indicate

elemental vectors and matrices, while upper case bold print is used to indicate their

global counterparts.

a gradient to yield function.

b gradient to plastic potential.

B strain-displacement matrix.

c soil cohesion.

cv coefficient of consolidation.

C global matrix containing stiffness and coupling matrices.

Ce global matrix containing elastic stiffness and coupling matrices.

Cep global matrix containing elastoplastic stiffness and coupling matrices.

De elastic stress-strain matrix.

Dep elastoplastic stress-strain matrix.

Dp plastic contribution to stress-strain matrix.

E Young’s modulus.

f yield function.

F Forcing function.

f ext, Fext elemental/global external force vectors.

f int, Fint elemental/global internal force vectors.

g plastic potential.

xii

h time step for consolidation.

h element flow matrix.

H global flow matrix.

k permeability matrix.

K extended global flow matrix for consolidation.

ke,Ke elemental/global elastic stiffness matrix.

kep,Kep elemental/global elastoplastic stiffness matrix.

kp,Kp plastic contribution to elemental/global stiffness matrix.

l,L elemental/global coupling matrix.

m {1, 1, 1, 0, 0, 0}T

M diagonal matrix with entries {1, 1, 1, 0.5, 0.5, 0.5}

Nu displacement shape function matrix.

Np pore pressure shape function matrix.

p total pore pressure.

pe excess pore pressure.

ps steady state pore pressure.

p,P elemental/global vector of total pore pressures.

t time.

T pseudo time.

u,U elemental/global vector of displacements.

X Combined displacement/pore pressure vector for consolidation.

λ.

plastic multiplier rate.

ν Poisson’s ratio.

φ friction angle of soil.

xiii

ψ dilation angle of soil.

γw unit weight of water.

εe, εp elastic/plastic strain vector.

σm, σ , θ stress invariants.

θ,φ1, φ2,φ3 integration parameters.

1Chapter 1

CHAPTER 1

INTRODUCTION AND HISTORICAL

REVIEW

2Chapter 1

1.1 INTRODUCTION

The use of the finite element method is now widespread amongst academics,

researchers and practitioners in all branches of engineering. Although analysis of

linear problems is considered routine, application of the technique to study

nonlinear behaviour is far more demanding. Indeed, to solve elastoplastic and

consolidation problems with any degree of confidence, it is usually necessary to

have a detailed understanding of the approximations that are inherent in most

nonlinear solution strategies. A primary aim of this Thesis is to reduce the

complexity of elastoplastic and consolidation analysis by the design of advanced

algorithms with automatic error control. This step is essential if nonlinear finite

element codes are to be used successfully by practising engineers.

Finite element analysis of nonlinear problems invariably uses a piecewise linear

approximation to model the solution. This linearisation divides the analysis into

a number of discrete increments, each of which is considered in turn, and has a

direct bearing on the accuracy of the solution. The use of finite increments has

a most pronounced effect in elastoplastic computations, since elastoplastic theory

is founded on the notion of infinitesimal increments. In most situations, the size

of the increments is chosen on the basis of experience, with little prior knowledge

of the likely accuracy of the results. A cautious user, for example, may choose

many more increments than is necessary and waste valuable manhours and

computing time. An inexperienced user, on the other hand, may choose

inadequate increment sizes and obtain a solution with large linearisation errors.

For elastoplastic and consolidation analysis, the size and distribution of increments

necessary to gain a solution of a desired accuracy is unknown. In most cases, a

costly trial-and-error procedure is required to ensure that the overall linearisation

error is below acceptable limits.

In elastoplastic problems, the stress-strain behaviour at each numerical integration

point is, by definition, nonlinear. To determine the stresses at the end of a given

3Chapter 1

displacement increment, it is necessary to integrate the stress-strain relationships

over a known strain interval. One method for doing this involves dividing the total

strain increment into a suitable number of subincrements and then linearising the

local constitutive matrix for each of these in turn. The size of the strain increments

necessary to obtain an accurate solution with this approach is dependent on the

local nonlinearity of the yield surface and the hardening law.

When the global response of elastoplastic solids is analysed using the finite

element method, the nonlinear load-displacement behaviour is linearised by

dividing the total load into a number of discrete increments. Each of these load

increments is applied in sequence until the total external load is in place. The

accuracy of the resulting load-displacement response is a function of the size of

the discrete increments used in the analysis. To obtain an accurate solution, it is

usually perceived that larger increments should be used at the beginning of an

analysis, with smaller increments being needed as collapse approaches. This

perception, while intuitively appealing, is not necessarily true and will be explored

in later Chapters of this Thesis.

The analysis of Biot consolidation is different to elastoplasticity in that a set of

coupled differential equations needs to be solved to obtain the unknown

displacements and excess pore water pressures. Consolidation is a transient

process and, even for an elastic soil, a nonlinear displacement/pore pressure

response is observed. In finite element analysis, the governing consolidation

equations are linearised over a number of discrete time increments, each of which

is analysed sequentially. Because the rate at which consolidation occurs is driven

by the excess pore water pressures, the size of the time steps required to gain an

accurate solution is strongly related to the excess pore water pressure gradients

in the soil. During the early stages of consolidation, pore water gradients are high

and small time increments are appropriate. As the process proceeds and the

excess pore water pressures dissipate, the pore water gradients are reduced and

4Chapter 1

relatively large time increments may be used without degrading the accuracy of

the analysis.

Although a variety of methods have been proposed for the automatic selection of

increment sizes in nonlinear finite element analysis, their primary goal is usually

to ensure convergence of various iterative solution schemes. The purpose of the

schemes developed in this Thesis is to control the errors resulting from the

linearisation. Note that the error caused by the use of discrete increments is

distinct from the spatial discretisation error. The latter reflects the number, type

and distribution of elements in a given finite element mesh and is a separate issue.

It is intended that the methods developed in this Thesis will enable the

linearisation error to be automatically limited to a prescribed tolerance, without

any prior knowledge of the nonlinearities in the system.

The research presented in this Thesis can be divided into three principle areas:

i) The development of an automatic strain subincrementation scheme for the

integration of elastoplastic constitutive laws. This work builds on the

algorithm of Sloan (1987), but incorporates a number of important new

refinements. It also includes the formulation of a rounded hyperbolic

approximation to the Mohr-Coulomb yield surface.

ii) The development of an automatic load subincrementation scheme for solving

elastoplastic load-deflection equations.

iii) The development of an automatic time subincrementation scheme for the

solution of the coupled differential equations of elastic and elastoplastic Biot

consolidation.

For each of the above cases, the governing equations are formulated as a system

of ordinary differential equations and are solved using adaptive integration

procedures. These methods have proved very successful in the field of

mathematical numerical analysis. All of the numerical schemes developed in this

Thesis are based on the strategy of automatically subdividing a series of

5Chapter 1

user-defined ‘coarse’ increments into a number of smaller subincrements. The size

of these subincrements is chosen so that a local error measure is constrained to

lie near a prescribed tolerance. By controlling the local error in this manner, it

is usually possible to control the final global solution error to within an order of

magnitude of the same tolerance. For each step, the local error is computed as

the difference between a low order solution and a high order solution. This means

that two different integration methods, whose order of accuracy differs by one,

need to be used over each subincrement. In order to be efficient, it is crucial for

these methods to minimise the number of evaluations and factorisations of the

governing matrix equations.

The structure of this Thesis reflects the three main topics listed above. Chapter

2 provides a background to some selected aspects of computational plasticity. It

begins with a derivation of the stress-strain and load-displacement relations that

form the basis of finite element elastoplasticity. The Chapter finishes with the

development of a rounded hyperbolic approximation to the Mohr-Coulomb yield

function. This yield surface is free of the gradient singularities associated with the

traditional Mohr-Coulomb surface, and is thus ideally suited for numerical

computation.

Chapter 3 is concerned with schemes for the integration of elastoplastic

stress-strain relations. The Chapter describes an improved version of the adaptive

explicit scheme of Sloan (1987), and compares its performance with that of several

implicit schemes. The refinements to Sloan’s original algorithm deal with a

number of key issues that are often overlooked in the numerical integration of

elastoplastic stress-strain relations, and result in a very efficient and robust

scheme.

In Chapter 4, an automatic load incrementation scheme for the solution of

elastoplastic load-deflection equations is developed. The strategy adopted here

is very similar to that used for integrating the stress-strain relations, and uses the

6Chapter 1

same pair of explicit solution techniques to provide a local estimate of the error

in the displacements. The algorithm assumes that a number of coarse load steps

are supplied, and then subincrements these to control the local error measure to

lie near a user-specified tolerance. In the last part of Chapter 4, detailed tests

on a range of boundary value problems are performed to establish the error

control properties and efficiency of the new scheme.

The final part of this Thesis focuses on the solution of the Biot consolidation

equations. The nature of these equations is very different to those of

elastoplasticity since they require the use of implicit integration schemes for

unconditional stability. In Chapter 5, the governing Biot equations which describe

the consolidation process, are formulated. Adaptive integration schemes for the

solution of these coupled equations are then developed for both elastic and

elastoplastic materials. These schemes control the time stepping error in the

computed displacements and excess pore pressures by automatically selecting

suitable time increments. The methods assume that a number of coarse time steps

have been supplied, together with a desired error tolerance. The performance of

these automatic algorithms is demonstrated in Chapter 6, where a variety of

numerical examples are considered.

1.2 HISTORICAL REVIEW

The fundamental theories of plasticity and consolidation evolved separately and

it was not until the 1970’s that the two were combined. Indeed, the study of

plasticity dates back to the 19th century, whereas the modelling of consolidation

did not receive significant attention until the work of Terzaghi (1923) and Biot

(1941a) in the middle of the 20th century.

A brief outline of some significant contributions to the development of plasticity

and consolidation theory is given below. Attention is focused primarily on

contributions made to the finite element method which are relevant to this Thesis.

7Chapter 1

1.2.1 Plasticity

The foundations of plasticity theory can be traced back to Tresca (1864), whose

studies on punching and extrusion led to the development of his well known yield

criterion. Other major contributions to the development of plasticity theory were

made by Saint-Venant (1870), Lévy (1870) and Mohr (1900). Mohr’s work on

describing the limits of elastic behaviour was used by other researchers to

determine the onset of plasticity. Mohr found that these limits were governed by

combinations of the shear and normal stresses. In 1913, von Mises suggested

another yield criterion which was suitable for metals. He also introduced the

concept that the direction of plastic deformation was related to the yield surface.

More recently, significant contributions to the development of plasticity theory

have been made by Henky (1924) and Prandtl (1924) amongst others. A detailed

account of the history of plasticity theory may be found in Hill (1950), who also

presents solutions to many classical problems.

Although Tresca’s work is usually considered to be the origin of plasticity theory,

Coulomb proposed a yield criterion almost a century earlier in 1773. Coulomb

also developed the notion that failure occurred along a plane and applied this to

study earth pressures on retaining walls. Further studies on earth pressures were

conducted by Rankine (1857) and Bell (1915), who invoked the concept of plastic

equilibrium.

The coupling of plasticity theory with the finite element method stems from the

early work of Marcal and King (1967). Yamada et al (1968), and later Zienkiewicz

et al (1969), followed this work and developed the governing elastoplastic relations

in a form suitable for finite elements. The incremental scheme of Yamada et al

(1968) forced the elements to yield one by one by restricting the size of each load

increment. After each element had yielded, the stiffness was changed and the next

load increment applied. While this approach is quite novel, and reduces the load

path error introduced by the use of discrete load increments, it is very inefficient

8Chapter 1

and unsuitable for complex yield criteria. The work of Zienkiewicz et al (1969)

used load increments of arbitrary size and proposed the initial stress iteration

technique. This scheme uses the elastic stiffness matrix to iterate to equilibrium

after each load increment has been applied, and was widely employed for many

years. The algorithm discussed in their paper included a number of refinements,

such as dividing the strain increment into elastic and plastic portions when a point

first undergoes plastic yielding. Although it is particularly robust, the initial stress

technique has a very slow convergence rate once significant plastic yielding has

occurred. In a landmark paper several years later, Nayak and Zienkiewicz (1972a)

extended this work and introduced the practice of subincrementing the strain

increments in order to evaluate the stresses more accurately. In the absence of

a better approach, they chose the number of subincrements on the basis of an

empirical rule and corrected the stresses back to the yield surface after each

substep. This paper also discussed the use of nonassociated flow rules, which are

important for modelling geomaterials, and also introduced a variety of iteration

schemes for solving the nonlinear stiffness equations.

Following the development of nonlinear finite element analysis, a significant

amount of attention has been focused on the design of solution strategies for the

governing equations. Much of this work has been driven by research on nonlinear

structures, but the methods are generally applicable to the study of nonlinear solid

behaviour as well. Broadly speaking, these techniques can be classified into the

categories of incremental or iterative methods. Incremental schemes approximate

the nonlinear response of a system by using a series of piecewise linear steps, and

are closely related to the large family of explicit methods which are used for

solving systems of ordinary differential equations. Provided the system of linear

equations to be solved in each step remains well conditioned, these techniques are

extremely robust. This property makes them attractive for geomechanics studies

which frequently employ very complex constitutive laws.

9Chapter 1

Rather than treating the governing relations as a system of ordinary differential

equations, iterative schemes attempt to solve the nonlinear equations directly.

Well known examples of iterative schemes include the Newton-Raphson, modified

Newton-Raphson, and initial stress methods. Iterative solution techniques for

nonlinear systems typically apply the unbalanced forces, compute the

corresponding displacement increments, and then repeat this procedure until the

drift from equilibrium is small. One major disadvantage of the Newton-Raphson

family of algorithms is that the iterations may not converge, particularly when the

behaviour is strongly nonlinear. To overcome this, various techniques have been

developed to stabilise and accelerate the convergence of Newton-Raphson

schemes. These include the line search techniques of Matthies and Strang (1979)

and Crisfield (1983,1984), as well as the arc length control procedures developed

by Wempner (1971) and Riks (1972,1979). Line search methods attempt to

stabilise Newton-Raphson iterations by shrinking or expanding the current

displacement increment to minimise the resulting unbalanced forces. In cases

where the current search direction is poor, or where the unbalanced forces are

nonsmooth functions of the displacements, line searches may be of limited use.

The philosophy behind arc length methods is to force the Newton-Raphson

iterations to remain within the vicinity of the last converged equilibrium point.

This means that the applied load must be reduced as the iterations proceed, but

greatly reduces the risk of divergence for strongly nonlinear problems. A detailed

discussion of various arc length methods, and their practical implementation, can

be found in Crisfield (1991). In a relatively recent development, Simo and Taylor

(1985) derived the consistent tangent technique for use with the Newton-Raphson

scheme. By incorporating high order terms that are usually ignored in the

standard form of the elastoplastic stiffness relations, this procedure gives a full

quadratic rate of convergence. Although powerful, the method is difficult to

implement for complex yield criteria because it is necessary to evaluate second

derivatives of the yield function.

10Chapter 1

To date, most automatic load incrementation algorithms have focused on ensuring

the convergence of various iterative solution schemes, with little attention being

given to the problem of controlling the overall load path error directly. Because

of the complex nature of elastoplastic stiffness equations, a variety of ad hoc

strategies and parameters have been used to decide when to increase or decrease

the increment size. Den Heijer and Rheinboldt (1981), for example, used the

curvature of the nonlinear path, while Bergan et al (1978) and Bergan and Soreide

(1978) developed the useful concept of the ‘current stiffness parameter’. More

recently, Crisfield (1981, 1991) recorded the number of iterations required to

restore equilibrium in each load increment and fed this into an empirical formula

to predict the size of the next increment. In a different approach, Schellekens et

al (1992) proposed a load incrementation method which is based on strain energy.

A major disadvantage of all empirical load incrementation schemes is that they

do not control the load path error directly. Even if the equilibrium iterations

converge satisfactorily, the error in modelling the load path may still be quite

large. Indeed, for problems with strongly nonlinear strain paths, the displacements

and stresses computed from analysis with large increments may differ greatly from

the correct displacements and stresses that would be found from analysis with very

small increments. For cases where the strain path is only moderately nonlinear,

large increments may be used with confidence.

1.2.2 Consolidation

The mathematical analysis of consolidation began with Terzaghi who, in 1923,

presented a model for one dimensional consolidation. In 1936, Rendulic proposed

a pseudo three dimensional theory of consolidation which, like Terzaghi’s theory,

had governing equations of the same form as the diffusion equation. Due to some

fundamental assumptions that were made about the behaviour of the stresses, this

model failed to couple the magnitude and rate of settlement properly for two or

three dimensions. Despite this shortcoming, Davis and Poulos (1972) used the

11Chapter 1

pseudo theory to obtain solutions for the rate of settlement of strip and circular

footings. A more rigorous three dimensional consolidation theory, which

overcomes the deficiencies of Rendulic’s formulation and provides compatibility

between the displacements and pore water pressures, was developed by Biot in

1941a. In a series of subsequent papers, Biot (1955, 1956a, 1956b, 1963) extended

this theory to include the effects of anisotropy, viscoelasticity, and initial stresses.

Relatively few analytical solutions have been obtained using Biot consolidation

theory due to the complex nature of the equations and boundary conditions. Biot

(1941b) himself considered the consolidation of a rectangular area, whilst Mandel

(1953) presented a solution for the consolidation between two rigid plates. A

decade later, Cryer (1963) formulated solutions for the consolidation of a sphere

using both Biot’s and Terzaghi’s theory. Both Mandel and Cryer predicted an

initial increase in pore water pressures, despite the load on the soil remaining

constant. This phenomena, which is caused by the redistribution of the total

stresses, has become known as the Mandel-Cryer effect. Solutions for the

consolidation of a semi-infinite mass under various load configurations have been

presented by de Josselin de Jong (1957), McNamee and Gibson (1960), Gibson

and McNamee (1963), and Schiffman et al (1969). Other consolidation solutions,

for footings resting on layers of finite depth, have been developed by Gibson et

al (1970) and Booker (1974). More recently, Chiarella and Booker (1975) derived

the consolidation solution for the settlement of a rigid die on a deep layer of clay.

The application of the finite element method to the solution of Biot’s

consolidation equations was first considered by Sandu and Wilson in 1969. Since

then, a number of finite element formulations for the consolidation of elastic

materials have appeared. These include the works of Christian and Boehemer

(1970), Yokoo et al (1971a,b), Hwang et al (1971), Krause (1978) and, most

recently, Borja (1986). A novel formulation, which is based on equilibrium

12Chapter 1

elements and uses total stresses and pore water pressures as the unknown

variables, has been developed by Cividini and Gioda (1982).

The extension of Biot’s equations to include elastoplastic behaviour was first

presented by Small et al (1976). This work used the Mohr-Coulomb yield criterion

and, for the two extreme cases of drained and undrained loading, verified the

consolidation solutions against the solutions obtained from straight elastoplastic

theory. At about the same time, Lewis et al (1976) employed a hyperbolic

stress-strain relationship and variable permeability to model nonlinear

consolidation. Other significant nonlinear studies include those of Ghaboussi and

Wilson (1973), who modelled the influence of pore fluid compressibility, and

Carter et al (1977, 1979) who incorporated the effects of finite deformations. More

recently, the text by Lewis and Schrefler (1987) covers many of the different types

of nonlinearities that may be associated with consolidation and also provides a

complete listing of a finite element program in FORTRAN.

Solution techniques for finite element analysis of Biot consolidation are usually

based on first order, implicit integration methods. The backward Euler scheme,

for example, is widely used in both linear and nonlinear studies. In the latter case,

a system of nonlinear equations must be solved in order to advance the solution

for each time step. The stability and accuracy of first order integration schemes

has been investigated by Booker and Small (1975). They proved that the

integration parameter must not be less than 0.5 in order for the solution scheme

to be unconditionally stable. In a different type of study, Vermeer and Verruijt

(1981) suggested that the time steps should not be made too small to avoid

oscillations in the pore pressures. The time stepping schemes commonly used for

elastic consolidation analysis are, in fact, identical to the family of techniques used

in the solution of first order differential equations. A vast amount of literature

exists on the accuracy and stability of these methods and an excellent summary

can be found in Wood (1990).

13Chapter 1

A number of general integration methods, which were developed for systems of

second order differential equations but are also applicable to systems of first order

differential equations, have been presented by Zienkiewicz et al (1984) and

Thomas and Gladwell (1988a). All of these schemes use an estimate of the local

truncation error to control the time step size and were primarily designed with

dynamics problems in mind. In the methods of Zienkiewicz et al, the local error

is found from a Taylor series expansion. Although the time steps may expand or

contract as the analysis proceeds, no effort is made to control the error in the

solution precisely. Thomas and Gladwell, on the other hand, use the difference

between solutions from pth and (p+1)th order schemes to estimate the local

truncation error. This error measure is used to adjust the size of every time step.

In the analysis of elastoplastic soils with implicit solution schemes, various

strategies have been adopted for solving the resulting systems of nonlinear

equations. Small et al (1976), for example, used an initial stiffness iteration

scheme with time-averaged values to calculate the unbalanced forces. Siriwardane

and Desai (1981) developed two different methods. The first of these, a tangent

stiffness scheme, employs no iteration and thus suffers from the disadvantage of

permitting the solution to drift from equilibrium. The second of their schemes,

an iterative initial stiffness solver, does not have this shortcoming since it attempts

to satisfy equilibrium over each time step. Other notable schemes for elastoplastic

consolidation include the predictor corrector methods of Prevost (1982,1983) and

the composite Newton and multi-step methods of Borja (1991a,b). Borja (1989)

also derived a consistent tangent algorithm which exploits the fast quadratic

convergence of the Newton-Raphson scheme. More recently, Bostrøm et al

(1995), who compared the suitability of various consolidation elements for

predicting collapse loads, implemented a cylindrical arc length method.

14Chapter 1

15Chapter 2

CHAPTER 2

GOVERNING EQUATIONS OF

ELASTOPLASTICITY

16Chapter 2

2.1 INTRODUCTION

Application of the finite element method to analysis of elastoplastic problems

involves the solution of two sets of ordinary differential equations, namely:

i) The incremental stress-strain relations.

ii) The global load-deflection equations.

The accurate solution of these differential equations is a key theme of this Thesis

and this Chapter begins by deriving their precise form.

The remainder of the Chapter is concerned with the development of a smooth

yield surface that eliminates all singularities from the Mohr-Coulomb yield

criterion. The new surface uses a hyperbolic approximation in the meridional

plane and a trigonometric rounding technique in the octahedral plane. It is both

continuous and differentiable for all stress states, and can be fitted to the

Mohr-Coulomb yield surface by adjusting two parameters.

2.2 GOVERNING STRESS-STRAIN RELATIONS

Depending upon its current stress state, an elastoplastic material is assumed to

behave either as an elastic solid or a plastic solid. The transition from elasticity

to plasticity is described by the yield criterion which forms a surface in three

dimensional principle stress space. Stress states lying within the yield surface are

regarded as elastic, while stress states lying on the yield surface are plastic. As

the material deforms plastically, the stresses must remain on the yield surface and

so stress states lying outside the yield surface are inadmissible. For an elastoplastic

material with isotropic hardening, the yield surface is described by a yield function

of the form f (σ, À), where σ is a vector of the current stresses and À is some

hardening parameter. If f (σ, À)< 0, the stress point lies within the yield surface

and the material behaves elastically according to

σ= De ε (2.1)

17Chapter 2

where De is the elastic stress-strain matrix, σ= σx , σy , σz , τxy , τxz , τyzT is a

vector of stress components, and ε= εx , εy , εz , γxy , γxz , γyzT is a vector of strain

components.

Once yielding takes place, f (σ, À)= 0 and the stresses remain on the yield surface

as plastic deformation occurs. Letting a superior dot denote a derivative with

respect to time, this constraint is enforced by the consistency condition

f.= ∂f∂σ

T

σ. +∂f∂À À

. = aTσ. +∂f∂À À

. = 0 (2.2)

where σ. is a vector of stress rates, À. is a hardening rate, and

a=∂f∂σ= ∂f∂σx , ∂f

∂σy,∂f∂σz,∂f∂τxy,∂f∂τxz,∂f∂τyz

T

is the gradient to the yield surface. At this stage, elastoplastic theory makes two

key assumptions. The first is that the total strain rate, ε. , can be expressed as the

sum of an elastic strain rate, ε. e, and a plastic strain rate, ε.p, according to

ε. = ε. e+ ε

.p (2.3)

The second is that the direction of the plastic strain rates is normal to a surface

called the plastic potential. This assumption, which is termed the flow rule, can

be expressed as

ε.p= λ

. ∂g∂σ= λ

.b (2.4)

where g is the plastic potential, λ.is a positive constant known as the plastic strain

rate multiplier, and

b= ∂g∂σ= ∂g∂σx , ∂g∂σy,∂g∂σz,∂g∂τxy,∂g∂τxz,∂g∂τyz

T

is the gradient to the plastic potential. For convenience, the plastic potential is

usually assumed to have a form similar to that of the yield criterion. When the

18Chapter 2

gradients to the plastic potential and the yield criterion are coincident, plastic flow

takes place in a direction which is normal to the yield surface and the flow rule

is said to be associated. Any other type of flow rule is said to be nonassociated.

Associated flow rules are often used in metal plasticity studies and a number of

important uniqueness theorems can be derived for them (Hill, 1950).

Differentiating (2.1) with respect to time and substituting equations (2.3) and (2.4)

gives

σ. = De ε

. − λ.De b (2.5)

Inserting (2.5) in the consistency condition (2.2), the plastic multiplier may be

written as

λ.=

aTDe ε.

A+ aTDeb(2.6)

where the parameter A is given by

A=−∂f∂ÀÀ.

λ. (2.7)

Substituting the expression for λ.from (2.6) into (2.5) furnishes the standard

elastoplastic stress-strain relations of the form

σ. = Dep ε

. (2.8)

where

Dep= De−De b aTDeA+ aTDeb

(2.9)

is known as the elastoplastic stress-strain matrix. Given that the strain rate ε. is

known, equation (2.8) describes a small system of ordinary differential equations

which can be integrated over a specified time interval to obtain the unknown

stresses and hardening parameter. The initial conditions for this system are the

known stresses and hardening parameter at the start of the time interval.

19Chapter 2

The elastoplastic stress-strain matrix (2.9) may also be expressed as a combination

of elastic and plastic components according to

Dep= De−Dp (2.10)

In this equation, De is the usual elastic stress-strain matrix and

Dp=Deb aTDeA+ aTDeb

represents the plastic contribution to the elastoplastic stress strain matrix. This

decomposition of the elastoplastic stress-strain matrix provides substantial

computational efficiencies, and will be discussed in a later Chapter.

The precise form of the parameter A depends on the type of hardening model that

is adopted. For an isotropic strain hardening model, the hardening parameter À

is assumed to be related to the equivalent plastic strains according to

À. = ε

.p= λ

. 23bTMb (2.11)

where M is the diagonal matrix

M=⎪⎪⎪⎪⎪

1110.50.50.5

⎪⎪⎪⎪⎪

Substituting (2.11) in (2.7) gives an explicit expression for A as

A=−∂f∂À

23bTMb

For an isotropic work hardening model, such as the one discussed by Hill (1950),

À is assumed to be related to the plastic work according to

À. = W

.

p= σTε.p= λ

.σTb (2.12)

20Chapter 2

In this case, equation (2.7) gives the parameter A as

A=−∂f∂Àσ

Tb

Both of the hardening models require integration over the strain path to give the

hardening parameter. Equations (2.11) and (2.12) define the ordinary differential

equations that need to be integrated over each specified time interval to give À.

2.3 GOVERNING LOAD-DEFLECTION EQUATIONS

Consider a body with volume V and surface area S. The stresses within the body

must be in equilibrium and this condition provides a starting point for formulating

the governing load-deflection equations in finite element analysis. The equations

of equilibrium for a three dimensional solid are

∂σx∂x +

∂τxy∂y +

∂τxz∂z + bx= 0

∂σy∂y +

∂τyx∂x +

∂τyz∂z + by= 0 (2.13)

∂σz∂z +

∂τzx∂x +

∂τzy∂y + bz= 0

where bx, by and bz are components of the body force in the x-, y- and z-directions

respectively. These equations can be expressed in the compact form

∇Tσ+ b= 0 (2.14)

where ∇ denotes the differential operator

∇T=⎪

⎪⎨

∂∂x

0

0

0

∂∂y

0

0

0

∂∂z

∂∂y∂∂x

0

∂∂z

0

∂∂x

0

∂∂z∂∂y⎪

⎪⎬

(2.15)

and b= bx , by , bzTrepresents a vector of body forces (not to be confused with

the plastic potential gradient b in the previous Section). By satisfying the

21Chapter 2

equilibrium equations throughout the body, together with any associated boundary

conditions, the stresses, strains, and deformations within the body can be

determined. One technique for expressing the equilibrium equations in an average

sense is the method of weighted residuals. When applied to equation (2.14), this

method requires that

W= V

wT∇Tσ+ bdV= 0

where w= wx , wy , wzTis a vector of arbitrary weighting functions with

components in the x- , y- and z-directions. Integrating the above equation by parts

using the Green-Gauss theorem gives the so-called weak form of the equilibrium

equations as

V

(∇w)Tσ dV−V

wTb dV−S

wTtdS= 0 (2.16)

where t= tx , ty , tzTis a vector of surface tractions which act over the boundary

surface S. These tractions must satisfy the boundary conditions

txtytz

===

σx nxτxy nxτxz nx

+++

τxy nyσy nyτyz ny

+++

τxz nzτyz nzσz nz

in which nx, ny and nz are direction cosines of the unit normal to the surface S.

An approximate form of equation (2.16) can be obtained by the application of the

finite element method. This involves dividing the body into a number of

sub-domains, known as elements, over which the displacements are approximated

by interpolation of the nodal displacements. For an element with n nodes, the

displacement field at any internal point is expressible in the form

d= Nu (2.17)

where d= u , v ,w is a displacement vector with components u, v and w in each

coordinate direction, N is a matrix of shape functions

22Chapter 2

N=⎪⎡

N100

0N10

00N1

N200

0N20

00N2

...

...

...

Nn00

0Nn0

00Nn⎪⎤

⎦(2.18)

and u is a vector of element nodal displacements

u= u1, v1, w1, u2, v2, w2, ... un, vn, wnT

(2.19)

The corresponding internal strains are given by differentiating (2.17) to give

ε= Bu (2.20)

where B is the element strain-displacement matrix

B= ∇N=

⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪

∂N1∂x

0

0

∂N1∂y∂N1∂z

0

0

∂N1∂y

0

∂N1∂x

0

∂N1∂z

0

0

∂N1∂z

0

∂N1∂x∂N1∂y

...

...

...

...

...

...

∂Nn∂x

0

0

∂Nn∂y∂Nn∂z

0

0

∂Nn∂y

0

∂Nn∂x

0

∂Nn∂z

0

0

∂Nn∂z

0

∂Nn∂x∂Nn∂y

⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪

Having defined the functional form of the displacements, the weighting functions

may be chosen as

w= δd= N δu (2.21)

where δu is a vector of arbitrary nodal displacements for an element. Using this

weighting scheme, the method of weighted residuals reduces to the classical

approach of Galerkin (1915). Substituting the weighting functions (2.21) into

(2.16), integrating over the element volume, V e, and surface area, Se, and

collecting terms furnishes

δuT⎪⎧⎩VeBTσ dV−

SeNTt dS−

VeNTb dV⎪⎫⎭

= 0 (2.22)

23Chapter 2

Since the displacements δu are arbitrary, it follows that

VeBTσ dV−

SeNTt dS−

VeNTb dV= 0 (2.23)

for (2.22) to be true in the general case. These equations, which govern the

behaviour of each finite element, are applicable to any constitutive relationship.

Since they describe the overall equilibrium conditions, they are often written in

the form

f ext− f int= 0 (2.24)

where the external forces are given by

f ext= VeNTb dV+

SeNTt dS (2.25)

and the internal forces are defined as

f int= VeBTσ dV (2.26)

The vector f ext comprises the nodal forces exerted on the element due to the

applied loading, while the vector f int comprises the nodal forces which are

supported by the internal stresses in the element.

For nonlinear problems it is necessary to develop an incremental or rate form of

equation (2.24). This may be obtained by differentiating equation (2.24) with

respect to time t and using the chain rule to give

dfdt

ext− dfdu

intdudt= 0 (2.27)

where the derivative of f int with respect to u is the Jacobian matrix

24Chapter 2

dfdu

int=

⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪

∂f int1∂u1

∂f int2∂u1⋮

∂f intn∂u1

∂f int1∂u2

∂f int2∂u2⋮

∂f intn∂u2

...

...

...

∂f int1∂un

∂f int2∂un⋮

∂f intn∂un

⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪

Now, from equation (2.25), it follows that

dfdu

int=

VeBT dσdudV=

VeBT dσdεdεdudV (2.28)

Neglecting terms involving second derivatives of the yield function with respect to

the stresses, the derivative of σ with respect to ε is

dσdε= Dep (2.29)

where Dep is the elastoplastic stress-strain matrix defined in (2.9). Similarly, the

derivative of the strains with respect to the nodal deflections is obtained by

differentiating the strain-displacement relations (2.20) to give

dεdu= B (2.30)

Substituting (2.29) and (2.30) in (2.28) leads to the definition of the elastoplastic

tangent stiffness matrix, kep, according to

dfdu

int=

VeBTDepB dV= kep (2.31)

Finally, combining equations (2.31) and (2.27) and rearranging gives

dudt= k–1ep dfdt

ext

or

25Chapter 2

u. = k–1ep f. ext (2.32)

where the superior dot again denotes a derivative with respect to time. The

relations (2.32) define a system of ordinary differential equations which govern the

load-deformation behaviour of a single elastoplastic element. Adding these

element contributions together in the usual way gives a system of ordinary

differential equations of the form

U.= K–1ep F

. ext (2.33)

where

Kep= elements

kep= elements

VeBTDepB dV

and

F. ext=

elements

f. ext=

elements

VeNTb

.dV+

elements

SeNT t

.dS

are, respectively, the global elastoplastic stiffness matrix and the global external

force rate vector. The relations (2.33) describe, in rate form, the global

load-displacement behaviour of a mesh of elastoplastic finite elements. The initial

conditions for these ordinary differential equations are the displacements, stresses

and hardening parameters which are known at the start of each time interval.

2.4 YIELD CRITERIA

The Mohr-Coulomb yield criterion, with either an associated or nonassociated

flow rule, is used widely in geotechnical analysis. Although more sophisticated

constitutive laws are available for predicting the behaviour of real soil, this simple

model has the important advantage that all of its parameters have direct physical

meanings and can be measured using conventional tests. Although it predicts an

excessive amount of dilation upon plastic shearing, the Mohr-Coulomb yield

26Chapter 2

criterion is traditionally employed with an associated flow rule. This type of model

has been assumed in many limit equilibrium and classical plasticity solutions

which, apart from being valuable in their own right, are especially useful for

validating finite element codes.

A plot of the Mohr-Coulomb yield surface in three dimensional stress space is

shown in Figure 2.1. When used in displacement finite element analysis, this

function presents a number of computational difficulties due to the gradient

discontinuities which occur at the tip and edges of the hexagonal yield surface

pyramid. For stress states lying precisely on these singularities, the gradient

vectors a and b are undefined and the elastoplastic constitutive matrix in equation

(2.9) cannot be computed. This prevents the elastoplastic stresses from being

found and also results in the elastoplastic tangent stiffness matrix of equation

(2.31) becoming undefined. Problems also arise if the stresses are in close

proximity to the singularities, since the gradients rapidly become ill-conditioned.

To avoid these difficulties, this Section describes the development of a smoothed

yield surface that removes all gradient singularities from the Mohr-Coulomb yield

criterion. The new surface uses a hyperbolic approximation in the meridional

Figure 2.1Mohr-Coulomb yield function in principal stress space.

− σ1

− σ3

− σ2

σ1= σ2= σ3hydrostatic axis

27Chapter 2

plane to eliminate the tip singularity and a trigonometric rounding in the

octahedral plane to eliminate the edge singularities. It is both continuous and

differentiable for all stress states, and can be fitted to the Mohr-Coulomb yield

surface by adjusting two parameters.

2.4.1 Rounded Mohr-Coulomb Yield Function

Defining tensile stresses as positive, the Mohr-Coulomb yield function may be

written as

f= (σ1− σ3)+ (σ1+ σ3) sinφ− 2c cosφ= 0 (2.34)

where the principal stresses are ordered so that σ1≥ σ2≥ σ3 and c and φ denote,

respectively, the cohesion and friction angle of the soil.

For computational convenience, the Mohr-Coulomb criterion can be expressed in

terms of the three stress invariants originally proposed by Nayak and Zienkiewicz

(1972b). These quantities are written as (σm, σ, θ) and are defined by

σm= 13 (σx+ σy+ σz)

σ= 12s2x+ s2y + s2z + τ2xy+ τ2yz+ τ2zx

θ= 13 sin–1− 3 32 J3

σ 3 , (− 30˚≤ θ≤ 30˚)

where

J3= sx sy sz+ 2 τxy τyz τzx− sx τ2yz− sy τ2xz− sz τ2xy

and

sx= σx− σm , sy= σy− σm , sz= σz− σm

In terms of these invariants, the principal stresses are given by

28Chapter 2

σ1=23σ sin(θ+ 120˚)+ σm (2.35)

σ2=23σ sin(θ)+ σm

σ3=23σ sin(θ− 120˚)+ σm (2.36)

Substituting (2.35) and (2.36) in (2.34), the Mohr-Coulomb yield criterion may be

expressed in the equivalent form

f = σm sinφ+ σ K(θ)− c cosφ = 0 (2.37)

where the function K is

K(θ)= cos θ− 13sinφ sin θ (2.38)

In the octahedral plane, defined by σm = constant, the shape of the yield function

is defined by the relationship between σ and θ. When viewed in this plane, the

Mohr-Coulomb surface has sharp vertices (and hence gradient discontinuities) at

θ= 30˚ as shown in Figure 2.2. It is necessary to permit the gradients to be

computed at these stress states since they are often encountered in finite element

analysis. Various techniques for dealing with these corners have been discussed

by Zienkiewicz and Pande (1977), Owen and Hinton (1980) and Sloan and Booker

(1986). Of these methods, the Sloan and Booker procedure has the advantage that

it uses a trigonometric rounding only in the vicinity of the vertices and thus models

the Mohr-Coulomb yield surface very closely. Because the modified yield surface

is internal to the Mohr-Coulomb criterion, this approximation also ensures that

the strength is modelled conservatively. Except for tensile hydrostatic stress states,

the Sloan and Booker (1986) surface is continuous and differentiable for all stress

states, and can be fitted to the Mohr-Coulomb surface as closely as desired by

adjusting a single parameter.

29Chapter 2

Figure 2.2Mohr-Coulomb yield function in octahedral plane.

σ1

σ3

σ2

σ1≥ σ2≥ σ3

θ= 30˚

θ=− 30˚

2 σ

θ

Sloan and Booker’s rounded Mohr-Coulomb yield surface retains the form of

equation (2.37), but redefines K(θ) in the vicinity of the vertices at θ= 30˚.

The rounded yield surface uses the modified form of K(θ) whenever |θ|> θT ,

where θT is a specified transition angle. Away from the vertices, where |θ|≤ θT,

Sloan and Booker’s yield surface is identical to the Mohr-Coulomb yield surface

so that K(θ) is given by equation (2.38). The complete yield surface is thus defined

by equation (2.37) with

K(θ)=⎨⎧⎩

(A− B sin 3θ) |θ|> θT

(cos θ− 13sinφ sin θ) |θ|≤ θT

(2.39)

and

30Chapter 2

A= 13 cos θT3+ tan θT tan 3θT+ 13 sign(θ)(tan 3θT− 3 tan θT) sinφ (2.40)

B= 13 cos 3θTsign(θ) sin θT+ 13 sinφ cos θT (2.41)

sign(θ)= + 1 for θ≥ 0˚− 1 for θ< 0˚

The value of the transition angle lies within the range 0≤ θT≤ 30˚, with larger

values giving better fits to the Mohr-Coulomb cross-section in the octahedral

plane. In practice, θT should not be too near 30˚to avoid ill-conditioning of the

approximation, and a typical value is 25˚. Once the transition angle is specified,

the coefficients A and B may be computed efficiently by evaluating all of the

constant terms in equations (2.40) and (2.41) respectively. A π-plane plot of Sloan

and Booker’s rounded surface, for θT= 25˚ and φ= 30˚, is shown in Figure 2.3.

Under a state of triaxial compression, the rounded surface underestimates the

θ= ---25˚

θ= 30˚

θ= ---30˚

θT= 25˚φ= 30˚

θ= 25˚

θ

Mohr-Coulomb

2 σc

Figure 2.3 Rounding of Mohr-Coulomb yield surface in the π-plane.

rounded Mohr-Coulomb

31Chapter 2

Mohr-Coulomb value for σ∕c by approximately 4.4 per cent. For a state of triaxial

extension, this difference is reduced to roughly 1.1 per cent.

2.4.2 Rounded Hyperbolic Mohr-Coulomb Yield Function

A comprehensive discussion of various smooth approximations to the

Mohr-Coulomb criterion has been given by Zienkiewicz and Pande (1977). They

noted that a hyperbolic approximation, as shown in Figure 2.4, can be used to

remove the singularity at the tip of the surface and requires only one extra

parameter. A further advantage of this type of approximation is that it asymptotes

rapidly to the Mohr-Coulomb yield surface as the compressive hydrostatic stress

increases. Because the hyperbolic surface is an internal approximation, the

predicted soil strength is always less than the strength that would be found from

an equivalent Mohr-Coulomb model with the same cohesion and friction angle.

Removing the apex singularity greatly improves the computational stability of

elastoplastic finite element analyses which involve tensile hydrostatic stress states.

These stress states often arise in the analysis of soils with significant friction angles

and low cohesions, and may cause the tangent stiffness matrix to become

ill-conditioned. They may also result in stress integration schemes becoming

unstable.

σ

σm

a

b

d

Hyperbolic approximation

Mohr-Coulomb

Figure 2.4 Hyperbolic approximation to Mohr-Coulomb yield function.

32Chapter 2

The relationship between σ and σm , for a constant θ, defines a meridional section

of the yield surface. For the Mohr-Coulomb criterion, this relationship can be

represented as a straight line in (σm , σ ) space as shown in Figure 2.4. The point

where the line cuts the σm-axis corresponds to the tip of the hexagonal

Mohr-Coulomb pyramid, and it is here that the gradient of the yield surface is

undefined.

The equation of the straight line defining the Mohr-Coulomb yield function in the

meridional plane can be determined directly from equation (2.37) as

σ = 1K(θ)

c cosφ− σm sinφ

The slope of this line is − sinφ∕K(θ) and it intercepts the σm-axis at

σm= c cotφ. Following Zienkiewicz and Pande (1977), a close approximation to

the straight line which defines the Mohr-Coulomb yield surface can be obtained

using an asymptotic hyperbola. The general equation of such a hyperbola, in

(σm , σ ) space, is

(σm− d )2

a2− σ

2

b2= 1 (2.42)

where a, b and d are the distances defined in Figure 2.4. The upper asymptote

to this hyperbola has slope − b∕a and crosses the σm-axis at σm= d. Equating

the slope and intercept of the Mohr-Coulomb surface to the slope and intercept

of the hyperbolic asymptote yields the two relations

ba=

sinφK(θ)

, d= c cotφ

Substituting these expressions into equation (2.42) gives the yield surface

f = σm+ σ 2K2(θ)+ a2 sin2φ − c cosφ= 0 (2.43)

where the negative branch of the hyperbola has been chosen. This function can

be made to model the Mohr-Coulomb yield function as closely as desired by

33Chapter 2

adjusting the parameter a. Moreover, the Mohr-Coulomb yield function is

recovered if a is set to zero. Various meridional sections of the hyperbolic yield

surface are plotted in Figure 2.5. For a≤ 0.25 c cotφ, the hyperbolic surface

closely represents the Mohr-Coulomb surface. In practice, setting a= 0.05 c cotφ

has been found to give results which are almost identical to those from the

Mohr-Coulomb model.

If (2.39) is used to define K(θ), equation (2.43) defines a hyperbolic yield function

which is rounded in both the meridional plane and the octahedral plane. The

resulting yield surface is continuous and differentiable for all stress states, and the

Mohr-Coulomb yield surface can be modelled as closely as desired by adjusting

the two parameters a and θT . Indeed, the Mohr-Coulomb function can be

recovered by substituting a=0 and θT= 30˚. A comparison between π-plane

sections of the rounded hyperbolic surface and the Mohr-Coulomb surface is

illustrated in Figure 2.6. For meridional and octahedral rounding parameters of

0.0

0.5

1.0

1.5

2.0

---2 ---1 0 1 2

a= 0.5c cotφ

a= 0.25c cotφ

σc

a= c cotφ

Mohr-Coulomb θ= 0˚φ= 30˚

Figure 2.5 Hyperbolic approximations to Mohr-Coulomb meridional section.

σmc

34Chapter 2

θ= ---25˚

a= 0.25c cotφ

θ= 30˚

θ= ---30˚

θT= 25˚φ= 30˚

θ= 25˚

a= 0.5c cotφ

θ

a= 0 (Mohr-Coulomb)

2 σc

Figure 2.6 Rounded hyperbolic yield surface in the π-plane.

a= 0.05c cotφ and θT= 25˚, respectively, the σ∕c values predicted by the

rounded hyperbolic surface differ from those of the rounded Mohr-Coulomb

surface by a maximum of 0.13 per cent. As the compressive mean normal stress

increases, this difference is reduced even further by the asymptotic nature of the

hyperbolic surface.

2.5 YIELD SURFACE GRADIENTS

The gradients of the yield surface and plastic potential play an essential role in

elastoplastic finite element analysis. These quantities are used to calculate the

elastoplastic stress-strain matrices and, in explicit stress integration schemes, to

correct for drift from the yield surface. As the gradients are usually calculated

many times in a single analysis, they need to be evaluated efficiently. One

convenient method for computing the gradient a of an isotropic yield function uses

the form

a=∂f∂σ= C1

∂σm∂σ + C2

∂σ∂σ+ C3

∂J3∂σ (2.44)

35Chapter 2

where

C1=∂f∂σm

C2=∂f∂σ−

tan 3θσ∂f∂θ

C3=−3

2 cos 3θσ3∂f∂θ

⎪⎪⎪

⎪⎪⎪⎬

∂σm∂σ =

13⎪⎪⎪

⎪⎪⎪⎨

1

1

1

0

0

0

⎪⎪⎪

⎪⎪⎪⎬

, ∂σ∂σ=12σ⎪⎪⎪

⎪⎪⎪⎨

sxsysz

2τxy

2τyz

2τxz

⎪⎪⎪

⎪⎪⎪⎬

,∂J3∂σ =⎪⎪⎪

⎪⎪⎪⎨

sysz− τyz2

sxsz− τxz2

sxsy− τxy2

2τyzτxz− szτxy2τxzτxy− sxτyz2τxyτyz− syτxz

⎪⎪⎪

⎪⎪⎪⎬

+ σ2

3 ⎪⎪⎪

⎪⎪⎪⎨

1

1

1

0

0

0

⎪⎪⎪

⎪⎪⎪⎬

(2.45)

and σ= σx , σy , σz , τxy , τyz , τxz T is the vector of stress components. This

arrangement, proposed by Nayak and Zienkiewicz (1972b), permits the gradients

for different yield criteria to be computed simply by evaluating the appropriate

coefficients C1, C2, and C3 . All of the other derivatives are independent of f and

are therefore the same for all yield criteria.

2.5.1 Rounded Mohr-Coulomb Gradients

Away from the corners of the Mohr-Coulomb yield criterion, the constants C1, C2,

and C3 are found by differentiating (2.37) to give

36Chapter 2

Cmc1 = sinφ

Cmc2 = K− tan 3θdKdθ

Cmc3 =−3

2 cos 3θσ2dKdθ

⎪⎪

⎪⎪⎬

|θ|≤ θT (2.46)

where K= K(θ) is defined by (2.38) and

dKdθ=− sin θ− 1

3sinφ cos θ

At a corner of the Mohr-Coulomb yield surface, θ= 30˚ so that tan 3θ=∞

and cos 3θ= 0. This implies that equations (2.46) cannot be used for a stress state

in the vicinity of a corner. Instead, the first of equations (2.39) should be

substituted into equations (2.46) to give the rounded form

Cmc1 = sinφ

Cmc2 = A+ 2B sin 3θ

Cmc3 =3 3 B2σ2

⎪⎪

⎪⎪⎬

|θ|> θT (2.47)

These terms are not singular at θ= 30˚ and should be used whenever

|θ|> θT.

The above coefficients, defined by either (2.46) or (2.47) together with (2.44), are

singular at the tip of the Mohr-Coulomb pyramid where σ= 0. Although this

problem cannot be avoided for this type of yield surface, the loss of accuracy as

σ approaches zero can be minimised by grouping terms so that σ is always divided

into quantities of similar magnitude. To this end, it is best to compute

37Chapter 2

Cmc3 = Cmc3 σ

2

and then evaluate the last term in (2.44) using

C3∂J3∂σ = C

mc3

⎪⎪⎪⎪⎪⎪⎪⎪⎪

1σ2⎪⎪⎪

⎪⎪⎪⎨

sysz− τyz2

sxsz− τxz2

sxsy− τxy2

2τyzτxz− szτxy2τxzτxy− sxτyz2τxyτyz− syτxz

⎪⎪⎪

⎪⎪⎪⎬

+ 13⎪⎪⎪

⎪⎪⎪⎨

1

1

1

0

0

0

⎪⎪⎪

⎪⎪⎪⎬

⎪⎪⎪⎪⎪⎪⎪⎪⎪

This procedure can be used for all values of σ and θ.

2.5.2 Rounded Hyperbolic Mohr-Coulomb Gradients

The coefficients C1, C2, and C3 for the hyperbolic yield surface are calculated

by differentiating equation (2.43). These can be expressed very simply in terms

of the above Mohr-Coulomb coefficients according to

Ch1 = Cmc1

Ch2 = α Cmc2

Ch3 = α Cmc3

⎪⎬

(2.48)

where Cmc1 , Cmc2 , and C

mc3 are given by equations (2.46) or (2.47), depending on

the value of θ, and

α= σK

σ 2K2+ a2 sin2 φ (2.49)

The above coefficients, defined by (2.48) together with (2.44), is still singular at

the tip of the hyperbolic Mohr-Coulomb pyramid where σ= 0. To avoid a loss

of accuracy in computing the gradient for small values of σ, all divisions by σ

38Chapter 2

should, wherever possible, be eliminated. When division by σ is unavoidable, the

terms can be grouped so that σ divides a quantity which is much smaller than itself.

In this way, all terms that involve a division by σ will approach zero for very small

values of σ. Let

Ch2 = α Cmc2

Ch3 = α Cmc3 σ

2

where

α= ασ=K

σ 2K2+ a2 sin2φ

Neither Ch2 nor C

h3 require division by σ in order to be computed. The division

by σ can now be avoided in the computation of the second term in (2.44) by using

Ch2∂σ∂σ=

12 C

h2

⎪⎪⎪

⎪⎪⎪⎨

sxsysz

2τxy

2τyz

2τxz

⎪⎪⎪

⎪⎪⎪⎬

(2.50)

Note that this contribution will approach zero as σ approaches zero and will

therefore be negligible in the proximity of the tip. Similarly, the third term in

(2.44) may be evaluated using

C3∂J3∂σ = C

h3

⎪⎪⎪⎪⎪⎪⎪⎪⎪

1σ⎪⎪⎪

⎪⎪⎪⎨

sysz− τyz2

sxsz− τxz2

sxsy− τxy2

2τyzτxz− szτxy2τxzτxy− sxτyz2τxyτyz− syτxz

⎪⎪⎪

⎪⎪⎪⎬

+ σ3⎪⎪⎪

⎪⎪⎪⎨

1

1

1

0

0

0

⎪⎪⎪

⎪⎪⎪⎬

⎪⎪⎪⎪⎪⎪⎪⎪⎪

(2.51)

39Chapter 2

For small values of σ, both of the terms inside the outer brackets, and hence the

overall gradient contribution, will approach zero. Since the contributions of (2.50)

and (2.51) both approach zero as σ approaches zero, this implies that

a≈sinφ3 ⎪⎪⎪

⎪⎪⎪⎨

1

1

1

0

0

0

⎪⎪⎪

⎪⎪⎪⎬

(2.52)

in the vicinity of the tip. This is precisely the gradient direction which points along

the positive hydrostatic axis, and is correct on physical grounds. In practical

computations σ is rarely found to be exactly equal to zero, and therefore the

quantities defined by (2.50) and (2.51) are well defined. In the unlikely event that

σ is found to be exactly zero, it can be reset to a very small positive value and

equations (2.50) and (2.51) can again be used. Thus, for very small or zero values

of σ, the gradient computation automatically converges to the correct value given

by (2.52).

2.6 GRADIENT DERIVATIVES

Many implicit stress integration methods, such as the backward Euler return

algorithm discussed by Crisfield (1991), require the derivatives of the gradient

vector with respect to the stresses. Since the implicit integration schemes to be

discussed later in this Thesis use these quantities, expressions for the gradient

derivatives of the rounded hyperbolic surface are now derived. For the sake of

simplicity, a two-dimensional stress vector is assumed with σ= σx , σy , σz , τxy T.

In general, the exact evaluation of the gradient derivatives involves a significant

amount of tedious algebra, especially for yield functions which are pressure

dependent, and this is perhaps one of the reasons why they have not been used

widely in geotechnical engineering codes.

40Chapter 2

Differentiating equation (2.44) gives

∂a∂σ=

∂C2∂σ∂σ∂σ+ C2

∂2σ∂σ +

∂C3∂σ∂J3∂σ + C3

∂2J3∂σ (2.53)

where ∂σ∂σ and∂J3∂σ are defined by (2.45) and

symmetric

∂2σ∂σ2= 1σ

⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪

13−sx sx4σ 2

− 16−sx sy4σ 2

− 16−sx sz4σ 2

−τxy sx2σ 2

13−sy sy4σ 2

− 16−sy sz4σ 2

−τxy sy2σ 2

13−sz sz2σ 2

−τxy sz2σ 2

1−τxy τxy

σ 2

⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪

symmetric∂2J3∂σ =

13⎪⎪⎪

sx− sy− sz2sz2sy2τxy

sy− sx− sz2sx2τxy

sz− sx− sy− 4τxy − 6sz

⎪⎪⎪

To complete the formation of the gradient derivatives, the derivatives of C1 and

C2 with respect to the stresses need to be evaluated for each type of yield function.

For the rounded Mohr-Coulomb criterion these derivatives are

∂Cmc2∂σ =

∂θ∂σ ∂K∂θ −∂2K∂θ2 tan 3θ− 3 ∂K∂θ sec2 3θ

∂Cmc3∂σ =

3

2σ 2 cos 3θ2σ∂K∂θ∂σ∂σ−

∂θ∂σ ∂2K∂θ2 + 3 ∂K∂θ tan 3θ⎪⎪

⎪⎪⎬

|θ|≤ θT (2.54)

41Chapter 2

∂Cmc2∂σ =−

3 3 Bσ 3∂J3∂σ − 3 J3σ ∂σ∂σ

∂Cmc3∂σ =−

3 3 Bσ 3

∂σ∂σ

⎪⎪

⎪⎪⎬

|θ|> θT (2.55)

where

∂θ∂σ=

− 3

2σ 3 cos 3θ∂J3∂σ − 3J3σ ∂σ∂σ (2.56)

and K= K(θ) is defined by (2.39) with

d2Kdθ2=− cos θ+ 1

3sinφ sin θ |θ|≤ θT (2.57)

Similarly, for the hyperbolic yield surface

∂Ch2∂σ = α

∂Cmc2∂σ + C

mc2∂α∂σ (2.58)

∂Ch3∂σ = α

∂Cmc3∂σ + C

mc3∂α∂σ (2.59)

where

∂α∂σ =

1− α2

σ 2K2+ a2 sin2 θ ∂σ∂σK+ σ ∂K∂θ ∂θ∂σ (2.60)

Thus the gradient derivatives for a hyperbolic surface with a rounded octahedral

cross-section is obtained from equations (2.53)---(2.60). In evaluating these terms,

it is again advisable to group the terms so that division by σ is avoided wherever

possible.

2.7 NUMERICAL IMPLEMENTATION

Two FORTRAN 77 subroutines, presented in Appendices 2A and 2B, illustrate how

the hyperbolic rounded surface may be implemented efficiently in a finite element

42Chapter 2

code. The subroutines, YIELD and GRAD, return the value of the yield function

and the gradient vector, respectively, for a specified stress state. They are

applicable to two-dimensional plasticity, with either associated or nonassociated

flow, and assume that the stress vector is σ ={σx , σy , τxy , σz}T . For the case of

a nonassociated flow rule, the gradients are found by assuming that the plastic

potential is of the same form as the yield function, with the only difference being

that the dilatancy angle replaces the friction angle. As well as incorporating the

rounded hyperbolic surface, the subroutines also model the usual Mohr-Coulomb

and Tresca yield criteria (both of which are rounded in the octahedral plane).

Since this code is executed a large number of times during the course of a typical

finite element computation, considerable attention has been paid to implementing

the models with a minimum amount of arithmetic.

43Chapter 2

44 Chapter 2

APPENDICES

45Chapter 2

2A SUBROUTINE “YIELD”SUBROUTINE YIELD(YLD,SIGXX,SIGYY,SIGXY,SIGZZ,MPROP,NMP,IOW)

******************************************************************************** PURPOSE:* --------* This subroutine returns the value of the yield function at the given* stress state for plane strain and axisymmetric plasticity. Smooth* approximations to the Mohr-Coulomb and Tresca criteria are used.** INPUT:* ------* YLD - Undefined* SIGXX - XX-component of normal stress* SIGYY - YY-component of normal stress* SIGZZ - ZZ-component of normal stress* SIGXY - XY-component of shear stress* MPROP - Vector of dimension (NMP)* - Contains material parameters* - MPROP(10) = (a*SIN(friction angle))**2* a = hyperbolic rounding parameter* - MPROP(11) = SIN(friction angle)* - MPROP(12) = COS(friction angle)* - MPROP(17) = Cohesion* - MPROP(20) = Value defining type of yield function* 0 = Elastic Material* 1 = Mohr-Coulomb rounded in octahedral plane* 2 = Hyperbolic Mohr-Coulomb rounded in octahedral* plane* 3 = Tresca rounded in octahedral plane* 4 = von-Mises* 5 = True Mohr-Coulomb* NMP - Parameter specifying number of material parameters* IOW - Unit number of output file** OUTPUT:* -------* YLD - Value of yield function* SIGXX - Unchanged* SIGYY - Unchanged* SIGZZ - Unchanged* SIGXY - Unchanged* MPROP - Unchanged* NMP - Unchanged* IOW - Unchanged** SUBROUTINES CALLED: NONE* -------------------** PROGRAMMER: Andrew Abbo* -----------** LAST MODIFIED: Oct 1995 Andrew Abbo* --------------******************************************************************************

INTEGER YFTYPE,NMP,IOW*

DOUBLE PRECISION STA,CTA,A,B,K,SGN,YLDDOUBLE PRECISION SIGXX,SIGYY,SIGZZ,SIGXYDOUBLE PRECISION DSIGX,DSIGY,DSIGZDOUBLE PRECISION CPHI,SPHI,COH,ASPHI2DOUBLE PRECISION THETA,J2,J3,S3TA,SIGM,SBAR

*DOUBLE PRECISION MPROP(NMP)

** Set constants*

INTEGER ELAST,RMC,HRMC,TR,VM,MCPARAMETER( ELAST=0,RMC=1,HRMC=2,TR=3,VM=4,MC=5 )

*DOUBLE PRECISION C00001,C004P5,C000P5,C00000,C00002PARAMETER( C000P5 = 0.5D0 , C004P5 = 4.5D0 )

46 Chapter 2

PARAMETER( C00001 = 1.0D0 , C00000 = 0.0D0 ,C00002=2.0D0 )*

DOUBLE PRECISION C000R3,CP3333,C00IR3PARAMETER( C000R3 = 1.732050807568877D0 )PARAMETER( CP3333 = 0.333333333333333D0 )PARAMETER( C00IR3 = 0.577350269189626D0 )

** Constants for rounded K function*

DOUBLE PRECISION A1,A2,B1,B2,ATTRAN** Rounding constants for theta > 25 degrees*

PARAMETER( A1 = 1.432052062044227D0 , A2 = 0.406941858374615D0 )PARAMETER( B1 = 0.544290524902313D0 , B2 = 0.673903324498392D0 )PARAMETER( ATTRAN = 0.436332312998582D0 )

** Calculate value of invariants*

SIGM = CP3333*(SIGXX+SIGYY+SIGZZ)DSIGX = SIGXX-SIGMDSIGY = SIGYY-SIGMDSIGZ = SIGZZ-SIGMJ2 = C000P5*(DSIGX*DSIGX+DSIGY*DSIGY+DSIGZ*DSIGZ) + SIGXY*SIGXYJ3 = DSIGZ*(DSIGX*DSIGY-SIGXY*SIGXY)SBAR = SQRT(J2)

*IF (J2.GT.C00000) THEN

** Calculate third stress invariant*

S3TA = -C004P5*J3/(C000R3*SBAR*J2)IF (S3TA.LT.-C00001) THENS3TA = -C00001

ELSEIF (S3TA.GT.C00001) THENS3TA = C00001

ENDIFTHETA = CP3333*ASIN(S3TA)

*ELSE

** Special case of zero deviatoric stress*

S3TA = C00000THETA = C00000

*ENDIF

** Extract form of yield function from MPROP vector*

YFTYPE = INT(MPROP(20))*

IF (YFTYPE.EQ.ELAST) THEN** Elastic analysis, set yield function negative*

YLD = -C00001*

ELSEIF ((YFTYPE.EQ.RMC).OR.(YFTYPE.EQ.HRMC)) THEN*-------------------------------------------------------------------------* Rounded or Rounded Hyperbolic Mohr-Coulomb yield function*

COH = MPROP(17)SPHI = MPROP(11)CPHI = MPROP(12)ASPHI2 = MPROP(10)

** Calculate K function*

IF (ABS(THETA).LT.ATTRAN) THEN** Calculate K function for unrounded region of octahedral plane*

STA = SIN(THETA)

47Chapter 2

CTA = COS(THETA)K = CTA-STA*SPHI*C00IR3

*ELSE

** Calculate K function for rounded region of octahedral plane*

SGN = SIGN(C00001,THETA)A = A1 + A2*SGN*SPHIB = B1*SGN + B2*SPHIK = A-B*S3TA

*ENDIF

** Calculate value of yield function*

IF (YFTYPE.EQ.HRMC) THEN** Hyperbolic Mohr-Coulomb surface*

YLD = SIGM*SPHI+SQRT((SBAR*SBAR*K*K)+ASPHI2)-COH*CPHI*

ELSE** Mohr-Coulomb surface*

YLD = SIGM*SPHI+SBAR*K-COH*CPHI*

ENDIF*

ELSEIF (YFTYPE.EQ.TR) THEN*------------------------------------------------------------------------* Tresca yield function*

COH = MPROP(17)*

IF (ABS(THETA).LT.ATTRAN) THEN** Calculate K function for unrounded region of octahedral plane*

K = COS(THETA)*

ELSE** Calculate K function for rounded region of octahedral plane*

SGN = SIGN(C00001,THETA)A = A1B = B1*SGNK = A-B*S3TA

*ENDIF

** Calculate value of yield function*

YLD = SBAR*K-COH*

ELSEIF (YFTYPE.EQ.VM) THEN*------------------------------------------------------------------------* Von Mises yield function*

COH = MPROP(17)*

YLD = C000R3*SBAR-C00002*COH*

ELSEIF (YFTYPE.EQ.MC) THEN*------------------------------------------------------------------------* Mohr-Coulomb yield function*

COH = MPROP(17)SPHI = MPROP(11)CPHI = MPROP(12)

** Calculate K function

48 Chapter 2

*STA = SIN(THETA)CTA = COS(THETA)K = CTA-STA*SPHI*C00IR3

** Calculate value of yield function*

YLD = SIGM*SPHI+SBAR*K-COH*CPHI*

ELSE** Invalid Yield Function type*

WRITE(IOW,’(’’ *** ERROR IN SUBROUTINE YIELD ***’’)’)WRITE(IOW,’(’’ INVALID YIELD FUNCTION - YFTYPE = ’’,I4)’)YFTYPESTOP

*ENDIF

*END

49Chapter 2

2B SUBROUTINE “GRAD”SUBROUTINE GRAD(GY1,GY2,GY3,GY4,GP1,GP2,GP3,GP4,SIGXX,SIGYY,SIGXY,+ SIGZZ,MPROP,NMP,IOW)

************************************************************************* PURPOSE:* --------** This subroutine returns the value of the gradient to the yield* surface and plastic potential at a given stress state for plane strain* and axisymmetric plasticity. Smooth approximations to the Mohr-Coulomb* and Tresca criteria are used. The routine is designed for both* associated and non-associated flow rules.** INPUT:* ------* GY1..GY4 - Undefined on entry* GP1..GP4 - Undefined on entry* SIGXX - XX-component of normal stress* SIGYY - YY-component of normal stress* SIGZZ - ZZ-component of normal stress* SIGXY - XY-component of shear stress* MPROP - Vector of dimension (NMP)* - Contains material parameters* - MPROP(8) = Value specifying type of flow rule* 0 = associated flow* 1 = non-associated flow* - MPROP(9) = (a*SIN(dilation angle))**2* a = hyperbolic rounding parameter* - MPROP(10) = (a*SIN(friction angle))**2* a = hyperbolic rounding parameter* - MPROP(11) = SIN(friction angle)* - MPROP(13) = SIN(dilation angle)* - MPROP(20) = Value defining type of yield function* 1 = Mohr-Coulomb rounded in octahedral plane* 2 = Hyperbolic Mohr-Coulomb rounded in octahedral* plane* 3 = Tresca rounded in octahedral plane* 3 = Von Mises* 3 = True Mohr-Coulomb* NMP - Parameter specifying number of material parameters* IOW - Unit number of output file** OUTPUT:* -------* GY1 - Component of gradient vector to yield function wrt SIGXX* GY2 - Comp. of gradient vector to yield function wrt SIGYY* GY3 - Comp. of gradient vector to yield function wrt SIGZZ* GY4 - Comp. of gradient vector to yield function wrt SIGXY* GP1 - Comp. of gradient vector to plastic potential wrt SIGXX* GP2 - Comp. of gradient vector to plastic potential wrt SIGYY* GP3 - Comp. of gradient vector to plastic potential wrt SIGZZ* GP4 - Comp. of gradient vector to plastic potential wrt SIGXY* SIGXX - Unchanged* SIGYY - Unchanged* SIGZZ - Unchanged* SIGXY - Unchanged* MPROP - Unchanged* NMP - Unchanged

50 Chapter 2

* FLAG - Unchanged* IOW - Unchanged** SUBROUTINES CALLED: NONE* -------------------** PROGRAMMER: Andrew Abbo* -----------** LAST MODIFIED: May 1993 Andrew Abbo* --------------*************************************************************************

INTEGER YFTYPE,NMP,IOWINTEGER FLOW

*DOUBLE PRECISION SIGXX,SIGYY,SIGZZ,SIGXYDOUBLE PRECISION SPHI,SPSI,ASPHI2,ASPSI2DOUBLE PRECISION DSIGX,DSIGY,DSIGZDOUBLE PRECISION THETA,SIGM,J2,J3,SBAR,ALPHADOUBLE PRECISION STA,CTA,C3TA,S3TA,T3TADOUBLE PRECISION A,B,K,DKDOUBLE PRECISION C1,C2,C3DOUBLE PRECISION GY1,GY2,GY3,GY4DOUBLE PRECISION GP1,GP2,GP3,GP4

*DOUBLE PRECISION MPROP(NMP)

** Set constants*

INTEGER RMC,HRMC,TR,ASSOC,VM,MCPARAMETER( RMC=1,HRMC=2,TR=3,VM=4,MC=5 )PARAMETER( ASSOC=0 )

*DOUBLE PRECISION TINYPARAMETER( TINY = 1.0D-10 )

*DOUBLE PRECISION J2TOLPARAMETER( J2TOL = 1.0D-20 )

*DOUBLE PRECISION C004P5,C000P5,CP3333PARAMETER( C004P5 = 4.5D0 )PARAMETER( C000P5 = 0.5D0 )PARAMETER( CP3333 = 0.333333333333333D0 )

*DOUBLE PRECISION C00000,C00001,C00002,C00003,C00004PARAMETER( C00000 = 0.0D0 )PARAMETER( C00001 = 1.0D0 )PARAMETER( C00002 = 2.0D0 )PARAMETER( C00003 = 3.0D0 )PARAMETER( C00004 = 4.0D0 )

*DOUBLE PRECISION C000R3,C00IR3,CP8660PARAMETER( C000R3 = 1.732050807568877D0 )PARAMETER( C00IR3 = 0.5773502691896258D0 )PARAMETER( CP8660 = 0.866025403784439D0 )

*DOUBLE PRECISION CRADPARAMETER (CRAD =0.017453292519943D0)

*

51Chapter 2

* Constants for rounded K function*

DOUBLE PRECISION A1,A2,B1,B2,ATTRAN** Rounding constants for theta > 25 degrees*

PARAMETER( A1 = 1.432052062044227D0 )PARAMETER( A2 = 0.406941858374615D0 )PARAMETER( B1 = 0.544290524902313D0 )PARAMETER( B2 = 0.673903324498392D0 )PARAMETER( ATTRAN=0.436332312998582D0 )

** Rounding constants for theta > 29.5 degrees** PARAMETER( A1 = 7.138654723242523D0 , A2 = 6.112267270920722D0 )* PARAMETER( B1 = 6.270447753139696D0 , B2 = 6.398760841429511D0 )* PARAMETER( ATTRAN = 0.514872129338327D0 )** Calculate value of invariants for the current stress state.*

SIGM = CP3333*(SIGXX+SIGYY+SIGZZ)*

DSIGX = SIGXX-SIGMDSIGY = SIGYY-SIGMDSIGZ = SIGZZ-SIGMJ2 = C000P5*(DSIGX*DSIGX+DSIGY*DSIGY+DSIGZ*DSIGZ) + SIGXY*SIGXYJ3= DSIGZ*(DSIGX*DSIGY-SIGXY*SIGXY)

*SBAR=SQRT(J2)

** Store type of flow rule* If MPROP(8)=1 then have non-associated flow rule* If have associated flow rule, then the gradients to the plastic* potential and the yield function will be set equal*

FLOW=INT(MPROP(8))** Extract form of yield function from MPROP vector*

YFTYPE=INT(MPROP(20))*

IF (YFTYPE.EQ.RMC) THEN*-----------------------------------------------------------------------

* Rounded Mohr-Coulomb yield function*

IF (J2.GT.J2TOL) THEN** Calculate third stress invariant*

S3TA = -C004P5*J3/(C000R3*SBAR*J2)IF (S3TA.LT.-C00001) THENWRITE(IOW,’(’’ *** S3TA set equal -1 ***’’,F12.5)’)S3TAWRITE(IOW,’(’’ J2,J3,SIGM ’’,4E14.5)’) J2,J3,SIGMWRITE(IOW,’(’’ J2,J3 ’’,4E14.5)’) DSIGX,DSIGY,DSIGZ,SIGXYS3TA = -C00001

ELSEIF (S3TA.GT.C00001) THENWRITE(IOW,’(’’ *** S3TA set equal +1 ***’’,F12.5)’)S3TAWRITE(IOW,’(’’ J2,J3,SIGM ’’,4E14.5)’) J2,J3,SIGMWRITE(IOW,’(’’ J2,J3 ’’,4E14.5)’) DSIGX,DSIGY,DSIGZ,SIGXY

52 Chapter 2

S3TA = C00001ENDIFTHETA = CP3333*ASIN(S3TA)

*ELSE

** Special case of zero deviatoric stress*

WRITE(IOW,’(’’ *** WARNING IN SUBROUTINE GRAD ***’’)’)WRITE(IOW,’(’’ ZERO DEVIATORIC STRESS STATE ’’)’)

*J2 = TINY*TINYSBAR = TINYTHETA = C00000S3TA = C00000

*ENDIF

** Set value of material parameters used in gradient calculations*

SPHI = MPROP(11)** Calculate gradient constants*

IF (ABS(THETA).LT.ATTRAN) THEN** Unrounded surface*

CTA = COS(THETA)C3TA = CTA*(C00004*CTA*CTA-C00003)T3TA = S3TA/C3TASTA = S3TA/(C00004*CTA*CTA-C00001)K = CTA-STA*SPHI*C00IR3DK = STA+CTA*SPHI*C00IR3

*C1 = SPHIC2 = K+T3TA*DKC3 = CP8660*DK/C3TA

*ELSE

** Rounded surface*

IF (THETA.GT.C00000) THENA = A1 + A2*SPHIB = B1 + B2*SPHI

ELSEA = A1 - A2*SPHIB = - B1 + B2*SPHI

ENDIF*

C1 = SPHIC2 = A + C00002*B*S3TAC3 = CP8660*C00003*B

*ENDIF

*ELSEIF (YFTYPE.EQ.HRMC) THEN

*-----------------------------------------------------------------------

53Chapter 2

* Hyperbolic Rounded Mohr-Coulomb yield function*

IF (J2.GT.J2TOL) THEN** Calculate third stress invariant*

S3TA = -C004P5*J3/(C000R3*SBAR*J2)IF (S3TA.LT.-C00001) THENWRITE(IOW,’(’’ *** S3TA set equal -1 ***’’,F12.5)’)S3TAWRITE(IOW,’(’’ J2,J3,SIGM ’’,4E14.5)’) J2,J3,SIGMWRITE(IOW,’(’’ J2,J3 ’’,4E14.5)’) DSIGX,DSIGY,DSIGZ,SIGXYS3TA = -C00001

ELSEIF (S3TA.GT.C00001) THENWRITE(IOW,’(’’ *** S3TA set equal +1 ***’’,F12.5)’)S3TAWRITE(IOW,’(’’ J2,J3,SIGM ’’,4E14.5)’) J2,J3,SIGMWRITE(IOW,’(’’ J2,J3 ’’,4E14.5)’) DSIGX,DSIGY,DSIGZ,SIGXYS3TA = C00001

ENDIFTHETA = CP3333*ASIN(S3TA)

*ELSE

** Special case of zero deviatoric stress*

WRITE(IOW,’(’’ *** WARNING IN SUBROUTINE GRAD ***’’)’)WRITE(IOW,’(’’ ZERO DEVIATORIC STRESS STATE ’’)’)

*J2 = TINY*TINYSBAR = TINYTHETA = C00000S3TA = C00000

*ENDIF

** Set value of material parameters used in gradient calculations*

SPHI = MPROP(11)ASPHI2 = MPROP(10)

** Calculate gradient constants*

IF (ABS(THETA).LT.ATTRAN) THEN** Unrounded surface*

CTA = COS(THETA)C3TA = CTA*(C00004*CTA*CTA-C00003)T3TA = S3TA/C3TASTA = S3TA/(C00004*CTA*CTA-C00001)K = CTA-STA*SPHI*C00IR3DK = STA+CTA*SPHI*C00IR3

*C1 = SPHIC2 = K+T3TA*DKC3 = CP8660*DK/C3TA

*ELSE

** Rounded surface*

54 Chapter 2

IF (THETA.GT.C00000) THENA = A1 + A2*SPHIB = B1 + B2*SPHI

ELSEA = A1 - A2*SPHIB = - B1 + B2*SPHI

ENDIF*

K = A-B*S3TA*

C1 = SPHIC2 = A + C00002*B*S3TAC3 = CP8660*C00003*B

*ENDIF

** Adjust coefficients for hyperbolic Mohr-Coulomb surface*

IF (ASPHI2.GT.C00000) THENALPHA = SBAR*KALPHA = K/SQRT(ALPHA*ALPHA + ASPHI2)C2 = C2*ALPHAC3 = C3*ALPHA

ENDIF*

ELSEIF ( YFTYPE.EQ.TR) THEN*-----------------------------------------------------------------------

* Tresca Yield Function*

IF (J2.GT.C00000) THEN** Calculate third stress invariant*

S3TA = -C004P5*J3/(C000R3*SBAR*J2)IF (S3TA.LT.-C00001) THENS3TA = -C00001

ELSEIF (S3TA.GT.C00001) THENS3TA = C00001

ENDIFTHETA = CP3333*ASIN(S3TA)

*ELSE

** Cannot have yielding at zero deviatoric stress for Tresca*

WRITE(IOW,’(’’ *** ERROR IN SUBROUTINE GRAD ***’’)’)WRITE(IOW,’(’’ ZERO J2 INVAR. FOR TRESCA YIELD FUNCTION’’)’)STOP

*ENDIF

** Calculate gradient constants*

IF (ABS(THETA).LT.ATTRAN) THEN** Unrounded surface*

CTA = COS(THETA)C3TA = CTA*(C00004*CTA*CTA-C00003)

55Chapter 2

T3TA = S3TA/C3TASTA = S3TA/(C00004*CTA*CTA-C00001)

*K = CTADK = STA

*C1 = C00000C2 = CTA+T3TA*STAC3 = CP8660*STA/C3TA

*ELSE

** Rounded surface*

A = A1IF (THETA.GT.C00000) THENB = B1

ELSEB = - B1

ENDIF*

C1 = C00000C2 = A + C00002*B*S3TAC3 = CP8660*C00003*B

*ENDIFFLOW=ASSOC

*ELSEIF ( YFTYPE.EQ.VM) THEN

*-----------------------------------------------------------------------* Von Mises Yield Function** Calculate gradient coefficients*

C1 = C00000C2 = C000R3C3 = C00000FLOW=ASSOC

*ELSEIF ( YFTYPE.EQ.MC) THEN

*-----------------------------------------------------------------------* Mohr-Coulomb yield function*

IF (J2.GT.C00000) THEN** Calculate third stress invariant*

S3TA = -C004P5*J3/(C000R3*SBAR*J2)IF (S3TA.LT.-C00001) THENS3TA = -C00001

ELSEIF (S3TA.GT.C00001) THENS3TA = C00001

ENDIFTHETA = CP3333*ASIN(S3TA)

*ELSE

** Special case of zero deviatoric stress*

WRITE(IOW,’(’’ *** WARNING IN SUBROUTINE GRAD ***’’)’)

56 Chapter 2

WRITE(IOW,’(’’ ZERO DEVIATORIC STRESS STATE ’’)’)J2 = TINYSBAR = TINY*TINYTHETA = C00000S3TA = C00000

*ENDIF

** Set value of material parameters used in gradient calculations*

SPHI = MPROP(11)*

CTA = COS(THETA)C3TA = CTA*(C00004*CTA*CTA-C00003)T3TA = S3TA/C3TA

** Calculate K function and its derivative wrt theta DK*

STA = S3TA/(C00004*CTA*CTA-C00001)K = CTA-STA*SPHI*C00IR3DK = STA+CTA*SPHI*C00IR3

** Calculate gradient coefficients for Mohr-Coulomb surface*

C1 = SPHIC2 = K+T3TA*DKC3 = (CP8660*DK)/C3TA

*ELSE

** Invalid yield function type*

WRITE(IOW,’(’’ *** ERROR IN SUBROUTINE GRAD ***’’)’)WRITE(IOW,’(’’ INVALID YIELD FUNCTION - YFTYPE = ’’,I4)’)YFTYPESTOP

*ENDIF

** Compose gradient to yield function*

C2 = C2*C000P5C1 = C1*CP3333IF ((YFTYPE.EQ.HRMC).AND.(ASPHI2.GT.C00000)) THENGY1=C1+C2*(DSIGX)+C3*((DSIGY*DSIGZ)/SBAR + CP3333*SBAR)GY2=C1+C2*(DSIGY)+C3*((DSIGX*DSIGZ)/SBAR + CP3333*SBAR)GY3=C00002*(C2*(SIGXY) - C3*(SIGXY*DSIGZ/SBAR))GY4=C1+C2*(DSIGZ)+C3*((DSIGX*DSIGY-SIGXY*SIGXY)/SBAR+CP3333*SBAR)

ELSEGY1=C1+C2*(DSIGX/SBAR)+C3*((DSIGY*DSIGZ)/J2 + CP3333)GY2=C1+C2*(DSIGY/SBAR)+C3*((DSIGX*DSIGZ)/J2 + CP3333)GY3=C00002*(C2*(SIGXY/SBAR) - C3*(SIGXY*DSIGZ/J2))GY4=C1+C2*(DSIGZ/SBAR)+C3*((DSIGX*DSIGY-SIGXY*SIGXY)/J2+CP3333)

ENDIF** Calculate gradient to potential for associated case*

IF (FLOW.EQ.ASSOC) THENGP1 = GY1GP2 = GY2GP3 = GY3

57Chapter 2

GP4 = GY4RETURN

ENDIF** If non-associated flow calculate gradient to plastic potential* Assume that the plastic potential has the same form as the yield* function except that the dilation angle is substituted for the* friction angle*

IF (YFTYPE.EQ.RMC) THEN*-----------------------------------------------------------------------

* Rounded Mohr-Coulomb plastic potential* Extract material parameters*

SPSI = MPROP(13)** Calculate K function and its derivative wrt theta*

IF (ABS(THETA).LT.ATTRAN) THEN** Unrounded surface*

K = CTA-STA*SPSI*C00IR3DK = STA+CTA*SPSI*C00IR3

*C1 = SPSIC2 = K+T3TA*DKC3 = CP8660*DK/C3TA

*ELSE

** Rounded surface*

IF (THETA.GT.C00000) THENA = A1 + A2*SPSIB = B1 + B2*SPSI

ELSEA = A1 - A2*SPSIB = - B1 + B2*SPSI

ENDIF*

C1 = SPSIC2 = A + C00002*B*S3TAC3 = CP8660*C00003*B

*ENDIF

*ELSEIF (YFTYPE.EQ.HRMC) THEN

*-----------------------------------------------------------------------

* Hyperbolic Rounded Mohr-Coulomb plastic potential* Extract material parameters*

SPSI = MPROP(13)ASPSI2 = MPROP(9)

** Calculate K function and its derivative wrt theta*

IF (ABS(THETA).LT.ATTRAN) THEN

58 Chapter 2

** Unrounded surface*

K = CTA-STA*SPSI*C00IR3DK = STA+CTA*SPSI*C00IR3

*C1 = SPSIC2 = K+T3TA*DKC3 = CP8660*DK/C3TA

*ELSE

** Rounded surface*

IF (THETA.GT.C00000) THENA = A1 + A2*SPSIB = B1 + B2*SPSI

ELSEA = A1 - A2*SPSIB = - B1 + B2*SPSI

ENDIFK = A-B*S3TA

*C1 = SPSIC2 = A + C00002*B*S3TAC3 = CP8660*C00003*B

*ENDIF

** Adjust coefficients for hyperbolic Mohr-Coulomb surface*

IF (ASPSI2.GT.C00000) THENALPHA = SBAR*KALPHA = K/SQRT(ALPHA*ALPHA + ASPSI2)C2 = C2*ALPHAC3 = C3*ALPHA

ENDIF*

ELSEIF ( YFTYPE.EQ.VM) THEN*-----------------------------------------------------------------------* Von Mises Plastic Potential** Calculate gradient coefficients*

C1 = C00000C2 = C000R3C3 = C00000

*ELSEIF (YFTYPE.EQ.MC) THEN

** Mohr-Coulomb plastic potential** Set value of material parameters used in gradient calculations*

SPSI = MPROP(13)*

CTA = COS(THETA)C3TA = CTA*(C00004*CTA*CTA-C00003)T3TA = S3TA/C3TA

*

59Chapter 2

* Calculate K function and its derivative wrt theta DK*

STA = S3TA/(C00004*CTA*CTA-C00001)K = CTA-STA*SPSI*C00IR3DK = STA+CTA*SPSI*C00IR3

** Calculate gradient coefficients for Mohr-Coulomb surface*

C1 = SPSIC2 = K+T3TA*DKC3 = (CP8660*DK)/C3TA

*ENDIF

** Compose gradient to plastic potential*

C2 = C2*C000P5C1 = C1*CP3333IF ((YFTYPE.EQ.HRMC).AND.(ASPSI2.GT.C00000)) THENGP1=C1+C2*(DSIGX)+C3*((DSIGY*DSIGZ)/SBAR + CP3333*SBAR)GP2=C1+C2*(DSIGY)+C3*((DSIGX*DSIGZ)/SBAR + CP3333*SBAR)GP3=C00002*(C2*(SIGXY) - C3*(SIGXY*DSIGZ/SBAR))GP4=C1+C2*(DSIGZ)+C3*((DSIGX*DSIGY-SIGXY*SIGXY)/SBAR+CP3333*SBAR)

ELSEGP1=C1+C2*(DSIGX/SBAR)+C3*((DSIGY*DSIGZ)/J2 + CP3333)GP2=C1+C2*(DSIGY/SBAR)+C3*((DSIGX*DSIGZ)/J2 + CP3333)GP3=C00002*(C2*(SIGXY/SBAR) - C3*(SIGXY*DSIGZ/J2))GP4=C1+C2*(DSIGZ/SBAR)+C3*((DSIGX*DSIGY-SIGXY*SIGXY)/J2+CP3333)

ENDIF*

END

61Chapter 3

CHAPTER 3

INTEGRATION OF STRESS-STRAIN

RELATIONS

62Chapter 3

3.1 INTRODUCTION

The first part of this Chapter discusses the practical implementation of an explicit

modified Euler scheme for integrating elastoplastic constitutive laws in finite

element analysis. This algorithm is based on the method of Sloan (1987) and

controls the error in the computed stresses by using a local error measure to

automatically subincrement and integrate the applied strain increment. It

incorporates a number of important new refinements which enhance the efficiency

and robustness of explicit subincrementation techniques. The next part of the

Chapter describes two implicit backward Euler schemes as described by Crisfield

(1991). These methods have been used widely in metal plasticity studies but have

had limited application to geotechnical constitutive models.

The last part of the Chapter compares the performance of the explicit modified

Euler scheme against the performance of the implicit backward Euler schemes.

To test the efficiency and robustness of these algorithms, the non-trivial boundary

value problem of a rigid strip footing resting on a Tresca or Mohr-Coulomb layer

is used. For the Mohr-Coulomb soil, results are presented for both associated and

nonassociated flow rules. The explicit modified Euler scheme with substepping

is shown to be very competitive with an implicit backward Euler return scheme,

and has the added advantage that the error in the computed stresses (for a given

mesh and load path) may be controlled to a desired level. A further attraction

of the explicit method is that it requires only first derivatives (with respect to the

stresses) of the yield surface and plastic potential. The second derivatives needed

for the implicit methods are both difficult and expensive to compute for many

geotechnical models. The results for the footing problems suggest that the implicit

schemes do not perform well in the vicinity of the corners of the Tresca and

Mohr-Coulomb yield criteria, even when they are rounded, and special strategies

may be required.

63Chapter 3

3.2 STRESS-STRAIN INTEGRATION

During a typical step or iteration of an elastoplastic finite element analysis, the

forces are applied in increments and the corresponding displacement increments

are found from the global stiffness equations. Once the nodal displacement

increments are known, the strain increments at a discrete number of integration

points within each element are determined using the strain-displacement relations

(2.20). If the stresses associated with an imposed strain increment cause plastic

yielding, it is necessary to solve the small system of ordinary differential equations

defined by (2.8) and (2.11) or (2.8) and (2.12). For either of these hardening

models, the governing relations may be written in the form

σ. = Depε

. (3.1)

À. = λ

.B (3.2)

where

Dep= De−Deb aTDeA + aTDeb

λ.=

aTDeε.

A+ aTDeb

and

B=− A∂f∕∂À

=⎪⎪⎨⎧

23bTMb

σTb

strain hardening

work hardening

To integrate these equations numerically, it is convenient to introduce a pseudo

time, T, defined by

T=t− t0Δt

64Chapter 3

where t0 is the time at the start of the load increment, t0+ Δt is the time at the

end of the load increment, and 0≤ T≤ 1. Since dT∕dt= 1∕Δt, application of

the chain rule to σ. and À. in (3.1) and (3.2) gives

dσdT= DepΔε= De− Deb aTDe

A + aTDebΔε= Δσe− ΔλDe b (3.3)

dÀdT= λ

.Δt B= ΔλB (3.4)

where

Δλ=aTDeΔεA+ aTDeb

=aTΔσe

A+ aTDeb(3.5)

Note that, in keeping with the philosophy of the static displacement finite element

procedure, the strain rate is assumed to be constant and equal to Δε∕Δt.

Equations (3.3) and (3.4) define a classical initial value problem which needs to

be integrated over the pseudo time interval from T= 0 to T= 1. The known

values in these relations are the imposed strain increments, Δε, together with the

stresses and hardening parameter at the start of the pseudo time increment. The

quantities a, b and B are functions of the stresses, while the parameter A is a

function of both the stresses and the hardening parameter.

To solve (3.3) and (3.4) for the unknown stresses σ and hardening parameter À

at the end of each pseudo time interval, a variety of numerical integration schemes

have been proposed. Because these equations need to be solved many times in

the course of a typical analysis, it is essential that the solution method is not only

accurate, but also efficient and robust. Two schemes that are used widely in

elastoplastic finite element codes are the explicit forward Euler algorithm and the

implicit backward Euler return algorithm. The former is one of a large family of

explicit methods and is often used with some form of subincrementation and stress

correction to improve its accuracy. Somewhat surprisingly, few published

65Chapter 3

comparisons are available on the relative performance of implicit and explicit

methods.

Advanced subincrementation methods, such as those discussed by Wissmann and

Hauck (1983) and Sloan (1987), are based on numerical procedures that have been

developed for integrating systems of ordinary differential equations. These

schemes are explicit, control the error in the computed stresses automatically, and

are often used in conjunction with a correction to return the stresses to the yield

surface during the integration process. Unlike implicit methods, explicit methods

do not require the solution of a system of nonlinear equations in order to compute

the stresses at each Gauss point. They do, however, need to compute the

intermediate stress state which lies on the yield surface if the stresses pass from

an elastic state to a plastic state. Explicit methods use the standard form of the

elastoplastic constitutive law and, thus, require only first derivatives of the yield

function and plastic potential.

Implicit backward Euler schemes are attractive because they do not require the

intersection with the yield surface to be computed if the stress point changes from

an elastic state to a plastic state. Furthermore, the resulting stresses will

automatically satisfy the yield criterion to a specified tolerance. In the most

general form of the implicit backward Euler scheme, which is commonly known

as the backward Euler return method, the elastoplastic stress increments are

obtained by solving a small system of nonlinear equations for each Gauss point.

Because these equations are usually solved using the Newton-Raphson algorithm,

considerable care must be taken to allow for possible non-convergence of the

resulting iteration scheme. The backward Euler return scheme has found wide

application in metal plasticity studies since it it provides all the information

required for the formation of the consistent tangent stiffness matrix. This matrix,

first identified by Simo and Taylor (1985), includes second order terms that are

usually ignored in the standard form of the elastoplastic constitutive relations, and

66Chapter 3

gives a quadratic rate of convergence for a Newton-Raphson solution of the global

stiffness equations. Although powerful, the backward Euler return method is

difficult to implement for complex constitutive relations because it is necessary to

evaluate second derivatives of the yield function and plastic potential.

Although the successful implementation of an elastoplastic model in a finite

element code is critically dependent on the choice of the stress integration scheme,

few comparisons have been published on the relative performance of implicit and

explicit methods. Direct comparisons are complicated by the fact that some

methods may perform well for a certain class of constitutive law but perform

poorly for others. Geotechnical constitutive models pose a severe test for many

stress integration schemes because they are typically complex in character and

depend on the hydrostatic pressure. This is in stark contrast to most metal

plasticity models, where the stress-strain behaviour is typically independent of the

hydrostatic pressure and governed by a relatively simple set of relations. Because

of the added complexity of many soil models, algorithms that perform well in

metal plasticity may prove unsatisfactory for geotechnical applications.

In two recent studies, Potts and Ganendra (1992, 1994) compared the

performance of the implicit return mapping algorithm of Ortiz and Simo (1986)

with the explicit subincrementation scheme of Sloan (1987). They used a critical

state soil model, which is typical of the complex constitutive laws used in

geotechnical studies, and concluded that the explicit subincrementation scheme is

more robust and efficient than the implicit return mapping algorithm.

Another study which investigated the relative performance of implicit and explicit

algorithms was published by Yamaguchi (1993). This work compared the explicit

Runge-Kutta and forward Euler techniques with the return mapping algorithm of

Ortiz and Simo (1986). Yamaguchi concluded that the Runge-Kutta scheme was

superior for problems with complicated constitutive laws or high accuracy

demands. It was also noted that the return mapping algorithm is effective for

67Chapter 3

simple material models but is less attractive for complex models. These

conclusions appear to confirm the findings of Potts and Ganendra (1992, 1994).

Some older finite element codes, such as the critical state implementation

described in Britto and Gunn (1987), still use a forward explicit scheme with a

single strain increment to integrate the stress-strain relations. This means that

many load increments need to be used in order to achieve an acceptable accuracy

in the solution, and is uncompetitive with the strategy of using an advanced stress

integration scheme with fewer load increments. The performance of single step

integration schemes, for both implicit and explicit methods, is discussed later in

this Chapter.

3.3 EXPLICIT INTEGRATION SCHEMES

In the substepping algorithms of Sloan (1987), the constitutive law is integrated

by automatically dividing the strain increment into a number of substeps. An

appropriate size for each substep is found through the use of modified Euler or

Runge-Kutta-Dormand-Prince formulae, which are specially constructed to

provide an estimate of the local error. A complete description of these techniques

may be found in Sloan (1987) and Sloan and Booker (1992).

The explicit substepping schemes used in this Thesis are based on the algorithms

of Sloan (1987) but include a number of enhancements to improve their accuracy,

efficiency and robustness. New algorithms for computing the yield surface

intersection, handling a negative plastic multiplier and correcting for drift from the

yield surface are described. Each of these aspects is now described in more detail.

3.3.1 Yield Surface Intersection

During a typical iteration or load increment of an elastoplastic analysis, the

incremental strains at each Gauss point are found from the incremental nodal

displacements using the strain-displacement relations. These may be written as

Δε= BΔu

68Chapter 3

where Δu denotes the nodal displacement increments, B the strain-displacement

matrix and Δε the vector of incremental strains. Once the strains have been

computed, the corresponding elastic stress increment is found using the elastic

stress-strain matrix De according to

Δσe= De Δε

Whether or not this increment causes a change from elastic to plastic behaviour

depends on the initial stresses σ0, the initial hardening parameter À0, the yield

function f, and the ‘elastic’ stresses σe= σ0+ Δσe. Such a change must occur if

f (σ0, À0)< 0 and f (σ0+ Δσe, À0)= f (σe , À0)> 0, and it is then necessary to

ascertain the fraction of Δσe which lies inside the yield surface. This situation,

shown in Figure 3.1, may arise many times during the course of an elastoplastic

Δσe

σe= σ0+ Δσe

σ0

f= 0

Figure 3.1 Yield surface intersection : Elastic to plastic transition.

f=+ FTOLf=− FTOL

σint= σ0+ αΔσe

finite element analysis and needs to be handled efficiently and accurately. Note

that, in Figure 3.1, the exact yield condition f (σ, À)= 0 has been replaced by the

approximation f (σ, À) ≤ FTOL, where FTOL is a small positive tolerance. This

allows for the effects of finite precision arithmetic and modifies the above

transition conditions to f (σ0, À0)<− FTOL and f (σe, À0)>+ FTOL. Suitable

values for the yield surface tolerance are typically in the range 10---6 to 10---9.

69Chapter 3

The problem of locating the stresses at the yield surface intersection point, σint,

and hence the elastic portion of Δσe, is equivalent to finding the scalar quantity

α which satisfies the nonlinear equation

f (σ0+ αΔσe, À0)= f (σint, À0)= 0 (3.6)

An α value of zero indicates that Δσe causes purely plastic deformation, while an

α value of unity indicates that Δσe causes purely elastic deformation. Thus, for

an elastic to plastic transition, α lies within the range 0< α< 1, and the elastic

part of the stress increment is given by αΔσe.

Since equation (3.6) defines a single nonlinear equation in the variable α, it can

be solved using the well known methods of bisection, regula-falsi, modified

regula-falsi, secant, and Newton-Raphson (see, for example, Conte and de Boor

(1980) for a detailed discussion of these algorithms). The first three schemes have

the advantage that they always bound the solution for α within a known interval

and, thus, are unconditionally convergent for continuous yield functions. The

Newton-Raphson and secant techniques, discussed by Sloan (1987), offer rapid

convergence rates but may diverge in some circumstances because they do not

constrain the solution. The modified regula-falsi procedure is ideally suited to

solving the yield surface intersection problem defined by (3.6) since it is

unconditionally convergent, does not require the use of derivatives, and typically

converges in four or five iterations (even when used with stringent values for the

tolerance FTOL). The complete modified regula-falsi algorithm is detailed below.

Modified Regula-falsi Intersection Scheme

1. Enter with stresses σ0 and hardening parameter À0, the stress increment

Δσe, initial values of α0 and α1 bounding the intersection with the yield

surface, and the maximum number of iterations MAXITS.

2. Set Fsave= f (σ0 , À0), F0= f (σ0+ α0Δσe ,À0) and F1= f (σ0+ α1Δσe,À0)

3. Do steps 4 to 7 MAXITS times

70Chapter 3

4. Calculate

α= α1− (α1− α0)F1

F1− F0

and set

Fnew= f (σ0+ αΔσe , À0)

5. If |Fnew|≤ FTOL go to step 9

6. If Fnew is of opposite sign to F0 then

Set α1= α and F1= Fnew

If Fnew is of the same sign as Fsave then set F0=F02

else

Set α0= α and F0= Fnew

If Fnew is of the same sign as Fsave then set F1=F12

7. Set Fsave= Fnew

8. Convergence not achieved afterMAXITS iterations, print error message and

stop.

9. Exit with α, the portion of Δσe that lies within the yield surface.

In the absence of better information, the algorithm is started by specifying α0= 0

and α1= 1. The maximum number of iterations permitted, MAXITS, is typically

set to ten and the procedure is terminated once the stresses satisfy the condition

f (σ0+ αΔσe, À0) ≤ FTOL.

3.3.2 Negative Plastic Multiplier

An elastic to plastic transition may also occur if a stress point, initially lying on

the yield surface, is subject to an elastic stress increment of the type shown in

Figure 3.2. This situation arises if the plastic multiplier, defined by equation (3.5),

is negative and f (σe, À0)>+ FTOL. The first of these conditions may be written

as

71Chapter 3

Δσe

f=+ FTOL

σe= σ0+ Δσe

σ0f= 0

f=−FTOL

Figure 3.2 Yield surface intersection: Negative plastic multiplier.

σint= σ0+ αΔσe

a0=∂f∂σ

θ

Δλ=aT0 Δσe

A0+ aT0Deb0< 0

where the quantities A0, a0, b0 are all evaluated at the initial stress state σ0.

Stress points with negative plastic multipliers can occur under monotonic loading

of the overall structure, particularly if the trial stress increment Δσe is large as a

result of the use of discrete load increments, and are often found near the tip of

the Mohr-Coulomb yield surface. Because the portion of the stress path that lies

within the yield surface is elastic, the elastoplastic constitutive law need only be

integrated beyond the last intersection point.

In practice, negative plastic multipliers can be detected by computing the cosine

of the angle between a0 and Δσe and checking whether

72Chapter 3

cos θ=aT0 Δσe a0 2 Δσe2

< LTOL

where LTOL is a suitable tolerance. This test is efficient since it avoids the need

to compute Δλ explicitly.

The procedure for finding the yield surface intersection for a negative plastic

multiplier is identical to that discussed in the previous Section, except that a

different set of starting values for α must be used. The situation is complicated

by the fact that the stress increment may in fact cross the yield surface twice, as

shown in Figure 3.2. This possibility is caused by the use of the tolerance FTOL,

which permits the stresses to lie just outside the yield surface, and must be

accounted for. To ensure the modified regula-falsi algorithm isolates the correct

crossing, it is sufficient to determine starting values, α0 and α1, which satisfy

f (σ0+ α0Δσe, À)<− FTOL and f (σ0+ α1Δσe, À)> FTOL. These conditions

guarantee that α0 and α1 bracket the second yield surface intersection. As in the

previous case, α= 0 indicates that Δσe causes purely plastic deformation while

α= 1 indicates that Δσe causes purely elastic deformation. Because it is assumed

that f (σe, À)>+ FTOL, α again must lie within the range 0< α< 1.

One strategy for locating the starting values which bracket the desired crossing is

based on breaking up the trial stress increment Δσe into a number of smaller

subincrements. Each of these is then scanned to see if the yield surface is crossed.

The number of subincrements used in the search, NSUB, is typically set to ten,

although numerical experiments suggest that as few as two subincrements may be

used with only a marginal increase in computation time. In the first iteration, the

subincrement size is set to Δσe∕NSUB, which corresponds to subincrements in α

of Δα= 1∕NSUB. A check is then made to see if the desired crossing lies in any

of the intervals defined by the pairs (αn−1 ,αn ), where αn= αn−1+ nΔα, α0= 0,

and n= 1, 2,. . ., NSUB. Such a crossing occurs if f (σ0+ αn−1Δσe, À)<− FTOL

and f (σ0+ αnΔσe , À)>+ FTOL. A geometric illustration of a successful search

73Chapter 3

with four subincrements is shown in Figure 3.3. In this example the required

σe= σ0+ Δσe

σ0

Figure 3.3 Starting values for yield surface intersection: Negativeplastic multiplier.

σint= σ0+ αΔσe

f=+ FTOL

f=− FTOL

f

α0

0.25 0.5 0.75 1

crossing between α= 0.75and α= 1

intersection with the yield surface lies between α= 0.75 and α= 1, and these two

limits serve as good starting values for the regula-falsi search.

If the stress increment Δσe is very large, or nearly tangent to the yield surface,

the initial subincrement size may not be small enough to detect the required

crossing. Although uncommon, this case can be checked by testing whether

f (σ0+ αn−1Δσe , À)≥− FTOL and f (σ0+ αnΔσe , À)> FTOL for each pair of

values (αn−1 ,αn ). If these two conditions are true, the crossing must lie in the

interval (0,αn ) and the search can be restarted using a smaller subincrement size

of Δα= αn∕NSUB. Because the benefit gained from each subsequent restart

diminishes fairly rapidly, these types of iterations should be limited in number.

Using the above strategy, the yield surface intersection point for a stress increment

with a negative plastic multiplier may be located as follows.

Modified Regula-Falsi Intersection Scheme for Negative Plastic Multiplier

1. Enter with initial stresses σ0, initial hardening parameter À0, and stress

increment Δσe.

74Chapter 3

2. Set α0= 0, α1= 1, F0= f (σ0,À0) and Fsave= F0

3. Do steps 4 to 5 MAXITS times

4. Calculate

Δα=α1− α0NSUB

5. Do steps 6 to 7 NSUB times

6. Calculate

σ1= σ0+ αΔσe

where

α= α0+ Δα

7. If f (σ1,À0)> FTOL, then

Set α1= α

If F0<−FTOL,

set F1= f (σ1,À0) and go to step 9.

else

set α0= 0 and F0= Fsave and exit loop over steps 6

and 7.

else

set α0= α and F0= f (σ1,À0).

8. Intersection not found after MAXITS iterations, print error message and

stop.

9. Exit with α0 and α1 bounding the yield surface intersection.

10. Call the modified regula-falsi algorithm with α0 and α1 to locate the yield

surface intersection.

In the above algorithm, the number of subincrements, NSUB, is typically set to ten,

while the maximum number of restart iterations, MAXITS, is typically set to three.

75Chapter 3

3.3.3 Correction of Stresses to Yield Surface

At the end of each subincrement in the integration process, the stresses may

diverge from the yield condition so that f (σ, À) > FTOL. The extent of this

violation, which is commonly known as yield surface ‘drift’, depends on the

accuracy of the integration scheme and the nonlinearity of the constitutive

relations. Sloan (1987) suggests that, provided the integration is performed

accurately, the extent of drift from the yield surface will tend to be small and no

remedial action is required. Potts and Gens (1985) and Crisfield (1991), on the

other hand, argue that some form of stress correction is advisable since the effect

of not satisfying the yield condition is cumulative. In this study, the stresses are

returned to the yield surface using a combination of two different methods.

Consider a point where the uncorrected stresses and hardening parameter, defined

by σ0 and À0, violate the yield condition so that f (σ0, À0) > FTOL. Ignoring

second order terms and above, f may be expanded in a Taylor series about this

initial stress point to give

f= f0+ aT0 δσ+

∂f∂À δÀ (3.7)

where δσ is a small stress correction, δÀ is a small hardening parameter correction,

f0= f (σ0, À0), and a0 is evaluated at σ0. In returning the stress state to the yield

surface, it is desirable that the total strain increment, Δε, remains unchanged,

since this is consistent with the philosophy of the displacement finite element

procedure. Inspection of equation (2.5) reveals that this requirement is satisfied

if the stress correction obeys the relation

δσ= − δλDe b0 (3.8)

where δλ is an unknown multiplier and b0 is evaluated at σ0. Using (3.2), the

hardening correction may be expressed as

δÀ= δλB0 (3.9)

76Chapter 3

where

B0=−A0∂f∕∂À

= ⎪⎪⎨⎧

23 bT0Mb0

σT0 b0

strain hardening

work hardening

Combining equations (3.7), (3.8) and (3.9) and setting f= 0 gives the unknown

multiplier as

δλ=f0

A0+ aT0De b0

This implies that the corrections to the stresses and hardening parameter are given

by

δσ= −f0De b0

A0+ aT0De b0

δÀ=f0B0

A0+ aT0De b0

and an improved stress state, which is closer to the yield surface, can be obtained

from

σ= σ0+ δσ

À= À0+ δÀ

This type of scheme, which is known as a consistent correction, may be applied

repeatedly until f (σ, À) ≤ FTOL. It has been used successfully for critical state

soil models by Potts and Gens (1985), and is also advocated by Crisfield (1991).

Under certain conditions, such as those that occur near the tip of the

Mohr-Coulomb surface for a material with a nonassociated flow rule, this

technique may not converge. Non-convergent behaviour is usually signalled when

77Chapter 3

the corrected stress state is further from the yield surface than the uncorrected

stress state. In these circumstances, the consistent return scheme may be

abandoned for one iteration and replaced with a correction which is normal to the

yield surface. This method, known as the normal correction, does not preserve

the total applied strain increment, but is very reliable and has been used

successfully by Nayak and Zienkiewicz (1972a), Owen and Hinton (1980), and

Sloan and Randolph (1982). With the normal correction scheme, equation (3.8)

is replaced by

δσ= − δλ a0

and it is assumed that the hardening parameter À0 remains unchanged. Using the

same argument as before, but neglecting any changes in À, it follows that

δσ= −f0 a0aT0 a0

(3.10)

This type of correction may also be applied iteratively until f (σ, À0) ≤ FTOL and

has proven to be very robust in practice.

The complete algorithm for returning the stresses to the yield surface and may be

summarised as follows.

Yield Surface Correction Scheme

1. Enter with uncorrected stresses σ0 and hardening parameter À0.

2. Do steps 3 to 6 MAXITS times.

3. Compute

δλ=f0

A0+ aT0Deb0

and then correct stresses and hardening parameter using

σ= σ0− δλDe b0

À= À0+ δλ B0

78Chapter 3

4. If |f (σ,À)|> |f (σ0,À0)|, then abandon previous correction and

compute

δλ=f0aT0 a0

σ= σ0− δλ a0

À= À0

5. If |f (σ,À)|≤ FTOL, then go to step 8.

6. Set σ0= σ and À0= À

7. Convergence not achieved after MAXITS steps, print error message and

stop.

8. Exit with stresses σ and hardening parameter À lying on the yield surface.

A suitable value for MAXITS, which denotes the maximum number of correction

iterations permitted, is typically between five and ten. Note that this correction

procedure will also be used for the single step backward Euler scheme described

in a later Section.

3.3.4 Modified Euler Scheme with Substepping

For a given strain increment, Δε, the constitutive relations to be integrated at each

Gauss point are described by equations (3.3) and (3.4) as

dσdT= DepΔε= Δσe− ΔλDeb (3.11)

dÀdT= ΔλB (3.12)

where

Δλ=aTDeΔεA+ aTDeb

=aTΔσe

A+ aTDeb

79Chapter 3

B=− A∂f∕∂À

=⎪⎪⎨⎧

23bTMb

σTb

strain hardening

work hardening

and the pseudo time lies in the range

0≤ T≤ 1

These equations describe a system of ordinary differential equations with known

initial conditions σ= σ0 and À= À0 at the start of the increment where T= 0

and t= t0. A wide variety of explicit methods may be used to integrate these

relations to give the stresses and hardening parameter at the end of the increment

where T= 1.

The approach used in this Thesis is based on the scheme of Sloan (1987). Sloan’s

method is attractive for finite element applications because it controls the errors

in the stresses and hardening parameter which are caused by the approximate

integration of the constitutive law. This error control is achieved by using a local

error measure to automatically subincrement the imposed strain increment Δε.

The local error measure is found by taking the difference between a modified

Euler solution, which is of second order accuracy, and an Euler solution, which

is of first order accuracy, for each subincrement. It thus corresponds to an

estimate of the local truncation error. Once the local error has been computed

for a given step, the size of the next step is determined by extrapolation of the

dominant error term. This means that the size of each subincrement may vary

throughout the integration process, depending on the nonlinearity of the

constitutive relations.

Consider a pseudo time subincrement in the range 0< ΔTn≤ 1 and let the

subscripts n− 1 and n denote quantities evaluated at the pseudo times Tn−1 and

Tn= Tn−1+ ΔTn. In the explicit Euler method, the solution for σ and À at the

end of a pseudo time step ΔTn is found from

80Chapter 3

σn= σn−1+ Δσ1

Àn= Àn−1+ ΔÀ1(3.13)

where

Δσ1= Dep (σn−1 , Àn−1 )Δεn

ΔÀ1= Δλ(σn−1 , Àn−1 ,Δεn)B(σn−1)(3.14)

and

Δεn= ΔTnΔε

A more accurate estimate of the stresses and hardening parameter at the end of

the interval ΔTn can be found using the modified Euler procedure. This gives

σ^n= σn−1+12 (Δσ1+ Δσ2)

À^n= Àn−1+12 (ΔÀ1+ ΔÀ2)

(3.15)

where Δσ1 and ΔÀ1 are computed from the Euler scheme and

Δσ2= Dep (σn−1+ Δσ1 , Àn−1 + ΔÀ1)Δεn

ΔÀ2= Δλ(σn−1 + Δσ1 , Àn−1+ ΔÀ1,Δεn)B(σn−1+ Δσ1)

Since the local truncation error in the Euler and modified Euler solutions is,

respectively, O(ΔT2) and O(ΔT3), the error in σn and Àn can be estimated from

⎪⎪⎨⎧

σ^n

À^n⎪⎪⎬⎫

⎭−⎪⎪⎨⎧

σn

Àn⎪⎪⎬⎫

⎭=⎪⎪⎨⎧

12 (Δσ2− Δσ1)

12 (ΔÀ2− ΔÀ1)

⎪⎪⎬⎫

Using any convenient norm, this quantity can be used to compute the relative error

measure

81Chapter 3

Rn= 12max Δσ2− Δσ1 σ^n ,ΔÀ2− ΔÀ1

À^n (3.16)

Note that the error in the stresses is treated separately from the error in the

hardening parameter to allow for differences of scale. Following Sloan (1987), the

current strain subincrement is accepted if Rn is not greater than some prescribed

tolerance, STOL, and rejected otherwise. Regardless of whether the subincrement

is accepted or rejected, the next pseudo time step is found from the simple relation

ΔTn+1= q ΔTn (3.17)

where q is chosen so that Rn+1 satisfies the constraint

Rn+1≤ STOL (3.18)

Now, since the local truncation error in the Euler method is O(ΔT2), it follows

from (3.17) that

Rn+1≈ q2Rn

Imposing the constraint (3.18) gives

q≤ STOL∕Rn

The above procedure for determining q is based on local extrapolation of the

dominant error term. Because local extrapolation may become inaccurate for

strongly nonlinear behaviour, it is wise to choose q conservatively to minimise the

number of rejected strain subincrements. Numerical experiments on a wide

variety of plasticity problems suggest that a suitable strategy for computing q is to

set

q= 0.9 STOL∕Rn (3.19)

and also constrain it to lie within the limits

0.1≤ q≤ 1.1 (3.20)

82Chapter 3

so that

0.1ΔTn−1≤ ΔTn≤ 1.1ΔTn−1

The coefficient of 0.9 acts merely as a safety factor, since it usually prevents the

step control mechanism from choosing strain subincrements which just fail to meet

the local error tolerance. Restricting the growth of consecutive strain

subincrements to ten percent also has this effect. Numerical experiments indicate

that increasing the maximum growth factor for consecutive subincrements to one

hundred percent has little influence on the performance of the algorithm.

Relaxing these constraints leads to larger subincrement sizes and hence fewer

strain subincrements overall, but this saving is counteracted by the increased

number of failed subincrements. Two final controls, of lesser importance than the

above refinements, impose a minimum absolute step size, ΔTmin, and prohibit the

step size from growing immediately after a failed subincrement. The first

condition is added merely for robustness, and will not be invoked unless the

constitutive law contains gradient singularities. The second condition ensures that

there are at least two strain subincrements of the same size following a failure,

and is useful for cases where the stress-strain path has sharp changes in curvature.

The integration scheme is started by applying (3.13) and (3.14) with the known

strains Δε , the initial stresses σ0, the initial hardening parameter À0, and an initial

pseudo time step ΔT1. In order to minimise the number of strain subincrements

for each Gauss point, ΔT1 is typically set to unity. If the relative error in the

resulting solution, as defined by equation (3.16), is not greater than the specified

tolerance STOL, then the current subincrement is accepted and the stresses and

hardening parameter are updated using either (3.13) or (3.15). In practice, it is

best to employ the higher order update rather than the lower order update, since

this is the most accurate of the two and has already been calculated. The extra

accuracy of the higher order update also compensates for the fact that (3.16) is

only a local and not a global error indicator. After a successful subincrement, the

83Chapter 3

new stresses and hardening parameter are corrected back to the yield surface using

the procedure described in Section 3.3.3. If the specified error tolerance is not

met, so that Rn> STOL,then the solution is rejected and a smaller step size is

computed using equations (3.19) and (3.20). The stage is then repeated and, if

necessary, the step size is reduced further until a successful subincrement size is

obtained. Regardless of whether the current subincrement is accepted or not, the

size of the next strain subincrement is found using (3.19) and (3.20). In successive

steps, the subincrements may become larger, smaller, or stay the same, depending

on the error that is calculated from equation (3.16). The end of the integration

procedure is reached when the entire increment of strain is applied so that

ΔTn= T= 1

The complete explicit modified Euler algorithm, which includes all of the

refinements described in Sections 3.3.1---3.3.3, may be summarised as follows.

Explicit Modified Euler Algorithm With Substepping

1. Enter with initial stresses σ0, initial hardening parameter À0, the strain

increment for the current step Δε, and the error tolerance for the stresses

STOL.

2. Compute the stress increment Δσe and the trial elastic stress state σe

according to

Δσe= DeΔε

σe= σ0+ Δσe

If f (σe,À0)≤ FTOL then the stress increment is purely elastic, so set

σ1= σe and À1= À0 and go to step 16.

3. If f (σ0,À0)<−FTOL and f (σe,À0)> FTOL then the stress point

undergoes a transition from elastic to plastic behaviour. Compute the

portion of Δσe that causes purely elastic deformation, α, using the modified

regula-falsi intersection scheme of Section 3.3.1 and go to step 5.

84Chapter 3

4. If |f (σ0,À0)|≤ FTOL and f (σe,À0)> FTOL then

Check for a negative plastic multiplier by computing the cosine of the

angle between a and Δσe from

cos θ= aTΔσe a 2 Δσe 2

where a is evaluated at the initial stress state.

If cos θ≥− LTOL then

The stress increment is purely plastic, so set α= 0.

else

Elastic unloading followed by plastic flow occurs. Compute the

portion of Δσe that causes purely elastic deformation, α, using

the modified regula-falsi intersection scheme for negative plastic

multipliers of Section 3.3.2.

else

The stress state is illegal.

5. Update the stresses at the onset of plastic yielding as σ0← σ0+ αΔσe.

Then compute the portion of Δσe that causes plastic deformation according

to Δσe← (1− α)Δσe.

6. Set T=0 and ΔT=1.

7. While T<1, do steps 8 to 15.

8. Compute Δσi and ΔÀi for i = 1 to 2 using

Δσi= ΔTΔσe − ΔλiDebi

ΔÀi= Δλi Bi

where

Δλi= max⎨⎧⎩ΔT aTi ΔσeAi+ aTi De bi

, 0⎬⎫⎭

85Chapter 3

Bi=−Ai∂f∕∂À= ⎪

⎪⎨⎧

23 bTi Mbi

σTi bi

strain hardening

work hardening

Ai=− (∂f∕∂À)Bi

ai= ∂f∂σi

bi= ∂g∂σi

are evaluated at (σi , Ài ), and

σ1= σT

σ2= σT+ Δσ1

À1= ÀT

À2= ÀT+ ΔÀ1

9. Compute the new stresses and hardening parameter and hold them in

temporary storage according to

σT+ΔT= σT+12 (Δσ1+ Δσ2)

ÀT+ΔT= ÀT+12 (ΔÀ1+ ΔÀ2)

10. Determine the relative error for the current substep from

RT+ΔT= max Δσ2− Δσ1 2 σT+ΔT , ΔÀ2− ΔÀ12ÀT+ΔT

, EPSwhere EPS is a machine constant indicating the smallest relative error

that may be calculated.

11. If RT+ΔT> STOL, then this substep has failed, so extrapolate to

obtain a smaller pseudo time step. First compute

q= max 0.9 STOL∕RT+ΔT , 0.1

and then set

86Chapter 3

ΔT← max { qΔT , ΔTmin }

before returning to step 8.

12. The substep is accepted, so update the stresses and the hardening

parameter according to

σT+ΔT= σT+ΔT

ÀT+ΔT= ÀT+ΔT

13. If f (σT+ΔT ,ÀT+ΔT) > FTOL, then correct σT+ΔT and ÀT+ΔT back

to the yield surface using the algorithm of Section 3.3.3.

14. Extrapolate to obtain the size of the next substep by computing

q= min 0.9 STOL∕RT+ΔT , 1.1

If previous step failed, limit growth of step size further by enforcing

q= min{q , 1}

Compute new step size and update pseudo time according to

ΔT← q ΔT

T← T+ ΔT

15. Ensure that next step size is not smaller than the minimum step size

and check that integration does not proceed beyond T=1 by setting

ΔT← max { ΔT , ΔTmin }

and then

ΔT← min { ΔT , 1− T }

16. Exit with stresses σ1 and hardening parameter À1 at end of increment,

where T=1.

An appropriate value for the tolerance LTOL, which is used for detecting elastic

unloading in step 4, is around 10---6. The tolerance EPS, which is used to define

the minimum relative error in step 10, is typically set to around 10---16 for double

precision arithmetic on a 32-bit machine.

87Chapter 3

The above scheme incorporates a several important refinements to the original

integration scheme of Sloan (1987). These include the improved regula-falsi

method for computing the elastic-plastic transition point on the yield surface, the

new scheme for dealing with negative plastic multipliers which cause plastic flow,

and the consistent yield surface correction procedure. Apart from these

algorithmic changes, a number of ‘tuning’ adjustments have also been made as a

result of extensive numerical experiments. The ‘safety factor’ coefficient on the

subincrement size, which was originally 0.8, has been increased to 0.9 to give better

control of the integration process. To minimise the number of rejected substeps,

the maximum growth in size between successive subincrements has been reduced

from one hundred percent to ten percent and no growth in the step size is now

permitted for the two subincrements that immediately follow a failed step. In the

event that a discontinuous yield surface is employed, a minimum step size has also

been introduced to force the integration of the constitutive equations. Typical

values for the minimum substep size ΔTmin are of the order of 10---4, which implies

that a maximum of 10,000 substeps may be used in the integration process.

3.3.5 Single Step Modified Euler Scheme

A single step modified Euler scheme can be derived from the previous algorithm

merely by setting the integration tolerance, STOL, to a large value. Results from

this type of method will be discussed in a later Section of this Chapter to provide

an indication of the accuracy of explicit integration schemes that do not employ

subincrementation.

3.3.6 Dormand-Prince Scheme with Substepping

The explicit Runge-Kutta-Dormand-Prince scheme is similar to the modified Euler

scheme described above except that a pair of four and fifth order integration

formulae are used to estimate the stresses and hardening parameter (Sloan and

Booker, 1992). With this high order algorithm, the stresses and hardening

parameter computed at the end of each subincrement are very accurate and

88Chapter 3

seldom need to be corrected back to the yield surface to satisfy the yield tolerance

of FTOL=10---9. This type of scheme is useful for checking the accuracy of lower

order methods, as it can be used with a stringent stress tolerance of STOL=10---9

or smaller.

When applied to (3.11) and (3.12) for a pseudo time step ΔTn, the fourth and fifth

order solutions for the stresses and hardening parameters, are given by

σn= σn−1+31540Δσ1+

190297Δσ3−

145108Δσ4+

351220Δσ5+

120Δσ6

Àn= Àn−1+31540ΔÀ1+

190297ΔÀ3−

145108ΔÀ4+

351220ΔÀ5+

120ΔÀ6

(3.21)

and

σ^n= σn−1+19216Δσ1+

10002079Δσ3−

125216Δσ4+

8188Δσ5+

556Δσ6

À^n= Àn−1+19216ΔÀ1+

10002079ΔÀ3−

125216ΔÀ4−

8188ΔÀ5+

556ΔÀ6

(3.22)

where

Δσi= Dep(σi, Ài)Δεn

ΔÀi= Δλ(σi, Ài,Δεn)B(σi)⎪⎪⎬⎫

⎭for i= 1, 2,. . ., 6

Δεn= ΔTn Δε

and

σ1= σn−1

À1= Àn−1(3.23)

89Chapter 3

σ2= σn−1+15Δσ1

À2= Àn−1+15ΔÀ1

(3.24)

σ3= σn−1+340Δσ1+

940Δσ2

À3= Àn−1+340ΔÀ1+

940ΔÀ2

(3.25)

σ4= σn−1+310Δσ1−

910Δσ2+

65Δσ3

À4= Àn−1+310ΔÀ1−

910ΔÀ2+

65ΔÀ3

(3.26)

σ5= σn−1+226729Δσ1−

2527Δσ2+

880729Δσ3+

55729Δσ4

À5= Àn−1+226729ΔÀ1−

2527ΔÀ2+

880729ΔÀ3+

55729ΔÀ4

(3.27)

σ6= σn−1−181270Δσ1+

52Δσ2−

266297Δσ3−

9127Δσ4+

18955Δσ5

À6= Àn−1−181270ΔÀ1+

52ΔÀ2−

266297ΔÀ3+

9127ΔÀ4+

18955 ΔÀ5

(3.28)

Subtracting equations (3.21) from equations (3.22) gives a fifth order estimate of

the local truncation error according to

⎪⎪⎨⎧

σ^n

À^n⎪⎪⎬⎫

⎭−⎪⎪⎨⎧

σn

Àn⎪⎪⎬⎫

⎭=⎪⎪⎨⎧

11360Δσ1−

1063Δσ3+

5572Δσ4−

2740Δσ5+

11280Δσ6

11360ΔÀ1−

1063ΔÀ3+

5572ΔÀ4−

2740ΔÀ5+

11280ΔÀ6

⎪⎪⎬⎫

The theory for implementing the explicit Dormand-Prince formulae is identical to

that for the modified Euler scheme. Steps 8, 9, 10, 11, and 14 of the algorithm

of Section 3.3.4 need to be changed as follows:

8. The variable i ranges from 1 to 6 and σi and Ài are given by equations

(3.23)---(3.28).

90Chapter 3

9. Compute the new stresses and hardening parameter and hold them in

temporary storage according to

σT+ΔT= σT+19216Δσ1+

10002079Δσ3−

125216Δσ4+

8188Δσ5

ÀT+ΔT= ÀT+19216ΔÀ1+

10002079ΔÀ3−

125216ΔÀ4−

8188ΔÀ5

10. Determine the relative error for the current substep from

RT+ΔT= max⎨⎧⎩ E σT+ΔT

σT+ΔT , E ÀT+ΔT

ÀT+ΔT, EPS⎬⎫⎭

where

E σT+ΔT

= 11360Δσ1−

1063Δσ3+

5572Δσ4−

2740Δσ5+

11280Δ

E ÀT+ΔT

= 11360ΔÀ1−

1063ΔÀ3+

5572ΔÀ4−

2740ΔÀ5+

11280Δ

and EPS is a machine constant indicating the smallest relative error

that may be calculated.

11. If RT+ΔT> STOL, then this substep has failed, so extrapolate to

obtain a smaller pseudo time step. First compute

q= max 0.9(STOL∕RT+ΔT)1∕5 , 0.1

and then set

ΔT← max { qΔT , ΔTmin }

before returning to step 8.

14. Extrapolate to obtain the size of the next substep by computing

q= min 0.9(STOL∕RT+ΔT)1∕5 , 1.1

If previous step failed, limit growth of step size further by enforcing

q= min{q , 1}

91Chapter 3

Compute new step size and update pseudo time according to

ΔT← q ΔT

T← T+ ΔT

The extra accuracy of the Dormand-Prince scheme is obtained at the expense of

additional evaluations of the constitutive relations for each subincrement. Six

evaluations per subincrement are needed, as opposed to two for the modified

Euler algorithm.

3.4 IMPLICIT INTEGRATION SCHEMES

A comprehensive discussion of various implicit stress integration methods has been

given by Crisfield (1991). Two schemes that have found wide application in metal

plasticity codes are forms of the backward Euler algorithm. The first of these,

which is termed a single step backward Euler method, is simple to implement since

it does not require second derivatives of the yield function or plastic potential.

The second method uses the well known backward Euler scheme for solving a

system of first order differential equations.

3.4.1 Single Step Backward Euler Scheme

The single step backward Euler scheme begins by computing the same trial elastic

stress state as before according to the sequence

Δε= BΔu

Δσe= DeΔε

σe= σ0+ Δσe

The stresses σe are then returned to the yield surface by using a single backward

Euler return and, if necessary, the correction process described in Section 3.3.3.

To derive the backward Euler return, f is expanded in a truncated Taylor series

about σe to give

92Chapter 3

f= fe+ aTe Δσ+∂f∂ÀΔÀ (3.29)

where Δσ is the elastoplastic stress increment, ΔÀ is the hardening parameter

increment, fe= f (σe , À0), and ae is evaluated at σe. In returning the stress state

to the yield surface, the total strain increment, Δε, must remain unchanged since

all of it has already been applied in computing the stress σe. Inspection of

equation (2.5) reveals that this condition is satisfied if the stress increment obeys

the relation

Δσ= − ΔλDe be (3.30)

where Δλ is an unknown multiplier and be is evaluated at σe. Using (3.2), the

hardening correction may be expressed as

ΔÀ= Δλ Be (3.31)

where

Be=−Ae∂f∕∂À= ⎪

⎪⎨⎧

23 bTeMbe

σTe be

strain hardening

work hardening

Substituting (3.30) and (3.31) in (3.29) and setting f= f (σ , À)= 0 gives the

unknown multiplier as

Δλ=fe

Ae+ aTe De be

This implies that the updated stresses and hardening parameter are

σ1= σe− ΔλDebe

À1= À0+ Δλ Be

If this updated stress state lies off the yield surface, so that f (σ1 , À1) > FTOL,

the correction algorithm of Section 3.3.3 is applied until f (σ1 , À1) ≤ FTOL. The

complete single step backward Euler procedure is described below.

93Chapter 3

Single Step Backward Euler Algorithm

1. Enter with initial stresses σ0, initial hardening parameter À0, and the

current strain increment Δε.

2. Compute the elastic stress increment Δσe and the trial elastic stress state

σe according to

Δσe= DeΔε

σe= σ0+ Δσe

If f (σe ,À0)≤ FTOL, then the increment is elastic so set σ1= σe and

À1= À0 and go to step 5.

3. Compute the plastic multiplier

Δλ=f (σe , À0)Ae+ aTe De be

and then update the stresses and hardening parameter using

σ1= σe− ΔλDe be

À1= À0+ Δλ Be

4. If f (σ1 ,À1) > FTOL, then restore stress state to yield surface using

algorithm of Section 3.3.3.

5. Exit with updated stresses σ1 and hardening parameter À1.

Note that the single step backward Euler method does not require the yield

surface intersection point to be found if the stress point undergoes a transition

from elastic to elastoplastic behaviour. This feature is an advantage over the

explicit schemes described previously. In cases where the strain increment is large

and the trial stress point, σe, is a long way outside the yield surface, the first

backward Euler return step may give a stress state for which f (σ1 ,À1)&FTOL.

Restoring the stress state to the yield surface then becomes vital and can cause

problems for complex constitutive models in which the gradients vary rapidly.

94Chapter 3

3.4.2 Backward Euler Return Scheme

The backward Euler return scheme involves solving a small system of nonlinear

equations at each Gauss point. These equations correspond to the incremental

form of the elastoplastic constitutive relations which, from (2.5), may be written

as

Δσ= Δσe− ΔλDeb

Letting σ0 denote the initial stress state, σe= σ0+ Δσe the trial elastic stress

state, and σ the final stress state, these equations may also be expressed in the

form

σ− σ0= σe− σ0− ΔλDe b

or

σ= σe− ΔλDeb

These equations are nonlinear because b is a function of the stresses and Δλ is

a function of both the stresses and the hardening parameter. To solve these

equations iteratively, it is convenient to define the residual

r= σ− σe+ ΔλDeb

which indicates how well the constitutive relations are satisfied at the current

values of σ and À. Noting that σe is constant and expanding r in a truncated Taylor

series gives

r= r0+∂σ∂σ δσ+ Δλ0De∂b∂σ0δσ+

∂(Δλ)∂(Δλ)

Deb0 δλ

where

r0= σ0− σe+ Δλ0Deb0

In the above, δσ is the iterative change in the elastoplastic stress increment, δλ

is the iterative change in the plastic multiplier, and the subscript 0 indicates

95Chapter 3

quantities that are evaluated using the stresses and hardening parameter at the

start of the iteration. Setting r= 0 and collecting terms, the iterative stress

increment is obtained as

δσ=−Q−10 r0+ δλDeb0 (3.32)

where

Q0= I+ Δλ0De∂b∂σ0and I is the identity matrix. To obtain the iterative change in the plastic multiplier,

δλ, the yield function f is expanded in a truncated Taylor series about the current

stress point according to

f= f0+ aT0 δσ+

∂f∂À δÀ= f0+ a

T0 δσ− A0δλ

where

δÀ= δλB0 (3.33)

and

B0=−A0∂f∕∂À

= ⎪⎪⎨⎧

23 bT0Mb0

σT0 b0

strain hardening

work hardening

Setting f= 0 and substituting (3.32) gives

δλ=f0− aT0 Q

−10 r0

A0+ aT0 Q−10 De b0

(3.34)

Equations (3.32)---(3.34) define the iterative change in the elastoplastic stress

increment and the hardening parameter. The stress state at the end of the

iteration is obtained from

96Chapter 3

σ= σ0+ δσ

À= À0+ δÀ

and the iterations are continued until f (σ ,À) ≤ FTOL.

As with the single step backward Euler method, the backward Euler return scheme

does not require the transition point between elastic and elastoplastic behaviour

to be isolated. This advantage is somewhat nullified by the fact that the gradient

derivative ∂b∕∂σ, which is needed to form the matrix Q, is often cumbersome to

enumerate for complex constitutive laws.

It should be noted that convergence of the backward Euler return scheme is not

guaranteed, particularly for large strain increments. To overcome this drawback,

it is advisable to implement a simple substepping strategy which automatically cuts

the size of the imposed strain increment if the stress state starts to diverge from

the yield surface between successive iterations. As a further safeguard, the strain

increment should also be reduced if convergence has not occurred within a

specified number of iterations.

The algorithm for the backward Euler return scheme, with substepping, is

summarised below.

Backward Euler Return Scheme

1. Enter with initial stresses σ0, initial hardening parameter À0, and the

current strain increment Δε.

2. Compute the elastic stress increment Δσe according to

Δσe= DeΔε

3. Set T=0 and ΔT=1.

4. While T<1, do steps 5---7 and steps 12---14 (substepping loop).

5. Compute

97Chapter 3

σe= σT+ ΔσeΔT

If f (σe,ÀT)≤ FTOL, then set σT+ΔT= σe and ÀT+ΔT= ÀT and go

to step 13.

6. First compute the multiplier

Δλ0T+ΔT=f (σe , ÀT)Ae+ aTe De be

Then set starting values for stresses and hardening parameter

according to

σ0T+ΔT= σe− Δλ0T+ΔTDe be

À0T+ΔT= ÀT+ Δλ0T+ΔT Be

If f (σ0T+ΔT , À0T+ΔT) ≤ FTOL, then set σT+ΔT= σ0T+ΔT and

ÀT+ΔT= À0T+ΔT and go to step 13.

7. Do steps 8 to 11 for i=1 to MAXITS (iteration loop).

8. Compute

δλi=f (σ i−1T+ΔT

, À i−1T+ΔT

)− aTsA+ aT t

where

s= Q–1 (σ i−1T+ΔT

− σe+ Δλi−1T+ΔTDeb)

t= Q–1Deb

Q= I+ Δλi−1T+ΔTDe ∂b∂σand A, a, b, and ∂b∕∂σ are all evaluated at (σ i−1

T+ΔT,À i−1T+ΔT

).

Then compute the iterative change in the stresses and hardening

parameter using

δσ i=− s− δλi t

δÀ i= δλi B

98Chapter 3

9. If f (σ i−1T+ΔT

+ δσ i,À i−1T+ΔT

+ δÀ i) > f(σ i−1T+ΔT

,À i−1T+ΔT

) , then goto step 12.

10. Update stresses, hardening parameter, and plastic multiplier

according to

σ iT+ΔT

= σ i−1T+ΔT

+ δσ i

À iT+ΔT

= À i−1T+ΔT

+ δÀ i

ΔλiT+ΔT= Δλi−1T+ΔT+ δλ

i

11. If |f (σ iT+ΔT

,À iT+ΔT

)|≤ FTOL, then set σT+ΔT= σiT+ΔT and

ÀT+ΔT= ÀiT+ΔT and go to step 13.

12. This substep failed to converge. Calculate new substep size according

to

ΔT← ΔT∕2

and return to step 5.

13. Check that the integration does not proceed beyond T=1 by setting

ΔT← min { ΔT , 1− T }

14. Update pseudo time according to

T← T+ ΔT

15. Exit with updated stresses σ1 and hardening parameter À1.

A typical value for the maximum number of backward Euler iterations, MAXITS,

is around five. If this limit is exceeded, the above scheme repeatedly halves the

applied strain increment until successful convergence is achieved.

3.5 COMPARISON OF INTEGRATION SCHEMES

To compare their numerical performance, each of the integration schemes

described in Sections 3.3 and 3.4 is used to predict the behaviour of a smooth rigid

strip footing resting on an elastoplastic soil mass. Due to the singularity at the

99Chapter 3

edge of the footing and the strong rotation of the principal stresses, this example

is a good test for assessing competing integration strategies.

The soil mass is modelled using the mesh of triangular cubic strain elements shown

in Figure 3.4. As discussed by Sloan and Randolph (1982), these elements are

smooth

smooth

smooth

B2

Figure 3.4 Mesh for footing analysis with various integration schemes.

5B

5B

72 elements

1166 degrees of freedom

capable of modelling incompressible plastic flow accurately and efficiently without

the need for reduced/selective integration or other numerical approximations. In

the first example, undrained loading of the soil is modelled using the rounded

Tresca yield described by Sloan and Booker (1986). For the remaining cases,

which simulate drained loading, the rounded hyperbolic Mohr-Coulomb yield

surface developed in Section 2.4.2 is employed. Elastic, perfectly plastic behaviour

is assumed.

In all cases, a non-iterative tangent stiffness method is used to solve the governing

stiffness equations and the footing is loaded by prescribed displacement

100Chapter 3

increments of equal size. The size of these increments is set so that the final

imposed displacement induces a state of collapse in the soil mass. At the end of

each displacement increment, the unbalanced nodal forces are calculated and

added to the next increment to minimise the drift from equilibrium. This type of

scheme corresponds to forward Euler integration with a load correction.

To assess the accuracy of each scheme, an estimate of the stress integration error

is is found directly from

σerror= stress error=

⎨⎧⎩mi=1[(σref− σ)T(σref− σ)]i⎬

⎫⎭1∕2

⎨⎧⎩mi=1[(σref)T(σref)]i⎬

⎫⎭1∕2

= σref− σ 2 σref 2

(3.35)

where the subscript i refers to each integration point, m is the total number of

integration points in the mesh, σ are the computed stresses, and the ‘reference’

stresses, σref , are calculated using the explicit Dormand-Prince integration scheme

with a stress tolerance of STOL=10---9. Note that the reference stresses provide

a very accurate set of stresses for the given mesh and loading sequence and all

values are computed at the end of the last load increment. A Euclidean norm is

used to measure the stress error because this matches the norm used in the explicit

subincrementation schemes. In all runs, an absolute yield surface tolerance of

FTOL=10---9 is employed. For the modified Euler runs with substepping, the error

computed from (3.35) may be compared directly with the stress tolerance STOL

to gauge the performance of the error control mechanism.

To provide an additional indication of the accuracy of each analysis, the numerical

collapse load is computed and compared with the exact solution from classical

plasticity theory. As loading is prescribed in the form of displacements, an

equivalent uniform pressure, p, is found by summing the appropriate nodal

reactions. At the end of each analysis, this uniform pressure is used to predict the

collapse load. It should be noted that the numerical collapse loads contain a

101Chapter 3

spatial discretisation error which reflects the size and distribution of elements in

the finite element mesh. Thus, even for analyses with very small stress integration

errors, the collapse loads will differ from the exact values.

All the timing data presented in this Section is for a HP710 workstation with the

HP FORTRAN 77 compiler and level 3 optimisation.

3.5.1 Rigid Strip Footing on Tresca Layer

The properties for the Tresca layer are summarised in Table 3.1. The rigid strip

footing is analysed with the mesh of Figure 3.4 using both 50 and 100 load

increments. The explicit modified Euler, explicit Dormand-Prince, implicit single

step backward Euler, and implicit backward Euler return integration schemes are

used to analyse each case. Results for the explicit modified Euler scheme are

presented for stress error tolerances ranging from STOL=1 to STOL=10---4, where

STOL=1 corresponds to a single step forward Euler scheme with no

subincrementation.

MaterialProperty

Tresca Mohr-Coulomb( associated )

Mohr-Coulomb( nonassociated )

Young’s Modulus E 298 1040 1040

Friction angle φ 0˚ 30˚ 30˚

Dilation angle ψ 0˚ 30˚ 20˚

Poisson’s Ratio ν 0.49 0.3 0.3

Cohesion c 1.0 1.0 1.0

Table 3.1 Material properties of soil layers.

The results for the analyses with 50 load increments are presented in Table 3.2.

For each of the integration schemes the total CPU time, collapse load, total

successful substeps, maximum number of successful substeps, and stress error, as

defined in equation (3.35), are recorded. Note that the total CPU time is for the

entire finite element analysis, not just the stress integration, since this quantity is

of primary interest in the design of finite element codes.

102Chapter 3

SchemeCPUtime(s)

Collapseload(p/c)

Totalsuccess.substeps

Max.success.substeps

Stresserror(σerror)

Modified Euler STOL=10---1 163.6 5.4102 14,082 1 0.16x10---2

STOL=10---2 164.2 5.4090 14,104 2 0.57x10---3

STOL=10---3 165.1 5.4079 15,635 5 0.71x10---4

STOL=10---4 164.7 5.4078 26,517 14 0.13x10---4

Single step modified Euler 163.6 5.4102 14,082 1 0.16x10---2

Dormand-Prince 165.6 5.4078 22,630 11 ---

Single step backward Euler 160.9 5.4075 14,059 1 0.96x10---3

Backward Euler return 163.0 5.4075 14,059 1 0.96x10---3

Table 3.2 Smooth rigid strip footing on Tresca layer (50 load steps).

The collapse loads for the explicit modified Euler scheme are similar for all of the

specified stress tolerances with values varying by less than 0.1 percent. They range

from p/c=5.4102 for STOL=10---1 to p/c=5.4078 for STOL=10---4, and are roughly

5 percent above Prandtl’s exact result of p/c=2+π. It is reassuring to note that

the collapse load obtained for STOL=10---4 is identical to the collapse load

obtained with the highly accurate Dormand-Prince method. The errors in the

computed stresses for the explicit modified Euler scheme, as defined by equation

(3.35), are always less than the integration tolerance STOL and thus give the

required error control. Because a minimal amount of substepping is performed

up to an integration tolerance of about 10---3, the error control appears

conservative with stress errors which are rather smaller than STOL. For the single

step modified Euler scheme, which has STOL=1, the stress error is 0.16× 10−2.

When STOL is tightened to STOL=10---2, the error is the same since no

substepping is required. With the most stringent tolerance of STOL=10---4, a

maximum of fourteen substeps are needed and the stress error is reduced by two

orders of magnitude to 0.13× 10−4. If the stress error tolerance in the modified

Euler scheme is tightened by a factor of ten, it is expected that the maximum

number of substeps would increase by a factor of approximately 10 . The

103Chapter 3

observed growth in the maximum number of substeps is somewhat less than this

prediction.

The implicit single step backward Euler and backward Euler return schemes

perform identically, predicting a collapse load of p/c=5.4075 which is 5.2 percent

above the exact Prandtl result. Interestingly, they both give the same stress error

of 0.96× 10−3 and never require iterations to ensure that the stresses satisfy the

yield criterion. This feature follows from the fact that, for stress states away from

the corners of the Tresca function, the gradient is constant along the return path

and the single step backward Euler procedure restores the stresses precisely to the

yield surface. Since the single step backward Euler method is used to initiate the

backward Euler return scheme, this implies that no iterations, and hence no

subincrements, are ever required with the latter algorithm. Both procedures may

be expected to give identical solutions under these conditions, and this is what is

observed.

SchemeCPUtime(s)

Collapseload(p/c)

Totalsuccess.substeps

Max.success.substeps

Stresserror(σerror)

Modified Euler STOL=10---1 322.0 5.4066 27,930 1 0.56x10---4

STOL=10---2 322.0 5.4066 27,930 1 0.56x10---4

STOL=10---3 320.1 5.4066 28,183 3 0.40x10---4

STOL=10---4 320.6 5.4066 35,319 7 0.98x10---5

Single step modified Euler 322.0 5.4066 27,930 1 0.56x10---4

Dormand-Prince 323.1 5.4066 32,043 6 ---

Single step backward Euler 320.6 5.4063 27,887 1 0.72x10---3

Backward Euler return 320.6 5.4063 27,887 1 0.72x10---3

Table 3.3 Smooth rigid strip footing on Tresca layer (100 load steps).

To investigate the influence of the load path on the performance of the various

integration schemes, the previous problem was also analysed using 100 load steps.

The statistics for this case are listed in Table 3.3. All the runs with the explicit

modified Euler scheme predict a collapse load of p/c=5.4066, regardless of the

104Chapter 3

stress error tolerance used. This value is identical to the collapse load from the

highly accurate Dormand-Prince method and suggests that the error due to the

discrete load increments is small. Because substepping is not required until

STOL=10---3, a single modified Euler step is often sufficient to meet the desired

level of accuracy and the observed stress errors are again substantially less than

the specified tolerances.

For reasons discussed previously, both of the implicit schemes return identical

results for the Tresca criterion. Their collapse loads and stress errors are very

similar to those of the 50 increment analysis.

Overall, the explicit modified Euler scheme is marginally more accurate than the

two implicit schemes. Since the total CPU time required for each of the stress

integration methods is remarkably similar, it would appear that there is little to

choose between explicit and implicit techniques for an elastic perfectly plastic

Tresca model.

3.5.2 Rigid Strip Footing on Associated Mohr-Coulomb

Layer

The material parameters describing the Mohr-Coulomb layer with an associated

flow rule are given in Table 3.1. As with the footing on the Tresca layer, the mesh

of Figure 3.4 is analysed using both 50 and 100 load increments.

Results for the 50 increment modified Euler analyses, shown in Table 3.4, give

collapse loads ranging from p/c=33.147 down to p/c=32.183, with the higher value

being for STOL=1 and the lower value being for STOL=10---4. These predictions

are, respectively, 10.0 percent and 6.7 percent above the exact Prandtl collapse

load of p/c=30.140. Overall, the results behave as expected, with the accuracy of

the collapse loads improving as the stress tolerance is tightened. As with the

Tresca analyses, the modified Euler collapse load for STOL=10---4 is the same as

the collapse load for the Dormand-Prince method.

105Chapter 3

SchemeCPUtime(s)

Collapseload(p/c)

Totalsuccess.substeps

Max.success.substeps

Stresserror(σerror)

Modified Euler STOL=10---1 168.2 32.582 26,541 20 0.52x10---1

STOL=10---2 169.0 32.242 28,725 91 0.17x10---1

STOL=10---3 169.9 32.185 40,556 280 0.21x10---2

STOL=10---4 174.2 32.183 78,819 290 0.20x10---3

Single step modified Euler 169.3 33.147 25,360 1 0.91x10---1

Dormand-Prince 180.2 32.183 66,575 166 ---

Single step backward Euler 166.5 32.957 25,288 1 0.69x10---1

Backward Euler return 173.0 32.175 28,484 256 0.27x10---2

Table 3.4 Smooth rigid strip footing on associated Mohr-Coulomb layer(50 load steps).

The observed stress errors for the explicit modified Euler scheme are always close

to the specified stress tolerance STOL. Indeed, the stress errors are generally

within an order of magnitude of the required levels, and thus display a ‘tolerance

proportionality’ which is certainly sufficient for controlling the stress integration

error in practical finite element analysis. For example, the most accurate modified

Euler integration, with STOL=10---4, gives a stress error of 2.0× 10−4, while

analysis with STOL=10---1 gives an error of 0.52× 10−1. Compared with the

corresponding results for the Tresca soil model, the Mohr-Coulomb stress-strain

relations typically require more than twice the total number of substeps to achieve

the same level of accuracy.

Generally speaking, the explicit and implicit single step integration schemes

perform in a like manner. The stress errors for these two methods are,

respectively, 0.91× 10−1 and 0.69× 10−1 and they both give similar collapse

loads. The stress error and collapse load for the backward Euler return scheme

indicate that it is significantly more accurate than the single step backward Euler

approach. Because the gradients along the return path are constant for the

Mohr-Coulomb relations, both versions of the backward Euler method do not

106Chapter 3

require iterations for stress points away from the corners. Near the corners,

however, the backward Euler return method may need a large number of substeps

in order to converge satisfactorily. This may be due to the fact that the rounded

Mohr-Coulomb surface has discontinuous second derivatives (with respect to the

stresses) at the transition angle θT. It could also be due to the fact that the

rounded surface has rapidly varying gradients. In any case, some form of

substepping is needed with the backward Euler return procedure when using the

rounded Mohr-Coulomb model. Alternatively, it may be better to employ the two

vector return strategy advocated by De Borst (1986) and Crisfield (1987). Rather

than smoothing the corners, this method invokes Koiter’s theorem for plastic flow

at yield surface discontinuities to construct a two stage backward Euler return.

This adds some complexity, but the results reported in Crisfield (1987) suggest that

the scheme works well.

SchemeCPUtime(s)

Collapseload(p/c)

Totalsuccess.substeps

Max.success.substeps

Stresserror(σerror)

Modified Euler STOL=10---1 332.4 32.378 51,765 11 0.31x10---1

STOL=10---2 329.2 32.175 53,992 58 0.73x10---2

STOL=10---3 330.5 32.155 60,255 149 0.12x10---2

STOL=10---4 334.7 32.154 89,530 150 0.12x10---3

Single step modified Euler 332.3 32.490 51,056 1 0.40x10---1

Dormand-Prince 339.6 32.154 82,556 95 ---

Single step backward Euler 324.9 32.592 50,199 1 0.52x10---1

Backward Euler return 342.0 32.156 54,273 127 0.55x10---3

Table 3.5 Smooth rigid strip footing on associated Mohr-Coulomb layer(100 load steps).

Results for the 100 increment analyses, shown in Table 3.5, indicate that all of the

stress integration methods give similar collapse loads to the 50 increment runs.

The errors in the collapse load for the modified Euler scheme range from 7.8

percent down to 6.7 percent, with the maximum error occurring for STOL=1 and

107Chapter 3

the minimum error occurring for STOL=10---4. For matching values of STOL, the

modified Euler collapse loads are very similar for the 50 increment and 100

increment analyses. With STOL=10---1, for example, the 50 and 100 increment

analyses give p/c=32.582 and p/c=32.378, a difference of only 0.6 percent.

Similarly, the corresponding values for STOL=10---4 are p/c=32.183 and

p/c=32.154, a difference of only 0.09 percent. This suggests that the modified

Euler error control mechanism is largely independent of the number of load

increments. Further support for this conclusion is available in the counts of the

total successful substeps. For STOL=10---4, where most Gauss points require more

than one substep, the 50 increment and 100 increment analyses give a similar total

number of substeps. In a similar manner to the 50 increment runs, the 100

increment analyses with the modified Euler scheme give stress errors which are

always close to the specified stress tolerance STOL.

Table 3.5 indicates that the collapse load for the single step backward Euler

method is altered little by increasing the number of load increments from 50 to

100. Indeed, doubling the number of increments only reduces the collapse load

error by 1.3 percent, from 9.4 percent to 8.1 percent. The corresponding stress

error, on the other hand, is reduced by roughly 25 percent. As in the 50 increment

analyses, the backward Euler return algorithm is more accurate than the single

step backward Euler method, but again requires substepping to obtain satisfactory

convergence. It is interesting to note that the collapse loads from the backward

Euler and modified Euler schemes are very similar for both sets of runs, provided

the latter algorithm is used with STOL=10---3 or smaller.

The CPU times given in Table 3.4 and Table 3.5 indicate that there is little to

choose between the explicit modified Euler method and the implicit backward

Euler methods in terms of computational efficiency. These conclusions echo those

for the Tresca criterion.

108Chapter 3

3.5.3 Rigid Strip Footing on Nonassociated Mohr-CoulombLayer

The material parameters for the Mohr-Coulomb layer with a nonassociated flow

rule are given in Table 3.1. Although the exact collapse load for a smooth strip

footing on this type of soil is unknown, the numerical results suggest that it is

probably very close to the associated flow rule value of p/c=30.140.

SchemeCPUtime(s)

Collapseload(p/c)

Totalsuccess.substeps

Max.success.substeps

Stresserror(σerror)

Modified Euler STOL=10---1 349.7 32.138 26,366 13 0.23x10---1

STOL=10---2 349.5 31.988 27,248 42 0.22x10---2

STOL=10---3 350.4 31.985 34,119 123 0.38x10---3

STOL=10---4 356.0 31.983 67,438 201 0.52x10---4

Single step modified Euler 350.1 32.198 25,661 1 0.40x10---1

Dormand-Prince 357.1 31.983 55,305 128 ---

Single step backward Euler 348.3 32.409 25,493 1 0.41x10---1

Backward Euler return 354.6 31.979 26,030 63 0.12x10---2

Table 3.6 Smooth rigid strip footing on nonassociated Mohr-Coulomb layer(50 load steps).

The results for the 50 and 100 increment analyses, shown in Table 3.6 and

Table 3.7 respectively, display similar trends to those for the associated flow rule

described previously. Indeed, many of the same observations and conclusions

apply. The observed stress errors for the modified Euler scheme are again within

an order of magnitude of the specified tolerances STOL for both sets of runs. This

indicates that the explicit error control also works well for a nonassociated

Mohr-Coulomb material.

The CPU times, collapse loads, and stress errors for the explicit and implicit single

step schemes are again similar for the 50 and 100 increment analyses. The

collapse loads for the backward Euler return and modified Euler schemes are also

very close, differing by less than 0.1 percent.

109Chapter 3

SchemeCPUtime(s)

Collapseload(p/c)

Totalsuccess.substeps

Maxsuccess.substeps

Stresserror(σerror)

Modified Euler STOL=10---1 692.3 32.079 51,715 7 0.16x10---1

STOL=10---2 694.5 31.976 52,454 21 0.41x10---2

STOL=10---3 693.0 31.954 55,539 100 0.77x10---3

STOL=10---4 695.0 31.951 80,145 108 0.79x10---4

Single step modified Euler 694.2 32.138 51,153 1 0.30x10---1

Dormand-Prince 698.6 31.951 72,214 75 ---

Single step backward Euler 689.2 32.252 50,671 1 0.31x10---1

Backward Euler return 698.8 31.948 51,764 64 0.60x10---3

Table 3.7 Smooth rigid strip footing on nonassociated Mohr-Coulomb layer(100 load steps).

3.6 CONCLUSIONS

The explicit modified Euler and implicit backward Euler return schemes both

provide an economical means of integrating rounded forms of the Tresca and

Mohr-Coulomb constitutive laws. Each procedure requires a similar amount of

CPU time and gives stresses which are of a similar accuracy. The explicit modified

Euler scheme with variable size substepping has the advantage of being able to

directly control the error in the stresses which would otherwise be unknown. This

is of particular benefit when analysing highly nonlinear problems where large

strain increments may be encountered. The implementation of the implicit

backward Euler scheme, with a simple substepping strategy to ensure convergence,

enables large strain increments to be integrated efficiently without fear of

numerical difficulty. When applied to the rounded Mohr-Coulomb yield function

without substepping, the implicit backward Euler method may require very small

load steps in order to ensure convergence of the iteration scheme.

Generally speaking, the explicit and implicit single step schemes integrate the

Tresca and Mohr-Coulomb constitutive relationships with the same degree of

accuracy and speed. Provided the load increments are not excessively large, each

110Chapter 3

of these methods also integrate the simpler Tresca yield criterion as accurately as

the more sophisticated modified Euler and backward Euler return procedures.

Consequently, these single step algorithms are viable options for the integration

of simple yield criteria.

111Chapter 4

CHAPTER 4

INTEGRATION OF

LOAD-DISPLACEMENT

RELATIONS

112Chapter 4

4.1 INTRODUCTION

This Chapter is concerned with the development of an algorithm for controlling

the error in nonlinear finite element analysis which is caused by the use of discrete

load steps. In contrast to most recent schemes, the proposed technique is

non-iterative and treats the governing load-deflection relations as a system of

ordinary differential equations. The procedure is very similar to the explicit stress

integration algorithm developed in Section 3.3.4 and is started by supplying a

number of coarse load increments. If necessary, these are automatically

subincremented to a size which is governed by the local truncation error. The

latter is measured by computing the difference between two estimates of the

displacement increments for each load step, with the initial estimate being found

from the first order Euler scheme and the improved estimate being found from

the second order modified Euler scheme. If the local truncation error exceeds a

specified tolerance, then the load step is abandoned and the integration is

repeated with a smaller load step whose size is found by local extrapolation. Local

extrapolation is also used to predict the size of the next load step following a

successful update. In order to control not only the local load path error, but also

the global load path error, the proposed scheme incorporates a correction for the

unbalanced forces. Overall, the cost of the automatic error control is modest and

compares favourably with that for traditional incremental methods. Because the

solution scheme is non-iterative and founded on successful techniques for

integrating systems of ordinary differential equations, it is particularly robust. To

illustrate the ability of the scheme to constrain the load path error to lie near a

desired tolerance, detailed results are presented in the last part of the Chapter.

Note that the load path error caused by the use of discrete load increments is quite

distinct from the spatial discretisation error due to the use of finite elements.

Although important, the latter issue is not addressed in this Thesis.

113Chapter 4

4.2 BACKGROUND

Techniques for solving the global equations associated with nonlinear finite

element analysis can be broadly classified as either iterative or incremental.

Iterative schemes treat the governing relations as a system of nonlinear equations

and attempt to solve them by applying the unbalanced forces, computing the

corresponding displacement increments, and iterating until the drift from

equilibrium is small. Newton-Raphson, modified Newton-Raphson, and initial

stress methods are all iterative techniques. One major disadvantage of the

Newton-Raphson family of algorithms is that the iterations may not converge,

particularly when the behaviour is strongly nonlinear. This may force various

stabilising measures to be used, such as line searches or arc length control, and

the procedures can rapidly become very complex in an effort to maintain

robustness. A second, equally serious, weakness of Newton schemes is that they

do not provide an estimate of the load path error. Although an iterative method

may converge, there is no guarantee that the final equilibrium state is sufficiently

close to the true equilibrium state. Satisfaction of this important condition can

only be ensured if small load increments are used, especially for plasticity

problems where significant load path errors can be caused by the use of coarse

load steps. Unfortunately, the size of the load increments required for an accurate

solution varies throughout the loading range and, not surprisingly, is highly

problem dependent. In general, accurate solutions with large load increments can

only be obtained for cases where the strain path is only mildly nonlinear, since it

is necessary to assume that the strain rate is constant throughout the iteration

process.

Explicit incremental schemes treat the governing relations as a system of ordinary

differential equations, involve no iteration, and generate the solution using a series

of piece-wise linear steps. Provided the global stiffness matrix remains well

conditioned, these techniques have proved to be very robust and are especially

114Chapter 4

useful for highly nonlinear problems involving complex constitutive behaviour.

With a sufficient number of load increments, they seldom fail to furnish a solution

of acceptable accuracy. One commonly perceived problem with explicit

incremental schemes is that they tend to ‘drift’ from equilibrium as the solution

proceeds. This effect can be minimised by calculating the unbalanced (or residual)

forces at the end of each load increment and adding these to the applied loads

for the next increment. This simple correction requires a relatively small amount

of computational effort yet ensures that equilibrium is approximately satisfied at

all times.

Somewhat surprisingly, the design of efficient, automatic load incrementation

strategies for nonlinear finite element analysis has not received wide attention in

the literature. Because of the inherent complexity of the problem, and the

tendency to focus on iterative solution schemes, most published techniques are

heuristic in nature and often require some intervention by the user. Some of the

more successful algorithms that have been developed use a variety of parameters

including the curvature of the nonlinear path (Den Heijer and Rheinboldt, 1981),

the ‘current stiffness parameter’ (Bergan et al, 1978 and Bergan and Soreide,

1978), and the number of iterations required to restore equilibrium (Crisfield,

1981). More recently Schellekens, Feenstra and de Borst (1992) proposed a

method based on strain energy. In general, the primary aim of these schemes is

to ensure convergence of various iterative procedures, rather than control the load

path error in the solution directly.

The method described in this Chapter is essentially an incremental scheme with

automatic step size control. The integration process selects each step so that the

local truncation error in the computed deflections is below a prescribed value, and

also includes an unbalanced force correction to prevent accumulation of global

error. A key feature of the scheme is that by automatically controlling the load

path, the load path error in the resulting final displacements can be constrained

115Chapter 4

to lie near a user specified tolerance. The scheme is particularly robust and

permits a broad class of load-deformation paths to be integrated with only a small

amount of drift from equilibrium. Since the method does not exploit any special

features of the governing equations, it can be used to deal with a wide range of

constitutive models. Moreover, complicated loading paths, such as those

associated with unloading and excavation sequences, do not present any special

problems. The motivation for the scheme comes from the successful application

by Sloan (1987) of a similar idea to the automatic integration of complex

constitutive laws.

4.3 EXPLICIT INCREMENTAL METHODS

In elastoplasticity, the system of differential equations to be solved for each load

increment can be expressed in rate form (see Section 2.3) as

U.= Kep(U)–1 F

.ext= Kep(U)–1 ΔFextΔt

(4.1)

where U.is a vector of unknown displacement rates, Kep is the tangent stiffness

matrix, and ΔFext is a vector of external force increments which are applied over

an arbitrary time interval Δt. For rate independent problems it is again convenient

to introduce the pseudo time, T, defined by

T=t− t0Δt

(4.2)

where t0 and t0+ Δt are, respectively, the times at the start and end of the load

increment and 0≤ T≤ 1. Using the chain rule for U.in (4.1) and substituting

for T from (4.2) yields

dUdT= Kep(U)–1ΔFext (4.3)

Equation (4.3) has the form of a classical initial value problem since ΔFext is

known, the right hand side is a function of U, and the initial conditions are the

known displacements, denoted as U0, at the start of the load increment where

116Chapter 4

T=0. The traditional and crudest method for solving such a system of differential

equations is the first-order forward Euler scheme. This explicit method calculates

the displacements at the end of the load increment, where T=1, using the

relationship

U1= U0+ Kep(U0 )–1ΔFext

Although the forward Euler scheme provides a simple means for solving the

governing load-deflection equations, it is accurate only for small load steps. The

numerical performance of the simple Euler scheme can be improved greatly by

first computing the unbalanced forces at the start of each increment according to

F unb(U0 )= Fext0−

V

BTσ 0dV (4.4)

and then adding these to the applied loads for the current load increment to give

U1= U0+ Kep(U0 )–1ΔFext+ Funb(U0 ) (4.5)

In (4.4), F ext0and σ0 are, respectively, the external forces and stresses at the start

of the current load increment. This simple procedure minimises the tendency of

the solution to drift from equilibrium as the integration proceeds. The accuracy

of the Euler solution can, of course, be further improved by dividing the applied

load increment into N subincrements of equal size. Equation (4.5) is then

replaced by the recurrence relation

UnΔT= U(n−1)ΔT+ Kep(U(n−1)ΔT )–11NΔF

ext+ Funb(U(n−1)ΔT )

where ΔTn= 1∕N is the pseudo time subincrement and n= 1, 2,. . ., N. The

initial conditions at the start of the load increment are U= U0 whilst at the end

of the load increment the displacements are given by U= U1. For this type of

subincrementation scheme to be efficient, it is necessary to be able to estimate the

117Chapter 4

number of load subincrements that are required to produce a solution of specified

accuracy. One such estimate, first proposed by Wissmann and Hauck (1983) for

the integration of the stress-strain relationships, can be obtained by using one

full-sized load increment and then two half-sized load increments. The difference

between these two solutions for the displacements provides a measure of the local

truncation error which, in turn, can be used to predict the number of

subincrements required to achieve a specified accuracy. Any of the well known

schemes for integrating systems of ordinary differential equations can be used with

this type of step size control, including the Euler method.

4.4 MODIFIED EULER SCHEME WITHSUBSTEPPING

Explicit adaptive integration schemes, which automatically adjust the step size to

keep the local truncation error near a specified tolerance, are standard methods

in numerical analysis for the solution of initial value problems. A variety of

explicit adaptive stress subincrementation schemes for integrating complex

constitutive relationships have been proposed by Sloan (1987) and Sloan and

Booker (1992). One of these variants, which is based on the Euler and modified

Euler methods, is discussed in Section 3.3.4 and is similar to the scheme described

here. Explicit substepping methods with automatic error control have proved to

be most effective in practice, and can be readily modified to integrate global finite

element equations. As described in Section 3.3.4, the key idea of these techniques

is to use two integration schemes, whose order of accuracy differs by one, to

predict the solution at the end of each step. The difference between the highest

order solution and the lowest order solution provides an estimate of the local

truncation error for the current step size. If this error is less than a specified level,

the solution is accepted and the next step size is predicted by using local

extrapolation of the dominant error term. Otherwise, the solution is rejected and

the stage is repeated with a smaller step whose size is again computed from local

118Chapter 4

extrapolation. In this way, the step size may increase or decrease, in accordance

with the local nonlinearity, as the integration proceeds.

Neglecting, for the moment, the effects of unbalanced forces, each stage of the

proposed scheme computes two estimates of the displacements which are based

on the Euler and modified Euler formulas. Consider a pseudo time subincrement

in the range 0< ΔTn≤ 1 and let the subscripts n− 1 and n denote quantities

evaluated at the pseudo times Tn−1 and Tn= Tn−1+ ΔTn. The Euler and

modified Euler schemes may be written, respectively, as

Un= Un−1+ ΔU1 (4.6)

U^n= Un−1+

12 (ΔU1+ ΔU2) (4.7)

where

ΔU1= Kep(Un−1 )–1ΔFextn (4.8)

ΔU2= Kep(Un−1+ ΔU1)–1ΔFextn (4.9)

and ΔFextn = ΔTnΔFext is the subincremental force vector. Since the local

truncation errors in Un and U^n are, respectively, O(ΔT2n) and O(ΔT3n), the

truncation error in Un can be estimated by subtracting the lower order solution

from the higher order solution to give

En ≈ 12 (ΔU2− ΔU1) (4.10)

where any convenient norm may be used. This quantity, which predicts the

absolute truncation error, can be divided by U^n to furnish the more usefuldimensionless relative error measure

Rn= En

U^n (4.11)

119Chapter 4

The current load subincrement is accepted if Rn is less than some specified

tolerance, DTOL, and rejected otherwise. In either case, the size of the next

pseudo time step ΔTn+1 is found from

ΔTn+1= qΔTn (4.12)

where q is a factor which is chosen to limit the predicted truncation error.

Following the argument of Section 3.3.4, the truncation error for the next load

subincrement, Rn+1, is related to the truncation error for the current load

subincrement, Rn , according to

Rn+1≈ q2Rn

and the required factor q is found by insisting that Rn+1≤ DTOL to give

q≤ DTOL∕Rn

As in Section 3.3.4, q is chosen conservatively to minimise the number of rejected

load subincrements. according to

q= 0.7 DTOL∕Rn (4.13)

with the additional constraint that

0.1≤ q≤ 1.1 (4.14)

The safety factor coefficient of 0.7 is less than the value of 0.9 used in the explicit

stress substepping scheme. This coefficient, which attempts to prevent the step

control mechanism from choosing load subincrements which just fail to meet the

local error tolerance, has been reduced because of the greater computational

penalty associated with a failed load subincrement. Note, however, that relaxing

this safety factor to 0.9, as well as increasing the upper limit on q to four in (4.14),

does not greatly influence the overall performance of the scheme. Although these

values may permit larger subincrements to be selected, this saving is often

120Chapter 4

counteracted by an increased number of failures. The strategy adopted in

equations (4.13) and (4.14) ensures that most of the substeps are successful

without making the step selection mechanism too conservative. As in the stress

substepping scheme of Section 3.3.4, it is also prudent to prohibit the step size

from growing immediately after a failed load subincrement. This ensures that

there are at least two load subincrements of the same size following a failure, and

is useful for cases where the load path has sharp changes in curvature.

The integration scheme is started by applying (4.6) and (4.7) with the known

incremental force vector ΔFext , the initial displacements U0, and an initial guess

for the pseudo time step ΔT1. For the first load increment ΔT1 is typically set

to unity, but in subsequent load increments ΔT1 may be initialised to the value

for the last successful subincrement. If the relative error in the resulting

displacements, as defined by equation (4.11), is less than or equal to the specified

tolerance DTOL, then the current load subincrement is accepted and the

displacements are updated using (4.6). The step size for the next load

subincrement is then found using equations (4.12)---(4.14). This may increase,

decrease, or stay the same, depending on the error that is calculated from equation

(4.11). If the specified error tolerance is not met, so that Rn> DTOL, then the

solution is rejected and a smaller step size is computed using equations

(4.12)---(4.14). The end of the integration procedure is reached when the entire

increment of load is applied so that

ΔTn= T= 1

A naïve analysis of the proposed algorithm would suggest that each successful load

subincrement requires two formations/factorisations of the stiffness matrix and two

equation solutions. With a minor change to the computation sequence, however,

only one stiffness matrix formation/factorisation and one equation solution are

needed, thus improving the efficiency of the scheme substantially. The key point

to note is that the displacements ΔU2, which are calculated for error control, can

121Chapter 4

be multiplied by the subincrement size factor q to provide the first order

displacements for the next load subincrement. Thus, after a successful load

subincrement, the first order displacements for the next load subincrement are

given by

ΔU1= q ΔU2 (4.15)

This feature is an important advantage of using the Euler-modified Euler pair for

error control. Compared with a simple forward Euler scheme using the same load

path, the adaptive method described above requires only one additional

formation/factorisation of the stiffness matrix and one extra equation solution for

each coarse load increment ΔFext. This extra work is minor for cases where the

number of subincrements in each coarse load increment is significant.

4.4.1 Correcting for Drift from Equilibrium

As mentioned previously, non-iterative incremental solutions tend to drift from

equilibrium as the integration proceeds and a load imbalance may develop

between the externally applied forces and the forces supported by the internal

stresses. One option for minimising this effect is to augment the externally applied

force vector for the current subincrement with the unbalanced forces at the end

of the previous subincrement. Equations (4.8) and (4.9) are then replaced by

ΔU1= Kep(Un−1 )–1ΔFextn + Funb(Un−1 ) (4.16)

ΔU2= Kep(Un−1+ ΔU1)–1ΔFextn + Funb(Un−1 + ΔU1) (4.17)

while the rest of the algorithm remains unchanged. Although simple and

seemingly efficient, equations (4.16) and (4.17) do not lead to a scheme with good

step size control. This is because the contribution of the unbalanced forces to the

deflections is independent of ΔTn . Even if ΔTn is reduced to zero, so that

ΔFextn = ΔTn ΔFext = 0, the error computed from (4.11) may still exceed the

122Chapter 4

specified tolerance due to the effect of the unbalanced forces. For strongly

nonlinear behaviour, this may result in the algorithm adopting tiny load

subincrements and not being able to advance the solution.

One strategy for avoiding this problem is to control the substep size by using only

the local truncation error due to the applied loads, as before, and add the effect

of the unbalanced forces separately. The subincremental displacements are then

divided into two parts, one part being due to the applied load and the other part

being due to the unbalanced force. Expanding equations (4.16) and (4.17), the

Euler and modified Euler updates may be rewritten as

Un= Un−1+ ΔU1+ ΔUunb1 (4.18)

U^n= Un−1+

12 (ΔU1+ ΔU2)+

12 (Δu

unb1 + ΔU

unb2 )

where

ΔU1= Kep(Un−1 )–1ΔFextn

ΔUunb1 = Kep(Un−1 )–1 Funb(Un−1 )

ΔU2= Kep(Un−1+ ΔU1+ ΔUunb1 )–1ΔFextn

ΔUunb2 = Kep(Un−1+ ΔU1+ ΔUunb1 )–1Funb(Un−1 + ΔU1+ ΔUunb1 )

To avoid the problems discussed above, the displacement contributions from the

unbalanced forces must be neglected when computing the local truncation error.

Thus equation (4.10) is still used to estimate En but the displacements are now

updated after a successful load subincrement using (4.18). The other change is

that (4.11) is no longer used to estimate the relative truncation error. Instead, this

quantity is computed from

Rn=‖ En ‖‖ Un ‖

123Chapter 4

Since the above modifications make it unnecessary to compute the quantities U^n

or ΔUunb2 , the revised scheme nominally requires two formations/factorisations of

the stiffness matrix and three equation solutions for each successful load

subincrement. As described in the previous Section, however, one of these

formations/factorisations and equation solutions can be removed by exploiting the

common evaluation points of the Euler and modified Euler formulas. The savings

occur immediately after a successful load subincrement, where the first order

displacements for the next subincrement can again be found from (4.15). When

compared with a simple forward Euler scheme using the same load subincrement

sizes, the adaptive method with an unbalanced force correction thus requires one

additional equation solution per subincrement. One additional formation

/factorisation of the stiffness matrix and one extra equation solution is also

required for each coarse load increment.

4.4.2 Prescribed Force Loadings

Due to the stiffness matrix becoming singular, difficulties may arise when collapse

studies are performed with prescribed force loading. No such problem, of course,

occurs for prescribed displacement analyses. Since the current scheme makes no

attempt to integrate post-peak behaviour with prescribed force loadings, a reliable

method for terminating the solution process gracefully is required. One simple

approach for detecting imminent collapse under these conditions is to monitor the

relative stiffness of the system. The current stiffness parameter of Bergan et. al.

(1978) is often used for this purpose, but it is best suited to proportional loading

and needs to be modified for the use of prescribed displacements.

An alternative measure of the relative stiffness can be found by first defining a

scalar which is a least squares fit to the incremental stiffness equations. Such a

scalar, denoted by Kn , minimises the quantity

KnΔUn− ΔFextn TKnΔUn− ΔFextn

124Chapter 4

where ΔFextn and ΔUn are, respectively, the incremental forces and displacements

for the nth step. Expanding this dot product and minimising with respect to Kn

gives

Kn=(ΔFextn )T ΔUn(ΔUn)TΔUn

At the end of the nth step, the relative stiffness can now be estimated as

K= KnK0=(ΔFextn )TΔUn(ΔUn)TΔUn

×(ΔU0)TΔU0(ΔFext0 )

T ΔU0(4.19)

where ΔFext0 and ΔU0 are the incremental forces and displacements for the first

step. The incremental stiffness parameter defined by equation (4.19) is applicable

at both the incremental and subincremental level, and can be used to monitor the

solution procedure. Once K falls below a small threshold value under force

prescribed loading, then the stiffness matrix is nearly singular and the analysis

should be terminated.

4.4.3 Efficient Formation of the Global Stiffness Matrix

Forming the elastoplastic global stiffness matrix afresh is computationally

expensive and should be avoided wherever possible. Fortunately, significant

savings can be obtained by separating the elastoplastic stiffness matrix into elastic

and plastic components and only recomputing the latter where necessary. To

derive this decomposition, the elastoplastic stress-strain matrix of equation (2.10)

is substituted into equation (2.31) to give the elastoplastic element stiffness matrix

as

kep= VeBT (De−Dp)BdV= ke− kp

where

ke= VeBTDeBdV

125Chapter 4

and

kp= VeBTDpBdV (4.20)

are the elastic and plastic element stiffness matrices respectively. This type of

subdivision naturally extends to the elastoplastic global stiffness matrix, Kep, which

is equal to the difference between the elastic global stiffness matrix, Ke, and the

plastic global stiffness matrix, Kp.

As the elastic global stiffness matrix remains unchanged during an analysis, it need

only be formed once and stored on disk. In order to generate Kep efficiently for

each load increment or subincrement, Ke is loaded into memory from disk and

the plastic element stiffness matrices kp are subtracted element by element. With

this approach, it is necessary only to consider the stiffness contributions from

plastic Gauss points. Gauss points which are elastic can be safely ignored as their

contribution to the stiffness matrix has already been accounted for.

The savings obtained from the above strategy are usually quite significant because,

in many elastoplastic problems, only a small proportion of the Gauss points

undergo plastic deformation. Even if this is not the case, and plastic deformation

is widespread, substantial savings are realised during the early stages of loading.

4.4.4 Implementation

The explicit modified Euler algorithm requires the user to specify a series of

coarse load increments that define ΔFext. These are then automatically

subincremented so that the relative load path error in the computed deflections

is close to a user-specified tolerance DTOL.

The complete load integration algorithm may be implemented as follows:

1. Enter with current stress state at each integration point (σ0 , À0), current

displacements U0, unbalanced force vector for current displacements

126Chapter 4

Funb(U0 ), external force increment ΔFext, previous subincrement size

ΔTlast , and displacement error tolerance DTOL.

2. Set T= 0 and ΔT= min {ΔTlast , 1}

3. Compute ΔU1 according to

ΔU1= ΔT Kep(U T)–1ΔFext

4. While T< 1 do steps 5 to 13.

5. Compute ΔUunb1 according to

ΔU unb1= Kep(UT)–1Funb(UT)

6. Calculate displacement update and hold it in temporary storage

according to

UT+ΔT= UT+ ΔU1+ ΔUunb1

For each integration point, compute strains Δε= BΔu1+ Δu unb1 and integrate constitutive law to find corresponding increments in

stresses, Δσ, and hardening parameter ΔÀ.

7. Compute ΔU2according to

ΔU2= ΔT Kep(UT+ΔT)–1ΔFext

8. Estimate local truncation error for current load subincrement using

ET+ΔT=12 (ΔU2− ΔU1 )

9. Compute relative error for current load subincrement using

RT+ΔT= maxEPS , ET+ΔT ∕ UT+ΔT

where EPS is a machine constant.

10. If RT+ΔT≤ DTOL go to step 11. Else current load subincrement has

failed, so estimate a smaller pseudo time step using

q= max{ 0.7 DTOL∕RT+ΔT , 0.1 }

127Chapter 4

ΔT← qΔT

Scale ΔU1according to

ΔU1← qΔU1

and return to step 6 to repeat the subincrement.

11. Current load subincrement is successful. Update displacements and

integration point stress states using

UT+ΔT= UT+ΔT

σT+ΔT= σT+ Δσ

ÀT+ΔT= ÀT+ ΔÀ

Then find new unbalanced forces Funb(UT+ΔT).

12. Compute least squares estimate of incremental stiffness from

Kn=(ΔFext)T ΔU1(ΔU1)TΔU1

If this is the first subincrement of the first coarse load increment, set

K0= Kn.

Calculate incremental stiffness parameter from

K= KnK0

If K ≤ KTOL and loading is by prescribed forces, exit to step 14 and

abandon subsequent coarse load increments.

13. Save size of last subincrement

ΔTlast= ΔT

and compute estimate of step size growth factor using

q= min 0.7 DTOL∕RT+ΔT , 1.1 , (1− T− ΔT )∕ΔT

If previous load subincrement was unsuccessful, do not allow

subincrement size to grow by enforcing

128Chapter 4

q← min{q , 1}

Update pseudo time and compute step size and first order

displacement prediction for next load subincrement according to

T← T+ ΔT

ΔT← qΔT

ΔU1= qΔU2

14. Exit with displacements U1 and integration point stress states (σ1, À1) at

T= 1, the end of the coarse load increment.

The variable ΔTlast stores the size of the second last subincrement that was used

successfully in the previous coarse load step. This is used as a trial value at the

start of each new coarse load step in order to minimise the number of rejected

subincrements. Note that it is necessary to store the second last subincrement size,

rather than the last subincrement size, since the latter may be severely truncated

to avoid overshooting the end of the integration interval.

In step 9, EPS again represents the smallest relative error that can be computed

on the host machine, and is typically set to around 10---16 for double precision

arithmetic on a 32-bit architecture. The value of KTOL, which is used in step 12

to detect singularity in the stiffness matrix for force prescribed loading, may be set

anywhere in the range 10---3 to 10---6. Extensive numerical experiments suggest that

a suitable value for detecting imminent collapse in plasticity problems is 10---4.

Finally, in step 13, care must be taken that the integration does not proceed

beyond T=1. This is implemented by insisting that the step growth factor, q, is

less than or equal to (1− T− ΔT)∕ΔT, where ΔT and T are, respectively, the

current subincrement size and the pseudo time at the start of the subincrement.

4.5 APPLICATIONS

In this Section, the automatic load stepping algorithm developed in the previous

Section is used to analyse the behaviour of four elastoplastic boundary value

129Chapter 4

problems. Detailed results are presented for the expansion of a thick cylinder, the

collapse of a rigid strip footing, the collapse of a flexible strip footing and the

stability of a trapdoor. Each of these boundary value problems is modelled using

fully-integrated cubic strain triangles, which are known to give reliable and

accurate results for plasticity problems (Sloan and Randolph, 1982). With the

exception of the trapdoor problem, which assumes a purely cohesive Tresca model,

all of the results are for cohesive-frictional constitutive behaviour and employ the

rounded hyperbolic Mohr-Coulomb yield surface described in Chapter 2. In the

case of the trapdoor problem, the corners of the Tresca criterion are also rounded

using the procedure described in Sloan and Booker (1986).

At the stress point level, the elastoplastic constitutive laws are integrated using an

the automatic stress integration scheme as described in Chapter 3. For all analyses

performed in this Section the constitutive laws are integrated very accurately by

using a stress error tolerance of STOL=10---6 in conjunction with an absolute yield

surface tolerance of FTOL=10---9.

As mentioned in Section 4.4, the new integration algorithm requires the loading

to be defined as a series of coarse load increments. For all of the problems, except

the flexible footing, the coarse load steps are applied as prescribed displacement

increments of equal size, such that the total imposed displacement induces a state

of collapse in the soil mass. In the case of the flexible footing, collapse is induced

by applying the coarse load steps as prescribed pressure increments of equal size.

To assess the performance of the automatic integration algorithm, the global load

path error in the final displacements u error is estimated using

uerror= displacement error= Uref− U ∞ Uref ∞

(4.21)

where Uref are a set of reference displacements. The reference displacements,

which have very small load path errors, are found using a first order Euler

130Chapter 4

algorithm with an equilibrium correction and a very large number of equal-size

load increments. Note that the error calculated from (4.21) may be compared

directly with the displacement error tolerance DTOL to assess the performance

of the automatic scheme. The max norm is used in (4.21) since this matches the

norm employed in step 9 of the algorithm of Section 4.4.4. As a further check

on the accuracy of the various analyses, the unbalanced forces are computed to

compare the equilibrium between the forces supported by the internal stresses and

the forces applied externally. Any drift from equilibrium is measured using

ferror= equilibrium error= F unb ∞ Fext ∞

(4.22)

where all values are computed at the end of each analysis. Equation (4.22) can

be used to confirm the accuracy of the reference solutions, which give

ferror ≤ 10−9 for all cases.

In the analyses using prescribed displacement loading, an equivalent uniform

pressure, p, is computed by summing the appropriate nodal reactions and dividing

by the displaced area. At the end of each analysis, this uniform pressure is used

to estimate the collapse load. It should be noted that these loads inevitably

contain a spatial discretisation error which is a function of the mesh refinement.

Thus, even for analyses with very small load path errors, the finite element

collapse loads will differ from the exact analytic values.

In the results that follow, various timing statistics are given to indicate the

efficiency of the proposed load incrementation scheme. All of these are for a

HP710 workstation with the HP FORTRAN 77 compiler and level 3 optimisation.

4.5.1 Thick Cylinder

Since it is essentially one dimensional in nature, the expansion of a thick cylinder

of cohesive-frictional material provides a simple test case for the automatic

integration algorithm. The mesh and material properties used to model this

131Chapter 4

problem are shown in Figure 4.1. The cylinder is loaded to failure by uniform

b

Ec = 10, 000 , ν= 0.3 , φ= 30˚b∕a = 2

smooth

smooth

uniformprescribeddisplacement

axis ofsymmetry

a

Figure 4.1 Expansion of thick cylinder of cohesive-frictional material.

prescribed displacements at its inner radius and an associated flow rule is used.

The analytic solution to this problem, which is useful for checking displacement

finite element codes, can be found in Yu (1992) and predicts that collapse occurs

when p/c=1.0174. In order to obtain a set of reference displacements which

contain very small load path errors, a forward Euler analysis with an equilibrium

correction and 100,000 load increments is used. These displacements, and the

corresponding stresses, are inserted in equations (4.21) and (4.22) to estimate the

displacement error, uerror, and the equilibrium error, ferror, at the end of each load

path.

To gauge the performance of a conventional integration scheme, the thick cylinder

is analysed using the forward Euler method with equilibrium correction and

various numbers of equal size load increments. The CPU times, collapse loads,

displacement errors and equilibrium errors for runs with 10, 100, 1,000 and

100,000 load steps are presented in Table 4.1. As expected, the displacement load

path errors for the Euler scheme decrease as the number of load increments is

increased. Due to the mildly nonlinear behaviour of this problem, only ten

132Chapter 4

increments are required to achieve a load path error of roughly 10---3 or better in

the final displacements. For a load path error of of around 10---5, about one

hundred increments are necessary. It is interesting to note that, with the simple

equilibrium correction, the global error of the Euler scheme is at least a quadratic

function of the increment size. Since the global error of the traditional Euler

scheme is a linear function of the step size, the merit of incorporating this

correction is clear. In a similar manner to the displacement error, the equilibrium

error also decreases as the number of increments is increased. With 100,000 load

increments, the Euler analysis predicts the exact collapse load of p/c=1.0174 and

gives an equilibrium error of less than 10---12.

Loadincrements

CPU time(sec)

Collapse load(p/c)

Displ. error(uerror)

Equil. error(ferror)

10 2 1.0277 1.9×10---3 0.36×10---2

100 9 1.0175 2.0×10---5 0.14×10---4

1000 84 1.0174 5.3×10---9 0.68×10---12

100,000 8,276 1.0174 --- 0.28×10---12

Table 4.1 Results for thick cylinder using Euler scheme.

To assess the performance of the automatic integration algorithm, the thick

cylinder is analysed with various values of the displacement error tolerance,

DTOL, ranging from 10---1 to 10---4. Each run with a fixed tolerance is performed

twice, once with five coarse load steps and once with ten coarse load steps, to test

the sensitivity of the results to this parameter. The selected load subincrements,

CPU times, collapse loads, displacement errors and equilibrium errors are

recorded in Table 4.2 for each analysis. When five or ten coarse load steps are

used, the scheme chooses not to subincrement for values of DTOL equal to 10---1

and 10---2. The computed load path errors in the displacements confirm that load

subincrementation is indeed unnecessary for these tolerances, since the maximum

error is only 0.7×10---2 for the five increment analysis. As the error tolerance is

tightened to 10---3 or 10---4, load subincrementation is required and the

133Chapter 4

performance of the automatic scheme can be assessed by comparing the computed

displacement errors with DTOL.

Errortol

Coarseload

Load subincr. CPUtime

Collapseload

Displ.error

Equil.errortol.

DTOLloadsteps succ. failed

time(sec)

load(p/c)

error(uerror)

error(ferror)

10---1 10 10 0 2.3 1.0277 0.2×10---2 0.36×10---1

5 5 0 1.7 1.0596 0.7×10---2 0.10

10---2 10 10 0 2.3 1.0277 0.2×10---2 0.36×10---1

5 5 0 1.7 1.0596 0.7×10---2 0.10

10---3 10 19 1 7 1.0195 0.4×10---3 0.16×10---1

5 18 2 6 1.0185 0.2×10---3 0.17×10---1

10---4 10 55 1 28 1.0176 0.5×10---4 0.40×10---2

5 56 2 29 1.0175 0.2×10---4 0.14×10---2

Table 4.2 Results for thick cylinder using automatic scheme.

Inspection of the results shown in Table 4.2 indicate that the resulting

displacement errors are all within an order of magnitude of the specified tolerance.

Analysis with ten coarse load steps, for example, gives a displacement error of

0.4×10---3 when an error tolerance of 10---3 is specified, while for a stricter

tolerance of 10---4, a displacement error of 0.5×10---4 results. As expected, the

equilibrium errors also decrease as the error tolerance is tightened but, in general,

these are two orders of magnitude larger than the displacement errors. It is

pleasing to note that, for all of the runs with the same error tolerance, the results

computed using five coarse load increments are very similar to those computed

using ten coarse load increments. Indeed, the errors for these two sets of runs

are of the same order of magnitude in each case, since the algorithm always

chooses a similar pattern of load subincrements. For typical error tolerances in

the range 10---2 to 10---3, the CPU time requirement of the automatic scheme is

very modest, and would certainly be competitive with that for a fast iterative

solution scheme.

134Chapter 4

The displacement load path errors cited in Table 4.1 and Table 4.2 are at the end

of the loading range, where collapse has taken place, and give no indication of how

the error varies during the analysis. This question is addressed in Figure 4.2, which

0.000001

0.000010

0.000100

0.001000

0.010000

40 50 60 70 80 90 100

automatic scheme with 10 coarseload steps and DTOL = 10--3

percentage of total displacement applied

10 Euler increments of equal size

100 Euler increments of equal size

displ.error(uerror)

Figure 4.2 Variation of displacement load path error with load level for thickcylinder.

shows the displacement error variation versus load level for a run with ten coarse

load steps and DTOL=10---3. In this analysis, the displacement error is calculated

at the end of each coarse load increment. For most of the loading range, it can

be seen that the automatic integration algorithm maintains an error of

approximately 0.5×10---3, which is equal to half the specified displacement

tolerance. Some oscillation in this type of plot is to be expected due to the effect

of truncated subincrements, which occur at the end of each coarse load step, as

well as the complexity of the differential equations being integrated. To give an

indication of the performance of a more conventional scheme, Figure 4.2 also

shows the displacement errors for the Euler algorithm with ten and one hundred

equal-size load increments. These results suggest that the use of fixed-size

135Chapter 4

increments is a good strategy when using the Euler method to analyse the thick

cylinder, since this keeps the load path error relatively constant.

The performance of the automatic algorithm is investigated further by applying all

of the load to the cylinder in a single coarse load step. The resulting

load-displacement curve, obtained with a load path error tolerance of 10---3, is

shown in Figure 4.3. Although an extreme test, the automatic scheme successfully

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

1.1

0.00 0.05 0.10 0.15 0.20 0.25 0.30

pc

1.017

displacement at inner radiusa × 103

Figure 4.3 Automatic load subincrement selection for thick cylinder with asingle coarse load step and DTOL=10---3.

reduces the initial increment size and then adjusts this throughout the integration

process to suit the nonlinearity of the behaviour. A more detailed plot of the load

subincrement sizes, shown in Figure 4.4, reveals that the scheme chooses three

large subincrements in the elastic range and then fifteen, roughly uniform,

subincrements in the elastoplastic range. The fact that the scheme selects load

steps of roughly equal size in the nonlinear range confirms the previous

observation that this is the optimum strategy for the Euler method when analysing

the thick cylinder problem.

136Chapter 4

0.00

0.01

0.02

0.03

0.04

0.05

0.06

0.07

0.08

0.09

0.10

0.11

0.12

1 5 10 15

load subincrement number

elastoplasticelastic

(ΔT)

loadsubincrement

size

Figure 4.4 Automatic load subincrement selection for thick cylinder with asingle coarse load step and DTOL=10---3.

4.5.2 Rigid Strip Footing

This section considers the problem of a smooth rigid strip footing resting on a

weightless cohesive-frictional soil with an associated flow rule. Due to the

singularity at the edge of the footing and the strong rotation of the principal

stresses, this case is a severe test for nonlinear solution schemes. The finite

element mesh and soil parameters used in the analysis are shown in Figure 4.5.

Vertical load is applied to the footing by a set of uniform prescribed displacements

and an equivalent pressure is again computed by summing the appropriate nodal

reactions. A total of 48 cubic strain triangles is used in the grid, and these are

concentrated under the edge of the footing in an effort to model the singularity.

The exact collapse load, derived by Prandtl, is given as p/c=30.1396. As with the

thick cylinder problem, the reference displacements are calculated using the Euler

scheme with an equilibrium correction and 100,000 load increments of equal size.

137Chapter 4

5B

B2

smooth

smooth

smooth

uniform prescribeddisplacement

5B

Figure 4.5 Smooth rigid strip footing on cohesive-frictional soil.

Gc = 400ν= 0.3φ= 30˚

The results shown in Table 4.3 summarise the performance of the Euler method

when it is used with an equilibrium correction and various numbers of equal-size

increments. Approximately one hundred steps are required to achieve a load path

error of 10---3 or better in the final displacements. For a load path error of 10---5,

more than one thousand increments are necessary.

Due to the influence of the simple equilibrium correction, the global load path

error of the Euler scheme appears to be a quadratic function of the increment size.

The equilibrium error for this problem shows a strong correlation with the

138Chapter 4

displacement error measure, and is negligible for an analysis with one thousand

load increments or more. With only ten load increments the error in the computed

collapse load is roughly eight percent. This is reduced to approximately two

percent when the analysis is performed with one thousand load increments.

Loadincrements

CPU time(sec)

Collapse load(p/c)

Displ. error(uerror)

Equil. error(ferror)

10 49 32.4508 5.5×10---1 0.31×100---

100 96 30.7530 6.4×10---3 0.80×10---3

1000 472 30.7522 3.2×10---5 0.57×10---5

100,000 45,835 30.7522 --- 0.42×10---9

Table 4.3 Results for smooth rigid strip footing using Euler scheme.

Errortol

Coarseload

Load subincr. CPUtime

Collapseload

Displ.error

Equil.errortol.

DTOLloadsteps succ. failed

time(sec)

load(p/c)

error(uerror)

error(ferror)

10---1 10 10 0 49 32.4509 5.5×10---1 0.31×100---

5 16 1 59 31.2404 2.7×10---1 0.26×100---

10---2 10 49 2 85 30.7724 3.2×10---2 0.41×10---2

5 44 3 78 30.7724 6.2×10---2 0.61×10---2

10---3 10 172 7 126 30.7525 2.0×10---3 0.43×10---3

5 167 8 127 30.7525 2.1×10---3 0.41×10---3

10---4 10 731 48 359 30.7522 1.5×10---4 0.11×10---3

5 744 50 369 30.7522 1.6×10---4 0.13×10---3

Table 4.4 Results for smooth rigid strip footing using automatic scheme.

The results for the footing analysis with the automatic integration algorithm are

presented in Table 4.4. Runs are performed with DTOL ranging from 10---1 to

10---4 and the loading is applied in five and ten coarse steps for each tolerance.

In all cases, the displacement load path error is well controlled by the automatic

scheme and is of the same order of magnitude as the specified error tolerance.

For example, with DTOL set to 10---1, the runs with five and ten coarse steps give,

respectively, observed load path errors of 2.7×10---1 and 5.5×10---1. With a tighter

139Chapter 4

tolerance of 10---4, the corresponding errors are 1.6×10---4 and 1.5×10---4. As in

the thick cylinder case, the load path subincrementation performed by the

automatic algorithm is largely independent of the number of coarse load steps

used in each analysis. For example, with an error tolerance of 10---4, the algorithm

creates 744 subincrements from five coarse load steps and 731 subincrements from

ten coarse load steps.

When ten coarse steps are used with DTOL=10---1 and DTOL=10---4, the error in

the respective collapse load ranges from eight percent to two percent. The timing

data in Table 4.4 indicates that, with roughly 750 load subincrements of variable

size, the automatic scheme takes approximately 0.50 seconds per subincrement.

This compares favourably with the performance of the Euler scheme which, from

the statistics shown in Table 4.3, takes approximately 0.47 seconds per load step

for a thousand steps of fixed size.

The variation of the load path error with load level for the strip footing analysis

is shown in Figure 4.6. Results are presented for analyses using five and ten coarse

load steps with displacement tolerances of 10---2 and 10---3. In each case the

displacement errors are seen to be within an order of magnitude of the specified

error tolerances. This plot also indicates that, for each tolerance, the load path

errors are similar regardless of the number of coarse load increments used in the

analysis. Figure 4.7 shows the load-displacement curve obtained from the

automatic scheme with a single coarse load step and an error tolerance of 10---2.

This case highlights the adaptive nature of the integration algorithm, which

automatically chooses small load subincrements in regions of highly nonlinear

behaviour. The slight kink in the load-deformation curve, just after yielding

commences, is detected by the scheme and small load subincrements are used.

The substeps then grow over the following portion of the curve which is nearly

linear. As a state of collapse is approached, the subincrement size is once again

reduced to enable accurate integration of the governing load-displacement

140Chapter 4

0.0001

0.0010

0.0100

0.1000

10 20 30 40 50 60 70 80 90 100

automatic scheme with DTOL = 10---3

displ.error(uerror)

percentage of total displacement applied

automatic scheme with DTOL = 10---2

5 coarse load increments10 coarse load increments

Figure 4.6 Variation of displacement load path error with load levelfor a rigid strip footing.

0

5

10

15

20

25

30

0.00 0.02 0.04 0.06 0.08 0.10 0.12 0.14 0.16

30.14

pc

footing displacementB

Figure 4.7 Automatic load subincrement selection for rigid footingwith a single coarse load step and DTOL=10---2.

141Chapter 4

0.00

0.01

0.02

0.03

0.04

0.05

0.06

1 5 10 15 20 25 30 35 40

loadsubincr.size

load subincrement number

(ΔT)

Figure 4.8 Automatic load subincrement selection for rigid footing withsingle coarse load step and DTOL=10---2.

equations. A more detailed picture of the variation of the load subincrement sizes

is presented in Figure 4.8.

4.5.3 Flexible Strip Footing

To demonstrate the ability of the automatic algorithm to analyse prescribed force

loading up to the point of collapse, the behaviour of a smooth flexible strip footing

resting on a cohesive-frictional soil is studied. The mesh, soil properties and

theoretical collapse load for this example are identical to those for the rigid footing

case, the only difference is that the footing is now loaded by nodal forces rather

than by nodal displacements. This problem is difficult because the stiffness matrix

becomes progressively ill-conditioned as collapse is approached. In order to

terminate the automatic scheme gracefully, the incremental stiffness parameter of

equation (4.19) is used to detect singularity with a threshold value of 10---4.

The only results presented for this case are the load-displacement curve of

Figure 4.9 and the plot of successful load subincrement sizes shown in Figure 4.10.

142Chapter 4

30.14

pc

0

5

10

15

20

25

30

0 0.020 0.040 0.060 0.080 0.100

centre-line displacementB

Figure 4.9 Automatic load subincrement selection for flexiblefooting with a single coarse load step and DTOL=10---2.

loadsubincr.size

load subincrement number

(ΔT)

0.00

0.02

0.04

0.06

0.08

0.10

0.12

1 5 10 15 20 25 30 35 40 45 50

0.0000

0.0002

0.0004

0.0006

0.0008

0.0010

35 40 45 50

Figure 4.10 Automatic load subincrement selection for flexible footing with asingle coarse load step and DTOL=10---2.

143Chapter 4

As before, these are obtained from analysis with a single coarse load increment

and a load path error tolerance of 10---2. Inspection of Figure 4.9 indicates that

the automatic scheme identifies the point of incipient collapse both correctly and

sharply. The load subincrement sizes, shown in Figure 4.10, clearly reflect the

nonlinearity of the load-displacement curve.

4.5.4 Rough Trapdoor

The undrained stability of a trapdoor provides another good test of the automatic

integration algorithm since the collapse mechanism is dominated by shear failure.

The mesh and soil properties for the problem are shown in Figure 4.11. To avoid

B2

B10

rough

rough

uniform prescribeddisplacement

5B

5B

smooth

Figure 4.11 Rough rigid trapdoor in purely cohesive soil.

Gc = 100

ν= 0.49

φ= 0˚

144Chapter 4

the development of a displacement discontinuity at the trapdoor edge, the element

side immediately adjacent to the trapdoor is subject to a linear variation of

imposed displacement. This displacement matches the trapdoor displacement at

one end of the side and decreases to zero at the other to satisfy the boundary

condition. Somewhat surprisingly, the exact collapse load for a trapdoor in purely

cohesive soil is still unknown, although rigorous upper and lower bounds have

been derived by Sloan et al (1990). For the trapdoor analysed in this paper, these

upper and lower bounds are respectively p/c=6.34 and p/c=5.77, where p is an

equivalent pressure. To obtain a set of reference displacements for this problem,

the corrected Euler method with 50,000 load increments is used.

Results for the corrected Euler method, with various numbers of fixed size

increments, are shown in Table 4.5. With this scheme, only ten increments are

required to achieve a load path error of 10---2 or better in the final displacements.

For a load path error of 10---4, around one hundred steps are necessary. Using

50,000 load increments, the Euler analysis predicts a collapse load of p/c=5.9316

which falls between the bounds of Sloan et al (1990). Since this analysis gives an

equilibrium error of less than 10---9, there would appear to be very little load path

error in the reference displacements. The influence of the equilibrium correction

on the performance of the Euler scheme is once again clearly apparent, as the

global load path error decreases quadratically with decreasing increment size.

Loadincrements

CPU time(sec)

Collapse load(p/c)

Displ. error(uerror)

Equil. error(ferror)

10 38 5.9455 2.0×10---2 0.25×10---1

100 160 5.9318 1.3×10---4 0.52×10---3

1,000 1,406 5.9316 9.6×10---7 0.90×10---6

50,000 72,826 5.9316 --- 0.35×10---9

Table 4.5 Results for rough trapdoor using Euler scheme.

145Chapter 4

Errortol

Coarseload

Load subincr. CPUtime

Collapseload

Displ.error

Equil.errortol.

DTOLloadsteps succ, failed

time(sec)

load(p/c)

error(uerror)

error(ferror)

10---1 10 10 0 48 5.9456 0.2×10---1 0.25×10---1

5 5 0 37 5.9348 0.6×10---1 0.10×100---

1 1 0 5 7.6139 1.0×10---1 0.33×100---

10---2 10 57 4 119 5.9344 0.06×10---2 0.44×10---2

5 53 4 106 5.9360 0.1×10---2 0.82×10---2

1 51 6 101 5.9373 0.1×10---2 0.57×10---2

10---3 10 133 4 224 5.9330 0.3×10---3 0.37×10---2

5 128 5 208 5.9349 0.6×10---3 0.63×10---2

1 130 6 208 5.9342 0.7×10---3 0.42×10---2

10---4 10 637 29 950 5.9319 0.3×10---4 0.52×10---3

5 650 34 955 5.9319 0.3×10---4 0.11×10---2

1 633 31 925 5.9319 0.3×10---4 0.30×10---2

Table 4.6 Results for rough trapdoor using automatic scheme.

The results for analysis of the trapdoor using the automatic integration algorithm

are shown in Table 4.6. Data for runs with error tolerances ranging from 10---1 to

10---4 are presented, with each tolerance being analysed with one, five and ten

coarse load steps.

In all of these analyses, the observed load path errors are considerably less than

the specified error tolerance, which indicates that the automatic scheme is rather

conservative in its step size control for this problem. With ten coarse load steps,

for example, a displacement error of 0.2×10---1 is obtained using an error

tolerance of 10---1. When the latter is tightened to a value of 10---4, the observed

displacement error is 0.3×10---4. As in the previous examples, the subincremental

strategy chosen by the algorithm is largely independent of the initial number of

coarse load steps. For coarse load steps of one, five and ten, the algorithm uses

a minimum of 633 and a maximum of 650 load subincrements to integrate to an

error tolerance of 10---4. The collapse loads for the trapdoor range from

p/c=7.6139 to p/c=5.9319, where the former value is computed using a single

146Chapter 4

coarse load step with a tolerance of 10---1 and the latter value is found from all

of the analyses with a tolerance of 10---4. The CPU times for the automatic scheme

are again competitive for typical tolerances of 10---2 to 10---3.

Figure 4.12 shows the load-displacement curve for a run performed with only a

0.0

1.0

2.0

3.0

4.0

5.0

6.0

0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40

pc

6.345.77

trapdoor displacementB

Figure 4.12 Automatic load subincrement selection for rigid trapdoor witha single coarse load step and DTOL=10---2.

single coarse load step and an error tolerance of 10---2. A more detailed profile

of the load subincrement sizes for this case, shown in Figure 4.13, clearly indicates

that a small step size is required at the beginning of the analysis while a larger

step size is permissible as the trapdoor collapses. This variation of load increments

is not intuitive and is unlikely to be chosen by even the most experienced analyst.

Figure 4.13 also suggests that the step size control during the analysis is restricted

by the rule which limits the growth of consecutive load subincrement sizes to ten

percent. Figure 4.14 shows the variation of the load path error with load level for

various analyses of the trapdoor. The results for the automatic scheme, using five

and ten coarse load increments with an error tolerance of 10---3, suggest that the

load path error is always kept below the desired threshold. The plot for the Euler

scheme, obtained using an equilibrium correction and one hundred fixed size

147Chapter 4

loadsubincr.size

load subincrement number

(ΔT)

0.00

0.01

0.02

0.03

0.04

0.05

0.06

0.07

0.08

0.09

1 5 10 15 20 25 30 35 40 45 50

Figure 4.13 Automatic load subincrement selection for rigidtrapdoor with a single coarse load step and DTOL=10---2.

0.00001

0.00010

0.00100

0.01000

10 20 30 40 50 60 70 80 90 100

displacementerror(uerror)

percentage of total displacement applied

100 Euler increments of equal size

automatic scheme with 5 coarseload increments and DTOL=10---3

automatic scheme with 10 coarseload increments and DTOL=10---3

Figure 4.14 Variation of displacement load path error with load level for rigidtrapdoor.

148Chapter 4

steps, indicates that the use of equal size increments leads to decreasing load path

error with increasing load level. This data supports the strategy of the automatic

scheme, which tries to keep the load path error constant by increasing the

subincrement size as the load is increased.

4.6 CONCLUSIONS

The explicit modified Euler algorithm is a robust and efficient method for

integrating the global stiffness equations in nonlinear finite element analysis. By

automatically subincrementing a number of user-defined coarse load steps, the

scheme is able to solve the governing relations so that the load path error in the

final displacements lies near a specified tolerance. The technique successfully

controls this error independently of the number of coarse load increments supplied

by the user, and may even be employed with a single coarse load step.

149Chapter 5

CHAPTER 5

INTEGRATION OF

CONSOLIDATION RELATIONS

150Chapter 5

5.1 INTRODUCTION

This Chapter describes a new finite element algorithm for solving elastic and

elastoplastic coupled consolidation problems. The procedure treats the governing

consolidation relations as a system of first order differential equations and is based

on the Backward Euler scheme with automatic subincrementation of a prescribed

series of time increments. The prescribed time increments, which are sometimes

called coarse time steps, serve to start the procedure and are chosen by the user.

In a similar manner to the load-stepping scheme described in Chapter 4, the

automatic consolidation algorithm attempts to choose the time subincrements such

that, for a given mesh, the time-stepping (or temporal discretisation) error in the

displacements lies close to specified tolerance.

Unlike existing solution techniques, the new algorithm computes not only the

displacements and pore pressures, but also their derivatives with respect to time.

These extra variables permit a family of unconditionally stable integration

algorithms to be constructed which automatically provide an estimate of the local

truncation error for each time step. This error estimate is very cheap to compute

and may be used to develop a simple and efficient automatic time stepping

mechanism. For the elastic case, the displacements and pore pressures at the end

of each subincrement may be solved directly without the need for iteration. For

elastoplastic behaviour, however, the governing relationships are nonlinear and a

system of nonlinear equations must be solved to compute the updates.

5.2 BACKGROUND

The finite element method can be used to model coupled consolidation using a

mixed formulation which incorporates displacement and pore pressure freedoms.

Even for elastic material behaviour, the resulting governing equations are

nonlinear due to the dissipation of excess pore water pressures. The solution of

these equations requires the linearisation of the time domain into a number a

discrete time increments, each of which is considered in sequence, and is difficult

151Chapter 5

to attempt by trial and error. The primary difficulty is due to the fact that an

acceptable increment size may vary by several orders of magnitude throughout the

analysis. During the early stages of consolidation, where excess pore water

pressure gradients are usually high, relatively small time increments are necessary

to obtain an accurate solution. As the analysis proceeds, however, much larger

time increments can be used with safety. Indeed, large increments are often

mandatory in order to obtain an efficient solution over a long period of

consolidation.

Terzaghi (1960) was the first to consider the analysis of consolidation and based

his formulation on the theory of diffusion. Nowadays, the numerical analysis of

consolidation is usually founded on the theory of Biot (1941), for which Sandu and

Wilson (1969) were the first to present a solution using finite elements. Since then

numerous researchers, including Christian and Boehemer (1970), Hwang et al

(1971), Yokoo, Yamagata and Nagaoka (1971a), Krause (1978), and Borja (1986)

have formulated the governing finite element equations for elastic materials. For

nonlinear materials, incremental solution strategies have been given by Lewis et

al (1976), Small et al (1976), Prevost (1982), and Borja (1989). Other solution

methods have been presented by Carter et al (1977 and 1979), who incorporated

the effects of finite deformations, while pore fluid compressibility was considered

by Gaboussi and Wilson (1973). In each of these formulations, whether linear or

nonlinear, the governing finite element relations can be expressed as a system of

coupled differential equations.

Techniques for the finite element solution of consolidation problems usually

involve a simple one dimensional time integration scheme, the stability and

accuracy of which has been investigated by Booker and Small (1975) and Vermeer

and Verruijt (1981). For elastic soils, the resulting time stepping schemes are

essentially the same as those used in the solution of first order systems of

differential equations. Since these types of equations arise in many areas of the

152Chapter 5

physical sciences, they have been studied extensively and a vast amount of

literature exists on their solution. An excellent summary of the stability and

accuracy of various algorithms can be found in Wood (1990). In order to solve

elastic coupled consolidation problems efficiently with the well known θ-method,

it is generally necessary to use an implicit time integration scheme with θ≥ 0.5.

With this choice of integration parameter, Booker and Small (1975) proved that

the solution process is unconditionally stable so that large time increments may

be used with safety. Explicit integration methods, which employ θ= 0, are only

conditionally stable and may require the use of very small time steps.

In the analysis of elastoplastic soils, the application of implicit time integration

schemes requires the solution of a system of nonlinear equations for each time

step. Small et al (1976) solved these nonlinear equations with an initial stiffness

iteration scheme which used averaged values of all time varying quantities. This

approach has been adopted by Siriwardane and Desai (1981), who also present an

alternative tangent stiffness update with no iteration.

Somewhat surprisingly, very little work has been done on the development of

automatic time stepping algorithms for finite element analysis of consolidation.

In a more general context, a number of schemes for controlling the increment size

during the solution of second order differential equations have been presented by

Zienkiewicz et al (1984) and Thomas and Gladwell (1988a). All of these methods

use an estimate of the local truncation error to regulate the step size and were

developed for the analysis of dynamic problems, such as those that occur in

earthquake loading. Although the schemes in Zienkiewicz et al (1984) use a Taylor

series expansion to estimate the local error, and permit the time increments to

expand or contract, they make no effort to control the error in the solution

directly. Thomas and Gladwell (1988a), on the other hand, take the difference

between solutions from p and p+1 stage schemes as an estimate of the local

truncation error and attempt to constrain this precisely by adaptive control of the

153Chapter 5

increment size. A key advantage of these methods is that they operate in single

step mode and, hence, do not need to use values generated in previous time

increments. Moreover, their local error estimators are embedded and can be

computed cheaply with no extra matrix factorisations. The two methods proposed

by Thomas and Gladwell (1988a) are aimed at systems of second order differential

equations and are based on a first order accurate 2---stage/second order accurate

3---stage pair and a second order accurate 3---stage/third order accurate 4---stage

pair. The latter is only conditionally stable but has a large region of stability. In

a companion paper, Gladwell and Thomas (1988b) discuss implementation issues

associated with their formulas and give an example code to illustrate the

computational detail.

The scheme presented in this Chapter differs from those of Thomas and Gladwell

(1988a) in that it uses a first order accurate 1-stage/second order accurate 2-stage

pair of integration schemes to estimate the local truncation error for systems of

first order differential equations. The proposed scheme enables the truncation

error to be estimated using a single factorisation and solution of the matrix

equations.

5.3 FORMULATION OF GOVERNING BIOTCONSOLIDATION EQUATIONS

Consider a porous body of volume V and surface area S whose pores are filled

with water. During the consolidation process, the distribution of the stresses

within the body is governed by the principles of effective stress, equilibrium, and

continuity of flow. These principles are all invoked to obtain the governing finite

element equations.

The effective stress principle assumes that the total stresses, σ, comprise the sum

of the effective stresses, σ′, and the pore water pressure, p, according to

σ= σ′ +m p (5.1)

154Chapter 5

where m= {1, 1, 1, 0, 0, 0}T and tensile stresses and pore pressures are taken as

positive. Differentiating this equation with respect to time gives

σ. = σ. ′ +m p. (5.2)

In geotechnical analysis, it is conventional to decompose the total pore pressure

into a steady state component, ps, and a time-varying excess component, pe,

according to

p= ps+ pe (5.3)

Noting that ps is constant, differentiation of (5.3) with respect to time gives

p. = p. e (5.4)

For applications which involve a horizontal phreatic surface, the steady state pore

pressure component corresponds to a hydrostatic stress distribution. To determine

the governing relations for finite element analysis of consolidation, the effective

stresses and pore pressures are treated separately and the primary nodal variables

are the displacement rates (velocities) and the pore pressure rates. Because of

the equivalence established in equation (5.4), there is little to choose between

working with total pore pressure rates or excess pore pressure rates. In this Thesis,

the formulation will be developed in terms of total pore pressure rates.

As described in Chapter 2, the method of weighted residuals can be used to

express the three dimensional equations of equilibrium in the weak form

V

(∇w)Tσ dV−V

wTbdV−S

wT tdS= 0 (5.5)

where ∇ denotes the differential operator defined by equation (2.15),

w= wx , wy , wzTis a vector of arbitrary weighting functions in the three

coordinate directions, σ= σx , σy , σz , τxy , τxz , τyzTare the total stresses within

the soil mass, b= bx , by , bzTare the applied body forces, and t= tx , ty , tz

Tis

a vector of external surface tractions acting over the boundary surface S.

155Chapter 5

Differentiating (5.5) with respect to time and substituting (5.2) furnishes the

equilibrium equations in rate form according to

V

(∇w)T (σ. ′ +mp. )dV−V

wTb.dV−

S

wT t.dS= 0 (5.6)

Following the procedure described in Chapter 2, a finite element solution to (5.6)

can be obtained by using shape functions to describe the variation of the velocities

and pore pressure rates throughout each element. For an element with n nodes,

the velocity field at any internal point is assumed to be of the form

d.= Nuu

.

where d.= u. , v. ,w. is a velocity vector with components u. , v. and w. in each

coordinate direction, Nu is a matrix of shape functions given by

Nu=⎪⎪⎡

Nu100

0Nu10

00Nu1

Nu200

0Nu20

00Nu2

...

...

...

Nun00

0Nun0

00Nun⎪⎪⎤

and u. = u. 1, v. 1, w. 1, u. 2, v. 2, w. 2, ... u. n, v. n , w. nTis a vector of element nodal

velocities.

Similarly, the field of pore pressure rates for an element with pore pressure

freedoms at m nodes is assumed to be of the form

p. = Npp. (5.7)

where

Np= Np1 Np2 ... Npm (5.8)

and

p. = p. 1, p.2, ... p

.mT

156Chapter 5

are, respectively, a matrix of shape functions and a vector of nodal pore pressure

rates. Note that, for generality, different sets of shape functions may be used to

describe the variation of the velocities and the pore pressure rates. This implies

that the nodes in the finite element mesh may have varying degrees of freedom,

with some being associated with velocities, some being associated with pore

pressure rates, and some being associated with both. In order for the pore

pressure rates to be consistent with the stress rates, it is usual to choose the

polynomial describing the pore pressure rates to be one order lower than the

polynomial describing the velocities. As discussed by a number of researchers,

including Yokoo, Yamagata and Nagaoka (1971a) and Sandhu, Lui and Singh

(1977), this approach leads to less accurate estimates of the settlements but much

smaller oscillations in the pore pressures.

Having defined the functional form of the velocities and pore pressure rates, the

Galerkin weighting functions may be chosen as

w= δd.= Nu δu

. (5.9)

where δu. is a vector of arbitrary nodal velocities for an element. Substituting

these weighting functions (5.9) into (5.6), integrating over the element volume, V e,

and surface area, Se, and collecting terms furnishes

δu. T⎪⎧⎩Ve

∇NuTσ. ′ dV+

Ve

∇NuTm p. dV−

SeNTu t

.dS−

VeNTu b

.dV⎪⎫⎭= 0

Since the velocities δu. are arbitrary, it follows that

VeBTuσ

. ′ dV+VeBTum p

. dV−SeNTu t

.dS−

VeNTu b

.dV= 0 (5.10)

where Bu is the strain rate-velocity matrix

157Chapter 5

Bu= ∇N=

⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪

∂Nu1∂x

0

0

∂Nu1∂y∂Nu1∂z

0

0

∂Nu1∂y

0

∂Nu1∂x

0

∂Nu1∂z

0

0

∂Nu1∂z

0

∂Nu1∂x∂Nu1∂y

...

...

...

...

...

...

∂Nun∂x

0

0

∂Nun∂y∂Nun∂z

0

0

∂Nun∂y

0

∂Nun∂x

0

∂Nun∂z

0

0

∂Nun∂z

0

∂Nun∂x∂Nun∂y

⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪

and the strain rates at any internal point are

ε. = ∇d

.= ∇Nu u

. = Buu. (5.11)

In the theory of elastoplasticity, the effective stress rates are assumed to be related

to the strain rates via the constitutive law

σ. ′ = Dep ε

.

where Dep is the elastoplastic constitutive matrix derived in Section 2.2.

Combining this equation with (5.11) permits the effective stresses to be expressed

in terms of the element nodal velocities according to

σ. ′ = Dep Buu

. (5.12)

Substituting (5.7) and (5.12) into (5.10) gives a weak statement of the conditions

of equilibrium for an element in rate form. These equations may be written as

kepu. + l p. = f

. ext (5.13)

where

kep= VeBTu Dep Bu dV (5.14)

l= VeBTu m Np dV (5.15)

158Chapter 5

are the elemental elastoplastic stiffness and coupling matrices and

f. ext=

VeNTu b

.dV+

SeNTu t

.dS (5.16)

is the elemental vector of external force rates.

In order to complete the mathematical description of the consolidation process,

it is necessary to consider the continuity of flow for an element of soil. Assuming

that the pore water and soil grains are incompressible, continuity of flow demands

that the rate at which water is drained from the soil skeleton must be equal to the

rate of volume decrease of the soil mass. This condition may be expressed

mathematically as

∂vx∂x +

∂vy∂y +

∂vz∂z =− (ε

.x+ ε

.y+ ε

.z )

or

div v+mT ε. = 0 (5.17)

where v= vx , vy , vzT denotes a vector of superficial (or bulk) fluid velocities and

ε. = ε. x , ε. y , ε. z , γ. xy , γ. xz , γ. yzT denotes a vector of strain rates for the soil. Thecontinuity equation may be expressed in terms of the pore water pressure using

Darcy’s law which, assuming a horizontal phreatic surface, states that the fluid

velocities are given by

vx=kxγw∂p∂x =

kxγw∂pe∂x

vy=kyγw∂p∂y− γw = kyγw ∂pe∂y

vz=kzγw∂p∂z =

kzγw∂pe∂z

where γw is the unit weight of water and kx, ky, kz are soil permeabilities in the

three coordinate directions. Note that, because compressive pore pressures are

159Chapter 5

taken as negative, the sign convention in the above equations is different to that

normally used in soil mechanics. Darcy’s law may also be written in the more

compact matrix form

v= kγw∇p− bw (5.18)

where bw= 0, γw , 0Tis a body force vector and k is a matrix of permeabilities

of the form

k=⎪⎡

kx00

0ky0

00kz⎪⎤

Inserting (5.18) into (5.17) gives the continuity equation in terms of the pore

pressures according to

div kγw ∇p− bw +mT ε. = div kγw ∇pe +mT ε. = 0

The weak form of this equation is obtained by applying the method of weighted

residuals with a Galerkin weight function, ω, according to

V

ωdiv kγw ∇p− bwdV+V

ωmT ε. dV= 0

Integrating the first term using the Green-Gauss theorem this becomes

S

ω ∇p− bwT kγw n dS−

V

(∇ω)T kγw∇p− bw dV+

V

ωmTε. dV= 0

or

V

(∇ω)T kγw ∇pdV−V

ωmTε. dV−S

ω qdS−V

(∇ω)T kγw bw dV= 0 (5.19)

where

160Chapter 5

q= ∇p− bwT kγwn= v

Tn

is a prescribed outward flow per unit area and n= nx , ny , nz Tis a vector of

direction cosines for the unit normal to S. To obtain the finite element counterpart

of (5.19), the pore pressure field for an element with m nodal pore pressures is

assumed to be of the form

p= Npp

where Np is given by (5.8) and

p= p1, p2, ... pmT

is a vector of nodal pore pressures. The Galerkin weight function is

ω= δp= Np δp (5.20)

where δp is an arbitrary set of nodal pore pressures for an element and

∇ω= ∇Np δp= Bp δp (5.21)

with

Bp = ∇Np=

⎪⎪⎪⎪⎪⎪⎪

∂Np1∂x∂Np1∂y∂Np1∂z

∂Np2∂x∂Np2∂y∂Np2∂z

...

...

...

∂Npm∂x∂Npm∂y∂Npm∂z

⎪⎪⎪⎪⎪⎪⎪

(5.22)

Substituting (5.11), (5.20), (5.21) and (5.22) in (5.19) and rearranging gives

δpT⎪⎧⎩VeNTp mTBudV u

. −VeBTp kγw Bp dV p−

SeNTp qdS−

VeBTp kγw bw dV⎪

⎫⎭= 0

where the integrations are now done over each element. As δpT is arbitrary, this

implies that

161Chapter 5

VeNTp mTBudV u

. −VeBTp kγw Bp dV p−

SeNTp qdS−

VeBTp kγw bw dV= 0

This equation expresses the continuity equation in weak form for each element

and may be written in the compact form

lTu. + hp= q (5.23)

where the coupling matrix l is given by (5.15), the flow matrix h is defined as

h=−VeBTp kγw Bp dV (5.24)

and

q= SeNTp qdS+

VeBTp kγw bw dV (5.25)

is a fluid supply vector. Equations (5.13) and (5.23), together with (5.14), (5.15),

(5.16), (5.24) and (5.25), define the governing relations for Biot consolidation at

the element level. Assembling the element matrices in the usual way produces

a global system of equations of the form

KepLTL

0U.P. +⎪⎡⎣

0

0

0

H⎪⎤⎦⎨⎧⎩U

P⎬⎫⎭= F

. ext

Q (5.26)

where

Kep= elements

kep= elements

VeBTu DepBudV

L= elements

l= elements

VeBTu mNp dV

H= elements

h=− elements

VeBTp kγw Bp dV

162Chapter 5

are the global elastoplastic stiffness, coupling and flow matrices and

F. ext=

elements

f. ext =

elements

VeNTu b

.dV+

elements

SeNTu t

.dS

Q= elements

q = elements

SeNTp qdS+

elements

VeBTp kγw bw dV

are the global external force rate and fluid supply vectors.

5.4 SOLUTION OF ELASTIC CONSOLIDATIONEQUATIONS

For the analysis of elastic solids with constant permeabilities, the relations (5.26)

constitute a system of linear first order differential equations of the form

CX.+ KX= F(t) (5.27)

where

C=⎪⎡⎣Ke

LT

L

0⎪⎤⎦

K=⎪⎡⎣0

0

0

H⎪⎤⎦

(5.28)

are matrices of constants, F(t) is a time dependent forcing function defined by

F(t)=⎨⎧⎩F. ext

Q⎬⎫⎭

(5.29)

and X= {U,P}T with X.= U. ,P. T. This type of system occurs in many areas of

engineering and science and has been widely studied. A very comprehensive

summary of the stability and accuracy of various solution strategies for solving

(5.27) may be found in Wood (1990). Since these methods are central to the work

undertaken in this Thesis, some of the more important techniques will now be

discussed.

163Chapter 5

5.4.1 Single-Step Schemes

The simplest strategy for solving (5.27) is commonly known as the θ-method. For

the nth time step ranging from tn−1 to tn= tn−1+ h, this algorithm may be

expressed in the form

[C+ θhK]Xn= C− (1− θ)hKXn−1+ h(1− θ) Fn+ θFn−1 (5.30)

where θ is an integration parameter in the interval 0≤ θ≤ 1, the subscripts n

and n− 1 denote, respectively, quantities evaluated at the start and end of the

step, and all values except Xn are known. The process assumes that X0 at time

t0 is known. For the case of an elastic soil with constant permeabilities, the

matrices C and K are independent of X and (5.30) defines a system of linear

equations which can be solved for Xn. Note that if h is kept constant, then the

matrix [C+ θhK] needs to be factorised only once to obtain the solution for all

time steps. Because θ may assume a range of values, equation (5.30) can be used

to generate a family of single-stage single-step schemes that all march the solution

X forward in time without the need for information from previous steps.

The θ-method is at least first order accurate and, provided θ≥ 0.5, is

unconditionally stable. Unconditionally stability is an essential characteristic for

an efficient consolidation scheme since it is often necessary to integrate over very

long time periods using large time steps. For the special case of θ= 0.5, the

θ-method is second order accurate and corresponds to the ubiquitous

Crank-Nicolson scheme. Although appealing because of its high accuracy, the

Crank-Nicolson method may generate spurious oscillations in the solution,

especially if there are abrupt changes in the forcing function, and often requires

special smoothing procedures such as those advocated by Wood and Lewis (1975)

and Wood (1977). Choosing a value of θ= 1 gives the well known backward

Euler scheme which is first order order accurate, unconditionally stable and

oscillation free (Wood, 1990). The last of these characteristics has led to the

164Chapter 5

Backward Euler method being widely used in finite element consolidation studies,

even though it is less accurate than the Crank-Nicolson scheme.

As an alternative to (5.30), the θ-method can be expressed in the more compact

form

[C+ θhK]V= (1− θ)Fn−1+ θFn−KXn−1 (5.31)

where

V= (Xn− Xn−1)∕h

is an average estimate of X.over the time step h and X is updated according to

Xn= Xn−1+ hV (5.32)

In the finite element literature, this form of the θ-method is often referred to as

the SS11 procedure, where the terminology SSpj stands for a Single-Step

algorithm which uses an approximation of degree p to solve a differential equation

of order j. These relations will be used to develop an automatic time stepping

scheme in later Sections of the Thesis.

5.4.2 Two-Step Schemes

A family of unconditionally stable second order accurate methods for solving

(5.27) can be derived by considering quantities generated over two consecutive

time steps. As discussed in Wood (1990), these schemes are of the general form

θ1hC+ 12 θ2h 2KXn= (1− θ1)hC+ 12 (2θ1− θ2− 1)h2KXn−2

+ (2θ1− 1)hC+ 12 (2θ2− 2θ1− 1)h2KXn−1

+12 (θ2− 2θ1+ 1) Fn−2+ 12 (2θ1− 2θ2+ 1) Fn−1+ 12 θ2Fn (5.33)

where θ1 and θ2 are specified, the subscripts n− 2 denote quantities evaluated

at tn−2= tn−1− h, and all values except Xn are assumed known. In order to

165Chapter 5

obtain unconditional stability, the integration parameters must be chosen such that

θ2> θ1≥ 0.5. For the special case of θ1− θ2= 0.5 and θ1> 0.5, the two-step

scheme is third order accurate but only conditionally stable. When C and K are

independent of X, equation (5.33) defines a system of linear equations which can

be solved to obtain Xn, the solution at the end of the current time step. As in

the case of the single-step schemes discussed previously, the two-step method

requires only one matrix factorisation if h is held constant throughout the solution

process.

One possible choice for the integration parameters is θ1= 0.5 and θ2= 2∕3.

These values, first suggested by Lees (1966), result in an algorithm that does not

require iteration for certain types of nonlinear problems, and have been used by

Lewis and Schrefler (1987) to model consolidation. According to the analysis

presented in Wood (1990), however, this method is subject to troublesome

oscillations and is not recommended.

Another possible alternative is the so-called backward difference scheme, which

is obtained by setting θ1= 1.5 and θ2= 2. Because of its ability to damp out

spurious high frequency components, this method is advocated by Richtmyer and

Morton (1967) and has also been used by Wood and Lewis (1975). In a detailed

study of various noise control strategies for the heat conduction equation, the

latter Authors conclude that, although the backward difference method does

indeed perform satisfactorily, an averaged form of the Crank-Nicolson scheme is

better still.

Other options that have been proposed for the integration parameters in (5.33)

include those of Liniger (1969), who suggests setting θ1= 1.218 and θ2= 1.292,

and Zlamal (1977), who recommends values of θ1= 5∕6 and θ2= 8∕9.

Unfortunately, no data is available to compare the performance of these schemes

in solving consolidation problems.

166Chapter 5

Although the two-step schemes discussed above have the advantage of being

second order accurate and unconditionally stable, they require special procedures

to generate the starting values X0 and X1. A further disadvantage is that the size

of the time step must be fixed for consecutive increments, thus making it difficult

to adjust the time step dynamically as the solution proceeds.

5.4.3 Two-Stage Single-Step Schemes

A general class of multistage single-step methods for solving systems of first order

differential equations, such as (5.27), has been proposed by Zienkiewicz et al

(1984). Of particular interest here is the two-stage single step scheme defined by

θ1hC+ 12 θ2h2KA= (1− θ1)Fn−1+ θ1Fn− CX.

n−1

−KXn−1+ θ1hX.

n−1 (5.34)

where

A= (X.n− X

.

n−1)∕h

is an average estimate of X..over the time step h and the updates are

Xn= Xn−1+ hX.

n−1+12 h2A (5.35)

X.n= X

.

n−1+ hA (5.36)

This scheme, which is commonly known as the SS21 algorithm, has the same

accuracy and stability properties as the two-step procedure described in the

previous Section. During a given time step, Xn and X.n are updated using (5.35)

and (5.36) after first solving for A in (5.34). Unlike the two-step method, the SS21

procedure advances the solution for both X and X.and only uses values from the

current time step. These characteristics explain why the algorithm is termed a

two-stage single-step method. The chief advantage of this type of scheme is that

167Chapter 5

the step size may be adjusted easily as the integration proceeds. The price of this

flexibility is the need to compute and store X.for each time step.

More recently, Thomas and Gladwell (1988a) have proposed a generalised form

of the SS21 algorithm. Their procedure uses three integration parameters instead

of two and may be written as

φ2hC+ φ3h2KA= Ft n−1+ φ1h − CX.

n−1−KXn−1+ φ1hX.

n−1 (5.37)

where A has the same meaning as before and the updates are again given by (5.35)

and (5.36). This scheme is second order accurate and unconditionally stable

provided 2φ3> φ1≥ 0.5 and φ2≥ 0.5. Using the fact that

Ft n−1+ φ1h = (1− φ1)Fn−1+ φ1Fn+O(h2) (5.38)

a comparison of equations (5.34) with (5.37) reveals that the SS21 algorithm of

Zienkiewicz et al (1984) is a special case of the Thomas and Gladwell algorithm

with θ1= φ1= φ2 and θ2= 2φ3. For cases where θ1= φ1= 1, this equivalence

holds without the need to approximate the forcing function by (5.38). Because of

the additional freedom that is introduced by having three integration parameters,

the Thomas and Gladwell method is ideally suited to the design of automatic

integration schemes with embedded error estimators. Indeed, it will be used for

this purpose in the next Section.

5.5 AUTOMATIC TIME STEPPING SCHEME FORELASTIC CONSOLIDATION

This Section describes an adaptive integration scheme that automatically adjusts

the time step according to a specified error criterion. For ease of use, the

algorithm assumes that a number of (coarse) time increments are defined and

automatically breaks these up into a number of smaller subincrements if necessary.

The coarse time step is assumed to start at t0 and end at t0 + Δt, and is thus of

size Δt. The nth time subincrement is assumed to be of size h, and ranges from

tn−1 to tn= tn−1+ h. This arrangement is shown schematically in Figure 5.1

168Chapter 5

t0 t0+ Δttn−1 tn−1+ h

Xn−1

X.

n−1

X.n

Xn

h

X.

X

X0

X.

0

time t

start of coarsetime step

end of coarsetime step

Figure 5.2 Coarse and subincremental time steps.

Δt

start ofsubstep

end ofsubstep

The adaptive procedure uses two integration methods, of different accuracy, to

provide an estimate of the local truncation error in the displacements and pore

pressures and is thus based on the same idea that has been exploited in Chapters

3 and 4. In this case, the SS11 version of the θ-method and the Thomas and

Gladwell method are employed, respectively, to generate first and second order

accurate solutions for the error control.

5.5.1 Theory

To derive an efficient solution scheme with an inbuilt error estimator, the

integration parameters for the SS11 and Thomas and Gladwell algorithms are

selected so that only one matrix factorisation is needed for each time step. The

constraints that this imposes on the integration parameters may be seen by

rewriting equations (5.31) and (5.37) as

169Chapter 5

CV+KX~ n−1+ θhV = (1− θ)Fn−1+ θFn (5.39)

CX. n−1+ φ2hA +KXn−1+ φ1hX.

n−1+ φ3h2A = Ftn−1+ φ1h (5.40)

where X~n−1 now denotes the estimate of X computed from the first order scheme.

Comparing (5.39) and (5.40), the SS11 and Thomas and Gladwell schemes give

rise to an identical system of equations if

V= X.

n−1+ φ2hA (5.41)

X~n−1+ θhV= Xn−1+ φ1hX

.

n−1+ φ3h2A (5.42)

and

(1− θ)Fn−1+ θFn = F t+ φ1h (5.43)

For the purposes of controlling the local error, it is assumed that the X values for

both methods are identical at the start of each time subincrement so that

X~n−1= Xn−1 (5.44)

in (5.39). This implies that the second order (rather than the first order) solution

is the one that is propagated throughout the analysis. It is also possible to design

and algorithm which marches forward with the first order solution, but this will

not be discussed here. Although it is natural to propagate the first order solution,

since this is the one the error is being estimated for, most modern algorithms for

solving systems of ordinary differential equations propagate the second order

solution. This approach compensates for the fact that the error measure is only

local, and leads to better control of the global temporal discretisation error which

accumulates over many time steps.

Neglecting, for the moment, the forcing function terms, the required constraints

on the integration parameters are obtained by substituting (5.41) and (5.44) in

(5.42) to give

170Chapter 5

θ= φ1= φ3∕φ2 (5.45)

Because the error estimator assumes that the SS11 method is first order accurate,

the Crank-Nicolson special case must be excluded so that

θ≠ 0.5 (5.46)

The constraints (5.45) and (5.46) ensure that first and second order accurate

solutions can be obtained, respectively, from the SS11 and Thomas and Gladwell

methods with only a single matrix factorisation for each time subincrement.

Combining these with the unconditional stability requirements

θ≥ 0.5

2φ3> φ1≥ 0.5

φ2≥ 0.5

gives the final set of constraints as

2φ3> θ= φ1= φ3∕φ2> 0.5 (5.47)

From (5.43), the forcing functions for the two schemes are identical if θ= φ1= 1.

For other choices of θ and φ1, it is necessary to use the second order

approximation for Ftn−1+ φ1h given in equation (5.38). With this substitution,

equation (5.43) becomes

(1− θ )Fn−1+ θFn = (1− φ1)Fn−1+ φ1Fn

and is automatically satisfied by the constraint (5.47).

During a typical time subincrement h, first and second order accurate estimates

for Xn may be found using the SS11 method of equation (5.39). This equation

is solved with a second order accurate starting value Xn−1 to give

V= [C+ θhK]−1(1− θ)Fn−1+ θFn−KXn−1 (5.48)

171Chapter 5

with the update of (5.32) being modified to

X~n= Xn−1+ hV (5.49)

A is then found from equation (5.41)

A=V− X

.

n−1φ2h

(5.50)

where X.

n−1 is assumed known and the second order updates (5.35) and (5.36) are

Xn= Xn−1+ hX.

n−1+12 h2A (5.51)

X.n= X

.

n−1+ hA (5.52)

Since the local truncation errors in the updates (5.51) and (5.49) are, respectively,

O(h3) and O(h2), the lower order estimate may be subtracted from the higher

order estimate to give the local truncation error measure

En= Xn− X~n= hX

.

n−1+12h2A− hV

Substituting equations (5.50) and (5.52), this estimator may be expressed in the

forms

En= h1− 12φ2X. n−1+ 12φ2− 1V

or

En= hφ2− 12X. n−1+12− φ2X. n

Note that the undesirable special case of φ2= 0.5, which gives a zero estimate

of the local error regardless of h, is automatically excluded by the constraints

(5.47). For the purposes of error control, En may be replaced by the more useful

dimensionless relative error measure

172Chapter 5

Rn= max Eun Un ,EpnPn (5.53)

where

Eun= hφ2− 12U. n−1+12− φ2U. n (5.54)

Epn= hφ2− 12P. n−1+12− φ2P. nand Un and Pn are the displacement and pore pressure components of Xn and U

.n

and P.n are the velocity and pore pressure rate components of X

.n. For cases where

a weightless soil is assumed, the total pore pressures correspond to the excess pore

pressures and Pn will approach zero in later stages of consolidation. To avoid

ill-conditioning of the relative error estimator in this situation, it is preferable to

measure the error in the displacements only and replace (5.53) by

Run= max Eun Un (5.55)

This approach, which is used throughout this Thesis, is sufficiently accurate for

practical computations and removes the need to employ a combination of absolute

and relative error indicators.

Once Run has been computed for the current time substep, the procedure for

adjusting the next time substep is very similar to that outlined in Section 4.4 of

Chapter 4. The current time subincrement is accepted if Run is less than some

specified tolerance on the local truncation error, DTOL, and rejected otherwise.

In either case, the size of the next time step hn+1 is found from

hn+1= qhn

where q is a factor which is chosen to limit the predicted truncation error. Since

the truncation error for the next time subincrement, Run+1, is approximately

related to Run by

173Chapter 5

Run+1≈ q2Run

the required factor q is found by insisting that Run+1≤ DTOL to give

q≤ DTOL∕Run

As in Section 4.4, q is chosen conservatively to minimise the number of rejected

time subincrements. according to

q= 0.9 DTOL∕Run (5.56)

with the additional constraint that

0.1≤ q≤ 2 (5.57)

The safety factor coefficient of 0.9 is greater than the value of 0.7 used in the

elastoplastic load substepping scheme because the nonlinearities for elastic

consolidation are less extreme. For the same reason, the maximum value for q,

which determines the maximum size of consecutive time substeps, has been

increased from 1.1 to 2. The value of the safety factor in (5.56), and the limits

in (5.57), were determined by numerical experiments on a wide variety of

examples and ensure that most of the substeps are successful without making the

step selection mechanism too conservative. For the reasons discussed in Section

4.4, the step size is again prohibited from growing immediately after a failed time

subincrement.

Assuming initial values for h and X0, with the latter typically zero, the integration

scheme is started by solving (5.48) for V. For the first coarse time increment, h

is typically set to Δt, but in subsequent coarse time increments it may be initialised

to the value that gave the last successful subincrement. In order to compute the

second order update for X using equation (5.51), a starting value for X.at t= 0

is needed. Assuming that the matrix C has an inverse, X.

0 may be found by solving

the governing differential equation (5.27) at t= 0 according to

174Chapter 5

X.

0= [C]–1F0−KX0 (5.58)

This type of procedure is valid for elements with a pore pressure expansion which

is one order lower than the displacement expansion. For elements where the

expansions are the same, C does not have an inverse and equation (5.58) cannot

be used. The simplest alternative in this case is to take a very small time step

δh and apply equations (5.48), (5.50) and (5.52) with X0= X.

0= 0 to give

V= [C+ θδhK]−1(1− θ)F0++ θF0+

X.

0+ = V∕φ2

where the subscript 0+ indicates quantities evaluated at time t= δh. For a

sufficiently small value of δh, X.

0+ may be used as the initial value for X.at time

t= 0.

At the start of a typical substep, V is found from (5.48) and the updates for X

and X.are computed using (5.50)---(5.52). The relative error Run is then

determined using equations (5.54) and (5.55). If this error is less than or equal

to the specified tolerance DTOL, then the current time subincrement is accepted

and the step size for the next time subincrement is found using (5.56) and (5.57).

If Run exceeds DTOL, then the solution is rejected and equations (5.56) and (5.57)

are used to predict a smaller step size that will hopefully satisfy the constraint on

the local error. In this case, the stage is repeated and, if necessary, the step size

is reduced further until a successful time substep size is obtained.

The discussion has, so far, assumed that it is convenient to evaluate the external

force rate, F. ext= dFext∕dt, analytically in the overall forcing function defined by

equation (5.29). For cases where this is not so, this derivative can be approximated

using discrete values of the external force vector. Examples of four useful

approximations are

F. extn = (Fextn − F

extn−1)∕h+ 0(h) (5.59)

175Chapter 5

F. extn−1= (F

extn − F

extn−1)∕h+ 0(h) (5.60)

F. extn = (3Fextn − 4F

extn−1∕2+ F

extn−1)∕h+ 0(h

2) (5.61)

F. extn−1= (− 3F

extn−1+ 4Fextn−1∕2− F

extn )∕h+ 0(h

2) (5.62)

where h is the current time step and the subscripts n− 1, n− 1∕2, and n denote

values computed at the times tn−1, tn−1∕2= tn−1+ h∕2 and tn= tn−1+ h.

Since the adaptive integration scheme described here is second order accurate,

equations (5.61) and (5.62) should be employed when the variation of the external

forcing function is nonlinear with time. For problems where the external loading

is piecewise linear with time, which covers most practical situations, the

approximations (5.59) and (5.60) are exact and therefore appropriate. These

approximations are used throughout this Thesis.

5.5.2 Scaling of Linear Equations

The automatic solution scheme described in the previous Section requires the

solution of the linear system of equations (5.48). These may be written in the

simple form

⎪⎡⎣Ke

LT

L

θhH⎪⎤⎦⎨⎧⎩U.

P.⎬⎫⎭= ⎨⎧⎩

Ru

Rp⎬⎫⎭

(5.63)

where U.and P

.are, respectively, average velocities and pore pressure rates over

a time step h, and Ru and Rp are arbitrary vectors. For small time steps h, this

system may become ill-conditioned as the diagonal terms in K can be many orders

of magnitude greater than the terms in θhH. The effects of ill-conditioning were

first noted by Ghaboussi and Wilson (1973), who developed a criterion for

selecting a minimum value of h. This criterion is valid for consolidation of an

elastic soil which is isotropic and homogeneous. Rather than limiting the size of

the time step, which can introduce another source of error due to the time

176Chapter 5

dependence of the governing equations, it is possible to scale the various terms

in (5.63) so that ill-conditioning is avoided. This approach, suggested by Reed

(1984) and used in Lewis and Schrefler (1987), has proven successful and is

adopted in this Thesis. The scaling preserves any symmetry of the original linear

equations and takes the form

⎪⎡

K

sLT

sL

s2θhH⎪⎤

⎦⎪⎪⎨⎧

U.

1s P.⎪⎪⎬⎫

⎭= ⎪⎪⎨⎧

Ru

sRp⎪⎪⎬⎫

where s is a scalar parameter which is chosen so as to roughly equate the size of

the diagonal terms in K and s2θhH. Ignoring the effects of the element geometry,

the diagonal terms of K are proportional to the terms of the elastic stress strain

matrix D, and hence are of the order of Young’s modulus E. Similarly, the terms

of the matrix H are approximately proportional to k∕γw, where k is a

representative value of the permeability. Equating these contributions and noting

that θ is of order 1, a suitable value for the scaling parameter can be estimated

as

s=Eγwhk

For problems involving a homogeneous isotropic soil, the choice of E and k in the

above equation is straightforward. In other situations, however, it is necessary to

choose representative values of these material parameters. One such scaling

strategy can be found in the code of Lewis and Schrefler (1987).

5.5.3 Implementation

In deciding upon an implementation of the automatic integration scheme

described in Section 5.5.1, it is necessary to choose specific values of the

integration parameters θ, φ1, φ2 and φ3 which satisfy the constraints (5.47). After

a series of numerical experiments covering a wide range of problems, these were

set to

177Chapter 5

θ= φ1= φ2= φ3= 1 (5.64)

The advantages of this choice are as follows:

i) Setting θ= 1 implies that the first order method corresponds to the backward

Euler scheme. This procedure is known to damp out unwanted oscillations

quickly and thus provides a reliable first order solution for estimating the local

truncation error.

ii) Selecting φ1= θ= 1 means that the forcing functions for the first order and

second order schemes are identical without any need to make the

approximation shown in equation (5.38). Moreover, there is no need to

evaluate the forcing function outside of the current time step.

iii) Setting φ2= 1 means that the vector V, computed from equation (5.48),

corresponds automatically to X.n, the value of X

.at the end of the current step

for the second order scheme. This feature, which can be seen by comparing

equations (5.41) and (5.52), results in a compact algorithm.

Other settings for these parameters are of course possible, and may lead to a

scheme with improved performance for certain cases. The values in (5.64),

however, will be shown in the next Chapter to give excellent results for a broad

range of practical problems.

As mentioned previously, the implementation of the adaptive integration

algorithm assumes that a series of coarse time increments have been specified.

These coarse increments are, if necessary, subincremented automatically to satisfy

a tolerance on the local truncation error.

The automatic time stepping algorithm for elastic consolidation may be

summarised as follows.

1. Enter with the time at the start of the coarse increment t0, the current

displacements and pore pressures Xt0, their corresponding derivatives X.

t0,

the coarse time increment Δt, the last successful time substep hlast, the

178Chapter 5

current effective stresses at each integration point σ′t0 , and the specified

displacement error tolerance DTOL. For the first coarse time step, set

hlast= Δt.

2. Set t= t0 and h= min {hlast ,Δt }

3. While t< t0+ Δt do steps 4 to 9.

4. Compute X.

t+h according to

X.

t+h= [C+ hK]–1Ft+h−KXt

where

Ft+h=⎨⎧⎩F. extt+h

Qt+h⎬⎫⎭= ⎨⎧⎩(Fextt+h− F

extt )∕h

Q t+h⎬⎫⎭

5. Estimate the local truncation error in the displacements for the current

subincrement using

Eut+h=12hU. t− U

.

t+h

where U.denotes the velocity component of X

..

6. Update the displacements and pore water pressures and hold them in

temporary storage according to

Xt+h= Xt+h2X. t+ X

.

t+h

7. Estimate the relative error for current subincrement using

Rut+h= maxEPS , Eut+h ∕ Ut+h

where Ut+h is the displacement component of Xt+h and EPS is a machine

constant.

8. If Rut+h> DTOL then go to step 9. Else this step is successful so update

displacements, pore pressures and integration point effective stresses

according to

179Chapter 5

Xt+h= Xt+h

σ′t+h = DBut+h

If the previous subincrement was successful, estimate a new subincrement

size by computing

q= min0.9 DTOL∕Rut+h , 2

and setting

h← qh

Store successful subincrement size hlast= h and, before returning to step

4, update time and check that integration does not proceed beyond t0+ Δt

by setting

t← t+ h

h← min h , t0+ Δt− t

9. This subincrement has failed, so estimate smaller time substep by

computing

q= max0.9 DTOL∕Rut+h , 0.1

and then setting

h← qh

before returning to step 4.

10. Exit with displacements and pore pressures, Xt0+Δt, their corresponding

rates, X.

t0+Δt, and integration point effective stresses, σ′t0+Δt , at end of

coarse time increment.

Any suitable norms may be used to estimate the error in steps 5 and 7. The max

norm is used throughout this Thesis. In step 7, EPS again represents the smallest

relative error that can be computed on the host machine, and is typically set to

around 10---16 for double precision arithmetic on a 32-bit architecture. Typical

180Chapter 5

values for the tolerance on the truncation error in the displacements, DTOL, are

in the range 10---2 to 10---4, with a value of 10---3 being adequate for most practical

computations.

Note that, in step 4 of the above algorithm, it it is generally necessary to form and

factorise the matrix [C+ hK] afresh for each subincrement in order to find X.

t+h.

This is because h may vary throughout the integration. Although the factorisation

cannot be avoided, the cost of the formation step can be reduced significantly by

computing

C=⎪⎡⎣Ke

LT

L

0⎪⎤⎦

only once, at the start of the analysis, and storing it on disk. After loading C into

memory, the contributions from

hK=⎪⎡⎣0

0

0

hH⎪⎤⎦

for each subincrement may then be added, element by element, to form [C+ hK].

Depending on the type of element used, further economies may be realised by

storing the element flow matrices, h, on disk to minimise the cost of recomputing

H.

5.6 AUTOMATIC TIME STEPPING SCHEME FORELASTOPLASTIC CONSOLIDATION

For the analysis of elastoplastic soils, the governing relations given by equation

(5.26) can be represented as a system of nonlinear equations of the form

R(X,X.)= F(t)− Cep(X)X

.−KX= 0 (5.65)

where

181Chapter 5

Cep(X)=⎪⎡⎣Kep(X)

LT

L

0⎪⎤⎦

K=⎪⎡⎣0

0

0

H⎪⎤⎦

F(t)=⎨⎧⎩F. ext

Q⎬⎫⎭

(5.66)

and X= {U,P}T with X.= U. ,P. T. It is again assumed that the permeabilities

are independent of time, so that the matrix K is constant. Due to the difficulty

in measuring the permeability of soil accurately, this is the usual assumption

adopted made in practice. The algorithm described in the following Sections can

be extended to deal with cases where K is time dependent, but this generalisation

will not be covered here. The major complication introduced by elastoplasticity

is that the matrix Cep is dependent on the current stress state (and hence the

displacements).

5.6.1 Theory

The theory for developing an automatic time stepping scheme for elastoplastic

soils is essentially the same as that used for the elastic case in Section 5.5.1. For

a given time step, the local truncation error is again measured by taking the

difference between a pair of first and second order solutions which, as before, are

provided by the SS11 and Thomas and Gladwell methods. The key change from

the elastic scheme is that it is now necessary to solve a system of nonlinear

equations in order to update the displacements and pore pressures. It is once

again assumed that a series of coarse time steps are defined which, if required,

are subdivided into substeps to keep the local truncation error below a specified

tolerance.

Applying the SS11 and Thomas and Gladwell algorithms to the system (5.65)

yields the pair of nonlinear equations

R1(V)= (1− θ)Fn−1+ θFn− Cep(X~n−1+ θhV)V−KX~ n−1+ θhV= 0 (5.67)

R2(A)= Ftn−1+ φ1h − Cep(Xn−1+ θhV)X.

n−1+ φ2hA

−KXn−1+ φ1hX.

n−1+ φ3h2A = 0

182Chapter 5

Following the procedure outlined in Section 5.5.1, these equations are identical

if the constraints (5.47) are satisfied and X~n−1 is replaced by the second order

solution Xn−1. Under these conditions, it is necessary to solve only (5.67) for V,

with the initial value Xn−1, in order to march the solution forward for each time

substep. As in the elastic case, the updates for the displacements and pore

pressures, Xn, and their corresponding rates, X.n, are found from equations

(5.50)---(5.52) and the local error estimator is given by (5.55). Thus, dropping the

subscript on R, the system of nonlinear equations to be solved for each time

substep may be written as

R(V)= (1− θ)Fn−1+ θFn− Cep(Xn−1+ θhV)V−KXn−1+ θhV = 0

or

R(V)= (1− θ)Fn−1+ θFn− Cep(X~)V− K X~ = 0 (5.68)

where

X~ = Xn−1+ θhV

and

V= X~.

The solution to the system (5.68) may be found using the Newton-Raphson

algorithm. Letting the superscript i denote iteration number, this scheme takes

the form

V i= V i−1+ δV i

X~ i= Xn−1+ θhV

i

where the iterative update for V i is

δV i= ∂R∂V−1R (V i−1 ) (5.69)

183Chapter 5

and the Jacobian matrix ∂R∕∂V is evaluated at X~ i−1. Suitable values for starting

the iterations may be obtained by setting

V 0= X.

n−1

X~ 0= Xn−1+ θhX

.

n−1

To complete the description of the Newton-Raphson algorithm, it is necessary to

evaluate ∂R∕∂V. Differentiating (5.68) and neglecting second derivatives with

respect to V gives the required Jacobian matrix as

∂R∂V = Cep(X~ )+ θhK (5.70)

Note that if the Kep component of Cep is formed using the Dep defined by (2.10),

then this Jacobian matrix is only an approximation to ∂R∕∂V and the rate of

convergence of the iteration scheme will be linear rather than quadratic. With the

proposed integration method, however, the number of iterations required for a

typical time substep is usually low due to the fact that the error control mechanism

automatically chooses small time steps in the vicinity of highly nonlinear

behaviour. At the cost of introducing complexity, quadratic convergence may be

obtained by using the so-called ‘consistent’ form of Dep introduced by Simo and

Taylor (1985). For cases which are only mildly nonlinear, it is possible to replace

(5.70) by the initial stiffness approximation

∂R∂V ≈ [C+ θhK]

where C is given by (5.28). Although much slower to converge, this approach does

not require a fresh factorisation for each iteration and always has a well

conditioned Jacobian matrix.

A convenient check for terminating the iteration procedure is to test whether the

relative change in the displacement component of X~is less than or equal to a

specified tolerance, ITOL. This can be expressed as

184Chapter 5

θhδU~. iU~ i ≤ ITOL

where U~corresponds to the displacement entries in X

~and U

~.

corresponds to the

velocity entries in V.

Depending on the values of the integration parameter θ, a number of alternative

strategies can be used to evaluate the residual R (V i−1 ) in equation (5.69). The

various options available become evident upon substituting equations (5.66) in

(5.68) to give the expanded form

R(V i−1 )= ⎪⎪⎨⎧

(1− θ)F. extn−1+ θF

. extn

(1− θ)Qn−1+ θQn⎪⎪⎬⎫

⎭−⎪⎪⎡

Kep(X~n)

LT

L

0

⎪⎪⎤

⎦⎪⎪⎨⎧

U~.i−1

P~.i−1⎪⎪⎬⎫

−⎪⎪⎡

0

0

0

H⎪⎪⎤

⎦⎪⎪⎨⎧

U~ i−1

P~ i−1⎪⎪⎬⎫

For values of the integration parameter in the range 0.5< θ< 1, it is easiest to

work with an analytic form for F. ext and perform the remaining matrix-vector

multiplications element by element.

For the backward Euler case where θ= 1, V and X~ contain values at the end of

the current time step and the residual becomes

R(V i−1n )= ⎪⎪⎨⎧

F. extn

Qn⎪⎪⎬⎫

⎭−⎪⎪⎪⎡

Kep(X~n)

LT

L

0

⎪⎪⎪⎤

⎦⎪⎪⎨⎧

U~.i−1n

P~.i−1n⎪⎪⎬⎫

⎭−⎪⎪⎪

0

0

0

H

⎪⎪⎪

⎦⎪⎪⎨⎧

U~ i−1n

P~ i−1n

⎪⎪⎬⎫

Noting that

F. intn =

V

BTσ. n dV= Kep(X~n)U~. i−1n + LP~

.i−1n

185Chapter 5

and using the first order approximations

F. intn = (F intn − F

intn−1)∕h+ 0(h)

F. extn = (F extn − F

extn−1)∕h+ 0(h)

this may written in the alternative form

R(V i−1n )= ⎪⎪⎨⎧

Fextn − Fintnh

Qn⎪⎪⎬⎫

⎭−⎪⎪⎨⎧

Fextn−1− Fintn−1

h

0⎪⎪⎬⎫

⎭−⎪⎪⎡

0

LT

0

H

⎪⎪⎤

⎦⎪⎪⎨⎧

U~.i−1n

P~ i−1n⎪⎪⎬⎫

⎭(5.71)

The first two terms on the right hand side of the above equation arise, respectively,

from the unbalanced forces at the end and the start of the time substep. Provided

the iteration tolerance ITOL is sufficiently small, the latter forces will be negligible

so that

F. unbn−1= (F

extn−1− F

intn−1)∕h≈ 0

Under these circumstances, equilibrium will also be satisfied at the end of the time

substep since the Newton-Raphson iterations enforce the condition

F. unbn = (Fextn − Fintn )∕h≈ 0 (5.72)

If ITOL is less stringent, so that the unbalanced forces at the start of the time

substep are not small, then the solution will tend to drift from equilibrium as it

is marched forward. Since the solution will be more accurate if equilibrium is

obeyed, it is always desirable to enforce the condition (5.72). This implies that

(5.71) may be replaced by

R(V i−1n )= ⎪⎪⎨⎧

Fextn − Fintnh

Qn⎪⎪⎬⎫

⎭−⎪⎪⎡

0

LT

0

H

⎪⎪⎤

⎦⎪⎪⎨⎧

U~.i−1n

P~ i−1n⎪⎪⎬⎫

To complete the residual evaluation for the case in which θ= 1, the matrix

multiplications

186Chapter 5

LT U~.i−1n

and

H P~ i−1n

may be performed element by element.

Once the solution for V has been obtained for each time substep, the error control

and step adjustment mechanism is almost identical to that discussed in Section

5.5.1. The only minor change is that the safety factor in (5.56) is lowered from

0.9 to 0.8 so that

q= 0.8 DTOL∕Run

This means the step control mechanism is slightly more conservative than that for

the elastic case, and allows for the material nonlinearity introduced by

elastoplasticity.

5.6.2 Implementation

The nonlinear automatic time stepping scheme developed in the previous Section

can be implemented in the same manner as the linear elastic scheme described

in Section 5.5.3. It is again assumed that a number of coarse time increments are

defined which, if necessary, are automatically subincremented so that the local

truncation error for each substep does not exceed a prescribed tolerance, DTOL.

Because of the advantages discussed in Section 5.5.3, the integration parameters

are again chosen to be

θ= φ1= φ2= φ3= 1

The automatic time stepping algorithm for elastoplastic consolidation may be

summarised as follows.

1. Enter with the time at the start of the coarse increment t0, the current

displacements and pore pressures Xt0, their corresponding derivatives X.

t0,

187Chapter 5

the coarse time increment Δt, the last successful time substep hlast, the

current stress state at each integration point (σ′t0 , Àt0), the iteration

tolerance ITOL, and the specified displacement error tolerance DTOL. For

the first coarse time step, set hlast= Δt.

2. Set t= t0 and h= min {hlast ,Δt }

3. While t< t0+ Δt do steps 4 to 8

4. Compute X.

t+h and X~

t+h using the Newton---Raphson or initial stiffness

algorithm. If the solution fails to converge, set

h← 0.25h

and try again.

5. Estimate the local truncation error in the displacements for the current

subincrement using

Eut+h=12hU. t− U

.

t+h

where U.denotes the velocity component of X

..

6. Estimate the relative error for current subincrement using

Rut+h= maxEPS , Eut+h ∕ U~ t+h where U

~

t+h is the displacement component of X~

t+h and EPS is a machine

constant.

7. If Rut+h> DTOL then go to step 8. Else this step is successful so compute

new displacements and pore pressures using the second order update

Xt+h= Xt+h2X. t+ X

.

t+h

For each integration point, compute the strains

Δε= Bh2 u. t+ u. t+hand integrate constitutive law to find corresponding increments in stresses,

Δσ′, and hardening parameter ΔÀ. Then update stress state according to

188Chapter 5

σ′t+h = σ′t + Δσ′

Àt+h= Àt+ ΔÀ

If the previous subincrement was successful, estimate a new subincrement

size by computing

q= min0.8 DTOL∕Rut+h , 2

and then setting

h← qh

Store successful subincrement size hlast= h and, before returning to step

4, update time and check that integration does not proceed beyond t0+ Δt

by setting

t← t+ h

h← min h , t0+ Δt− t

8. This subincrement has failed, so estimate smaller time subincrement by

computing

q= max0.8 DTOL∕Rut+h , 0.1

and then setting

h← qh

before returning to step 4.

9. Exit with displacements and pore pressures, Xt0+Δt, their corresponding

rates, X.

t0+Δt, and integration point stress state, (σ′t0+Δt , Àt0+Δt ), at end of

coarse time increment.

The Newton-Raphson procedure for solving the solving the nonlinear equations

(5.68) may be summarised as

1. Enter with the current displacements and pore pressures Xt, their

corresponding derivatives X.

t, the current step size h, the iteration tolerance

ITOL, and the maximum number of iterations MAXITS.

189Chapter 5

2. Compute estimate of new displacements/pore pressures and their

corresponding rates using

X~ 0t+h= Xt+ hX

.

t

X. 0t+h= X

.

t

3. Set

α0= hU. t ∕ U~ 0t+h where U

.

t is the velocity component of X.

t and U~ 0t+h is the displacement

component of X~ 0t+h.

4. Repeat steps 5 to 8 for i=1 to MAXITS

5. Compute residual vector

Ri=⎪⎪⎨⎧

Fextt+h− F

intt+h

h

Qt+h⎪⎪⎬⎫

⎭−⎪⎪⎡

0

LT

0

H

⎪⎪⎤

⎦⎪⎪⎨⎧

U. i−1t+h

P~ i−1t+h⎪⎪⎬⎫

and solve for δX. i using

δX. i= Cep+ hK

−1Ri (tangent stiffness)

or

δX. i= [C+ hK]−1Ri (initial stiffness)

where Fintt+h and Cep are evaluated at X~ i−1t+h.

6. Update the displacements and pore pressures and their rates

according to

X. it+h= X

. i−1t+h+ δX

. i

X~ it+h= Xt+ hX

. it+h

7. Compute convergence criterion

αi= hδU. i ∕ U~ it+h

190Chapter 5

where δU. i is the velocity component of δX

. i and U~ it+h is the

displacement component of X~ it+h. If α

i< ITOL then go to step 10.

8. For tangent stiffness only, check rate of convergence. If αi∕αi−1> 0.5

for more than 2 consecutive iterations then exit with “no convergence”

warning.

9. Maximum number of iterations exceeded. Exit with “no convergence”

warning.

10. Exit with displacements/pore pressures, X~

t+h= X~ it+h, and their rates,

X.

t+h= X. it+h, at time t+h.

Typical values for the iteration tolerance, ITOL, are in the range 10---3 to 10---6,

with the lower limit ensuring that the drift from equilibrium is very small. The

maximum number of iterations permitted for each subincrement, MAXITS, is

typically set to around 15. To minimise the high cost associated with failed

substeps, and to allow for possible divergence of the iteration scheme, h is

automatically reduced by a factor of four if no convergence is obtained within this

limit. This feature may result in an excessive number of substeps if MAXITS is

set to a very low value. The optimum setting for MAXITS is dependent on the

specified value of DTOL, since loose values of this tolerance may give large time

steps and hence large numbers of iterations. It is also a function of the solution

scheme used to solve the governing nonlinear equations. The tangent stiffness

algorithm generally requires far fewer iterations than the initial stiffness algorithm,

especially if the behaviour is highly nonlinear.

For the initial stiffness method, the matrix [C+ hK] needs to be formed and

factorised afresh only once per subincrement to obtain the iterates δX. i. The cost

associated with each formation phase may be reduced significantly using the

procedure outlined in Section 5.5.3. With the tangent stiffness Newton-Raphson

scheme it is necessary to reform and refactorise the Jacobian matrix [Cep+ hK]

once per iteration, since Cep is dependent on the displacements held in X~ i−1t+h.

191Chapter 5

In this case, it is possible to exploit the strategy discussed previously in Section

4.4.3. Using the decomposition

Cep=⎪⎡⎣Kep

LT

L

0⎪⎤⎦= C− Cp=⎪⎡⎣

Ke

LT

L

0⎪⎤⎦−⎪⎡⎣Kp

0

0

0⎪⎤⎦

it is evident that the matrix C need only be formed once and stored on disk. In

order to generate Cep efficiently for each iteration, C is loaded into memory from

disk and the plastic element stiffness matrices kp, defined by (4.20), are subtracted

element by element. Since only plastic Gauss points contribute to Kp, the effort

required to form Kep, and hence Cep, is usually small for much of the loading

range.

Finally, it is also possible to hold the Jacobian matrix constant for several

subincrements and only refactorise it when convergence becomes slow or when the

step size h changes significantly. These options have not been pursued in this

Thesis, but may lead to reductions in the overall CPU time.

192Chapter 5

193Chapter 6

CHAPTER 6

CONSOLIDATION APPLICATIONS

194Chapter 6

6.1 INTRODUCTION

In this Chapter, the automatic time stepping algorithms developed in Chapter 5

are used to study the behaviour of several problems involving the consolidation

of porous media. The aim of these analyses is to demonstrate that the new

procedures:

¯ Can control the global temporal discretisation (or time-stepping) error in the

displacements to lie near a prescribed error tolerance.

¯ Automatically choose a suitable number of time subincrements to achieve the

desired accuracy, regardless of the number of coarse time increments that are

specified initially.

¯ Are robust and efficient.

The first part of this Chapter considers the consolidation of porous elastic media.

Analyses are performed for the consolidation of a layer under one dimensional

loading, the consolidation of a layer between two rigid plates, and the

consolidation of a flexible strip footing resting on a layer of finite depth. In these

examples, the performance of the automatic algorithm is investigated using various

numbers of coarse time increments and a range of error tolerances. Where

appropriate, the efficiency of the automatic algorithm is measured against the

efficiency of the conventional backward Euler method by comparing the CPU

times required to generate solutions of comparable accuracy. To gauge the ability

of the new algorithm to control the level of temporal discretisation error, the

results from the automatic scheme are compared against those from a second

order accurate scheme using very small time steps. The latter solutions contain

negligible temporal discretisation errors, and thus serve as useful benchmarks.

Note that no attempt is made to measure the spatial discretisation error, which

is governed by the mesh configuration.

The next part of this Chapter investigates the ability of the elastoplastic

consolidation formulation to predict drained and undrained failure modes in a soil

195Chapter 6

mass. These cases illustrate two extremes of consolidation behaviour and are thus

useful checks on the accuracy of the finite element technique. The problems used

in these analyses are the expansion of a thick cylinder and the collapse of a flexible

strip footing.

The final part of the Chapter considers the consolidation of a flexible strip footing

resting on an elastoplastic soil layer. The layer is modelled using a rounded

Mohr-Coulomb yield surface with either an associated or a nonassociated flow

rule. For both the associated and nonassociated cases where where the dilation

angle is nonzero, the automatic time integration procedure is implemented using

a Newton-Raphson (or tangent stiffness) iteration scheme to solve the nonlinear

incremental equations for each time step. For the special case of a nonassociated

model with a zero dilation angle, an initial stiffness iteration scheme is employed

to solve these equations. As in the elastic consolidation examples, the results from

the automatic scheme are compared with those from the conventional backward

Euler scheme to assess its accuracy and efficiency. These analyses are also used

to investigate the effect of the iteration tolerance on the accuracy of the

displacements at various stages of consolidation.

In each of the problems to be analysed, whether elastic or elastoplastic, the soil

mass is modelled using six-noded triangles with a quadratic displacement

expansion for the displacements and a linear expansion for the pore water

pressures. This element avoids the spurious oscillations associated with elements

which use the same order of expansion for the displacements and pore pressures

(see, for example, Reed, 1984 and Kanok-Nukulchal and Suaris, 1982) and is

simple to implement.

For all problems considered in this Chapter, a ramp loading is imposed over the

time period t0 as shown in Figure 6.1. Following conventional practice, the rate

of loading is frequently expressed in terms of the time factor Tv, rather than the

196Chapter 6

t0 , Tv0

load

time, dimensionless time

Figure 6.1 Load vs time.

q0

actual time t, since this quantity is dimensionless. In most cases, Tv is defined to

be equal to

Tv=cv tH2

(6.1)

where cv is a coefficient of consolidation and H is a measure of the length of the

drainage path. Note, however, that the coefficient of consolidation may be either

one or two dimensional, depending on the problem, and the length measure used

may also vary. The precise form of Tv will be defined in the preamble to each

problem.

To gauge the performance of the automatic time stepping algorithm, the global

temporal error in the transient displacements Ut are estimated using

uerror=Ut− Uref ∞Uref ∞

(6.2)

where Uref are a set of reference displacements calculated at the corresponding

time. These reference displacements are computed using the second order scheme

of Thomas and Gladwell (1988a), described in Section 5.4.3, with the three

integration parameters set to φ1= φ2= φ3= 1. These solutions have a very

small temporal discretisation error, since they are obtained using a very large

number of time increments. Using the reference displacements and equation

(6.2), uerror gives an approximate estimate of the global time-stepping error, and

197Chapter 6

may be compared directly against the specified tolerance DTOL to ascertain the

performance of the error control strategy. Ideally, the observed value of uerror will

lie reasonably close to DTOL, at least to within an order of magnitude. It is also

desirable that, as the tolerance is tightened, the observed time-stepping errors will

be reduced by a commensurate amount.

In the results that follow, various timing statistics are given to indicate the

efficiency of the proposed automatic time incrementation scheme. All of these

are for a Sun Ultra 170 workstation with the Sun FORTRAN 77 compiler and level

3 optimisation.

6.2 ELASTIC CONSOLIDATION

For each of the problems considered in this Section, the soil is modelled as an

elastic, isotropic, weightless medium with a uniform permeability. The properties

of the soil are thus completely defined by its drained Youngs modulus, E′, its

drained Poisson’s ration, ν′, and its permeability k.

In all of the elastic analyses, a three-point scheme is used to integrate the element

stiffness, coupling and flow matrices for the six-noded triangle. This rule is exact

for a straight-sided plane strain triangle with a quadratic expansion for the

displacements and a linear expansion for the pore pressures (see, for example,

Laursen and Gellert, 1978) and is the most efficient method available for

computing the stiffness and coupling matrices. Note that slightly greater

economies could be achieved by employing a one-point rule to evaluate the

element flow matrices, h, since all of their terms are constants. This was not done

in the current study because the additional savings are only marginal.

6.2.1 One Dimensional Compression of a Finite Layer

An analytical solution for the one dimensional consolidation of an elastic porous

layer under a uniform surface pressure has been presented by Terzaghi (1923).

The mesh and boundary conditions for this problem are shown in Figure 6.2. The

198Chapter 6

H

1

Figure 6.2 Uniform mesh for one dimensional consolidation of finite layer.

impermeable

permeable q0

soil layer is assumed to be of thickness H and loaded by a uniform surface pressure

of q0 . As indicated in Figure 6.1, the finite element analysis assumes that a ramp

load is imposed over the dimensionless time period Tv0= 0.0001, where

Tv0=cvt0H2

and cv, the one dimensional coefficient of consolidation, is given by

cv=kE′(1− ν′)

γw(1+ ν′)(1− 2ν′)(6.3)

In the above equation E′, ν′ and k denote, respectively, the drained Young’s

modulus, the drained Poisson’s ratio, and the permeability of the soil, while γw is

the unit weight of water. After the total pressure q0 has been applied over the

period Tv0, the layer is allowed to consolidate over a dimensionless time factor

increment of ΔTv= 1.20. Thus, at the end of the analysis, the total dimensionless

time is given by Tv= 0.0001+ 1.20= 1.2001.

To measure the accuracy of various algorithms, the global time-stepping errors in

the displacements are estimated using equation (6.2). The reference

199Chapter 6

displacements for this case are calculated using the second order accurate scheme

of Thomas and Gladwell (1988a), with 1000 equal size increments to apply the

load and 10,000 equal size time increments to model the subsequent consolidation.

DTOLNo. coarse time No. time subincrements* CPU

DTOLNo. coarse timeincrements* Successful Failed

CPUtime (s)

10---2 1+1=2 9+48=57 2+3=5 1.1

1+6=7 9+50=59 2+3=5 1.51+12=13 9+53=62 2+3=5 1.9

10---3 1+1=2 27+129=156 4+3=7 2.6

1+6=7 27+133=160 4+3=7 2.91+12=13 27+133=160 4+3=7 3.2

10---4 1+1=2 85+379=464 5+3=8 7.0

1+6=7 85+382=467 5+3=8 7.3

1+12=13 85+386=471 5+3=8 7.5

Table 6.1 Results for one dimensional consolidation using automatic scheme.(* no. in loading stage + no. in consolidation stage = total no.)

Results for analyses using the automatic time stepping scheme of Section 5.5.3 are

shown in Table 6.1, Figure 6.3 and Figure 6.4. Data are presented for

DTOL=10---2, 10---3 and 10---4, which are typical of the tolerance values that would

be used in practice. To test the sensitivity of the automatic scheme to the starting

conditions, three runs are performed for each tolerance using 2, 7 and 13 coarse

time steps. In each case, all of the load is applied in the first coarse time step

which has a time factor increment of ΔTv0= 0.0001. The remaining coarse

increments are of uniform size and apply a total time factor increment of

ΔTv= 1.2. Note that entries in Table 6.1 of the form i+j=k indicate that i steps

occurred in the loading phase, j steps occurred in the consolidation phase, and k

steps occurred overall.

Figure 6.3 compares the numerical consolidation curve, obtained using the

automatic scheme with 2 coarse time steps and DTOL=10---2, against the analytical

solution derived by Terzaghi (1923). This analysis generates a total of 56 substeps

and its predictions are in excellent agreement with the exact results.

200Chapter 6

0.00001 0.00010 0.00100 0.01000 0.10000 1.00000

Figure 6.3 Degree of consolidation versus time factor for one dimensionalconsolidation.

U

Tv=cvtH2

Terzaghi (1923)

0.2

0.4

0.6

0.8

1.0

0.0

finite element(2 coarse time steps, DTOL=10---2)

The results in the Table 6.1 indicate that, for each value of DTOL, the automatic

scheme always chooses a similar number of subincrements, regardless of the

number of coarse time increments that are specified initially. With DTOL=10---2,

for example, the automatic scheme selects 57, 59 and 62 time substeps when 2,

7, and 13 coarse time steps are specified. Each of these analyses automatically

selects 9 substeps during the loading phase. Because of the design of the

algorithm, it is usual for the last substep in each coarse time step to be truncated.

For a fixed value of DTOL, this causes the total number of substeps to increase

slightly as the number of coarse steps is increased. For all values of DTOL, the

number of failed substeps is a small proportion of the total number of successful

substeps. This suggests that the adaptive substepping strategy is correctly tuned

and does not suffer from spurious oscillations.

The variation of the temporal discretisation error during each of the automatic

analyses with 13 coarse time increments is shown in Figure 6.4. In each case, the

201Chapter 6

DTOL=10---2

Figure 6.4 Temporal discretisation error in displacements versus time factorfor one dimensional consolidation.

1300 increments

130 increments

automatic (13 coarse time steps)

backward Euler

DTOL=10---3

DTOL=10---4

13000 increments

Tv=cvtH2

10−2

10−3

10−4

0.0 0.2 0.4 0.6 0.8 1.0 1.2

uerror

10–5

maximum temporal discretisation errors are just below the specified tolerance

DTOL. With DTOL=10---3, for example, the maximum temporal error occurs at

Tv≈ 1.0 and is approximately equal to 5¢10---4. These results suggest that the

automatic scheme is able to constrain the global time-stepping error to lie near

the specified tolerance DTOL. Because the automatic scheme increases the step

size as consolidation takes place, the temporal error in the displacements is

roughly constant over the last half of the time interval.

202Chapter 6

To assess the performance of a traditional solution method, this problem was also

analysed using the backward Euler scheme with various numbers of equal-size

time increments. Since the material is elastic, the backward Euler method

requires only two assemblies and two factorisations of the global equations to

complete each analysis. These assemblies and factorisations occur at the start of

the loading and consolidation phases. The CPU times and temporal discretisation

errors for the various backward Euler runs are shown, respectively, in Table 6.2

and Figure 6.4.

No. time incrementsCPU time (s)

Loading Consolidation TotalCPU time (s)

10 120 130 1.4

100 1200 1300 4.2

1000 12000 13000 34

Table 6.2 Backward Euler results for one dimensional consolidation.

The results in Table 6.2 suggest that, for the analyses with up to around a thousand

time steps, the bulk of the computational work occurs in the assembly and

factorisation stages and the CPU time is not proportional to the number of time

steps used. For the runs with very small time steps, the assembly and factorisation

times are less dominant and the overall CPU time grows in the manner expected.

Figure 6.4 indicates that the temporal discretisation error in the displacements

decreases as the number of time increments is increased. For all the backward

Euler analyses, the temporal discretisation error in the displacements is greatest

at Tv≈ 0.01 and drops off significantly in later stages of consolidation. The runs

with 130, 1300 and 13,000 time steps give maximum time-stepping errors in the

displacements of roughly 1.2¢10---2, 1.2¢10---3 and 1.2¢10---4. These results clearly

exhibit the first order accuracy of the backward Euler scheme.

The efficiency of the backward Euler and the automatic schemes can be compared

using the data in Table 6.1, Table 6.2 and Figure 6.4. Inspection of the latter

indicates that the 1300 increment backward Euler analysis gives a maximum

203Chapter 6

temporal discretisation error which is close to that of the automatic analysis with

DTOL=10---3. The data in Table 6.2 and Table 6.1 reveal that the CPU times for

these two runs are, respectively, 4.2 and 3.2 seconds. For the most accurate

analysis with DTOL=10---4, the automatic scheme generates a total of 471 substeps

and requires a maximum of 7.5 seconds of CPU time. This compares very

favourably with the 13000 increment backward Euler run, whose result is of similar

accuracy but uses 34 seconds of CPU time. It could be argued that it is

inappropriate to compare the automatic scheme against the backward Euler

method with fixed increment sizes, and that a better performance measure would

be obtained by using the latter method with varying increment sizes. The problem

then, though, is how to choose the optimal size and distribution of time increments

for the backward Euler scheme. In practice, it may be necessary to perform

several analyses in order to ascertain an efficient time stepping sequence. The

automatic scheme overcomes this difficulty, since it requires only a single analysis

to obtain a solution whose accuracy is known approximately.

The analyses shown in Table 6.1 which use just one coarse time increment for the

loading phase and one coarse time increment for the consolidation phase are

included to highlight the robustness of the proposed algorithm. A bar chart of the

successful substeps chosen by the automatic scheme, for the case of DTOL=10---2,

is presented in Figure 6.5. This indicates that the time step grows by almost five

orders of magnitude, from an initial value of Tv≈ 5× 10−6 to a maximum value

of Tv≈ 0.17. As expected, the automatic scheme selects very small increments

in the early stages of the analysis where the rate of pore pressure dissipation is

the greatest. Using a log-log plot of the increment size ΔTv versus Tv, the growth

in step size for all of the automatic runs is shown in Figure 6.6. The dramatic

increases indicated in this plot highlight the inefficiency of using uniform time

steps for this particular problem, and further demonstrates the benefits obtained

by the use of an adaptive integration scheme.

204Chapter 6

Figure 6.5 Subincrement size selection for analysis of one dimensionalconsolidation

10 20 30 40 50

successful subincrement number

ΔTv

2 coarse time stepsDTOL=10---2

10---1

10---2

10---3

10---4

10---5

1

10---6

Figure 6.6 Subincrement size versus time factor for analysis of onedimensional consolidation.

ΔTv

Tv= cv tH2

10010---110---210---4 10---310---510---6

10---1

10---2

10---3

10---4

10---5

10---6

Tv0

DTOL= 10−4DTOL= 10−3DTOL= 10−2

2 coarse time steps

10---7

10---7

205Chapter 6

A number of researchers, including Sandhu et al (1977), Kanok-Nukulchal and

Suaris (1982) and Reed (1984), have noted that small initial time step may produce

spatial oscillations in the pore pressures near free draining boundaries. Because

of this phenomenon, Vermeer and Verruijt (1981) recommend that the step size

should not be reduced below a threshold value. Using an uncoupled diffusion

model for one dimensional consolidation, they proposed that the minimum time

step is given by Δtmin= l2∕(6θcv), where l is the length of the shortest element

immediately adjacent to a free draining boundary, cv is the one dimensional

coefficient of consolidation given by equation (6.3), and θ is an integration

parameter. This relation may also be written in the dimensionless form

ΔT minv = 16 lH2 (6.4)

and is applicable to any one dimensional element with a linear pore pressure

expansion. It should be stressed that (6.4) is not strictly valid for coupled Biot

consolidation problems in one or two dimensions. Vermeer and Verruijt (1981)

suggest, however, that similar relationships will hold for these cases and give some

numerical evidence to support their claim.

Although the practice of using large time steps in the early stages of consolidation

may help to reduce the pore pressure oscillations, it will also have the undesirable

effect of increasing the temporal discretisation error in the solution. This error

tends to dissipate as consolidation nears completion, but may be very significant

during, and immediately after, the loading phase. Large pore pressure oscillations

which are adjacent to free draining boundaries can, alternatively, be viewed as a

signal that the mesh needs to be refined in these zones. Indeed, for most practical

problems, judicious refinement of the mesh near free draining boundaries will

often drastically reduce troublesome oscillations. This is believed to be a better

strategy than imposing an artificially large time step constraint on the solution

process, since it addresses the cause of the oscillations directly and does not

206Chapter 6

introduce additional sources of error. In view of these arguments, the automatic

time incrementation schemes developed in this Thesis do not impose a minimum

on the size of the time step.

To illustrate the effect of mesh refinement on the pore pressures throughout a one

dimensional layer, consider the uniform mesh of Figure 6.2 and the graded mesh

of Figure 6.7. Both of these grids have the same number of elements and the same

number of degrees of freedom, the only difference is that the latter mesh is highly

refined in the vicinity of the top drainage boundary. The pore pressure isochrones

for these two meshes, obtained from the automatic scheme with DTOL=10---2 and

2 coarse time steps, are presented in Figure 6.8. Also shown on this plot are the

exact solutions derived by Terzaghi (1923). For the uniform mesh, oscillations in

the pore pressures are clearly evident in the very early stages of the analysis but

dissipate quickly with time. The oscillations are most pronounced at the end of

the loading phase, where Tv= 10–4 , and arise in spite of the fact that nine time

subincrements have been used up to this point. It is interesting to note that these

observations are in accordance with equation (6.4), which predicts that oscillations

H

1

Figure 6.7 Graded mesh for one dimensional consolidation of finite layer.

impermeable

permeable

z

q0

207Chapter 6

0.0 0.5 1.0 1.50.000 0.500 1.000

zH

pq0

Tv= 1.0

10−1

10−4

10−2

10−3

Figure 6.8 Pore pressure isochrones for one dimensional consolidation.

graded mesh

pq0

Tv= 1.0

10−1

10−4

10−2

uniform mesh

1.0

0.9

0.8

0.7

0.6

0.5

0.4

0.3

0.2

0.1

0.0

finite elementTerzaghi (1923)

10−3

will occur for the uniform mesh if ΔTv≤ 4.167× 10–4. The results for the graded

mesh are in excellent agreement with Teraghi’s exact solution, even for small

values of Tv. This analysis uses 40 time subincrements during the dimensionless

time period of Tv= 10–4, with the first (and smallest) being equal to

ΔTv= 0.36× 10–7. Although this value is smaller than the minimum of

ΔTminv = 1.042× 10–6 predicted by equation (6.4), any oscillations in the pore

pressures are now small and dissipate extremely quickly.

208Chapter 6

6.2.2 Finite Layer Compressed Between Two Rigid Plates

Consider the consolidation of an elastic plane strain layer, compressed between

two smooth rigid plates, as shown in Figure 6.9. This problem has been solved

smooth rigidF= 2aq0impermeable plates

2a

flowflow

Figure 6.9 Consolidation of layer between two rigid plates (Mandel 1953).

a

permeableboundary

x

q0

analytically by Mandel (1953), and thus serves as a useful benchmark for checking

two dimensional finite element formulations. In the finite element model of

Figure 6.9, load is applied to the plates in the form of prescribed pressures and

the rigid boundary is modelled by constraining the nodal displacements along the

plate interface to be equal. The dimensionless time factor for this problem is

Tv=cvt

3a2

where a is the half-width of the plates and cv is the one dimensional coefficient

of consolidation defined by (6.3). The load is ramped, as illustrated in Figure 6.1,

and reaches its maximum value after the dimensionless time period Tv0= 0.001.

Once the total pressure, q0, is applied, consolidation is analysed over a

dimensionless time factor increment of ΔTv= 1.2. Therefore, at the conclusion

of the analysis, the dimensionless time is given by Tv= 0.001+ 1.200= 1.201.

209Chapter 6

The reference solution for this example, which provides a benchmark to compute

the global time-stepping errors for various other runs, is found using the second

order scheme of Thomas and Gladwell (1988a). To ensure that the temporal error

is minimised, 1000 and 10,000 equal size increments are used over the loading and

consolidation phases respectively.

DTOLNo. coarse time No. subincrements CPU

DTOLNo. coarse timeincrements Successful Failed

CPUtime (s)

10---2 1+1=2 1+23=24 0+1=1 1.2

1+5=6 1+26=27 0+1=1 1.51+10=11 1+29=30 0+1=1 1.9

10---3 1+1=2 9+60=69 4+2=6 3.0

1+5=6 9+61=70 4+2=6 3.3

1+10=11 9+64=73 4+2=6 3.610---4 1+1=2 30+173=203 5+3=8 7.9

1+5=6 30+174=204 5+3=8 8.2

1+10=11 30+178=208 5+3=8 8.6

Table 6.3 Results for consolidation of layer between rigid plates using automaticscheme.

Results for the automatic time incrementation scheme are shown in Table 6.3.

Data are presented for DTOL values ranging from 10---2 to 10---4, with each

tolerance being run using 2, 6 and 11 coarse time increments. In each analysis,

all of the load is applied in the first coarse time step which has a time factor

increment of ΔTv= 0.001. The remaining coarse increments are of uniform size

and give a total time factor increment of ΔTv= 1.2.

The results in the Table 6.3 indicate that the behaviour of the automatic scheme

is largely independent of the coarse time steps that are specified initially. For a

fixed value of the tolerance DTOL, it always chooses a similar number of

subincrements. With DTOL=10---2, for example, the automatic scheme selects 24,

27 and 30 time substeps when 2, 6, and 11 coarse time steps are specified. Note

that no subincrementation is required with this tolerance during the loading phase,

210Chapter 6

as each of the analyses generates only a single substep. As in the one dimensional

consolidation example, the number of failed substeps is a small proportion of the

total number of successful substeps.

A typical plot of the transient pore pressure variation at the centre of the layer

is shown in Figure 6.10. This particular curve was obtained using two coarse time

increments and a tolerance of DTOL=10---2. With these settings, the finite

element analysis generates a total of 24 successful substeps and predicts pore

pressures which are in excellent agreement with the exact results of Mandel

(1953). A detailed bar chart of the time steps that were used to construct

Figure 6.10 is shown in Figure 6.11.

Figure 6.12 illustrates the temporal discretisation errors at various stages of the

runs with 11 coarse time increments. In each case, the maximum temporal

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.0001 0.0010 0.0100 0.1000 1.0000

Figure 6.10 Pore pressure (at centre of layer) versus time factor forconsolidation of layer between rigid plates.

pq0

Tv=cvt3a2

Mandel (1953)

finite element(2 coarse time steps, DTOL=10---2)

211Chapter 6

Figure 6.11 Subincrement size selection for consolidation of layer betweenrigid plates.

0.0010

0.0100

0.1000

1.0000

10.0000

5 10 15 20

ΔTv

successful subincrement number

2 coarse time stepsDTOL=10---2

discretisation error is just below the specified tolerance DTOL and is roughly

constant for the last half of the consolidation period. With DTOL=10---3, for

example, the maximum temporal error occurs at Tv≈ 2.0 and is approximately

equal to 6¢10---4. These results again suggest that the automatic scheme is able

to constrain the global time-stepping error to lie near the specified tolerance

DTOL.

To investigate the performance of a traditional solution scheme, this problem was

also analysed using the backward Euler algorithm with various numbers of

equal-size time increments. The CPU times and temporal discretisation errors for

these runs are shown, respectively, in Table 6.4 and Figure 6.12. These results

confirm the expected first order accuracy of the backward Euler scheme. For the

runs with 110, 1100 and 11000 uniform increments, the corresponding maximum

temporal errors are of the order of 10---2, 10---3 and 10---4 respectively.

212Chapter 6

DTOL=10---2

Figure 6.12 Temporal discretisation error in displacements versus time factorfor consolidation of layer between rigid plates.

1,100 increments

110 increments

automatic (11 coarse time steps)

backward Euler

DTOL=10---3

Tv=cvt3a2

10−2

10−3

10−4 DTOL=10---4

11,000 increments

0.0 0.2 0.4 0.6 0.8 1.0

uerror

10–5

No. time increments CPU timeLoading Consolidation Total (s)

10 100 110 1.0

100 1000 1100 3.6

1000 10000 11000 30

Table 6.4 Backward Euler results for consolidation of layer between rigid plates.

213Chapter 6

The relative efficiency of the automatic and backward Euler method can be

estimated using the data in Table 6.3, Table 6.4 and Figure 6.12. For example, with

DTOL=10---3, the maximum time-stepping error for the automatic analysis is

roughly equal to that for the 1100 increment backward Euler analysis. The CPU

times for these two runs are very similar, but the automatic scheme achieves this

accuracy with a maximum of 73 substeps. For higher accuracies, the efficiency of

the automatic scheme increases relative to that of the backward Euler method.

As discussed in the previous Section, the use of small initial time increments may

cause spurious oscillations in the pore pressures in the vicinity of free draining

boundaries. To investigate the influence of mesh refinement on these oscillations,

the two grids shown in Figure 6.13 are used to model the plate consolidation

problem. The first mesh is uniform, with an element length of l= 0.1a, while the

second mesh is graded such that the size of the element adjacent to the drained

boundary is l= 0.00625a. Using equation (6.4), the minimum initial time steps

predicted by the Vermeer and Verruijt condition are ΔTminv = 5.6× 10–4 for the

uniform mesh and ΔTminv = 2.17× 10–6 for the graded mesh. Both of the grids

are analysed using the automatic scheme with DTOL=10---3 and one coarse time

step in the loading and consolidation phases. Unlike the other analyses performed

in this Section, which assume that Tv0= 0.001, the loading for this case is ramped

Figure 6.13 Uniform and graded meshes for consolidation of layer betweentwo rigid plates.

uniform mesh

a

graded mesh

214Chapter 6

over the period of Tv0= 0.0001. The pore pressure isochrones for the uniform

mesh, constructed from nodal values at the centre of the layer, are presented in

Figure 6.14. A similar plot for the graded grid is given in Figure 6.15. These

xa

pq0

Tv= 1.0

Tv= 0.1

Tv= 0.0001

Tv= 0.01

Tv= 0.001

Figure 6.14 Pore pressure isochrones for consolidation between rigid platesusing uniform mesh and DTOL=10---3.

0.6

0.4

0.2

0.00.0 0.2 0.4 0.6 0.8 1.0

0.7

0.5

0.3

0.1

finite element

Mandel (1953)

xa

pq0

Tv= 1.0

Tv= 0.1

Tv= 0.0001

Tv= 0.01

Tv= 0.001

Figure 6.15 Pore pressure isochrones for consolidation between rigid platesusing graded mesh and DTOL=10---3.

0.6

0.4

0.2

0.00.0 0.2 0.4 0.6 0.8 1.0

0.5

0.3

0.1

finite elementMandel (1953)

215Chapter 6

results indicate that the uniform mesh produces significant oscillations in the pore

pressures up to a time factor of around Tv= 0.01. After this point, the pore

pressures behave as expected and are close to zero by the time Tv= 1. In

contrast, the numerical predictions from the graded mesh and are in good

agreement with Mandel’s solution and display no pore pressure oscillations for all

values of Tv greater than 10---4. Although some oscillations may occur for

Tv≤ 10−4, these are generally small and have little effect on the overall quality

of the solution.

6.2.3 Flexible Strip Footing on Finite Layer

In this Section the automatic time stepping scheme is used to analyse the

consolidation of a rough flexible strip footing resting on a porous elastic layer.

The mesh and boundary conditions for the problem considered are shown in

Figure 6.16. In this example, the ramp load is applied to the footing over the

smooth/impermeable

smooth/impermeable

rough / impermeable

permeable

10B

H=5B

B

Figure 6.16 Flexible strip footing on elastic layer.

q0

initial period Tv0= 0.0001 and the time factor is given by

Tv=cv tH2

(6.5)

216Chapter 6

where H is the depth of the soil layer and cv is the one dimensional coefficient

of consolidation defined by (6.3). To study the behaviour of the automatic and

backward Euler schemes under fully drained conditions, the consolidation process

is modelled up to a time factor of Tv= 10. This guarantees a fully drained state,

as all of the excess pore pressures have essentially dissipated when Tv≈ 2.

The reference displacements in this case are calculated differently to the preceding

examples, with 1000 equal size time increments being used to model the loading

phase and 900 uniform increments per log cycle being used to model the

consolidation phase. Since Tv= 10 at the end of the analysis, the total number

of increments employed in computing the reference solutions is equal to

1, 000+ 5× 900= 5, 500. The efficiency of using a logarithmic incrementation

scheme is discussed in more detail later in this Section.

DTOLNo. coarse time No. subincrements CPU

DTOLo. coa se t eincrements Successful Failed

C Utime (s)

10---1 1+1=2 1+19=20 0 5.0

1+5=6 1+22=23 0 5.9

1+10=11 1+25=26 0 7.410---2 1+1=2 3+36=39 2+1=3 9.1

1+5=6 3+39=42 2+1=3 9.9

1+10=11 3+42=45 2+1=3 11.510---3 1+1=2 12+89=101 3+3=6 20

1+5=6 12+91=103 3+3=6 21

1+10=11 12+95=107 3+3=6 23

10---4 1+1=2 39+247=286 4+3=7 531+5=6 39+250=289 4+3=7 55

1+10=11 39+253=292 4+3=7 56

Table 6.5 Results for elastic strip footing using automatic scheme and uniformcoarse time increments.

Results for various footing analyses with the automatic scheme are shown in

Table 6.5. Data are presented for DTOL values ranging from 10---1 to 10---4, with

each tolerance being run using 2, 6 and 11 coarse time increments. For all

217Chapter 6

analyses, the load is applied in the first coarse time step which has a time factor

increment of ΔTv0= 0.0001. The remaining coarse increments are of near

uniform size and advance the solution to Tv= 10. As in previous examples, these

results indicate that the automatic scheme chooses a similar number of substeps

for each value of DTOL, regardless of the initial coarse time step size. With

DTOL=10---2, for example, the new algorithm generates 39, 42 and 45 substeps

for runs with 2, 6 and 11 initial coarse time steps. The bulk of these substeps occur

in the consolidation phase, with only three substeps being generated during the

application of the load.

To illustrate the accuracy of the automatic scheme, Figure 6.17 shows two plots of

the degree of consolidation at the centre of the footing versus the time factor.

These results were generated using two coarse time steps and DTOL values of

10---1 and 10---2. The finite element predictions for both of these tolerances are

very similar and match the analytic solution of Booker (1974) over all of the

loading range. The small amount of deviation indicated is attributable to the

spatial discretisation error. It is interesting to note that, on the scale of

Figure 6.17, the results for the most stringent tolerance of DTOL=10---4 are

indistinguishable from those for DTOL=10---2 . This suggests the latter value is

a practical starting point for analysing the behaviour of elastic two dimensional

consolidation problems. To gain some insight into the step selection philosophy

of the automatic scheme, Figure 6.18 shows the successful time step sizes for the

analysis with DTOL=10---2. As in previous examples, the step size is small at the

start of the analysis and increases consistently throughout the entire consolidation

process. The step size ranges from a minimum value of ΔTv= 3× 10–5 to a

maximum of ΔTv= 3.32 and, on average, grows by an order of magnitude over

6 or 7 consecutive substeps. An alternative picture of the substep size variation,

for all values of DTOL, is shown in Figure 6.19. This plot indicates that all of the

analyses adopt a similar step size once consolidation is complete, regardless of the

value of the error tolerance. This is a direct result of the governing differential

218Chapter 6

0.00001 0.00010 0.00100 0.01000 0.10000 1.00000 10.00000

0.00001 0.00010 0.00100 0.01000 0.10000 1.00000 10.00000

Figure 6.17 Degree of consolidation versus time factor for elastic strip footing.

U

Tv=cvtH2

U

Tv=cvtH2

0.0

0.2

0.4

0.6

0.8

1.0

0.0

0.2

0.4

0.6

0.8

1.0

Finite element (DTOL=10---1)

Booker (1974)

Finite element (DTOL=10---2)

Booker (1974)

ν= 0.0

ν= 0.0

219Chapter 6

Figure 6.18 Subincrement size selection for consolidation of elastic stripfooting.

0.000010

0.000100

0.001000

0.010000

0.100000

1.000000

10.000000

5 10 15 20 25 30 35

ΔTv

subincrement number

2 coarse timestepsDTOL=10---2

Figure 6.19 Subincrement size versus time factor for consolidation of elasticstrip footing.

Tv0

Tv

DTOL= 10−4DTOL= 10−3DTOL= 10−2DTOL= 10−1

10010---110---210---4 10---310---510---6

10---1

10---2

10---3

10---4

10---5

10---6

100

10+1

101

ΔTv

2 coarse time steps

220Chapter 6

equations becoming uncoupled and, in general, the step size is limited only by the

growth control factors discussed in Section 5.5.1.

Time factor increment (ΔTv) Total time factor (Tv)

Loading 0.0001 0.00010.0009 0.001

0.009 0.01

Consolidation 0.09 0.1

0.9 1.09.0 10.0

Table 6.6 Log cycles for analysis of strip footing.

To further investigate the step control behaviour of the automatic algorithm, an

additional set of footing analyses are performed in which each log cycle of the time

factor is used as a single coarse time step. The coarse time steps adopted in this

type of analysis are shown in Table 6.6. As before, all of the load is imposed over

the interval Tv0= 0.0001 and the total time factor at the end of the run is

Tv= 10. It is pleasing to note that the results of these computations, shown in

Table 6.7, are very similar to those generated using uniform coarse steps

(Table 6.5). With DTOL=10---2, for example, an average of 42 substeps are

generated in the analyses with uniform coarse steps, while 43 substeps are

generated by the analysis with logarithmically varying coarse steps. This reinforces

the conclusion from the previous examples that the automatic step control

mechanism is largely insensitive to the starting conditions.

DTOLNo. coarse time No. subincrements CPU

DTOLo. coa se t e

incs. per log cycle Successful FailedC Utime (s)

10---1 1 1+23=24 0 6.7

10---2 1 4+39=43 2+2=4 11.7

10---3 1 13+94=107 3+3=6 26

10---4 1 40+258=298 4+3=7 68

Table 6.7 Results for elastic strip footing using automatic scheme andlogarithmic coarse time steps.

221Chapter 6

Figure 6.20 shows the temporal discretisation errors at various stages of the runs

with the logarithmic coarse time step variation. These results are plotted only up

to Tv= 1, as beyond this point the fully drained state is approached and the errors

become very small. For all cases, the maximum temporal discretisation errors are

held just below the specified tolerance DTOL and are roughly constant over much

of the consolidation period. These results confirm that the automatic scheme is

0.0001 0.0010 0.0100 0.1000 1.0000

Figure 6.20 Temporal discretisation error in displacements versus timefactor for elastic strip footing (logarithmic coarse time steps).

900 incs. per log cycle

90 incs. per log cycle

9 incs. per log cycle

1 inc. per log cycle

DTOL=10---4

DTOL=10---3

DTOL=10---2

DTOL=10---1

automatic (6 coarse log steps)

Backward Euler

Tv=cvth2

10---1

10---2

10---3

10---4

10---5

10---6

uerror

222Chapter 6

able to constrain the global time-stepping error to a desired level for the case of

logarithmically varying coarse time steps.

Time increment size Total no. CPULoading Consolidation increments time (s)

0.0001 0.1 1+20=21 3.5

0.00001 0.01 10+200=210 5.7

0.000001 0.001 100+2000=2100 270.0000001 0.0001 1000+20000=21000 241

Table 6.8 Backward Euler results for elastic strip footing using uniformincrement sizes.

To gauge the performance of a traditional solution method, the footing problem

was also analysed using a backward Euler scheme. The results for these runs are

shown in Table 6.8. These studies employed 1, 10, 100 and 1,000 equal size time

increments to model the loading phase and, correspondingly, 20, 200, 2000 and

20000 equal size increments to model the consolidation phase. Unlike the other

analyses in this Section, which are solved to Tv= 10, these analyses only compute

the solution to Tv= 2. A plot of the time-stepping error versus time factor for

these runs, shown in Figure 6.21, indicates that the error decays by several orders

of magnitude as consolidation proceeds. It also clearly confirms the expected first

order accuracy of the backward Euler scheme. Figure 6.21 shows that the 2100

increment backward Euler run gives a maximum time-stepping error of roughly

10---2 and requires 27 seconds of CPU time. For a similar accuracy, the data in

Table 6.5 reveals that the automatic scheme, with DTOL=10---2 and two uniform

coarse load steps, uses 9.1 seconds of CPU time.

Although it is computationally convenient, the backward Euler method need not

be used with time increments of equal size. A simple alternative, which will now

be investigated, is to adopt a fixed number of increments over each log cycle of

the time factor. Although this strategy leads to abrupt changes in step size

between adjacent log cycles, it provides a simple hand method for increasing the

223Chapter 6

Figure 6.21 Temporal discretisation error in displacements versus time factorfor elastic strip footing using backward Euler scheme.

1000 incs.per log cycle

logarithmic increment sizesuniform increment sizes

1 inc.per log cycle

Tv=cvtH2

10---2

10---3

10---4

10---5

10---61.00.001 0.01 0.1

10 incs.per log cycle

100 incs.per log cycle

21 incs.

210 incs.

2100 incs.

21000 incs.

uerror

time increments as consolidation proceeds. To assess the performance of such a

scheme, the backward Euler algorithm was run with the time step regimes shown

in Table 6.6 and Table 6.9. In these analyses, the loading and consolidation stages

were modelled using 1, 10, 100 and 1,000 increments per log cycle of the time

factor. The CPU times and error data for these runs are shown in Table 6.9,

Figure 6.20, and Figure 6.21, respectively. The error plots indicate that a

logarithmic variation of the step size results in the time-stepping errors being held

224Chapter 6

essentially constant over most of the consolidation period. This is in stark contrast

to the backward Euler results for a uniform step size, which display a marked

decay with time. Figure 6.20 shows that the error control characteristics of the

backward Euler method, when used with a logarithmic step size variation, are very

similar to those of the automatic scheme.

No. time increments per log cycle CPUtime

Loading Consolidation Totaltime(s)

1 1 6 2.6

10 10 60 3.5

100 100 600 121000 1000 6000 97

Table 6.9 Backward Euler results for strip footing using logarithmic incrementsizes.

In comparing the performance of the various strategies, it can be seen from

Figure 6.20 that the backward Euler run with nine fixed size time increments per

log cycle is of comparable accuracy to the automatic analysis with six coarse steps

per log cycle and DTOL=10---2. For this case, the backward Euler algorithm is

over three times faster than the automatic scheme. The automatic scheme,

however, is much more competitive for runs where greater accuracy is required.

The analysis with six coarse steps and DTOL=10---4, for example, is only

marginally less accurate than the backward Euler analysis with 900 increments per

log cycle, but uses thirty percent less CPU time.

6.3 ELASTOPLASTIC CONSOLIDATION

In this Section, a range of elastoplastic consolidation problems are considered.

The overall aim of the studies is to assess the efficiency and accuracy of the

nonlinear consolidation algorithm algorithm developed in Chapter 5. In each of

the problems discussed, the soil is again modelled as a weightless medium so that

the total pore pressure is equal to the excess pore pressure. An elastic perfectly

225Chapter 6

plastic model is assumed for the soil skeleton, and is used in conjunction with the

rounded Mohr-Coulomb yield surface described in Chapter 2. Unless noted

otherwise, the elastoplastic constitutive laws are integrated using the automatic

stress integration scheme described in Chapter 3 with a stress error tolerance of

STOL=10---6 and a yield surface tolerance of FTOL=10---9. These tolerances are

set stringently for the purposes of error checking and benchmarking, and should

be relaxed for practical computations. Setting STOL=10---3 and FTOL=10---6 will

provide sufficient accuracy in most applications and will also lead to substantial

reductions in CPU time.

In all of the elastoplastic analyses, a six-point integration scheme is used to the

evaluate the element stiffness, coupling, and flow matrices. This rule is used in

preference to the three point rule because it improves the stability of the tangent

stiffness iteration algorithm when large plastic strain increments are encountered.

Further efficiencies in this area could be realised by using a three point scheme

for the coupling matrices and a one point scheme for the flow matrices.

As described in Section 5.6.2, the nonlinear equations which govern elastoplastic

consolidation are solved using either an initial stiffness or a Newton-Raphson

iteration algorithm. Because the latter proved be unstable for some problems

involving elastoplastic soils with a zero dilation angle, the initial stiffness algorithm

is generally employed for these cases. Unless stated otherwise, the initial stiffness

and Newton-Raphson schemes are used with iteration tolerances of ITOL=10---3

and ITOL=10---6. The looser tolerance is needed for the initial stiffness scheme

because of its much slower rate of convergence. In solving the nonlinear equations

for each time step, no restriction is placed on the maximum number of iterations

that can be performed. This results in the time step size being governed

completely by the local error estimator, so that the global time-stepping error in

the displacements is due solely to the step control mechanism used in the

automatic scheme. Under these circumstances, the global time-stepping errors

226Chapter 6

may be compared directly against the specified error tolerance DTOL to ascertain

the performance of the error control strategy. Note that for practical

computations that are not concerned with benchmarking, MAXITS would typically

be set in the range 5---10 to avoid significant numbers of wasted iterations.

6.3.1 Drained and Undrained Analysis of Thick Cylinder

Drained and undrained loading conditions represent extremes of consolidation

behaviour and can be used to validate finite element models. For real soils, these

two modes of deformation are caused, respectively, by extremely slow and

extremely fast loading rates. In this context, the terms “slow” and “fast” have

different meanings for different materials and need to be defined relative to the

soil permeability.

Following Small (1977), the drained and undrained predictions of an elastoplastic

consolidation formulation may be verified by using exact analytical solutions for

the expansion of a thick cylinder of soil. Under undrained loading, the cylinder

deforms at constant volume and its behaviour corresponds to that of an

elastoplastic Tresca material. The complete load-deformation response and

internal stress distribution for this condition has been given by Hill (1950). More

recently, the fully drained analytical solution, which assumes a Mohr-Coulomb

material and contains the Hill solution as a special case, has been presented by

Yu (1992). As discussed in detail by Small (1977), the material properties for the

two different types of loading are not independent and must satisfy the relations

Eu= 3E′2(1+ ν′) (6.6)

cu∕c′ = 2 Nφ ∕(1+Nφ) (6.7)

where the subscript u denotes an undrained quantity and

Nφ= (1+ sinφ)∕(1– sinφ)

227Chapter 6

These equations, together with the incompressibility condition, govern the

parameters that must be used in the Hill solution when it is compared to the

undrained consolidation results obtained with a fast loading rate. It is also

important to note that the undrained consolidation analysis must be performed

with a zero dilation angle in order to avoid large strength gains which are caused

by excessive volume changes.

The geometry, boundary conditions, and axisymmetric finite element mesh used

to model the thick cylinder are shown in Figure 6.22. The drained parameters

smooth / impermeable

smooth / impermeable

axis ofsymmetry

Figure 6.22 Expansion of thick cylinder.

q

a

b= 2a

undrained: inner and outer boundaries impermeabledrained: inner and outer boundaries permeable

assumed in the finite element study are

E′∕c′ = 200 ν′ = 0.0 φ′ = 30˚ ψ′ = 0˚

Equations (6.6) and (6.7), together with the constant volume condition, give the

undrained parameters required for the Hill (1950) solution as

Eu∕cu= 346.4 νu= 0.49999 φu= 0˚ ψu= 0˚

For this set of material properties, the drained and undrained collapse pressures

of the cylinder are given, respectively, by the expressions q∕c′ = 1.02 and

q∕c′ = 1.2 (or q∕cu= 1.4), where q is the uniform pressure applied to the inner

surface of the cylinder.

228Chapter 6

To account for the effects of the soil permeability, the rate at which the load is

imposed on the cylinder is defined in terms of the dimensionless quantity

ω=Δq∕c′ΔTv

where

ΔTv=cvΔta2

In the above equation, cv is the usual one dimensional consolidation coefficient

and a is the internal radius of the cylinder.

To compute the time-stepping error in the displacements for the various

consolidation runs, a set of reference displacements were computed for both the

drained and undrained analyses. These were found using the second order method

of Thomas and Gladwell (1988a) with 10,000 equal size time increments and a

Newton-Raphson iteration tolerance of ITOL=10---8.

No. No. subincrements No. iterations CPUDTOL

No.coarse timeincrements

Successful FailedSuccessfulsteps

Failedsteps

Max*

CPUtime(s)

10---2 1 18 10 50 40 7 2.1910 23 6 62 20 4 2.51

10---3 1 56 18 129 53 7 3.94

10 63 21 141 58 3 4.7010---4 1 250 95 496 198 7 13.9

10 248 100 494 203 3 14.4

Table 6.10 Results for undrained loading of thick cylinder using automaticscheme and uniform coarse increments.

(* max no. iterations in successful or unsuccessful time steps)

In the first set of analyses, the automatic algorithm is used to predict the undrained

response of the thick cylinder. These runs impose a rapid loading rate of ω= 104

to simulate undrained conditions and use the Newton-Raphson iteration scheme

229Chapter 6

to solve the incremental equations for each time step. As shown in Table 6.10,

results are generated for analyses using 1 and 10 coarse time steps with error

tolerances of DTOL=10---2, 10---3 and 10---4. In all cases, the automatic algorithm

selects a similar number of subincrements for analyses performed with the same

displacement tolerance. With DTOL=10---2, for example, the runs with 1 and 10

coarse time steps generate, respectively, 18 and 23 successful subincrements.

Similarly, for a tolerance of DTOL=10---3, the analyses employs 56 and 63

successful subincrements.

The relatively high number of failed subincrements for this example is a

consequence of the fact that the undrained deformation response of the cylinder

is particularly sensitive to the effects of Gauss points turning plastic. Because of

the abrupt change in behaviour that is imposed by an elastic perfectly plastic

model, an elastic-plastic transition for a single Gauss point has a pronounced affect

on the value of the local error indicator for this particular problem. This impacts

on the present scheme as consecutive substeps are allowed to double in size to

enable rapid growth of the time step during later stages of the consolidation

process. One possible strategy for reducing the number of failed substeps is to

limit this growth factor to a lower value of around ten percent, as was done in the

automatic load stepping algorithm of Chapter 4. Since most of the failed steps

will occur during application of the load, this restriction would not need to be

enforced during the consolidation phase. Other cases considered later in this

Section do not exhibit a large proportion of failed steps, so this refinement has

not been incorporated in this Thesis.

On average, only two to three iterations are required for each successful substep

of the automatic scheme. This reflects the ability of the Newton-Raphson

algorithm to provide rapid convergence when it used with an appropriate time

step. As expected, the maximum number of iterations for a given analysis is

highest for cases where the maximum load is applied in a single coarse time step.

230Chapter 6

Somewhat surprisingly, the average number of iterations for a failed step is fairly

low and typically lies somewhere between two and three.

The load displacement curve for the undrained analysis with a single coarse time

increment and a displacement tolerance of DTOL=10---2 is plotted in Figure 6.23.

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

1.8

2.0

0.0 0.5 1.0 1.5 2.0 2.5 3.0

qc′

100(inner radius displ.∕a)

Figure 6.23 Pressure versus displacement for drained/undrained loading ofthick cylinder.

1.02c′

1.2c′

Eu∕cu= 346.4, νu= 0.49

undrained ω= 104

undrained Hill (1950)

DTOL=10---2, single coarse step

φu= 30,ψu= 0drained Yu (1992)

drained ω= 10−2E′∕c′ = 200, ν′ = 0φ′ = 30,ψ′ = 0 ⎨⎧⎩

Although this is a particularly severe test, the algorithm successfully reduces and

then adjusts the subincrement size to reflect the load-displacement behaviour of

the cylinder. The numerical results match the analytical solution of Hill (1950)

with acceptable accuracy over all of the loading range and predict the exact

collapse pressure precisely. The small oscillations which occur upon initial

yielding of the cylinder may be eliminated by using a tighter value of DTOL. A

detailed picture of the subincrement sizes adopted by the automatic algorithm for

231Chapter 6

this example is shown in Figure 6.24. As expected, the algorithm initially chooses

0

0.030

0.060

0.090

0.120

0.150

0.180

0.210

0.240

5 10 15

ΔTv

Figure 6.24 Subincrement size selection for undrained loading ofthick cylinder.

subincrement number

1 coarse time stepDTOL=10---2

large subincrements in the elastic range and then reduces the step size after the

onset of plasticity.

The variation of the time-stepping error as the cylinder is loaded, for runs with

10 coarse time steps and various values of DTOL, is shown in Figure 6.25. In this

plot, initial plastic yielding of the cylinder occurs at q∕c′ = 0.65 which corresponds

to 54 percent of the total pressure applied. Prior to this threshold being reached,

the behaviour is elastic and the algorithm does not need to subincrement the

coarse load steps. This results in a constant value of the time-stepping error which

is independent of DTOL. After the onset of plastic yielding, the time-stepping

error in the displacements grows to a level which is close to the desired tolerance.

In the case of the drained analysis of the thick cylinder, a much slower loading

rate of rate of ω= 10–2 is used to apply the internal pressure. As in the undrained

example, the automatic algorithm is employed to predict the response using 1 and

10 coarse time steps with tolerances of 10---2, 10---3 and 10---4. Results for these

232Chapter 6

10 20 30 40 50 60 70 80 90 100

percentage of total pressure applied

Figure 6.25 Variation of displacement load path error with load level forundrained loading of a thick cylinder.

DTOL= 10−3

DTOL= 10−2

DTOL= 10−4

10---2

10---3

10---4

10---5

10---6

uerror

analyses are summarised in Table 6.11. The observations to be made from these

statistics are similar to those made for the undrained case, except that roughly

double the number of substeps are required and the proportion of failed substeps

is smaller.

No. No. subincrements No. iterations CPUDTOL

o.coarse timeincrements

Successful FailedSuccessfulsteps

Failedsteps

Max

C Utime(s)

10---2 1 38 15 92 49 9 3.1910 41 12 101 30 4 3.45

10---3 1 115 33 267 99 9 7.53

10 116 25 267 68 4 7.3210---4 1 411 97 815 209 9 19.8

10 421 103 834 210 3 20.6

Table 6.11 Results for drained loading of thick cylinder using automatic schemeand uniform size coarse increments.

Figure 6.23 indicates the drained numerical deformation response for the case of

a single coarse load step with DTOL=10---2. This is in good agreement with the

233Chapter 6

analytical solution of Yu (1992) over all of the loading range and accurately

predicts the exact collapse pressure. As in the undrained case, some oscillations

are observed immediately after the onset of plastic yielding, but these may be

eliminated by using a smaller value of DTOL. Figure 6.26 illustrates the

10 20 30 40 50 60 70 80 90 100

percentage of total pressure applied

Figure 6.26 Variation of displacement load path error with load level fordrained loading of thick cylinder.

DTOL= 10−3

DTOL= 10−2

DTOL= 10−4

10---2

10---3

10---4

10---5

10---6

10---7

10---8

uerror

time-stepping errors in the displacements for the runs using 10 coarse load steps

and various values of DTOL. Under drained conditions, initial yielding occurs at

q∕c′ = 0.58 which corresponds to 56.8 percent of the total applied pressure. Prior

to this point, where the behaviour is elastic, the automatic scheme chooses very

small subincrements, which causes the time-stepping errors to lie well below their

desired tolerances. These small substeps are clearly seen in Figure 6.27, which

shows a bar chart of the size of each substep. The use of such small subincrements

in the elastic range is surprising, but is explained by the fact that the numerical

response is always undrained in the first load step. This may be seen from the

governing finite element equations, and is a direct consequence of assuming that

the initial displacements and pore pressures are equal to zero at t= 0. The net

234Chapter 6

0

5

10

15

20

25

30

10 20 30

ΔTv

Figure 6.27 Subincrement selection for drained loading of thick cylinder.

subincrement number

result of this is that the automatic scheme adopts small steps during the transition

from the undrained state to the drained state, after which it behaves as expected.

This phenomenon occurs only when a consolidation analysis is performed with an

extremely slow loading rate in an effort to mimic drained behaviour. Following

the onset of plastic yielding, the time-stepping errors shown in Figure 6.26 are very

close to their desired tolerances.

6.3.2 Undrained Analysis of Strip Footing

The intent of this Section is to investigate the ability of the consolidation

formulation to predict the undrained deformation response, and hence the

ultimate collapse load, for a smooth flexible strip footing. The study is motivated

by the work of Small (1977) who noted that a Biot consolidation formulation, when

used with a simple elastoplastic Mohr-Coulomb model, is unable to model

undrained behaviour accurately unless a zero dilation angle is used. This

observation follows from the fact that a finite dilation angle inevitably causes a

change in volume, and a consequent gain in strength, upon plastic shearing. In

235Chapter 6

finite element consolidation analysis, this effect is manifested by a “hardening”

deformation response which does not exhibit a precise failure load or agree with

the predictions from a simple elastoplastic computation.

The finite element mesh and boundary conditions used to model the flexible strip

footing are shown Figure 6.28. The drained Mohr-Coulomb parameters assumed

undrained:impermeabledrained: permeable

16B

8B

B

Figure 6.28 Flexible strip footing on elastoplastic layer.

smooth/impermeable

smooth/impermeable

smooth / impermeable

q

in the consolidation analyses are

E′∕c′ = 200 ν′ = 0.3 φ′ = 20˚ ψ′ = 0˚–20˚

The conventional elastoplastic analysis for this problem models undrained

behaviour using a Tresca yield criterion with a zero friction angle and a zero

dilatancy angle. From equations (6.6) and (6.7), it follows that the undrained

parameters for this case must be set as

Eu∕cu= 245.6 νu= 0.499 φu= 0˚ ψu= 0˚

The undrained collapse pressure for the footing is given by the well known Prandtl

formula q∕cu= 5.14 (or q∕c′ = 4.83), where q denotes the applied pressure on

the footing.

236Chapter 6

For the consolidation analyses, the rate at which the load is imposed on the footing

is defined in terms of the dimensionless quantity

ω=Δq∕c′ΔTv2

(6.8)

where the time factor is

ΔTv2=cv2ΔtB2

(6.9)

In the above equation, cv2 is the two dimensional consolidation coefficient defined

by

cv2=kE′

2γw(1+ ν′)(1− 2ν′)

and B is the half-width of the footing.

The undrained response of the footing, obtained from a conventional elastoplastic

analysis with the automatic load-stepping algorithm described in Chapter 4, is

shown in Figure 6.29. This solution clearly asymptotes toward the exact collapse

pressure of q= 4.83c′ = 5.14cu and may be compared against the results of

various consolidation analyses which are also plotted. The latter were performed

using the automatic scheme with a rapid loading rate of ω= 150 to simulate

undrained conditions. In the special case of ψ′ = 0˚, the consolidation analysis

was run with the initial stiffness algorithm, an iteration tolerance of ITOL=10---4,

and a displacement tolerance of DTOL=10---2. This solution closely matches that

from the conventional undrained analysis and accurately predicts the exact

collapse pressure of q= 5.14cu. The analyses for the remaining dilation angles

of 1˚, 5˚ and 20˚ were conducted using tangent stiffness iteration with tolerances

of ITOL=10---6 and DTOL=10---3. No detailed statistics for these runs are

presented, but they typically generated between 100 and 200 time steps when

performed with a single coarse time step. Figure 6.29 confirms the findings of

Small (1977), who concluded that consolidation analyses with a nonzero dilation

237Chapter 6

0

1

2

3

4

5

6

7

8

9

0 5 10 15

qc′

centre-line displacement∕B

Figure 6.29 Pressure versus displacement for undrained loading of flexiblestrip footing on elastoplastic layer.

ψ′ = 20˚

ψ′ = 1˚

ψ′ = 5˚

consolidation, ν′ = 0.3 , E′∕c′ = 200φ′ = 20˚, ω= 150

ψ′ = 0˚

elastoplastic, νu= 0.499 ,Eu∕cu= 245.6φu= ψu= 0˚

q= 4.83c′ = 5.14cu

angle may seriously overestimate the undrained collapse load. Even for a dilation

angle of 1 ,̊ there is a noticeable “hardening” of the load-deformation response

with no discernible point of failure.

6.3.3 Strip Footing with Associated Flow Rule

All of the elastoplastic problems considered so far in this Chapter have been

concerned with the material response only over the loading range. No cases have

been analysed in which the load is applied and then held constant while

consolidation takes place. This Section is concerned with the consolidation

238Chapter 6

behaviour of a smooth flexible strip footing resting on an elastoplastic layer. The

layer is modelled by a Mohr-Coulomb yield criterion with an associated flow rule.

Although the inadequacy of such a flow rule was highlighted in the previous

example, the analyses are performed to enable a direct comparison of the results

with those of Manoharan and Dasgupta (1995). Detailed statistics are given for

both the automatic and backward Euler methods, so that their accuracy and

efficiency can be compared.

The mesh and material properties are identical to those considered in the previous

Section (see Figure 6.28), except that the top boundary is now free to drain and

the friction and dilation angles are set to φ′ = ψ′ = 20˚. The load is applied to

the footing as a uniform prescribed pressure and is imposed over the initial period

of Tv20= 0.01, where Tv2 is the time factor defined by equation (6.9). The

subsequent consolidation is modelled up to a total time factor of Tv2= 1000.

Unless noted otherwise, all analyses are performed using a total load of

q0∕c′ = 10, where q0 is the maximum pressure applied to the footing at time

Tv2= Tv20. The reference solutions for this case were again computed using the

second order scheme of Thomas and Gladwell (1988a). This analysis used 1,000

equal size time steps over each log cycle of the time factor and an iteration

tolerance of ITOL=10---7. The total time factor increments adopted for each log

cycle are shown in Table 6.12.

Time factor increment (ΔTv2) Total time factor (Tv2)

Loading 0.001 0.0010.009 0.01

0.09 0.1

0.9 1.0Consolidation 9.0 10.0

90.0 100.0

900.0 1000.0

Table 6.12 Log cycles for analysis of elastoplastic strip footing.

239Chapter 6

In the first series of runs with the automatic scheme, three different load levels

are analysed using a displacement tolerance of DTOL=10---2. The load levels of

q0∕c′ = 5 , q0∕c′ = 10 and q0∕c′ = 15 are all applied to the footing in a single

coarse step. A single coarse step is also used to model the subsequent

consolidation.

q0∕c′ No. subincrements No. iterations CPUSuccessful Failed Successful* Failed* Max time (s)

5 13+30=43 6+0=6 51+88=139 42+0=42 13 113

10 17+27=44 6+0=6 78+102=180 35+0=35 12 167

15 21+30=51 6+0=6 101+164=265 39+0=39 11 273

Table 6.13 Results for strip footing using automatic scheme with 2 coarse timeincrements and DTOL=10---2.

(* no. iterations in successful and failed subincrements)

Detailed statistics for these analyses, and a plot of their corresponding transient

displacements, are shown in Table 6.13 and Figure 6.30 respectively. The latter

indicates that the numerical settlement predictions from the automatic method are

almost identical to the finite element results recently published by Manoharan and

Dasgupta (1995). These authors used a mesh composed of 8-noded quadrilaterals

with a total of 70 elements and 245 nodes. The mesh used here, which is shown

in Figure 6.28, was constructed from their model by replacing each quadrilateral

with four triangles to give a grid of 280 elements and 595 nodes.

Even when it is used with a single coarse load step to model the loading and

consolidation phases of the footing, the automatic scheme generates few failed

substeps. Table 6.13 indicates that the ratio of failed to successful steps is typically

twelve to fourteen per cent, which is an acceptable rejection rate for practical

computations. Interestingly, all of the failures occur in the loading phase and are

primarily generated when the algorithm tries to establish an initial step size.

The iteration counts presented in Table 6.13 highlight the effectiveness of

combining the automatic step control mechanism with the Newton-Raphson

240Chapter 6

Figure 6.30 Settlement versus time factor for elastoplastic strip footing

Tv2=cv2t

B2

q0∕c′ = 5

q0∕c′ = 15q0∕c′ = 10

Manoharan and Dasgupta (1995)

10010---110---2 10+210---3 10+1 10+3

0.0

10

20

30

10---4

40

50

100(settlement/B)

ν′ = 0.3 , E′∕c′ = 200φ′ = ψ′ = 20˚

2 coarse steps, DTOL=10---2

solution strategy. Each successful substep requires, on average, about five

iterations. Not surprisingly, the average number of iterations for failed substeps

is greater than this due to the fact that these steps usually occur when the plastic

strain increments are large. Since the average number of iterations for a failed

substep is about seven, and the maximum number of iterations for any substep

always exceeds ten, the efficiency of the automatic scheme would be improved

considerably by setting the maximum iteration limit, MAXITS, to about five. As

discussed in Section 6.3, MAXITS is currently set to a large number in this Thesis

so that it does not interfere with the step size predicted by the local error

estimator.

The distribution of subincrement sizes for the analysis with q0∕c′ = 10 is shown

in Figure 6.31. As expected, the smallest substep of ΔTv2= 1.3× 10−4 occurs

241Chapter 6

Figure 6.31 Subincrement size selection for consolidation of elastoplastic stripfooting

0.0001

0.0010

0.0100

0.1000

1.0000

10.0000

100.0000

10 20 30 40

ΔTv

subincrement number

during the loading phase, where the changes in the deformations and pore

pressures are rapid. The substeps are also small immediately after the maximum

load is reached, but then grow quickly once the excess pore pressures are partially

dissipated. Toward the end of the consolidation period, where the excess pore

pressures are close to zero, the automatic scheme selects the largest time step of

ΔTv2= 403. Thus, over the total consolidation interval considered, the time step

is increased by more than six orders of magnitude. Time step distributions such

as the one indicated in Figure 6.31 are difficult to select by hand, and would not

be obvious even to an experienced analyst.

To further investigate the performance of the automatic scheme, the footing was

analysed for the case of q0∕c′ = 10 with a range of displacement tolerances and

a logarithmically varying set of coarse time steps. The latter are listed in

Table 6.12 and the results from these runs are shown in Table 6.14. For the case

of DTOL=10---2, the automatic scheme generates 49 successful substeps, 5 failed

242Chapter 6

substeps, and uses 163 seconds of CPU time. With the same value of DTOL, but

using two coarse time steps, the corresponding statistics shown in Table 6.13 are

44 successful steps, 6 failed steps and 167 seconds of CPU time. This comparison

again confirms that the performance of the automatic scheme is largely

independent of the size and distribution of the initial coarse time steps.

DTOLNo. coarsetime incs per

No. subincrements No. iterations insuccessful

CPUtimeDTOL time incs. per

log cycle Successful* Failed*successfulsubincrements

time(s)

10---2 1 19+30=49 5+0=5 83+114=197 16310---3 1 51+58=109 5+0=5 158+167=325 21510---4 1 193+149=342 25+3=28 429+342=771 485

Table 6.14 Results for elastoplastic strip footing using automatic scheme withlogarithmic coarse time steps.

The data in Table 6.14 shows that reducing the error tolerance DTOL reduces the

average number of iterations required for each time step. This is expected, as

smaller values of DTOL lead to smaller time steps. In this example, cutting DTOL

from 10---2 to 10---4 results in the average number of iterations being decreased

from just under four to just over two. For the same reason, smaller values of

DTOL also lead to a reduced proportion of failed substeps. These two effects

explain why the CPU time does not increase dramatically when DTOL is tightened

by an order of magnitude. The time-stepping errors in the displacements for each

of the analyses in Table 6.14 are plotted in Figure 6.32. For all values of DTOL,

the automatic scheme again successfully constrains the error to an approximately

constant value which lies near the desired tolerance.

To assess the performance of a more traditional solution method, the footing

response is now analysed using the backward Euler scheme for the load case of

q0∕c′ = 10. In this example, the time factor increments, listed in Table 6.12, are

subdivided using 1,10, 100 and 1,000 equal-size intervals to give the time steps

shown in Table 6.15.

243Chapter 6

Tv2=cv2 tB2

100 incs.per log cycle

10 incs.per log cycle

1000 incs. per log cycle

Figure 6.32 Temporal discretisation error in displacements versus time factorfor elastoplastic strip footing (logarithmic coarse time steps).

10010---110---2 10+210---3 10+1 10+3

10---2

10---3

10---4

10---5

10---6backward Euler

automatic

uerror

DTOL=10---4

DTOL=10---3

DTOL=10---2

10---1

1 inc. per log cycle

No. incrementsper log cycle

Total no.increments

No. of iterationsCPUtime

Loading Consol.increments

Total Max. (s)1 1 1+6=7 25+33=58 23 88

10 10 10+60=70 72+145=217 7 153

100 100 100+600=700 367+917=1284 4 7271000 1000 1000+6000=7000 2564+5180=7744 3 4370

Table 6.15 Backward Euler results for elastoplastic strip footing usinglogarithmic increment sizes.

The results for these analyses, also summarised in Table 6.15, show that the

Newton-Raphson scheme may require a large number of iterations if the time

244Chapter 6

discretisation is very coarse. For the crudest analysis with one time step in the

loading phase and six time steps in the consolidation phase, up to 23 iterations are

needed to reach convergence and the average iteration count is just over eight.

This is in contrast to the 7000 increment run which gives maximum and average

iteration counts of three and one respectively. As expected, the largest average

iteration counts for all the analyses occur during the loading phase. Figure 6.32

summarises the temporal discretisation error at the end of each log cycle of the

time factor for the various runs. This plot indicates that even with the crudest

discretisation of one time step per log cycle, the backward Euler method gives a

time-stepping error of just two percent. The results also confirm the expected first

order accuracy of this type of solution strategy.

The results in Table 6.14, Table 6.15 and Figure 6.32 can be used to compare the

relative efficiency of the automatic and backward Euler methods for the footing

problem. The Figure indicates that these two schemes, when used with

DTOL=10---2 and one increment per log cycle, respectively, give similar maximum

time stepping errors. The CPU time for the backward Euler method, however,

is only 88 seconds as opposed to 163 seconds for the automatic scheme. For more

accurate analyses, the relative performance of the automatic scheme improves

significantly and is competitive with that of the backward Euler algorithm. This

can be seen by comparing the timing statistics for the automatic analysis with

DTOL=10---4 against the timing statistics for the 700 increment backward Euler

run. The former, although marginally more accurate, requires only 485 CPU

seconds as opposed to 727 CPU seconds for the latter, a saving of approximately

33 percent. When making this type of comparison, it should be remembered that

the backward Euler scheme will usually need to be run with a range of different

time steps in order to determine when the time-stepping error is negligible. The

automatic scheme, however, need only be run once.

245Chapter 6

6.3.4 Strip Footing with Nonassociated Flow Rule

The results shown in Section 6.3.2 highlighted the need to employ a nonassociated

flow rule when using a simple Mohr-Coulomb model with a Biot consolidation

formulation. In particular, it is necessary to use a non-dilatant flow rule to obtain

solutions which match those from a conventional elastoplastic analysis for the

limiting case of undrained loading. This Section examines the ability of the

consolidation formulation to accurately predict the drained collapse pressure for

a flexible strip footing subjected to a very slow rate of loading. The important

influence of the iteration tolerance on the efficiency of the initial stiffness

algorithm is also investigated.

The geometry, boundary conditions and finite element mesh for the footing are

identical to those used in the preceding example (see Figure 6.28). The drained

soil parameters adopted for the Mohr-Coulomb model are similar to those

employed in Section 6.3.2, except that the dilation angle is always zero. Thus

E′∕c′ = 20 ν′ = 0.3 φ′ = 20˚ ψ′ = 0˚

and the corresponding Prandtl collapse pressure is q∕c′ = 14.83. The

conventional elastoplastic analysis for this problem uses the above parameters, and

the rounded Mohr-Coulomb yield surface described in Chapter 2, to model the

drained state. One thousand equal size pressure increments are imposed on the

footing and an initial stiffness algorithm is employed to solve the governing

equations. The automatic load-stepping scheme of Chapter 4 could not be used

for this case, as the tangent stiffness matrix becomes ill-conditioned when ψ′ = 0.

The consolidation analyses are performed with loading rate values ranging from

ω= 0.015 to ω= 150, where ω is again defined by equation (6.8). The former

case generates very small excess pore pressures and thus models the drained

condition very closely. As shown in Section 6.3.2, the higher setting accurately

simulates undrained loading. All of the consolidation runs are performed with two

246Chapter 6

coarse load steps, DTOL=10---2, and an initial stiffness iteration tolerance of

ITOL=10---4.

The results for the analyses, shown in Figure 6.33, indicate that the drained

0

2

4

6

8

10

12

14

16

0 10 20 30 40 50 60 70 80 90 100

elastoplastic

ν′ = 0.3 , E′∕c′ = 20φ′ = 20˚, ψ′ = 0˚

consolidation

qc′

Figure 6.33 Pressure versus displacement for flexible strip footing onelastoplastic layer with varying load rates.

undrained

drained

ω= 150

ω= 15

ω= 0.015

ω= 1.5

centre-line displacement∕B

⎨⎧⎩

q= 14.83c′

pressure-displacement responses from the consolidation and conventional

elastoplastic methods are in close agreement over all of the loading range. Both

techniques accurately predict the Prandtl collapse pressure of q∕c′ = 14.83, which

is valid for a soil with an associated flow rule. As expected, the consolidation

results give a stiffer response as the loading rate is reduced. For the undrained

case, which is modelled using the maximum loading rate of ω= 150 , the results

are the same as those discussed in Section 6.3.2.

247Chapter 6

As mentioned previously, the initial stiffness variant of the consolidation algorithm

is useful for soil models which involve nonassociated flow rules, where the tangent

stiffness matrix may become ill-conditioned. When performing consolidation

analyses with the initial stiffness method, the setting of the iteration tolerance

ITOL has a dramatic effect on the total CPU time. It is natural to suggest that

the value of this parameter should, in some way, be linked to the value of DTOL,

since there is little sense in performing very accurate iterations if the chief source

of error is due to the use of large time steps.

DTOL ITOLNo. subincrements No. iterations

in successfulCPUtimeDTOL ITOL

Successful Failedin successfulsubincrements

time(s)

10---2 10---2 15+47=62 5+3=8 20+48=60 9910---3 18+41=59 6+0=6 109+83=192 14610---4 19+38=57 7+0=7 325+116=441 27610---5 19+38=57 7+0=7 670+234=904 497

10---3 10---3 50+72=122 11+0=11 65+95=160 13010---4 52+71=123 13+0=13 323+148=471 19110---5 53+71=124 15+0=15 1051+204=1255 34210---6 53+71=124 16+0=16 2108+308=2416 557

10---4 10---4 264+184=448 90+5=95 346+325=671 49610---5 203+175=378 59+1=60 1189+393=1582 52610---6 206+173=379 59+0=59 3936+501=4437 86610---7 207+173=380 50+0=50 8250+721=8971 1356

Table 6.16 Results for strip footing on nonassociated layer using automaticscheme with initial stiffness iteration.

To investigate this question, the footing problem described above is analysed using

a range of iteration tolerances. The geometry, boundary conditions and mesh are

unchanged, as are the material properties. The only difference is that the

maximum load of q0∕c= 4.2 is now applied over the time step ΔTv2= 0.01 and

consolidation is allowed to take place until Tv2= 1000. In addition, one step per

log cycle is used to specify the size of each initial coarse time increment. The

reference displacements for this case are obtained using the automatic scheme

248Chapter 6

with tolerances of ITOL=10---7 and DTOL=10---6. These results contain very small

time-stepping errors because of the large number of time steps that are generated.

The results of the iteration study for the footing are summarised in Table 6.16.

Data is presented for DTOL settings of 10---2, 10---3 and 10---4, with each of these

tolerances being analysed for a range of ITOL values. Because the rate of

convergence of the initial stiffness method is only linear, the iteration tolerance

is observed to have a marked influence on the CPU times recorded. In the case

with DTOL=10---2, for example, the CPU time increases from 99 seconds for

ITOL=10---3 to 497 seconds for ITOL=10---5. Similar growth factors are observed

for all other values of DTOL considered. The extra expense incurred by using a

stringent iteration tolerance is justified only if the accuracy of the resulting solution

is greatly improved. To investigate whether this is the case, plots of the errors for

the analyses with DTOL equal to 10---2 and 10---3 are shown, respectively, in

Figure 6.34 and Figure 6.35. Both of these plots demonstrate that the

Tv2=cv2 tB2

Figure 6.34 Temporal discretisation error in displacements versus time factorfor elastoplastic strip footing using automatic algorithm and initial stiffness

iteration.

ITOL= 10–4

ITOL= 10–5

ITOL= 10−2

ITOL= 10−3

10010---110---2

100

10---1

10---2

10---3

101 10310210---4

q0∕c′ = 4.2

uerror

ν′ = 0.3 , E′∕c′ = 200φ′ = 20˚, ψ′ = 0˚

1 coarse step per log cycle, DTOL=10---2

249Chapter 6

Figure 6.35 Temporal discretisation error in displacements versus time factorfor elastoplastic strip footing using automatic algorithm and initial stiffness

iteration.

ITOL= 10–6

ITOL= 10−4

ITOL= 10−5

Tv2=cv2 tB2

ITOL= 10−3

ITOL= 10−1

ITOL= 10−2

10010---110---2

10---1

10---2

10---3

10---4101 103102

uerror

q0∕c′ = 4.2ν′ = 0.3 , E′∕c′ = 200φ′ = 20˚, ψ′ = 0˚

1 coarse step per log cycle, DTOL=10---3

displacement error does not continue to decrease as the iteration tolerance is

tightened. In the analysis with DTOL=10---2, for example, there is no consistent

improvement in the accuracy of the analysis once ITOL is reduced below a value

of around ITOL=10---4. Similarly, for DTOL=10---3, there is no discernible

reduction in the displacement error for iteration tolerances tighter than 10---5.

These results suggest that the time-stepping error typically reaches a minimum

value when ITOL≈ DTOL∕100. Choosing a value below this threshold does not

reduce the overall solution error, and only serves to increase the CPU time. The

results for various runs with ITOL set to DTOL/100 are shown in Figure 6.36. In

all cases, the displacement error is held approximately constant over the

250Chapter 6

Tv2=cv2 tB2

Figure 6.36 Temporal discretisation error in displacements versus time factorfor elastoplastic strip footing using automatic algorithm and initial stiffness

iteration.

DTOL= 10–3

DTOL= 10–2

DTOL= 10–4

10010---110---2

10---1

10---2

10---3

10---4

10---5101 103102

uerror

q0∕c′ = 4.2ν′ = 0.3 , E′∕c′ = 200φ′ = 20˚, ψ′ = 0˚

1 coarse step per log cycle, ITOL=DTOL/100

consolidation interval and lies near the desired tolerance DTOL. Although no

results are presented here, this method of setting the iteration tolerance has also

been found to work well for consolidation analyses which employ Newton-Raphson

iteration.

6.4 CONCLUSIONS

The automatic time incrementation scheme has been used successfully to predict

the behaviour of a number of problems involving the consolidation of elastic and

elastoplastic materials. These applications prove that the proposed algorithm is

not only robust and efficient, but also able to constrain the temporal discretisation

error to lie near a desired tolerance. A major advantage of the new algorithm is

that it removes the need to select the time increments in consolidation analysis

by trial and error. Moreover, the behaviour of the scheme is largely insensitive

251Chapter 6

to the size and distribution of the coarse time steps that are required to start the

analysis.

The pitfalls of adopting an associated Mohr-Coulomb model in consolidation

analysis have been highlighted. For rapid rates of loading, this yield surface should

be used with a nonassociated flow rule and a zero angle of dilatancy to obtain

reliable predictions of soil behaviour.

Finally, it was shown that there is little benefit in using excessively stringent

iteration tolerances in elastoplastic consolidation calculations. For the automatic

scheme, a simple rule is proposed which ties the value of the iteration tolerance

to the displacement error tolerance. This rule helps to minimise the number of

wasted iterations, yet ensures that the desired accuracy requirements are met.

252Chapter 6

253Chapter 7

CHAPTER 7

CONCLUDING REMARKS

254Chapter 7

7.1 SUMMARY

In this Thesis, various algorithms for finite element analysis of consolidation and

elastoplasticity have been developed. This includes the derivation of a smooth

approximation to the Mohr-Coulomb yield criterion which has no singularities in

its gradient. The main thrust of this Thesis, however, has been concerned with

the development of techniques for controlling the linearisation error which is

caused by the use of finite size steps in the solution of nonlinear problems. This

work can be divided in to three distinct areas; namely, the integration of

elastoplastic stress-strain laws, the solution of the governing load-deflection

relations of elastoplasticity, and the solution of the coupled equations of

elastoplastic consolidation.

7.2 ROUNDED APPROXIMATION TO THEMOHR-COULOMB YIELD CRITERION

In Chapter 2, a fully rounded approximation to the Mohr-Coulomb yield criterion

was derived. The new yield criterion, which uses a trigonometric approximation

in the octahedral plane and a hyperbolic approximation in the Meridional plane,

eliminates all gradient singularities from the Mohr-Coulomb yield criterion. The

resulting yield surface is both continuous and differentiable at all stress states and

is able to model the true Mohr-Coulomb yield surface as closely as desired by

adjusting only two parameters. The approximation is designed so that the true

Mohr-Coulomb yield criterion can be recovered as a special case.

The use of the smooth yield criterion in finite element computations avoids many

of the numerical difficulties that arise due to discontinuous gradients. In

particular, the integration of the constitutive relationships, the correction of stress

states to the yield surface, and the calculation of the stiffness matrices is

straightforward.

255Chapter 7

7.3 INTEGRATION OF ELASTOPLASTICCONSTITUTIVE LAWS

In Chapter 3, a number of important refinements to enhance the efficiency and

robustness of the explicit integration scheme of Sloan (1987) were developed.

These enhancements include improved algorithms for determining the intersection

to the yield surface, handling a negative plastic multiplier, and the correction of

stresses which drift from the yield surface during integration. By using a measure

of the local truncation error, the explicit modified Euler scheme automatically

subincrements the applied strain increments if a specified error tolerance is

exceeded. The advantage of this approach is that the error in the integration of

the constitutive law can be controlled using a rational method.

In the remainder of Chapter 3, the performance of the adaptive modified Euler

scheme is compared to that of the implicit backward Euler return scheme. As part

of this comparison, the performance of an explicit single step forward Euler

scheme and a single step implicit backward Euler scheme were also considered.

Both of these schemes are common in the analysis of elastoplasticity.

The performance of the single step schemes for the Tresca and Mohr-Coulomb

constitutive relationships is similar, and thus there is little to distinguish between

the two. The accuracy of these schemes is also comparable to the more

complicated modified Euler and backward Euler return procedures when the

strain increments are not excessively large. Hence, for simple yield criteria with

small load increments, explicit and implicit single step algorithms provide an

accurate and efficient solution. These methods, however, are unsuitable for

problems that are loaded in the fully plastic range.

The adaptive modified Euler and implicit backward Euler return schemes both

provide an economical means of integrating rounded Tresca and Mohr-Coulomb

constitutive laws. Each procedure requires a similar amount of CPU time and give

stresses of a similar accuracy. The explicit modified Euler scheme, however, has

256Chapter 7

the added advantage that the error in the computed stresses may be controlled to

a desired level. This is of particular benefit when analysing highly nonlinear

problems where large strain increments may be encountered.

A further advantage of the explicit modified Euler scheme is its robustness.

Because the technique is non-iterative and unconditionally stable, it can be used

to integrate complicated constitutive laws with large strain increments. In contrast,

the implicit backward Euler return scheme is less reliable and convergence

problems can and do arise in the iterative process. A robust implementation of

this scheme must also use a simple substepping strategy to ensure convergence.

The substepping strategy used in this Thesis enabled large strain increments to be

integrated efficiently without fear of numerical difficulty. When applied to the

rounded Mohr-Coulomb yield function without substepping, the implicit backward

Euler method may require very small load steps in order to ensure convergence

of the iteration scheme. The results for the footing problems suggest that the

implicit schemes do not perform well in the vicinity of the corners of the Tresca

and Mohr-Coulomb yield criteria, even when they are rounded, and special

strategies may be required.

A final attraction of the adaptive explicit method is that it requires only first

derivatives (with respect to the stresses) of the yield surface and plastic potential.

The second derivatives needed for the implicit methods are both difficult and

expensive to compute for many geotechnical models.

7.4 SOLUTION OF ELASTOPLASTICLOAD-DISPLACEMENT RELATIONS

In Chapter 4, a scheme for the automatic solution of the governing load deflection

equations in elastoplasticity was developed. The aim of this scheme was to control

the linearisation error that arises due to the use of finite size load increments.

The scheme, which is non-iterative and treats the governing relationships as a

system of ordinary differential equations, is structured so that user defined coarse

257Chapter 7

load increments are automatically subdivided into smaller subincrements. The

size of these subincrements is chosen so as to hold the local truncation error in

each step of the solution to below a user specified tolerance. The truncation error

is measured by taking the difference between incremental solutions obtained from

the first order accurate Euler scheme and the second order accurate modified

Euler scheme. By extrapolating the local error estimate, the algorithm is able to

expand or contract the size of the subincrements to reflect the nonlinearity of the

solution. For subincrements in which the local truncation error exceeds the

specified tolerance, the load step is abandoned and the integration repeated using

a smaller subincrement. Otherwise, the subincrement is accepted, the solution

advanced, and a new subincrement size predicted.

The performance of the explicit modified Euler scheme was demonstrated through

the analysis of several elastoplastic boundary value problems. The scheme was

able to solve the governing relations so that the load path error in the final

displacements lies near the desired tolerance. Generally, the error in the

displacements was constrained to within an order of magnitude of the desired

tolerance. The technique controlled this error independently of the number of

coarse load increments supplied by the user. Even for the most severe analyses

in which the total load was specified as a single coarse increment, the scheme

successfully adjusted the subincrement sizes to achieve the specified accuracy in

the solution.

The speed of the automatic error control scheme compared favourably with the

conventional forward Euler scheme. Indeed, the average CPU time per step for

these two methods differed only marginally. The chief benefit of the automatic

scheme is that it removes the guess work involved in specifying the load increments

by hand. Moreover, it only uses small load increments where necessary. The

benefits of using an adaptive incrementation scheme were demonstrated by

considering the size and distribution of the automatically generated subincrements.

258Chapter 7

In the analysis of the thick cylinder, the automatic scheme maintained a constant

error in the solution by using uniform size increments throughout the range of

plastic deformation. The analysis of the trapdoor problem, however, resulted in

a series of subincrements which expanded in size throughout the loading range.

The new automatic scheme can be applied to a wide range of nonlinear problems,

not just those that occur in plasticity. The algorithm is particularly robust, and

can only fail if the tangent stiffness matrix becomes singular.

7.5 SOLUTION OF THE GOVERNING EQUATIONS INCONSOLIDATION

Chapter 5 details the development of automatic time stepping algorithms for

solving elastic and elastoplastic coupled consolidation problems. Unlike the other

governing equations considered in this Thesis, the governing equations of

consolidation are coupled and require the use of implicit integration methods to

ensure unconditionally stability. The procedures developed in Chapter 5 treat the

governing consolidation relations as a system of first order differential equations

and are based on the backward Euler method and the Thomas and Gladwell

(1988a) method. By choosing the integration parameters for these two methods

judiciously, the local truncation error for each time step can be computed very

cheaply. This permits the time steps to be controlled in a manner similar to that

developed for the governing equations of elastoplasticity. The automatic

consolidation algorithm operates by subdividing a series of user defined coarse

time increments and limits the local truncation error in the displacements to less

than a user specified tolerance. For the elastic case, the displacements and pore

pressures at the end of each subincrement are solved directly without the need for

iteration. For elastoplastic behaviour, however, the governing relationships are

nonlinear and a system of nonlinear equations must be solved to compute the

updates.

259Chapter 7

In Chapter 6, the automatic consolidation algorithm was applied to several

problems concerning the consolidation of elastic and elastoplastic materials. In

all cases, the scheme was able to constrain the global temporal error in the

displacements to lie near the desired tolerance. The elastoplastic thick cylinder

analysis confirmed the accuracy of the consolidation formulation for the limiting

cases of undrained and undrained behaviour. The example of a rapidly loaded

strip footing highlighted the need to use the Mohr-Coulomb yield surface with a

nonassociated flow rule for undrained conditions. In the last part of the Chapter,

other strip footing examples further confirmed the accuracy and speed of the

automatic consolidation formulation. All of the cases considered prove that the

behaviour of the automatic procedure is largely insensitive to the size and

distribution of the initial coarse time steps.

The performance of the automatic consolidation scheme compares favourably to

that of the conventional backward Euler scheme. To achieve solutions of similar

accuracy, the automatic and backward Euler schemes use a similar amount of

computational effort. The backward Euler scheme is marginally faster for a crude

analysis while the automatic scheme is much faster when an accurate solution is

required. The chief advantage of the automatic method is that it removes the need

to determine the time stepping error by an empirical trial and error procedure.

260Chapter 7

261References

REFERENCES

262References

Bell, A.L., ‘The lateral pressure and resistance of clay and the supporting power

of clay foundations’, Proc. I.C.E., 199, 223, 1915

Bergan, P.G, Horrigmoe, G., Krakeland, B. and Soreide, T.H., ‘Solution techniques

for nonlinear finite element problems’, International Journal for

Numerical Methods in Engineering, 12, 1677-1696, 1978

Bergan, P.G. and Soreide, T., ‘Solution of large displacement and instability

problems using the current stiffness parameter’, in Finite Elements in

Nonlinear Mechanics (Eds P.G. Bergan et al.) Tapir Press, 1978

Biot, M.A., ‘General theory of three dimensional consolidation’, Journal of Applied

Physics, 12, 155-164, 1941a

Biot, M.A., ‘Consolidation under a rectangular load distribution’, Journal of

Applied Physics, 12, 426-430, 1941b

Biot, M.A., ‘Theory of elasticity and consolidation for a porous anisotropic solid’,

Journal of Applied Physics, 26, 182-185, 1955

Biot, M.A., ‘General solution of the equation of elasticity and consolidation for

a porous material’, Journal of Applied Mechanics, 23, 91-96, 1956a

Biot, M.A., ‘Theory of deformation of a porous viscoelastic anisotropic solid’,

Journal of Applied Physics, 27, 459-467, 1956b

Biot, M.A., ‘Theory of stability and consolidation of a porous medium under initial

stress’, Journal of Mathematics and Mechanics, 12, 521-541, 1963

Booker, J.R. ‘The consolidation of a finite layer subject to surface loading’,

International Journal for Solids and Structures, 10, 1053-1065, 1974

Booker, J.R. and Small, J.C., ‘An investigation of the stability of numerical

solutions of Biot’s equations of consolidation’, International Journal of

Solids and Structures, 11, 907-917, 1975

263References

Borja, R.I., ‘Finite element formulation for transient pore pressure dissipation: A

variational approach’, International Journal of Solids and Structures, 22,

1201-1211, 1986

Borja, R.I., ‘Linearisation of elasto-plastic consolidation equations’, Engineering

Computations, 6, 163-168, 1989

Borja, R.I., ‘One-step and linear multistep methods for nonlinear consolidation’,

Computer Methods in Applied Mechanics and Engineering, 85, 239-272,

1991

Borja, R.I., ‘Composite Newton-PCG and quasi-Newton iterations for nonlinear

consolidation’, Computer Methods in Applied Mechanics and

Engineering, 86, 27-60, 1991

Bøstrøm, B., Svanø, G. and Nordal, S., ‘Computation of collapse load using mixed

finite element analysis’, in Proceedings of the Fifth International

Conference on Numerical Models in Geomechanics, Davos (Switzerland),

397-402, 1995

. Britto, A. and Gunn, M., Critical State Soil Mechanics Via Finite Elements, Ellis

Horwood, Chichester, 1987

Carter, J.P., Small, J.C. and Booker, J.R., ‘A theory of finite elastic consolidation’,

International Journal of Solids and Structures, 13, 467-478, 1977

Carter, J.P., Small, J.C. and Booker, J.R., ‘The analysis of finite elasto-plastic

consolidation’, International Journal for Numerical and Analytical

Methods in Geomechanics, 3, 107-129, 1979

Chiarella, C. and Booker, J.R., ‘The time-settlement behaviour of a rigid die

resting on a deep clay layer’, Quarterly Journal of Mech. and Applied

Mathematics, XXVIII, 317-328, 1975

264References

Christian, J.T. and Boehmer, J.W., ‘Plain strain consolidation by finite elements’,

Journal of the Soil Mechanics and Foundation Division of ASCE, SM4,

1435-1457, 1970

Cividini, A. and Gloda, G., ‘Compatible and equilibrium F.E for soil consolidation:

a preliminary comparison’, in Proceedings of the fourth International

Conference on Numerical Methods in Geomechanics, Edmonton,

Canada, 297-306, 1982

Conte, S.D. and de Boor, C., Elementary Numerical Analysis, An Algorithmic

Approach, 3rd edition, McGraw-Hill, 1980

Coulomb, C.A., ‘Application des règles de maximus au minimus à quelques

problèmes de statique relatifs à l´architecture’, Mém. Math. et Phys., 7,

p. 343, 1773

Crisfield, M.A., ‘A fast incremental/iterative solution procedure that handles

‘snap-through’ ’, Computers and Structures, 13, 55---62, 1981

Crisfield, M.A., ‘An arc-length method including line searches and accelerations’,

International Journal for Numerical Methods in Engineering, 19,

1269-1289, 1983

Crisfield, M.A., ‘Accelerating and damping the modified Newton-Raphson

method’, Computers and Structures, 18, 395-407, 1984

Crisfield, M.A., ‘Plasticity computations using the Mohr-Coulomb yield criterion’,

Engineering Computations, 4, 300-308, 1987

Crisfield, M. A., Non-linear Finite Element Analysis of Solids and Structures,

Volume 1, John Wiley and Sons, Chichester, 1991

Cryer, C.W., ‘A comparison of the three-dimensional consolidation theories of Biot

and Terzaghi’, Quarterly Journal of Mech. and Applied Mathematic, XVI,

401-412, 1963

265References

Davis, E. H. and Poulos, H.G., ‘Rate of settlement under two- and

three-dimensional conditions’, Geotechnique, 22, 95-114, 1972

. De Borst, R., Nonlinear Analysis of Frictional Materials, PhD thesis, University of

Delft, 1986

Den Heijer, C. and Rheinboldt, W.C., ‘On step length algorithms for a class of

continuation methods’, SIAM journal of Numerical Analysis, 18,

925---948, 1981

Galerkin, B.G., ‘Series solution of some problems of elastic equilibrium of rods

and plates’, (Russian), Vestn. Inzh. Tech., 19, 897-908, 1915

Ghaboussi, J. and Wilson, E.L., ‘Flow of compressible fluid in porous elastic

media’, International Journal for Numerical Methods in Engineering, 5,

419-442, 1973

Gibson, R.E., and McNamee, J., ‘A three dimensional problem of the

consolidation of a semi-infinite clay stratum’, Quarterly Journal of Mech.

and Applied Mathematic, 16, 115-127, 1963

Gibson, R.E., Schiffman, R.L. and Pu, S.L., ‘Plane strain and axially symmetric

consolidation of a clay layer on a smooth impervious base’, Quarterly

Journal of Mech. and Applied Mathematic, XXIII, 505-519, 1970

Gladwell, I. and Thomas, R.M., ‘Variable-order variable step algorithms for

second-order systems. Part 2: The codes’, International Journal for

Numerical Methods in Engineering, 26, 55-80, 1988b

Hencky, H., ‘Zur Theorie plastischer Deformationen und der hierurch im Meterial

hervorgerufenen Nebenspannungen’, in Proccedings of the 1st

International Congress on Applied Mechanics, Delft, 43-54, 1924

Hill, R., The Mathematical Theory of Plasticity, Clarendon Press, Oxford, 1950

266References

Hwang, C.T., Morganstern, N.R. and Murray, D.W., ‘On solutions of plane strain

consolidation problems by finite elements methods’, Canadian

Geotechnical Journal, 8, 109-118, 1971

de Josselin de Jong, ‘Applications of stress functions to consolidation problems’,

in Proceedings of the 4th International Conference on Soil Mechanics and

Foundation Engineering, London, 320-323, 1957

Kanok-Nukulchal, W. and Suaris, V.W., ‘An efficient finite element scheme for

elastic porous media’, International Journal of Solids and Structures, 48,

37-49, 1982

Krause, G., ‘Finite element schemes for porous media’, Journal of the Engineering

Mechanics Division, ASCE, 104(EM3), 605-620, 1978

Laursen, M. E. and Gellert, M., ‘Some criteria for numerically integrated matrices

and quadrature formulas for triangles’, International Journal for

Numerical Methods in Engineering, 12, 67-76, 1978

Lees, M., ‘A linear three level difference scheme for quasilinear parabolic

equations’, Mathematics of Computation, 20, 516-622, 1966

Lévy, M., ‘Memoire sur les équations générales des mouvements intérieurs des

corps solides ductiles au delà des limites où l’élasticité pourrait les

ramener à leur premier état’, Compt. Rend., 70, 1323-1325, 1870

Lewis, R.W., Roberts, G.K. and Zienkiewicz, O.C., ‘A nonlinear flow and

deformation analysis of consolidated problems’, in Proceedings of the

2nd International Conference on Numerical Methods in Geomechanics,

ASCE, New York, 1976

Lewis, R.W. and Schrefler, B.A., The Finite-Element Method in the Deformation and

Consolidation of Porous Media, Wiley, Chichester, 1987

267References

Liniger, W., ‘Global accuracy and A-stability of of one and two step integration

formulae for stiff ordinary differential equations’, in Proceedings of

Conference on Numerical Solution of Differential Equations, Dundee

University, Springer, 1969

Mandel, J. ‘Consolidation des sols’, Geotechnique, III, 287-299, 1953

Manoharan, N. and Dasgupta, S.P., ‘Consolidation analysis of elastoplastic soil’,

Computers and Structures, 54, 1005-1021, 1995

Marcal, P.V. and King, I.P., ‘Elastic-plastic analysis of two-dimensional stress

systems by the finite element method’, International Journal of

Mechanical Sciences, 9, 143-155, 1967

Matthies, H. and Strang, G., ‘The solution of non-linear finite element equations’,

International Journal for Numerical Methods in Engineering, 14,

1613-1626, 1979

McNamee, J. and Gibson, R.E., ‘Plane strain and axially symmetric consolidation

of a semi-infinite clay stratum’, Quarterly Journal of Mech. and Applied

Mathematic, 13, 210-227, 1960

Mohr, O., Zeits. Ver. deutsch. Ing., 1900

Mohr, O., Abhandlungen aus dem Gebiete der technischen Mechanik, 2nd edition,

Ernest, Berlin, 192-235, 1914

Nayak, G.C. and Zienkiewicz, O.C., ‘Elasto-plastic stress analysis. A generalisation

for various constitutive relations including strain softening’,

International Journal for Numerical Methods in Engineering, 5, 113-135,

1972a

Nayak, G. C. and Zienkiewicz, O. C., ‘Convenient form of stress invariants for

plasticity’, Journal of the Structural Division, ASCE., 98, 949-954, 1972b

268References

Oritz, M. and Simo, J.C., ‘An analysis of a new class of integration algorithm for

elasto-plastic constitutive relations’, International Journal for Numerical

Methods in Engineering, 23, 353-366, 1986

Owen, D. R. J. and Hinton, E., Finite Elements in Plasticity: Theory and Practice,

Pineridge Press, Swansea, 1980

Potts, D.M. and Ganendra, D., ‘A Comparison of solution strategies for non-linear

finite element analysis of geotechnical problems’, in Proceedings of the

Third International Conference on Computational Plasticity, Barcelona,

803-814, 1992

Potts, D.M. and Ganendra, D., ‘An evaluation of substepping and implicit stress

point algorithms’, Computer Methods in Applied Mechanics and

Engineering, 119, 341-354, 1994

Potts, D.M. and Gens, A., ‘A critical assessment of methods of correcting for drift

from the yield surface in elastoplastic finite element analysis’,

International Journal for Numerical and Analytical Methods in

Geomechanics, 9, 149-159, 1985

Prandtl, L., ‘Spannungsverteilung in plastischen Koerpern’, in Proceedings of the

1st International Congress on Applied Mechanics, Delft, 43-54, 1924

Prevost, J.H., ‘Nonlinear transient phenomena in saturated porous media’,

Computer Methods in Applied Mechanics and Engineering, 20, 3-18, 1982

Prevost, J.H., ‘Implicit-explicit schemes for nonlinear consolidation’, Computer

Methods in Applied Mechanics and Engineering, 39, 225-239, 1983

Rankine, W., ‘On the stability of loose earth’, Phil. Trans R. Soc. London, 147, 9-27,

1857

Reed, M.B., ‘An investigation of numerical errors in the analysis of consolidation

by finite element analysis’, International Journal for Numerical and

Analytical Methods in Geomechanics, 8, 243-257, 1984

269References

Rendulic, L., ‘Porenziffer und Poren Wasserdruck in Tonen’, Bauingenieur, 17,

559-564, 1936

Richtmyer, R.D. and Morton, K.W., Difference Methods for Initial Value Problems,

Wiley, New York, 1967

Riks, E., ‘The application of Newtons method to the problem of elastic instability’,

Journal of Applied Mechanics, 39, 1060-1066, 1972

Riks, E., ‘An incremental approach to the solution of snapping and buckling

problems’, International Journal of Solids and Structures, 15, 529-551,

1979

de Saint-Venant, B., ‘Memoire sur l’établissement des équations différentielles des

mouvements intérieurs opérés dans les corps solides ductiles au delà

des limites où l’élasticité pourrait les ramener à leur premier état’,

Compt. Rend., 70, 473-480, 1870

Sandhu, R.S. and Wilson, E.L. ‘Finite element analysis of seepage in elastic

media’, Journal of Engineering Mechanics. ASCE, 95, 641-652, 1969

Sandhu, R.S., Lui H. and Singh, K. J. ‘Numerical performance of some finite

element schemes for analysis of seepage in porous elastic media’,

International Journal for Numerical and Analytical Methods in

Geomechanics, 1, 177-194, 1977

Schellekens, J.C.J, Feenstra, P.H. and de Borst, ‘A self-adaptive load estimator

based on strain energy’, in Proceedings of the Third International

Conference on Computational Plasticity, Barcelona, 187-198, 1992

Schiffman, R.L., Chen, A.T. and Jordan, J.C., ‘An analysis of consolidation

theories’, Journal of the Soil Mechanics and Foundation Division of

ASCE, 95, (SM1), 285-213, 1969

270References

Simo, J.C. and Taylor, R.L., ‘Consistent tangent operators for rate-independent

elastoplasticity’, Computer Methods in Applied Mechanics and

Engineering, 48, 101-118, 1985

Siriwardane, H.J. and Desai, C.S., ‘Two numerical schemes for nonlinear

consolidation’, International Journal for Numerical Methods in

Engineering, 17, 405-426, 1981

Sloan S.W., ‘Substepping schemes for the numerical integration of elastoplastic

stress-strain relations’, International Journal for Numerical Methods in

Engineering, 24, 893-911, 1987

Sloan, S. W., Assadi, A. and Purushothaman, N. ‘Undrained stability of a

trapdoor’, Geotechnique, 40, 45-62, 1990

Sloan, S. W. and Booker, J. R., ‘Removal of singularities in Tresca and

Mohr-Coulomb yield functions’, Communications in Applied Numerical

Methods, 2, 173-179, 1986

Sloan, S.W. and Booker, J.R., ‘Integration of Tresca and Mohr-Coulomb

constitutive relations in plane strain elastoplasticity’, International

Journal for Numerical Methods in Engineering, 33, 163-196, 1992

Sloan, S.W. and Randolph, M.F., ‘Numerical prediction of collapse loads using

finite element methods’, International Journal for Numerical and

Analytical Methods in Geomechanics, 33, 163-196, 1982

Small, J.C., Elasto-plastic consolidation of Soils, PhD thesis, University of Sydney,

1977

Small, J.C., Booker, J.R. and Davis, E.H., ‘Elasto-plastic consolidation of soil’,

International Journal of Solids and Structures, 12, 431-448, 1976

271References

Terzaghi, K. ‘Die Berechnung der Durchlassigkeitsziffer des Tones aus dem

Verlauf der hydrodynamischen Spannungsersceinungen’, Originally

published in 1923 and reprinted in From Theory to Practice in Soil

Mechanics, John Wiley and Sons, New York, 133-146, 1960

Thomas, R.M. and Gladwell, I., ‘Variable-order variable step algorithms for

second-order systems. Part 1: The methods’, International Journal for

Numerical Methods in Engineering, 26, 39-53, 1988a

Tresca, H, ‘Sur l’ecoulement des corps solids soumis à de fortes pression’, Comptes

Rendus Acad. Sci. Paris, 59, p. 754, 1864

Vermeer, P.A. and Verruijt, A., ‘An accuracy condition for consolidation by finite

elements’, International Journal for Numerical and Analytical Methods in

Geomechanics, 5, 1-14, 1981

von Mises, R., Mechanik der festen ‘Koerper im plastisch deformablen Zustard’,

Goettinger Nachr., math---phys. K1, 582-592, 1913

Wempner, G.A., ‘Discrete approximations related to nonlinear theories of solids’,

International Journal of Solids and Structures, 7, 1581-1599, 1971

Wissmann, J.W. and Hauck, C., ‘Efficient elastic-plastic finite element analysis

with higher order stress point algorithms’, Computers and Structures, 17,

89---95, 1983

Wood, W.L. ‘Control of Crank-Nicholson noise in the numerical solution of the

heat conduction equation’, International Journal for Numerical Methods

in Engineering, 11, 1059-1065, 1977

Wood, W.L. and Lewis, R. W., ‘A comparison of time-marching schemes for the

transient heat conduction equation’, International Journal for Numerical

Methods in Engineering, 9, 679-689, 1975

Wood, W.L. Practical Time Stepping Schemes, Clarendon Press, Oxford, 1990

272References

Yamada, Y., Yishimura, N. and Sakurai, T., ‘Plastic stress-strain matrix and its

application for the solution of elastic-plastic problems by the finite

element method’, International Journal of Mechanical Sciences, 10,

343-354, 1968

Yamaguchi, E., ‘A comparative study of numerical methods for computing stress

increments in elastic-plastic materials’, in Proceedings of the Asia-Pacific

Symposium on Advances in Plasticity and its Applications, Hong Kong,

625-630, 1993

Yokoo, Y., Yamagata, K. and Nagaoka, H., ‘Finite element method applied to

Biot’s consolidation theory’, Soils and Foundations (Japanese Society

for Soil Mechanics and Foundation Engineering), 11(1), 29-46, 1971a

Yokoo, Y., Yamagata, K. and Nagaoka, H., ‘Variational principles for

consolidation’, Soils and Foundations (Japanese Society for Soil

Mechanics and Foundation Engineering), 11(4), 25-35, 1971b

Yu, H.S, ‘Expansion of a thick cylinder of soil’, Computers and Geotechnics, 14,

21-41, 1992

Zienkiewicz, O. C. and Pande, G. N., ‘Some useful forms of isotropic yield surfaces

for soil and rock mechanics’, in Finite Elements in Geomechanics (ed.

G. Gudehus), Wiley, Chichester, 179-190, 1977

Zienkiewicz, O.C., Valliapan, S. and King, I.P., ‘Elasto-plastic solutions of

engineering problems. Initial stress, finite element approach’,

International Journal for Numerical Methods in Engineering, 1, 75-100,

1969

Zienkiewicz, O.C., Wood, W.L., Hine, N.W. and Taylor, R.L., ‘A unified set of

single step algorithms. Part 1: General formulation and applications’,

International Journal for Numerical Methods in Engineering, 20,

1529-1552, 1984

273References

Zlamal, M., ‘Finite element methods in heat conduction problems’, in The

Mathematics of Finite Elements and Application (ed. J.R. Whiteman),

Academic Press, London, 85-104, 1977