Dynamics of Ice Sheets and Glaciers

Similar documents
Advanced Course in Theoretical Glaciology

PEAT SEISMOLOGY Lecture 2: Continuum mechanics

Getting started: CFD notation

Chapter 1. Continuum mechanics review. 1.1 Definitions and nomenclature

Fundamentals of Fluid Dynamics: Elementary Viscous Flow

V (r,t) = i ˆ u( x, y,z,t) + ˆ j v( x, y,z,t) + k ˆ w( x, y, z,t)

Mathematical Background

Dynamics of Glaciers

Chapter 9: Differential Analysis

Chapter 3. Forces, Momentum & Stress. 3.1 Newtonian mechanics: a very brief résumé

Mathematical Tripos Part IA Lent Term Example Sheet 1. Calculate its tangent vector dr/du at each point and hence find its total length.

Chapter 9: Differential Analysis of Fluid Flow

Chapter 2: Fluid Dynamics Review

Multiple Integrals and Vector Calculus (Oxford Physics) Synopsis and Problem Sets; Hilary 2015

Lecture 3: 1. Lecture 3.

Mechanics of solids and fluids -Introduction to continuum mechanics

Review of fluid dynamics

Introduction to Continuum Mechanics

Introduction to Fluid Mechanics

Continuum Mechanics Fundamentals

CH.9. CONSTITUTIVE EQUATIONS IN FLUIDS. Multimedia Course on Continuum Mechanics

Mechanics of materials Lecture 4 Strain and deformation

Continuum Mechanics. Continuum Mechanics and Constitutive Equations

Numerical Heat and Mass Transfer

A CONTINUUM MECHANICS PRIMER

KINEMATICS OF CONTINUA

Navier-Stokes Equation: Principle of Conservation of Momentum

Chapter 2: Basic Governing Equations

Chapter 5. The Differential Forms of the Fundamental Laws

Multiple Integrals and Vector Calculus: Synopsis

Introduction to Seismology Spring 2008

Vector and Tensor Calculus

INDEX 363. Cartesian coordinates 19,20,42, 67, 83 Cartesian tensors 84, 87, 226

Chapter 3 Stress, Strain, Virtual Power and Conservation Principles

A Study on Numerical Solution to the Incompressible Navier-Stokes Equation

Elements of Rock Mechanics

Lecture 8: Tissue Mechanics

Dynamics of the Mantle and Lithosphere ETH Zürich Continuum Mechanics in Geodynamics: Equation cheat sheet

CONTINUUM MECHANICS. lecture notes 2003 jp dr.-ing. habil. ellen kuhl technical university of kaiserslautern

EKC314: TRANSPORT PHENOMENA Core Course for B.Eng.(Chemical Engineering) Semester II (2008/2009)

Chapter 7. Kinematics. 7.1 Tensor fields

Artificial Intelligence & Neuro Cognitive Systems Fakultät für Informatik. Robot Dynamics. Dr.-Ing. John Nassour J.

Basic hydrodynamics. David Gurarie. 1 Newtonian fluids: Euler and Navier-Stokes equations

Computational Fluid Dynamics Prof. Dr. Suman Chakraborty Department of Mechanical Engineering Indian Institute of Technology, Kharagpur

Chapter 0. Preliminaries. 0.1 Things you should already know

Mathematical Preliminaries

III.3 Momentum balance: Euler and Navier Stokes equations

2 GOVERNING EQUATIONS

Continuum mechanism: Stress and strain

Chapter 1. Governing Equations of GFD. 1.1 Mass continuity

AE/ME 339. Computational Fluid Dynamics (CFD) K. M. Isaac. Momentum equation. Computational Fluid Dynamics (AE/ME 339) MAEEM Dept.

Chapter 1 Fluid Characteristics

3 2 6 Solve the initial value problem u ( t) 3. a- If A has eigenvalues λ =, λ = 1 and corresponding eigenvectors 1

Soft Bodies. Good approximation for hard ones. approximation breaks when objects break, or deform. Generalization: soft (deformable) bodies

NDT&E Methods: UT. VJ Technologies CAVITY INSPECTION. Nondestructive Testing & Evaluation TPU Lecture Course 2015/16.

Cartesian Tensors. e 2. e 1. General vector (formal definition to follow) denoted by components

Rotational motion of a rigid body spinning around a rotational axis ˆn;

Math Review: Vectors and Tensors for Rheological Applications

Kinematics. B r. Figure 1: Bodies, reference configuration B r and current configuration B t. κ : B (0, ) B. B r := κ(b, t 0 )

7 The Navier-Stokes Equations

2. Conservation Equations for Turbulent Flows

12. Stresses and Strains

ENGR Heat Transfer II

Review of Vector Analysis in Cartesian Coordinates

Entropy generation and transport

CVEN 7511 Computational Mechanics of Solids and Structures

CE-570 Advanced Structural Mechanics - Arun Prakash

Macroscopic theory Rock as 'elastic continuum'

3D Elasticity Theory

Convection Heat Transfer

Numerical methods for the Navier- Stokes equations

2. FLUID-FLOW EQUATIONS SPRING 2019

CHAPTER 8 ENTROPY GENERATION AND TRANSPORT

Professor George C. Johnson. ME185 - Introduction to Continuum Mechanics. Midterm Exam II. ) (1) x

1. Matrix multiplication and Pauli Matrices: Pauli matrices are the 2 2 matrices. 1 0 i 0. 0 i

A SHORT INTRODUCTION TO TWO-PHASE FLOWS Two-phase flows balance equations

UNIVERSITY of LIMERICK

Course Syllabus: Continuum Mechanics - ME 212A

Math 302 Outcome Statements Winter 2013

2.20 Fall 2018 Math Review

In this section, mathematical description of the motion of fluid elements moving in a flow field is

Fundamentals of Linear Elasticity

Physical Conservation and Balance Laws & Thermodynamics

EULERIAN DERIVATIONS OF NON-INERTIAL NAVIER-STOKES EQUATIONS

Game Physics. Game and Media Technology Master Program - Utrecht University. Dr. Nicolas Pronost

Tensors, and differential forms - Lecture 2

Basic Equations of Elasticity

Candidates must show on each answer book the type of calculator used. Log Tables, Statistical Tables and Graph Paper are available on request.

Elements of Continuum Elasticity. David M. Parks Mechanics and Materials II February 25, 2004

Covariant Formulation of Electrodynamics

APPENDIX A. Background Mathematics. A.1 Linear Algebra. Vector algebra. Let x denote the n-dimensional column vector with components x 1 x 2.

Quick Recapitulation of Fluid Mechanics

A short review of continuum mechanics

Modelling of turbulent flows: RANS and LES

Coordinate systems and vectors in three spatial dimensions

Lecture Notes Introduction to Vector Analysis MATH 332

1 Gauss integral theorem for tensors

4. The Green Kubo Relations

Module 2: Governing Equations and Hypersonic Relations

Module-3: Kinematics

Transcription:

Dynamics of Ice Sheets and Glaciers Ralf Greve Institute of Low Temperature Science Hokkaido University Lecture Notes Sapporo 2004/2005

Literature Ice dynamics Paterson, W. S. B. 1994. The Physics of Glaciers. Pergamon Press, Oxford etc., 3rd edition. Van der Veen, C. J. 1999. Fundamentals of Glacier Dynamics. A. A. Balkema, Rotterdam. Hutter, K. 1983. Theoretical Glaciology: Material Science of Ice and the Mechanics of Glaciers and Ice Sheets. D. Reidel Publishing Company, Dordrecht etc. Zotikov, I. A. 1986. The Thermophysics of Glaciers. D. Reidel Publishing Company, Dordrecht etc. Greve, R. 2000. Large-scale glaciation on Earth and on Mars. Habilitation thesis, Department of Mechanics, Darmstadt University of Technology (available online at http://hgxpro1.lowtem.hokudai.ac.jp/ greve/papers/2000/greve00c.pdf). Continuum mechanics, mathematics Hutter, K. and K. Jöhnk. 2004. Continuum Methods of Physical Modeling. Springer- Verlag, Berlin etc. Greve, R. 2003. Kontinuumsmechanik. Eine Einführung für Ingenieure und Physiker. Springer-Verlag, Berlin etc. Liu, I-S. 2002. Continuum Mechanics. Springer-Verlag, Berlin etc. Heinbockel, J. H. 1996. Introduction to tensor calculus and continuum mechanics. Department of Mathematics and Statistics, Old Dominion University, Norfolk, Virginia (free, available online at http://www.math.odu.edu/ jhh/counter2.html). i

Bronshtein, I. N., K. A. Semendyayev, G. Musiol and H. Muehlig. 2004. Handbook of Mathematics. Springer-Verlag, Berlin etc., 4th edition. Any textbook on elementary calculus and linear algebra of the distinguished reader s preference... ii

Contents 1 Vectors, tensors and their representation 1 1.1 Definition of a vector, basic properties.................. 1 1.2 Representation of vectors as number triples................ 3 1.3 Tensors................................... 5 2 Elements of continuum mechanics 7 2.1 Bodies and configurations......................... 7 2.2 Kinematics................................. 8 2.2.1 Deformation gradient, stretch tensors............... 8 2.2.2 Velocity, acceleration, velocity gradient.............. 11 2.3 Balance equations.............................. 15 2.3.1 Reynolds transport theorem.................... 15 2.3.2 General balance equation...................... 17 2.3.3 General jump condition...................... 18 2.3.4 Mass balance............................ 20 2.3.5 Momentum balance......................... 21 2.3.6 Balance of angular momentum................... 24 2.3.7 Energy balance........................... 26 2.4 Constitutive equations........................... 29 2.4.1 Simple bodies with fading memory................ 29 2.4.2 Linear-elastic solid......................... 30 2.4.3 Newtonian fluid........................... 32 3 Large-scale dynamics of ice sheets 35 3.1 Constitutive equations for polycrystalline ice............... 35 3.1.1 Microstructure of ice........................ 35 3.1.2 Creep of polycrystalline ice..................... 36 iii

Contents 3.1.3 Flow relation............................ 37 3.1.4 Heat flux and internal energy................... 41 3.2 Full Stokes flow problem.......................... 42 3.2.1 Field equations........................... 42 3.2.2 Boundary conditions........................ 45 3.2.3 Ice-thickness equation....................... 50 3.3 Hydrostatic approximation......................... 51 3.4 Shallow-ice approximation......................... 53 3.5 Vialov profile................................ 60 4 Large-scale dynamics of ice shelves 65 4.1 Full Stokes flow problem, hydrostatic approximation........... 65 4.2 Shallow-shelf approximation........................ 68 5 Dynamics of glacier flow 75 6 Glacial isostasy 77 6.1 General remarks............................... 77 6.2 Structure of the Earth........................... 79 6.3 Simple isostasy models........................... 80 6.3.1 LLRA model............................ 80 6.3.2 ELRA model............................ 81 6.3.3 LLDA model............................ 83 6.3.4 ELDA model............................ 86 6.3.5 Performance of the simple models................. 86 iv

1 Vectors, tensors and their representation 1.1 Definition of a vector, basic properties In mathematics, a vector is defined as an element of a vector space, and a vector space is a commutative (Abelian) group with a scalar multiplication. This is an abstract definition which has many possible realizations (numbers, functions, geometric objects and so on). For our purposes, it is sufficient to consider one of them, namely the geometric object of an arrow in the three-dimensional, Euklidian, physical space E. Therefore, in our sense a vector a E is an arrow which is characterized by a length and a direction. Physical quantities which can be described by such vectors are, for instance, velocity, acceleration, momentum and force. By contrast, scalars are simple numbers and characterize physical quantities without a direction, like mass, density, temperature etc. We will usually denote vectors by bold-face symbols like a, b, c... The sum s = a + b (1.1) of two vectors is then obtained by the parallelogram construction, and the scalar multiplication p = λa, λ R (1.2) (R denotes the set of real numbers) follows from multiplying the length of a by λ (Fig. 1.1). The length of the vector a is written as a (absolute value, norm), and its direction can be characterized by the unit vector (length equal to one) e a = a/ a. Further, the dot product (inner product) δ = a b (1.3) of two vectors is equal to the scalar given by a b cos ϕ (where ϕ is the angle between 1

1 Vectors, tensors and their representation Figure 1.1: Sum s = a + b and scalar multiplication p = λa of vectors. the two vectors), and the cross product (outer product) c = a b (1.4) is equal to the vector with length a b sin ϕ and direction perpendicular to the plane spanned by a and b, such that a, b and c form a right-hand system (Fig. 1.2). Note that a = a a, (1.5) a b = 0 a b, (1.6) a b = b a (1.7) and a a = 0, (1.8) where 0 denotes the vector of length zero ( zero vector ). Figure 1.2: Cross product c = a b of vectors. Finally, the dyadic or tensor product a b (sometimes denotes as a b) is the linear transformation which, when applied to an arbitrary vector x, obeys the relation (a b) x = a(b x), (1.9) 2

1.2 Representation of vectors as number triples where (b x) means the dot product (1.3). In other words, the transformation a b maps the vector x onto the vector which has the direction of a and (b x) times the length of a. 1.2 Representation of vectors as number triples Let {e i } i=1,2,3 be a set of unit vectors which are perpendicular to each other and form a right-hand system. In other words, e i e j = δ ij, (1.10) where δ ij is the Kronecker symbol defined as 1, for i = j, δ ij = 0, for i j, (1.11) and e i e j = e k, (i, j, k) {(1, 2, 3), (2, 3, 1), (3, 1, 2)}. (1.12) We will refer to such a set {e i } as an orthonormal basis (also Cartesian basis). An arbitrary vector a can then uniquely be written as 3 a = a 1 e 1 + a 2 e 2 + a 3 e 3 = a i e i, (1.13) i=1 where the a i are real numbers. With Einstein s summation convention, which says that double indices (here i) automatically imply summation, this can be written in compact form as a = a i e i. (1.14) Since the coefficients a i are unique for a given basis {e i }, it is possible to represent the vector a by these coefficients. It is usual to arrange them in a column (number triple) and write a {ei } = a 1 a 2, (1.15) a 3 which is to say, the vector a is represented by the components a i with respect to the basis {e i }. Of course, when a different orthonormal basis {e i } is used, the representation 3

1 Vectors, tensors and their representation of the vector a will change: a = a i e i, (1.16) or a {e i } = a 1 a 2 a 3. (1.17) Note that the vector a is still the same object (arrow in space), whereas its components have changed. It is therefore of great importance to distinguish between vectors themselves and their representation as number triples. Mixing up these two different things is a notorious source of confusion. Only when a single basis {e i } is defined from the outset, and all vectors are expressed in this basis, it is allowed to write sloppily a = but this requires a good deal of care. a 1 a 2, (1.18) a 3 In components with respect to a given basis {e i }, the dot product (1.3) can be evaluated as a b = a i b i, (1.19) and the ith component of the cross product (1.4) is (a b) i = ε ijk a j b k (1.20) (summation over j and k). In the latter expression, ε ijk is called the Levi-Civita symbol, defined as ε ijk = 1, for (i, j, k) {(1, 2, 3), (2, 3, 1), (3, 1, 2)}, 1, for (i, j, k) {(1, 3, 2), (3, 2, 1), (2, 1, 3)}, (1.21) 0, otherwise (two or three indices equal). The dyadic product defined in (1.9) is expressed as a b = (a i e i ) (b j e j ) = a i b j e i e j, (1.22) (summation over i and j), where e i e j is the dyadic product of the respective basis vectors. 4

1.3 Tensors 1.3 Tensors A tensor A of rank two (often simply called a tensor) is defined as a linear transformation which maps vectors onto vectors: y = A x. (1.23) We have already encountered special tensors of rank two, namely the dyadic products between two vectors introduced in (1.9). Their expression with respect to an orthonormal basis {e i } was given by (1.22), and similarly a general tensor of rank two can be written as A = A ij e i e j. (1.24) Evidently, the tensor A is represented by the components A ij, and in analogy to (1.15) this can be noted as A 11 A 12 A 13 A {ei } = A 21 A 22 A 23, (1.25) A 31 A 32 A 33 where the components have been arranged into a quadratic matrix. Again, if a different basis {e i } is used, the representation will change, A 11 A 12 A 13 A {e i } = A 21 A 22 A 23, (1.26) A 31 A 32 A 33 so that tensors and matrices must be distinguished in the same way as vectors and number triples. With the expression (1.24), the transformation (1.23) reads y = (A ij e i e j ) (x k e k ) = A ij x k e i (e j e k ) = A ij x k e i δ jk = A ij x j e i, (1.27) or y i = A ij x j. (1.28) Evidently, this is nothing else but the matrix-column product y 1 A y 2 = 11 A 12 A 13 x A 21 A 22 A 23 1 x 2 y 3 A 31 A 32 A 33 x 3 (1.29) 5

1 Vectors, tensors and their representation expressed in (Cartesian) index notation. Index notation is a very efficient method of carrying out computations in vector/tensor algebra and analysis [see also the expressions (1.19) and (1.20) for the dot product and the cross product, respectively], and we will use it frequently. An important example for a tensor of rank two is the unity tensor 1, which provides the identical transformation x = 1 x. Its components in any orthonormal basis {e i } are given by the Kronecker symbol δ ij, that is, 1 = δ ij e i e j, (1.30) so that its matrix representation is given by the unity matrix, 1 0 0 1 {ei } = 0 1 0. (1.31) 0 0 1 Evidently, when expressed in components, a tensor of rank two is a quantity with two indices [see (1.24)]. As we have seen in Sect. 1.2, the expression of a vector in components leads to a quantity with one index (for instance, a i ), and of course a scalar quantity does not have any indices at all. Therefore, vectors and scalars are also refered to as tensors of rank one and zero, respectively. As a generalisation of (1.23), tensors A [r] of rank r > 2 can now be defined inductively as linear transformations which map vectors x onto tensors Y [r 1] of rank r 1, Y [r 1] = A [r] x. (1.32) Such tensors can be written in components as A [r] = A i1 i 2...i r e i1 e i2... e ir (1.33) (summation over the r indices i 1, i 2,..., i r ). As an example, the Levi-Civita symbol (1.21) can be interpreted as the components of a third-rank tensor ε [3], ε [3] = ε ijk e i e j e k, (1.34) which is known as the epsilon or permutation tensor. Tensors of rank four play a role in the theory of elasticity. 6

2 Elements of continuum mechanics 2.1 Bodies and configurations Continuum mechanics is concerned with the motion and deformation of continuous bodies (for instance, a glacier). A body consists of an infinite number of material elements, called particles. For any time t, each particle is identified by a position vector x (relative to a prescribed origin O) in the physical space E, and the continuous set of position vectors for all particles of the body is called a configuration κ of the body. If t is the actual time, the corresponding configuration is called the present configuration κ t. In addition, we define a reference configuration κ r which refers to a fixed (or initial) time t 0. Position vectors in the reference configuration will be written in capitals like X; they can be used for identifying the individual particles of the body independent of the actual time. Note that different sets of basis vectors ({E A } A=1,2,3, {e i } i=1,2,3 ) and different origins may be used in the two configurations (Fig. 2.1). Figure 2.1: Bodies, reference configuration κ r and present configuration κ t. The mapping χ which provides the position x of any particle at time t as a function of its reference position X is called motion: χ : κ r κ t X x = x(x, t). (2.1) 7

2 Elements of continuum mechanics It is assumed that the motion x(x, t) is continuously differentiable in the entire body (with the possible exception of singular lines or surfaces), so that the inverse mapping χ 1 exists: χ 1 : κ t κ r x X = X(x, t). (2.2) The displacement is defined as the connecting vector between a given particle in the reference and present configuration. If the connecting vector between the two origins of the basis systems is denoted by B, then u = x X + B (2.3) holds. The above relationships are illustrated in Fig. 2.1. Of course, in a deformable body the displacement at time t will in general be different for the different particles, so that it can be written as the vector field u = u(x, t). However, this is not the only possibility. Equation (2.2) shows that X can be expressed in terms of x and t, so that we can also assume the displacement field as a function of x and t, that is, u = u(x, t). These two possibilities hold also for other field quantities ψ (density, temperature, velocity etc.), and we call ψ(x, t) the Lagrangian or material description, whereas ψ(x, t) is refered to as the Eulerian or spatial description. Most frequently, for solid bodies the Lagrangian description is used, whereas the Eulerian description is more appropriate for problems of fluid dynamics (like glacier flow). 2.2 Kinematics 2.2.1 Deformation gradient, stretch tensors The deformation gradient F is defined as the material gradient (gradient with respect to X) of the motion (2.1), or in components F = Grad x(x, t), (2.4) F ia = x i(x, t) X A = x i,a, (2.5) where F = F ia e i E A, the operator Grad ( ) is the material gradient, and the notation ( ),A means the partial derivative ( )/ X A. It is a tensor of rank two. Note that small indices refer to the present configuration and capital indices to the reference configuration. 8

2.2 Kinematics According to the definition (2.4), the deformation gradient F can be interpreted as the functional matrix of the motion function (2.1). It transforms line elements from the reference configuration (dx) to the present configuration (dx), dx = F dx, or dx i = F ia dx A, (2.6) which is illustrated in Fig. 2.2. Figure 2.2: Deformation gradient: Transformation between line and volume elements in the reference and present configuration. The determinant of the deformation gradient, called the Jacobian, is given as J = det F, or J = 1 6 ε ABCε ijk F ia F jb F kc ; (2.7) in the component form on the right the Levi-Civita symbol, which was defined in (1.21), appears. Since we have demanded that the motion function is invertible, J must be different from zero, and the inverse deformation gradient F 1 exists. Further, real motions cannot invert the orientation, so that J > 0 (2.8) must hold. The Jacobian determines the local volume change due to the motion, dv = J dv, (2.9) where dv is the volume element in the reference configuration which may be spanned by three line elements dx (1), dx (2), dx (3), and dv is the volume element in the present configuration spanned by dx (1), dx (2), dx (3) (Fig. 2.2). The theorem of polar decomposition tells us that, like any tensor with positive deter- 9

2 Elements of continuum mechanics minant, the deformation gradient F can be uniquely decomposed according to F = R U = V R, (2.10) where R is a proper orthogonal tensor (R R T = R T R = 1 and det R = +1), and the tensors U und V are symmetric (U = U T, V = V T ) and positive definite ( x 0: x U x > 0, x V x > 0). The tensors U and V are called the right and left stretch tensor, respectively, and R is the rotation tensor. The polar decomposition of F can be obtained as follows. With (2.10) we compute F T F = U T R T R U = U 1 U U 2 = F T F (2.11) and F F T = V R R T V T = V 1 V V 2 = F F T, (2.12) which provides the stretch tensors U and V. The rotation tensor R follows then from (2.10) as R = F U 1, or R = V 1 F. (2.13) Note that (2.10) also implies the relation V = R U R T, (2.14) which means that the two stretch tensors are connected by a similarity transformation. The polar decomposition of the deformation gradient F allows the interpretation of an arbitrary deformation as a sequence of a local rigid-body rotation followed by a stretching, or vice versa. With (2.6) we can write dx = R U dx = V R dx. (2.15) Since U is symmetric, there exists a special set of orthonormal basis vectors {ē i }, called the principal axes, for which the matrix U {ēi } is diagonal, that is, U {ēi } = diag(λ 1, λ 2, λ 3 ) = λ 1 0 0 0 λ 2 0 0 0 λ 3. (2.16) The λ i, i = 1... 3, are the eigenvalues of U, and due to the positive definiteness they are all positive. This holds also for V, and because of the similarity transformation 10

2.2 Kinematics (2.14) U and V even have the same eigenvalues. Further, the proper orthogonal tensor R is geometrically a rigid-body rotation. Figure 2.3: Polar decomposition of the deformation gradient. Let us now consider an infinitesimal cube (volume element) in the reference configuration, the edges ds of which shall be aligned with the principal axes of U (Fig. 2.3). The sequence of transformations dx = R U dx stretches in the first step the edges of the cube by the factors λ i (eigenvalues of U), which deforms the cube to a parallelepiped element. In the second step, this element is rotated by the orthogonal transformation R. The alternative sequence dx = V R dx leads to the same result, but here the initial cube is first rotated by R and then stretched by V. Of course, since F, U, V and R are in general functions of X and t (or x and t), this decomposition is only local. In other words, a volume element at a different position will experience a different stretching and a different rotation. 2.2.2 Velocity, acceleration, velocity gradient As usual, we define the velocity v as the first time derivative of the position x [motion (2.1)], v = v(x, t) = and the acceleration a as its second time derivative, a = a(x, t) = 2 x(x, t) t 2 = x(x, t), (2.17) t v(x, t). (2.18) t 11

2 Elements of continuum mechanics Evidently, this yields the velocity and acceleration fields in Lagrangian description. By inserting the inverse motion (2.2) one can readily obtain the corresponding Eulerian descriptions v(x, t) and a(x, t). The time derivatives in (2.17) and (2.18) are taken for fixed material position vectors X. We will call this the material time derivative, denoted briefly by the operators d( )/dt or ( ). Therefore, for any field quantity ψ, ψ = dψ dt ψ(x, t) =. (2.19) t For example, (2.17) and (2.18) read v = ẋ = dx dt, a = ẍ = d2 x dt 2 dv = v = dt. (2.20) By contrast, the operator ( )/dt shall denote the local time derivative for fixed spatial position vector x, ψ t ψ(x, t) =. (2.21) t With the chain rule, the relation between material and local time derivative follows as dψ dt = d ψ(x(x, t), t) dt ψ(x, t) = + grad ψ(x, t) t = ψ t dx(x, t) dt + (grad ψ) v, (2.22) where the operator grad ( ) is the spatial gradient, the components of which are the partial derivatives ( ),i = ( )/ x i. One says that the material time derivative dψ/dt is composed by a local part ψ/dt and an advective part (grad ψ) v. Relation (2.22) is equally valid if ψ is a vector or tensor field. Therefore, for the acceleration expressed by (2.20) 2, The tensor quantity a = v t + (grad v) v. (2.23) L = grad v = v i x j e i e j = v i,j e i e j (2.24) which appears in (2.23) is called the velocity gradient. It is related to the material time 12

2.2 Kinematics derivative of the deformation gradient as follows: F ia = 2 x i (X, t) = v i(x, t) = v i(x, t) x j (X, t) = v i(x, t) F ja t X A X A x j X A x j Ḟ = L F, or L = Ḟ F 1. (2.25) Without a formal proof (which is a bit tedious) we note for the material time derivative of the Jacobian J = J div v = J tr L (2.26) (div v = v i,i = L ii = tr L, where the operator tr means the trace of a tensor). In words, the divergence of the velocity field determines local volume changes [see also Eq. (2.9)], which is a very intuitive result. Like any arbitrary tensor, the velocity gradient can be additively decomposed into a symmetric and an antisymmetric part: L = D + W, (2.27) with D = 1 2 (L + LT ) ( strain-rate tensor ), W = 1 2 (L LT ) ( spin tensor ). (2.28) Evidently, D = D T (symmetry) and W = W T (antisymmetry) are fulfilled. In order to give an interpretation of the elements of the matrix of the strain-rate tensor D (with respect to a given orthonormal basis {e i }), we now compute the material time derivative of line elements dx in the present configuration: (dx) = Ḟ dx = L F dx = L dx (2.29) [where Eqs. (2.6), (2.25) and (dx) = 0 were used]. For the scalar product between two line elements dx (1) and dx (2), this yields (dx (1) dx (2) ) = (dx (1) ) dx (2) + dx (1) (dx (2) ) = (L dx (1) ) dx (2) + dx (1) (L dx (2) ) = dx (1) L T dx (2) + dx (1) L dx (2) = 2 dx (1) D dx (2). (2.30) 13

2 Elements of continuum mechanics Let us assume dx (1) = n (1) ds (1), dx (2) = n (2) ds (2), n (1) n (2) = cos((π/2) γ) = sin γ, (2.31) where n (1), n (2) are unit vectors, and γ is the deviation of the angle between n (1) and n (2) from a right angle. Equation (2.30) then reads (sin γ ds (1) ds (2) ) = 2ds (1) ds (2) n (1) D n (2) γ cos γ + sin γ ( (ds (1) ) ) ) + (ds(2) = 2n (1) D n (2). (2.32) ds (1) ds (2) We first make the special choice n (1) = n (2) = e x (γ = 90 ) and ds (1) = ds (2) = ds (Fig. 2.4, left part). Then, 2 (ds) ds = 2e x D e x D xx = (ds) ds. (2.33) Analogous results are found for the y- and z-direction. Therefore, the elements D xx, D yy, D zz on the main diagonal of the matrix of D are equal to the dilation rates (ds) /ds in x-, y- and z-direction, respectively. Figure 2.4: Dilation (left) and shear (right) of line elements in the present configuration. Second, we choose n (1) = e x and n (2) = e y, so that γ = 0 (Fig. 2.4, right part). This yields γ = 2e x D e y D xy = γ 2. (2.34) Analogous relations can be obtained for the two other off-diagonal elements D xz and D yz, and so these elements denote (apart from a factor 2) the shear rates γ, that is, the temporal change of right angles formed by the coordinate directions. As for the spin tensor W, we note that its matrix has only three independent elements (this holds for any antisymmetric tensor). Without loss of generality, it can therefore 14

2.3 Balance equations be written as W = 0 w 3 w 2 w 3 0 w 1 w 2 w 1 0 (2.35) [see also Eq. (1.18) and the discussion there]. The w i arranged in the above form are the components of the dual vector w = dual W = w 1 w 2, (2.36) w 3 with which the linear transformation W a (arbitrary vector a) can be expressed as a cross product, W a = w a. (2.37) Thus, Eq. (2.29) yields (dx) = D dx + w dx. (2.38) The first summand on the right-hand side describes the strain (deformation) part of the motion, the second summand the local rigid-body rotation with angular velocity w. This justifies the names strain-rate tensor and spin tensor for D and W, respectively. 2.3 Balance equations 2.3.1 Reynolds transport theorem We consider a material volume ω κ t in the present configuration. Material means that the volume consists of the same particles for all times, ω denotes the boundary of ω, v the velocity field of the body and n the normal unit vector on ω (see Fig. 2.5). For an arbitrary scalar, vector or tensor field quantity ψ(x, t) (e.g., mass density ρ, momentum density ρv) the term (d/dt) ω ψ dv, that is, the temporal change of the field quantity integrated over the volume ω shall be computed. To this end, we transform the integration variable to material coordinates X, which changes the integration domain ω to the volume Ω κ r in the reference configuration: d ψ(x, t) dv = d ψ(x(x, t), t) J(X, t)dv. (2.39) dt ω dt Ω 15

2 Elements of continuum mechanics n v t e 3 e 2 e 1 Figure 2.5: On the Reynolds transport theorem: Material volume ω in the present configuration κ t. For the transformation of the volume element Eq. (2.9) was used. Since Ω, as a material volume in the reference configuration, must be time-independent, differentiation and integration can be exchanged on the right-hand side: d ψ(x, t) dv = dt ω = = Ω Ω ω ( ψj + ψj) dv ( ψ + ψ div v) JdV ( ψ + ψ div v) dv. (2.40) In the second step, Eq. (2.26) was used, and in the last step the integral was transformed back to spatial coordinates. The result can be further rewritten: d ψ(x, t) dv = dt ω = ω By using Gauss integral theorem, we obtain ω ( ) ψ + (grad ψ) v + ψ div v dv t ( ) ψ + div (ψv) t dv. (2.41) d ψ ψ(x, t) dv = dt ω ω t ω dv + ψv n da, (2.42) which is known as the Reynolds transport theorem. It says that the temporal change of the integral ω ψ dv over the material volume ω is caused by two contributions, (i) the local change ψ/ t within ω, and (ii) the advective flux ψv in normal direction n across the boundary ω. Note that, if ψ is a tensor field of rank r 1, ψv is a tensor product which yields a tensor of rank r + 1. 16

2.3 Balance equations 2.3.2 General balance equation Let G(ω, t) be an arbitrary physical quantity of the entire material volume ω (e.g., its mass, momentum or internal energy). We assume that the change of G with time may be due to three different processes, namely the flux F( ω, t) of G across the boundary ω, the production P(ω, t) of G within the volume ω, the supply S(ω, t) of G within the volume ω. Therefore, we can balance dg/dt as follows: d G(ω, t) = F( ω, t) + P(ω, t) + S(ω, t), (2.43) dt where positive fluxes have been defined as outflows of the volume, so that the flux term has a negative sign. Conserved quantities shall be characterised by a vanishing production. In order to reformulate this statement, we assume that G, P and S can be expressed as volume integrals of corresponding densities g, p and s (additivity), G(ω, t) = ω g(x, t) dv, g : density of the quantity G, P(ω, t) = ω p(x, t) dv, p : production density of G, (2.44) S(ω, t) = ω s(x, t) dv, s : supply density of G, and that F can be obtained as the surface integral of a flux density φ, F( ω, t) = ω φ(x, t) n da, (2.45) where da is the scalar surface element. Note that, if G is a tensor quantity of rank r 0 (scalar, vector etc.), then the rank of g, p and s is also equal to r, whereas the rank of φ is r + 1. Inserting the expressions (2.44) and (2.45) in Eq. (2.43) yields the general balance equation in integral form, d g(x, t) dv = φ(x, t) n da + p(x, t) dv + s(x, t) dv. (2.46) dt ω ω ω ω Provided that all fields in this equation are continuously differentiable, this equation can be localized as follows. We apply the Reynolds transport theorem (2.42) to the 17

2 Elements of continuum mechanics left-hand side (with ψ = g), transform all surface integrals to volume integrals with Gauss integral theorem and assemble all terms on the left-hand side: ω ( g t + div (gv) + div φ p s ) dv = 0. (2.47) This relation must hold for any arbitrary material volume ω, which is only possible, if the integrand itself vanishes. Thus, g t = div (φ + gv) + p + s, (2.48) which is the general balance equation in local form. It balances the local change of the density g with the production and supply densities and the negative divergence of two flux terms, the actual flux density φ and the advective (or convective) flux density gv. 2.3.3 General jump condition The local balance equation (2.48) is only valid for those parts of the material volume ω for which all fields are continuously differentiable. We now consider the case that there exists an oriented singular surface σ within ω for which this is not fulfilled. In particular, the density g may be discontinuous on σ. The normal unit vector on σ shall be n. We define the side of ω into which n points as ω + and the other side as ω. The singular surface need not to be material, that is to say, it travels with its own velocity w which may differ from the particle velocity v (Fig. 2.6). Figure 2.6: Singular surface σ within the material volume ω. The values which a field quantity ψ(x, t) assumes when the point x σ (on the singular surface) is approached on an arbitrary path in ω or ω + shall be denoted as ψ and ψ +, respectively: x σ : ψ (x, t) = lim y x, y ω ψ(y, t), ψ + (x, t) = lim y x, y ω + ψ(y, t). (2.49) 18

2.3 Balance equations Of course, this requires that the limits exist and are finite. We define the jump [ψ ] of ψ on σ as x σ : [ψ ] (x, t) = ψ + (x, t) ψ (x, t). (2.50) If [ψ ] 0, the quantity ψ experiences a discontinuity on the singular surface. We now motivate a balance equation similar to (2.46) for the pill-box volume ν around the singular surface σ with basal area S B und mantle area S M, which is also indicated in Fig. 2.6. It is hereby assumed that S B and S M are very small, so that the curvature of σ can be neglected. This entails a very small volume of the pill-box, so that the three volume integrals in (2.46) will be negligible compared to the surface integral of the flux density φ. So the first guess for the balance equation of the pill-box volume ν would be ν φ(x, t) n da = 0. However, in general we are dealing with a non-material volume here, so that, in addition to the actual flux density φ an advective flux density g(v w) due to the particle motion (velocity v) relative to the motion of the singular surface (velocity w) must be taken into account. Thus, the correct form of the balance equation is ν (φ + g(v w)) n da = 0. (2.51) The geometry of the pill-box volume is such that S M S B, so that all fluxes through S M can be neglected in comparison with fluxes through S B. Therefore, only the basal surfaces with area S B on the ω + and ω side of σ (denoted as S B + and SB, respectively) contribute to the surface integral (2.51): S + B (φ + g(v w)) n da + S B (φ + g(v w)) ( n) da = 0 (2.52) (note that the outer normal unit vector is n on S B + and n on SB ). Since all field quantities are virtually constant on the very small surfaces S B + and SB, this can be written as S B (φ + + g + (v + w)) n + S B (φ + g (v w)) ( n) = 0 φ + n φ n + (g + (v + w)) n (g (v w)) n = 0, (2.53) and with the definition (2.50) we obtain [φ n] + [g ((v w) n)] = 0. (2.54) This is the general jump condition on singular surfaces. For a formal derivation of 19

2 Elements of continuum mechanics this jump condition it is refered to the textbooks on continuum mechanics given in the literature list. 2.3.4 Mass balance If the physical quantity G is identified with the total mass M of the material volume ω, it is clear that dm/dt = 0, because the mass of a material volume cannot change (as long as no relativistic velocities are involved). With the (mass) density ρ this can be expressed as d ρ dv = 0. (2.55) dt ω By comparison with the general balance equation (2.46) we find immediately g = ρ, φ = 0, p = 0, s = 0. (2.56) With these densities, the local balance equation (2.48) reads ρ t + div (ρv) = 0. (2.57) This is the mass balance, also known as continuity equation. An equivalent form can be derived by differentiating the product, ρ t + (grad ρ) v + ρ div v = 0 ρ + ρ div v = 0. (2.58) An important special case is that of a density-preserving material (also refered to as incompressible, which is less precise, though), defined by a constant density, that is, ρ = const or ρ = 0. For this case, (2.58) simplifies to div v = 0, (2.59) which is the mass balance or continuity equation for density-preserving materials. Evidently, the corresponding velocity field is source-free (solenoidal). By inserting the densities (2.56) in the general jump condition (2.54), we obtain the mass jump condition on singular surfaces as [ρ ((v w) n)] = 0. (2.60) 20

2.3 Balance equations It simply states that the mass inflow at one side of the singular surface must equal the mass outflow at the other side. Let us come back to the general balance equation (2.48) and write the density g as g = ρg s, (2.61) where g s denotes the physical quantity under consideration per mass unit (which, of course, only makes sense if the quantity is not the mass itself). This yields (ρg s ) + div (φ + ρg s v) = p + s t ρ g s t + g ρ s t + div φ + g s div (ρv) + (grad g s ) ρv = p + s { } { } gs ρ ρ t + (grad g s) v + g s + div (ρv) t = div φ + p + s. (2.62) The first term in curly brackets is the material time derivative of g [see (2.22)], the second term vanishes because of the mass balance (2.57). It remains ρ dg s dt = div φ + p + s (2.63) as an alternative representation of the general balance equation, which can be used for any quantity except mass. 2.3.5 Momentum balance Let us now identify the physical quantity G with the total momentum P (vector!) of the material volume ω. Since momentum is equal to mass times velocity, the momentum density can be expressed as mass density times velocity, g = ρv, (2.64) and the total momentum P is equal to ω ρv dv. Following Newton s Second Law, its temporal change dp /dt must be given by the sum of all forces, F, which act on the volume ω. These forces can either be external volume forces f(x, t) which act on any volume element within ω (e.g., the gravity field), or internal stresses t n (x, t) which are cut free on the boundary surface ω (Fig. 2.7). The latter do not only depend 21

2 Elements of continuum mechanics on position x and time t, but also on the orientation of the surface, expressed by the normal unit vector n. Figure 2.7: Volume forces f and surface forces t n which act on a material volume ω in the present configuration κ t. The total force acting on the material volume ω is therefore F = and Newton s Second Law reads ω t n (x, t) da + ω f(x, t) dv, (2.65) d ρ(x, t)v(x, t) dv = t n (x, t) da + f(x, t) dv. (2.66) dt ω ω ω Except for the surface integral (flux term), this has the form of the general balance equation (2.46). By comparing the flux terms, we infer that the stress vector t n must be a linear function of n, that is, t n (x, t) = t(x, t) n, (2.67) where t(x, t) is a tensor field of rank two which is called the Cauchy stress tensor. Now we can identify g = ρv, φ = t, p = 0, s = f (2.68) (the volume force is interpreted as a supply and not a production term because it is assumed to have an external source), and from (2.48) we obtain the local form of the momentum balance as (ρv) t + div (ρv v) = div t + f. (2.69) 22

2.3 Balance equations With the specific momentum (momentum per mass unit) g s = g ρ = v (2.70) and the representation (2.63) of the general balance equation, an equivalent form of the momentum balance is ρ dv dt = div t + f. (2.71) Note that if the momentum balance is formulated in a non-inertial system (for instance, the rotating Earth), the volume force f contains contributions from inertial forces (centrifugal force, Coriolis force etc.). The momentum jump condition on singular surfaces is readily obtained from (2.54) and (2.68), [t n] [ρv ((v w) n)] = 0. (2.72) It relates the jump of the stress vector (t n) to the jump of the advective momentum flux across the interface. In case of a material singular surface (v + n = v n = w n) the stress vector is continuous, [t n] = 0. (2.73) t zz t xz t yz t yy t xx t zx t zx t yx t zy t xy t yy e z t xy t zy t xx t xz t yx e x e y t yz t zz Figure 2.8: Components of the Cauchy stress tensor. With respect to a given orthonormal basis {e i }, the matrix of the Cauchy stress 23

2 Elements of continuum mechanics tensor t defined by (2.67) is t = t xx t xy t xz t yx t yy t yz t zx t zy t zz. (2.74) The elements of this matrix can be interpreted as follows. For a cut along the xy plane, that is, with normal unit vector n = ±e z (the sign depends on the orientation of the plane), the stress vector t ±ez = ±t e z = ± t xz t yz t zz (2.75) is obtained. Evidently, the diagonal element t zz is perpendicular to the cut plane, whereas the off-diagonal elements t xz and t yz are parallel to the plane. The same result is found for cuts along the xz and yz planes. Therefore, the three diagonal elements (t xx, t yy, t zz ) are refered to as normal stresses, and the six off-diagonal elements (t xy, t yx, t xz, t zx, t yz, t zy ) are called shear stresses. The meaning of these components is illustrated in Fig. 2.8. 2.3.6 Balance of angular momentum In point mechanics, the angular momentum L of a mass point is given by L = x P (cross product of position vector and momentum), and the torque M is defined as M = x F (cross product of position vector and force acting on the mass point). In view of the momentum densities (2.68) for the continuous body, this motivates the following identifications for the densities of angular momentum: g = x ρv, φ = x t, p = 0, s = x f. (2.76) Inserting this in (2.48) yields the balance of angular momentum, or in index notation (x ρv) t + div [(x ρv) v] = div (x t) + x f, (2.77) t (ρ ε ijkx j v k ) + (ρ ε ijk x j v k v l ),l = (ε ijk x j t kl ),l + ε ijk x j f k. (2.78) 24

2.3 Balance equations With the momentum balance (2.69), this can be drastically simplified. We compute x (2.69) in index notation, t (ρv k) + (ρv k v l ),l = t kl,l + f k ε ijk x j t (ρ ε ijkx j v k ) + x j (ρ ε ijk v k v l ),l = x j (ε ijk t kl ),l + ε ijk x j f k t (ρ ε ijkx j v k ) + (ρ ε ijk x j v k v l ),l ρ ε ijk v k v l x j,l = (ε ijk x j t kl ),l ε ijk t kl x j,l + ε ijk x j f k, (2.79) and subtract this from the balance of angular momentum (2.78), which leaves ρ ε ijk v k v l δ jl = ε ijk t kl δ jl ρ ε ijk v k v j = ε ijk t kj 1(ρ ε 2 ijkv k v j + ρ ε ikj v j v k ) = 1(ε 2 ijkt kj + ε ikj t jk ) ρ ε ijk (v k v j v j v k ) = ε ijk (t kj t jk ) ε ijk (t kj t jk ) = 0. (2.80) Evaluation of this result for i = 1 yields ε 123 (t 32 t 23 ) + ε 132 (t 23 t 32 ) = 0 (t zy t yz ) (t yz t zy ) = 0 t yz = t zy. (2.81) Similarly, for i = 2 and 3 one finds t xz = t zx and t xy = t yx, respectively. The balance of angular momentum thus reduces to the statement that the Cauchy stress tensor is symmetric, t = t T, or t ij = t ji. (2.82) By contrast, an independent jump condition of angular momentum does not exist; it is fully equivalent to the momentum jump condition (2.72). 25

2 Elements of continuum mechanics 2.3.7 Energy balance Balance of kinetic energy We now compute the dot product of the momentum balance (2.71) and the velocity v in index notation: dv k ρv k dt = t kl,lv k + f k v k ρ d ( ) vk v k = (t kl v k ),l t kl v k,l + f k v k dt 2 = (t v) l,l (t L) ll + f k v k ρ d ( ) v 2 = div (t v) tr (t L) + f v (2.83) dt 2 [in the step from line 2 to 3 the symmetry of t, Eq. (2.82), has been used, and in the last line we have introduced the speed v = v, which is the absolute value of the velocity]. For the second summand on the right-hand side, tr (t L), we apply the decomposition (2.27) of L, the symmetry of t and the antisymmetry of W, so that tr (t L) = tr (t D) + tr (t W ) = tr (t D) + t ij W ji = tr (t D) + 1 2 (t ijw ji + t ji W ij ) = tr (t D) + 1 2 (t ijw ji t ij W ji ) = tr (t D), (2.84) ρ d ( ) v 2 = div (t v) tr (t D) + f v. (2.85) dt 2 Since the kinetic energy of a mass m is given by mv 2 /2, the term v 2 /2 denotes the specific kinetic energy (per unit mass) of a continuous body. Comparison of the above result with the general balance equation (2.63) shows that it can be interpreted as the balance of kinetic energy, where g = ρv 2 /2 (kinetic energy density), g s = v 2 /2 (specific kinetic energy), φ = t v (power of stresses), p = tr (t D) ( p: dissipation power), s = f v (power of volume forces). (2.86) 26

2.3 Balance equations The attribution of the dissipation power as a production term and the power of volume forces as a supply term was done because the former is only due to intrinsic quantities, whereas in the latter the volume force occurs which has an external source. Thus, in contrast to mass, momentum and angular momentum, the kinetic energy has a non-zero production density, which means that it is not a conserved quantity. Energy balance, balance of internal energy The balance of kinetic energy, Eq. (2.85), is not an independent statement, but a mere consequence of the momentum balance (2.71). However, classical mechanics and thermodynamics tells us that the kinetic energy is only one part of the total energy of a system (here: continuous body), and that the total energy is a conserved quantity (no production). In order to formulate the (total) energy balance, we thus extend (2.86) by introducing an internal energy, a heat flux and a radiation power and setting the production to zero: g = ρ(u + v 2 /2) (u: specific internal energy), g s = u + v 2 /2, φ = q t v (q: heat flux), p = 0, s = ρr + f v (r: specific radiation power). (2.87) By inserting these densities in (2.48), the energy balance [ ρ ( u + v2 ) ] [ + div ρ ( ] u + v2 ) v t 2 2 = div q + div (t v) + ρr + f v (2.88) is obtained, which is also known as the First Law of Thermodynamics. An alternative form follows from (2.63), ρ d dt ( v 2 ) u + = div q + div (t v) + ρr + f v. (2.89) 2 27

2 Elements of continuum mechanics This can be simplified further: ρ du dt + ρv dv k k dt = q k,k + (t kl v k ),l + ρr + f k v k = q k,k + t kl,l v k + t kl v k,l + ρr + f k v k ρ du { dt + v k ρ dv } k dt t kl,l f k = q k,k + t lk L kl + ρr. (2.90) The term in curly brackets vanishes identically because of the momentum balance (2.71), and the term t lk L kl = tr (t L) can again be replaced by tr (t D) [see Eq. (2.84)], so that we obtain ρ du dt = div q + tr (t D) + ρr. (2.91) Evidently, this is the balance of internal energy in the form (2.63), with the corresponding densities g = ρu, g s = u (specific internal energy), φ = q (heat flux), p = tr (t D) (dissipation power), s = ρr (r: specific radiation power). (2.92) In contrast to the total energy, the internal energy is not a conserved quantity. Its production density is equal to the dissipation power, which already appeared in the balance of kinetic energy (2.85) with a negative sign. The name dissipation power results from the fact that it annihilates kinetic energy and changes it into internal energy. In other words, macroscopic mechanical energy is transformed into heat (microscopic, unordered motion). Therefore, the dissipation power can also be interpreted as heat production due to internal friction. From (2.54) and (2.87) we obtain the energy jump condition [q n] [v t n] + [[ ρ ( u + v2 2 ) ((v w) n) ]] = 0. (2.93) In case of a material singular surface (v + n = v n = w n) the third summand vanishes, and because of the continuity of the stress vector t n [Eq. (2.73)] it can be 28

2.4 Constitutive equations factored out in the second summand: [q n] [v ] t n = 0 [q n] [v ] t n [[ ]] v t n = 0 [q n] [[ ]] v t n = 0 (2.94) [where v = v + v, with the normal component v = (v n)n and the tangential component v = v (v n)n; the jump of v vanishes because of v + n = v n]. Only under the additional assumption of no-slip conditions, that is, [[ ]] v = 0, the normal component of the heat flux (q n) becomes continuous. 2.4 Constitutive equations 2.4.1 Simple bodies with fading memory The balance equations of mass, momentum and internal energy derived in Sect. 2.3 read dρ dt ρ dv dt ρ du dt = ρ div v, (2.95) = div t + f, (2.96) = div q + tr (t D) + ρr (2.97) [see Eqs. (2.58), (2.71) and (2.91); the balance of angular momentum is implicitly included in the symmetry of t]. They constitute evolution equations for the unknown fields ρ, v and u; however, on the right-hand sides the fields t and q are also unknown. The supply terms f and r are assumed to be prescribed as external forcings. Therefore, in components we have 1 + 3 + 1 = 5 equations (the momentum balance is a vector equation!) for the 1 + 3 + 6 + 1 + 3 = 14 unknown fields ρ, v, t (symmetric tensor!), u and q, and the system is highly under-determined. Therefore, additional closure relations between the field quantities are required. These closure relations describe the specific behaviour of the different materials (whereas the balance equations are universally valid), and they are called constitutive equations or material equations. The general theory of constitutive equations is a long story and beyond the scope of this text. For a comprehensive treatment the reader may see the textbooks on 29

2 Elements of continuum mechanics continuum mechanics given in the literature list. Here, we confine ourselves to a simple class of materials, the so-called simple bodies with fading memory. This will be sufficient for our purpose of describing ice-dynamic processes. A simple body with fading memory is defined as a material whose constitutive equations are functions of the form t = t (F, Ḟ, T, T, grad T ), q = q (F, Ḟ, T, T, grad T ), u = u (F, Ḟ, T, T, grad T ) (2.98) (sometimes also higher-order time derivatives of the deformation gradient F are considered), where the temperature T (x, t) has been introduced as an additional scalar field quantity. Hence, the Cauchy stress tensor t, the heat flux q and the specific internal energy u are understood as the dependent material quantities, and the material shows neither non-local nor memory effects. Note that due to Eq. (2.25) the dependency on Ḟ can also be expressed as a dependency on the velocity gradient L. In the following, we will discuss two examples of simple bodies with fading memory relevant for ice dynamics, namely the linear-elastic solid (Hookean body) and the Newtonian fluid. 2.4.2 Linear-elastic solid An elastic body is defined as a material for which the stress tensor depends on the deformation gradient only: t = t(f ). (2.99) In particular, this excludes any temperature dependencies, so that the problem is purely mechanical, and the energy balance (2.97) need not be taken into account. For many practical applications, it is sufficient to assume small deformations, that is, F 1. If we use the same origins (B = 0) and bases (e i = δ ia E A ) for the reference and the present configuration (see Fig. 2.1), then the displacements u = x X will be small, that is, x X. The reference configuration and the present configuration virtually fall together. It is then no longer necessary to distinguish between material and spatial derivatives ( / x i / X A for i = A). For this situation, the displacement gradient H is defined as H = Grad u = F 1, (2.100) 30

2.4 Constitutive equations and the infinitesimal strain tensor ε is the symmetric part of it, ε = 1 2 (H + HT ), or ε ij = 1 2 (u i,j + u j,i ). (2.101) The constitutive equation of the isotropic, linear-elastic solid for small deformations, also known as Hookean body, is now t = (λ tr ε) 1 + 2µ ε. (2.102) This material equation is called Hooke s law, and the two coefficients λ, µ are the Lamé parameters. For the Hookean body, the mass balance (2.95) can be integrated directly. With Eq. (2.26) and F 1 J 1, we calculate ρ + ρ div v = 0 ρ ρ + J J = 0 dρ ρ + dj J = 0 t ln ρ ρ 0 + ln J = 0 ρ ρ 0 J = 1 ρ = ρ 0 J ρ 0, (2.103) where t 0 is the initial time which defines the reference configuration, and ρ 0 is the constant density in the reference configuration. We now insert Hooke s law (2.102) in the momentum balance (2.96). For the divergence of the stress tensor we find t 0 (div t) i = t ij,j = (λ ε kk δ ij + 2µ ε ij ),j = (λ u k,k δ ij ),j + µ u i,jj + µ u j,ij = λ u k,kj δ ij + µ u i,jj + µ u j,ji = (λ + µ) u k,ki + µ u i,jj div t = (λ + µ) grad div u + µ 2 u, (2.104) where the Laplace operator 2 is defined as 2 = div grad. With this, Eq. (2.103) and v = ẋ = (u + X) = u we obtain ρ 0 d 2 u dt 2 = (λ + µ) grad div u + µ 2 u + f. (2.105) This is the equation of motion for the Hookean body, and it is known as the Navier equation. It consists of three component equations for the three unknown displacement 31

2 Elements of continuum mechanics components u x, u y und u z, so that indeed a closed system is at hand. 2.4.3 Newtonian fluid Compressible Newtonian fluid A material is called viscous if the material function (2.98) 1 for the stress tensor t contains an explicit dependency on Ḟ or L. The most important realisation of a viscous material is the Newtonian fluid (also called linear-viscous fluid), which can either be compressible or density-preserving (incompressible). For the compressible case, t depends linearly on the strain-rate tensor D (symmetric part of L) and the density ρ by the following material function, t = p(ρ) 1 + (λ(ρ) tr D) 1 + 2η(ρ) D, (2.106) where p is the pressure, and λ and η are the coefficients of viscosity. The parameters η and κ = λ + 2η/3 are also known as the shear viscosity and the bulk viscosity, respectively. Except for the pressure term, the material function (2.106) corresponds largely to the Hooke s law (2.102), and by a computation similar to (2.104) we find for the divergence of the stress tensor div t = grad p + (λ + η) grad div v + η 2 v. (2.107) With this result, the momentum balance (2.96) yields the equation of motion ρ dv dt = grad p + (λ + η) grad div v + η 2 v + f, (2.108) which is the Navier-Stokes equation for the case of a compressible Newtonian fluid. Note the formal similarity to the Navier equation (2.105) for the Hookean body. Together with the mass balance (2.95), we now have four component equations for the four unknown field components v x, v y, v z and ρ, which is again a closed system. Density-preserving (incompressible) Newtonian fluid For the density-preserving (incompressible) Newtonian fluid, ρ = const holds, so that the mass balance reduces to div v = 0 (2.109) 32

2.4 Constitutive equations [see Eq. (2.59)]. It is then convenient to split up the stress tensor according to t = p 1 + t D, (2.110) where p = 1 tr t (2.111) 3 denotes the pressure, and t D is the traceless stress deviator (tr t D = 0). Now the material function (2.98) 1 only determines the stress deviator t D and reads t D = 2η D, (2.112) where the coefficient η is again the shear viscosity (or simply the viscosity). In order to derive the equation of motion for the density-preserving case, we compute the divergence of the stress tensor with the decomposition (2.110), the material function (2.112) and the mass balance (2.109): div t = grad p + div t D, (2.113) where (div t D ) i = 2η D ij,j = η (v i,jj + v j,ij ) = η (v i,jj + v j,ji ) = η [v i,jj + (div v),i ] = η v i,jj = η ( 2 v) i. (2.114) Insertion of these results in the momentum balance (2.96) yields the Navier-Stokes equation for the density-preserving Newtonian fluid, ρ dv dt = grad p + η 2 v + f. (2.115) Note that here the pressure p appears as a free field, so that we have the four component equations (2.109) and (2.115) for the four unknown field components v x, v y, v z and p. If the viscosity η is temperature-dependent, then a thermo-mechanically coupled problem is at hand, for which the energy balance (2.97) must be solved in addition. This requires to specify the material functions (2.98) 2 and (2.98) 3 for the heat flux and the internal energy, respectively. Insertion in the energy balance yields the missing evolution equation for the temperature. For instance, if the heat flow is given by Fourier s law of heat conduction, q = κ grad T (2.116) 33

2 Elements of continuum mechanics (κ: heat conductivity), and the internal energy is proportional to the temperature, u = ct (2.117) (c: specific heat), then the energy balance (2.97) results in ρc dt dt = κ div grad T + tr [( p1 + 2ηD) D] + ρr = κ 2 T p tr D + 2η tr D 2 + ρr ρc dt dt = κ 2 T + 2η tr D 2 + ρr (2.118) (note that tr D = div v = 0). Now we have the five equations (2.109), (2.115) and (2.118) which govern the evolution of the five fields v x, v y, v z, p and T. 34

3 Large-scale dynamics of ice sheets 3.1 Constitutive equations for polycrystalline ice 3.1.1 Microstructure of ice The modification of H 2 O ice which exists under the pressure and temperature conditions encountered in ice sheets and glaciers is called ice Ih. It forms hexagonal crystals, that is, the water molecules are arranged in layers of hexagonal rings (Fig. 3.1). The plane of such a layer is called the basal plane, which actually consists of two planes shifted slightly (by 0.0923 nm) against each other. The direction perpendicular to the basal planes is the optic axis or c-axis, and the distance between two subsequent basal planes is 0.276 nm. Figure 3.1: Structure of an ice crystal. The circles denote the oxygen atoms of the H 2 O molecules. (a) Projection on the basal plane. (b) Projection on plane indicated by the broken line in (a). Figure by Paterson (1994), modified. Owing to this relatively large distance, the basal planes can glide on each other when a shear stress is applied, comparable to the deformation of a deck of cards. To a much lesser extent, gliding is also possible in the prismatic and pyramidal planes (see 35

3 Large-scale dynamics of ice sheets Fig. 3.2). This means that the ice crystal responds to an applied shear stress with a permanent deformation, which continues as long as the stress is applied (creep, fluidlike behaviour) and depends on the direction of the stress relative to the crystal planes (anisotropy). Figure 3.2: Basal, prismatic and pyramidal glide planes in the hexagonal ice-ih crystal. Measurements have shown that ice crystals show some creep even for very low stresses. In a perfect crystal such a behaviour would not be expected. However, in real crystals dislocations occur, which are defects in its structure. These imperfections make the crystal much more easily deformable, and this is enhanced even more by the fact that during creep additional dislocations are generated. This creep mechanism is consequently called dislocation creep. 3.1.2 Creep of polycrystalline ice Naturally, ice which occurs in ice sheets and glaciers does not consist of a single ice crystal. It is rather composed of a vast number of individual crystals, the typical size of which is of the order of millimeters to centimeters. Such a compound is called polycrystalline ice. An example is shown in Fig. 3.3. The c-axis orientations of the individual crystals (grains) in polycrystalline ice differ from one another. In the following, we will assume that the orientation distribution is completely at random. In this case, the anisotropy of the single crystals averages out in the compound, so that its macroscopic behaviour will be isotropic. In other words, the material properties of polycrystalline ice do not show any directional dependence. Let us assume to conduct a shear experiment with a small sample of polycrystalline ice as sketched in Fig. 3.4 (left panel). The shear stress τ is assumed to be held constant, and the shear angle γ is measured as a function of time. The resulting creep curve γ(t) is shown schematically in Fig. 3.4 (right panel). An initial, instantaneous elastic 36

3.1 Constitutive equations for polycrystalline ice Figure 3.3: Thin-section of polycrystalline glacier ice regarded between crossed polarization filters. The individual ice crystals are clearly visible, and their apparent colours depend on the c-axis orientation. deformation of the polycrystalline aggregate is followed by a phase called primary creep during which the shear rate γ decreases continuously. This behaviour is related to the increasing geometric incompatibilities of the deforming single crystals with different orientations. After some time, a minimum shear rate is reached which remains constant in the following, so that the shear angle increases linearly with time. This phase is known as secondary creep. Especially in case of rather high temperatures ( 10 C), at a later stage dynamic recrystallisation (nucleation and growth of crystals which are favourably oriented for deformation) sets in, which leads to accelerated creep and finally a constant shear rate (linear increase of the shear angle with time) significantly larger than that of the secondary creep. This is called tertiary creep. Figure 3.4: Shear experiment for a sample of polycrystalline ice. τ denotes the applied shear stress, γ the shear angle and t the time. 3.1.3 Flow relation In good approximation we can assume that in deforming ice masses like ice sheets and glaciers secondary creep prevails for low temperatures ( 10 C), whereas tertiary 37