Crystal Plasticity PDF

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Exercises EM Course on

Multi-scale and Micromechanics – 2020

Crystal Plasticity

Introduction
In these exercises, a simple crystal plasticity model will be used. In this model, the deformation gradient
tensor F is decomposed into a plastic and an elastic component, denoted by the subscripts “p” and “e”,
respectively:

F = F e · F p. (1)

The velocity gradient tensor L = FÛ · F −1 is decomposed according to:

L = Le + F e · Lp · F e −1 with Le = FÛ e · F e −1 ; Lp = FÛ p · F p −1 , (2)

respectively. For plastic deformation by crystallographic slip, the term Lp in Equation (2) equals:
N
Õ
Lp = γÛ α s®0α n®α0 , (3)
α =1

with γÛ α the shear rate on the slip system given by the slip plane normal n®α and the slip direction s®α .
The plastic shear rates are related to the stress state, in particular the shear stress on the slip plane in
the direction of slip, the so-called resolved shear stress τ α . This shear stress is defined such that the
plastic work per volume in the intermediate configuration equals
N
Õ
τ α γÛ α = J σ : (L − Le ). (4)
α =1

Consequently, the resolved shear stress can be written as:

τ α = Jσ : Pα . (5)

where P α = s®α n®α denotes the non-symmetric Schmid tensor of the αth slip system. Then, a constitutive
law of the form γÛ α (τ α ) needs to be formulated. Here, plastic deformation on a slip system is modeled
by a visco-plastic (rate-dependent) power law relation of the form:
n−1
α τ α τ α
γÛ = γÛ0 α α . (6)
τ0 τ0

The slip resistance τ0α is assumed to be constant. Furthermore, the following elastic law is used:

τ = 4C : Ee, (7)

with E e the elastic Green-Lagrange strain tensor and τ the elastic second Piola–Kirchhoff stress tensor.
By using this rate-independent elastic law, both the total deformation gradient tensor and the Cauchy

1
stress tensor can be expressed in terms of the elastic deformation gradient tensor, if an integration law
for the plastic flow rule FÛ p = Lp · F p is used. If Lp is assumed to be constant during an increment with
length ∆t, the plastic deformation gradient at the end of this increment, at time level tn+1 = tn + ∆t, is
given by

F p (tn+1 ) = exp(∆tLp ) · F p (tn ). (8)

For the evaluation of the tensor exponential, the Padé approximation is used:
  −1  
∆t ∆t
F pinc = exp(∆tLp ) ≈ I − Lp · I + Lp . (9)
2 2

For finite increments, the determinant of the approximated exponential term may deviate slightly from
unity, which is corrected by the following normalization:
− 31
F p (t n+1 ) = F̃ pinc · F p (t n ) with F̃ pinc = [ det(F pinc ) ] F pinc . (10)

Then, by substituting equations (1, 3, 5–7, 8–10) into each other, both the Cauchy stress tensor σ and
the deformation gradient tensor F can be expressed in F e . When F is prescribed, a set of nonlinear
equations in terms of the elastic deformation gradients is obtained. This set of equations is solved by a
Newton–Raphson procedure.

Material parameters
This crystal plasticity model will be used to describe the mechanical behavior of fictitious cubic crystals.
The material parameters for these crystals are given in Table 1, where the shear strength τ0 of each slip
system is assumed to be constant.

γÛ0 [s−1 ] τ0 [MPa] n [−] E [MPa] ν [−]


0.01 1 40 10 0.3

Table 1. Material parameters.

The Miller index notation will be used for slip systems. The slip plane normal n® and slip direction s® of
the slip system (hkl)[uvw] in a cubic crystal are given by:

e1 + k e®2 + l e®3
h® ue®1 + v e®2 + w e®3
n® = √ ; s® = √ , (11)
h2 + k 2 + l 2 u2 + v2 + w 2
where the coordinate system e®i is defined by the principal crystallographic axes. Crystallographic
orientation are shown in pole figures, which are explained in Appendix A.

Matlab code
The crystal plasticity framework is implemented in a number of Matlab functions. For this purpose,
a series of time-increments are defined. For each increment, the elastic deformation gradient tensor
of a crystal is determined with the function celpl for a given total deformation gradient tensor F.
This function requires as input the elastic and viscoplastic material parameters of the crystal, its initial
orientation and Schmid tensors, the magnitude of the time step, the plastic deformation gradient at the
beginning of the increment, the total deformation gradient at the end of the increment, and an initial
estimate for the elastic deformation gradient at the end of the increment. The calculated elastic and

2
plastic deformation gradients at the end of the increment are inserted into the function at the beginning
of the next increment. Besides these quantities, also the Cauchy stress tensor, the resolved shear rates,
and a new crystal orientation are given as output.
In the Matlab script emcp.m that will be used for the first exercise, a single crystal with one slip
system is considered. The slip system, which is chosen such that all deformation is in the 12-plane,
is defined in the function slipone. With the rotation tensor Q, an in-plane rotation with angle ϕ
can be applied. Alternatively, random orientations can be generated with the function randq. During
the simulation, rigid body rotation of the crystallographic [100]-direction is shown in a pole figure.
The rotations of the crystallographic directions
q [11̄0] and [110] are displayed in another pole figure.
Furthermore, the equivalent stress σ eq = 32 σ d : σ d is plotted versus the stretch ratio λ. Finally, also
the magnitude of the resolved shear rates is shown versus the stretch ratio.
In the exercises, the script will first be expanded to multiple slip systems and then to multiple crystals,
where a Taylor interaction law will be used to specify the deformation of each crystal.

Getting started

• Run the Matlab script startup.m. This sets the required paths.
• Set as current working directory, the directory problems. This directory contains all Matlab
scripts that need to be modified. The main file is startup.m.

1 Single crystal: planar deformation

In this exercise, assume a single crystal with one slip system. Let the crystallographic axes e®i be aligned
with the axes of the global coordinate system ϵ®i . Let the slip system be given by (11̄0)[110].

1.1 What is the initial orientation of the slip plane normal and the slip direction? What is the corre-
sponding symmetric Schmid tensor?

Let the deformation of this crystal be given by the deformation gradient tensor:
F = λϵ®1ϵ®1 + λ−1ϵ®2ϵ®2 + ϵ®3ϵ®3 , (12)
where the stretch ratio is given by:
λ = exp(εÛ0t) ; εÛ0 = 0.01 s−1 ; 0 ≤ t ≤ 100 s. (13)
This deformation-history represents isochoric plane strain deformation with extension in e®1 -direction.
For the simulations, use the Matlab script emcp.m.

1.2 How do the orientations of the slip plane normal and the slip direction evolve with time? How
does the shear rate evolve with time? What is the consequence for large deformations? Explain
the stress-strain curve that you find (keep in mind that no hardening is present in the crystal
plasticity model).

1.3 What happens if the crystallographic axes are initially not aligned with the global coordinate
system (i.e. ϕ , 0◦ )? Use different initial orientations (−45◦ ≤ ϕ ≤ 45◦ ).

1.4 Add the slip system (110)[11̄0] and repeat the previous questions (modify the Matlab script
emcp.m so that the function sliptwo.m is used, where the slip system can be added). Let
the crystallographic axes again be aligned with the global coordinate system.

3
1.5 In this model, the resolved shear rate of a slip system is related to the resolved shear stress by the
power law relation in equation (6). What is the influence of the parameters n and τ0 ?

1.6 Add the slip systems (100)[010] and (010)[100]. Subject this crystal (with now four slip systems)
to plane strain tension for ϕ = 0◦ . Which slip systems are active (and why)?

1.7 Subject this crystal with ϕ = 0◦ to simple shear:

F = I + γ ϵ®1ϵ®2 ; γ = γÛ0 t ; 0 ≤ t ≤ 50 s and γÛ0 = 0.01 s−1 . (14)

Which slip systems are now active?

2 Single crystal: kinematic constraints

Five independent slip systems are required in order to be able to accommodate any arbitrary volume-
invariant plastic deformation by the mechanism of crystallographic slip.

2.1 Why do five independent slip systems suffice for any arbitrary volume-invariant plastic deforma-
tion?

Consider the following set of slip systems for a hypothetical cubic crystal:

No. Slip system


1 (100)[010]
2 (100)[001]
3 (100)[011]
4 (010)[001]
5 (110)[11̄0]

Table 2. Slip systems for a hypothetical crystal.

2.2 Derive the symmetric Schmid tensors for these slip systems.

2.3 Can this crystal undergo an arbitrary isochoric plastic deformation (why)? If not, what is the
additional kinematic constraint (besides plastic incompressibility)?

2.4 Apply volume-invariant uniaxial deformation (with λ(t) = exp(0.01t)) to this crystal (slipfive.m).
Use various initial orientations (3d random initials orientations can be generated with the function
randq). What is the effect of the initial orientation?

2.5 Add the required slip system(s) that will allow any arbitrary plastic deformation by crystallo-
graphic slip. What is the difference, when subjected to volume-invariant uniaxial deformation?

4
3 Single crystal: FCC lattice

The general arrangement of the atoms in metals with a Face Centered Cubic (FCC) structure, like alu-
minum and copper, can be described by specifying the atom positions in a repeating unit cell as is
represented in Figure 1(a). One atom is situated at each corner and at the center of each face of the

(a) (b)

Figure 1. (a) FCC unit cell and (b) the (111) plane and its slip directions.

unit cell. This arrangement allows four planes to be most densely packed, the {111}-family of planes,
which will act as slip planes. Combined with the most densely packed directions, the h11̄0i-family, 12
slip systems can be formed, which are summarized in Table 3. In Figure 1(b), a slip plane and its slip
directions are shown. The FCC slip systems are implemented in the Matlab-script slipfcc.m.

Slip plane Slip directions

(111) [101̄] [011̄] [11̄0]


(111̄) [011] [101] [11̄0]
(11̄1̄) [101] [011̄] [110]
(11̄1) [011] [101̄] [110]

Table 3. Slip systems in an FCC lattice.

3.1 Do the 12 slip systems form a set of 5 independent slip systems? What does this imply for the
ductility of a material composed of FCC crystals?

3.2 Apply volume-invariant uniaxial deformation to the crystal with various initial orientations. How
much does the global response vary with initial orientation? How does this compare to the crystal
in exercise 2.

3.3 How many slip systems are typically active for one crystal?

5
4 Multiple crystals: Taylor approach & FCC lattice

The crystal plasticity framework provides a constitutive model for single crystals. This constitutive
model could for example be used within a finite element (FE) context, where each grain is subdivided
into several elements (see Figure 2). In this approach, the average element size typically is smaller than

Figure 2. Finite element mesh of a crystalline structure.

the average grain size. Another approach uses a polycrystalline model for an aggregate of crystals to
describe the constitutive behavior within a material point of a FE model. In the latter approach, the
element or aggregate size is considerably larger than the average grain size. In a polycrystalline model,
an interaction law is used to relate the deformation and stress fields of a material point to the local
deformation and stress fields of the individual crystals.
The most widely used interaction law in polycrystal plasticity is the Taylor interaction law. This in-
teraction law assumes that the local deformations for each crystal equal the macroscopically imposed
deformation of the aggregate. The macroscopic Cauchy stress tensor is taken to be the volume-averaged
Cauchy stress of the aggregate. The Taylor interaction model is the viscoplastic equivalent to the Voigt
average model that is frequently used for elastic polycrystals. The one-dimensional analogue for this
model is a parallel connection of the microscopic elements, see Figure 3.

Figure 3. Schematic illustration of the Taylor model.

6
4.1 Modify the Matlab script emcp.m such that it can be used for multiple crystals. Use the Taylor
model to specify the interactions between the crystals. Use random initial orientations and FCC
slip systems. Prescribe isochoric uniaxial deformation for the aggregate of crystals.

4.2 How does the global response relate to the responses of the individual crystals in the previous
exercise?

4.3 How does the texture evolve?

5 Multiple crystals: Taylor approach & kinematic constraints


5.1 Use the Taylor approach for an aggregate of crystals with slip systems as given in Table 2. What
is the consequence of using the Taylor interaction model for these crystals?

5.2 Are there alternative approaches that are more suitable for these crystals?

7
A Pole figures

To obtain a graphical representation of orientations of for example crystals, the three-dimensional ori-
entations must be converted to a two-dimensional graph. A commonly used method for this is repre-
sentation by pole figures.
A direction in the three-dimensional space can be considered as a vector, which can be positioned with
its origin in the origin of the global coordinate system. If this vector is normalized to unit-length, the
end-point will be located on the unit sphere with its origin in the origin of the coordinate system, as is
graphically shown in Figure 4(a). A pole figure is then constructed by projecting the end-point of the
unit vector towards a particular cross-section through the origin of the unit sphere (see Figure 4(b)).
This plane is referred to as the equatorial plane, i.e. the plane between the north and the south pole of
the unit sphere. A two-dimensional representation of the equatorial plane then forms the constructed
pole figure (see Figure 4(c)). A location in the pole figure is given by its radial coordinate ρ and the
angle φ.

3 3

2 2
1 1

(a) (b)
√ √
2 3 2ρ 2

1
ρ

(c) (d)

Figure 4. Construction of pole figures. (a) vector with unit length, (b) projection to equatorial plane,
(c) pole figure (equatorial plane), (d) equal area projection.

For the actual projection of a unit vector towards the equatorial plane, different techniques exist, re-
sulting in slightly different pole figures. The equal area projection technique results in a homogeneous
distribution for a set of random orientations. With this technique (see Figure 4(d)), the radial coordinate
ρ in the pole figure is obtained by the length of the line connecting the location of the vector
√ end-point
on the unit sphere with the pole on the same hemisphere, down-scaled by a factor of 2, which is the
length of the line connecting the pole with a location on the edge of the equatorial plane.
In Figure 5 some examples of three-dimensional vectors and the corresponding marks in a pole figure
are given. A vector pointing towards either the north or the south pole results in a mark in the center of
the pole figure, whereas a vector within the equatorial plane corresponds to a mark on the outer edge of
the pole figure (see Figure 5(a) and (b)). All vectors within a plane through the origin of the unit sphere
and perpendicular to the equatorial plane will be located on a straight line in the pole figure (Figure 5(c)
and (d)). Two vectors that are mirrored by the equatorial plane will result in the same position in the

8
pole figure. Therefore, from the pole figure, it cannot be deduced in which hemisphere the vector is
located.
2
3
2
1 1

(a) (b)

2
3

2
1 1

(c) (d)

Figure 5. Examples of the construction of pole figures.

With this graphical representation technique, the orientation of a single vector can be displayed. How-
ever, to record the complete orientation of a crystallographic lattice (i.e. the orientation of a coordinate
system), the orientations of at least two directions are needed. Therefore, for a complete orientation
description, a combination of at least two pole figures is required.

You might also like