Sie sind auf Seite 1von 12

Powder Technology 345 (2019) 589–600

Contents lists available at ScienceDirect

Powder Technology

journal homepage: www.elsevier.com/locate/powtec

Prediction of the anisotropic mechanical properties of compacted


powders
Peter Loidolt a, Manfred H. Ulz b, Johannes Khinast a,⁎
a
Institute of Process and Particle Engineering, Graz University of Technology, Graz 8010, , Austria
b
Institute of Strength of Materials, Graz University of Technology, Graz 8010, Austria

a r t i c l e i n f o a b s t r a c t

Article history: The multi-particle finite element method (MPFEM) was used to test the anisotropic elastic and plastic properties
Received 25 July 2018 of compacted powders with cohesive contacts. A representative volume element (RVE) of monodisperse, spher-
Received in revised form 9 January 2019 ical, deformable particles was used to investigate the powder properties after compaction. Efficient periodic
Accepted 20 January 2019
boundary conditions and an RVE of only 50 particles allowed extensive parameter studies. During parameter
Available online 24 January 2019
studies the relative density after compaction, the contact cohesion strength and the strain path during compac-
Keywords:
tion were varied. The strain paths were characterized by the ratios of the applied principal strains during compac-
Powder compaction tion that results in different Dirichlet boundary conditions on the RVE. Seven different strain paths were
Multi-particle finite element method considered including the practically important isostatic and closed die compaction.
Cohesive particle contact The outcome of the parameter study were the elastic constants of an orthotropic material model, the uniaxial
Anisotropy yield strength for tension and compression, and the yield surfaces for general load cases. No anisotropy was ob-
Elasticity served for isostatic compaction but increasing anisotropy was observed with increasing ratio of the principal
Yield surface strains during compaction. Regression curves were generated to describe the mechanical properties as a function
of the model parameters. In this way, continuous functions were obtained which were capable to describe the
distribution of the mechanical material properties in a FEM model of a heterogeneous compacted powder part.
© 2019 Published by Elsevier B.V.

1. Introduction real system that requires the identification of assumedly negligible ef-
fects on the system to set up a suitable model. Suitable models are not
Compaction of powders is a very common industrial process. It is well established for powder materials. There are some very basic phe-
used for producing green bodies before sintering a metallic or ceramic nomenological models available which correlate the applied pressure
part, to produce pellets of minerals or animal food or to manufacture with the relative density or porosity of a powder (compressibility),
tablets in the pharmaceutical industry [1–4]. The requirements depend e.g. the Heckel [5] and Kawakita [6] equations. Other models
on the application: for example, the total mass and its uniformity is cru- correlate the relative density of a powder compact with its strength
cial in the pharmaceutical industry, whereas shape of the parts is essen- (compactability) [7]. The constants of these models differ from powder
tial for structural parts in mechanical engineering. The properties of the to powder and have to be estimated during compaction experiments by
compacted powder depend on the particle properties, such as particle force measurement. The benefit of these models is limited, since there is
size and shape, the mechanical features of the particles and the no spatial resolution which may be crucial for a complex stress distribu-
particle-surface properties. In addition to that the process conditions tion inside the part. Furthermore, the gain of mechanical understanding
during compaction play a major role. The geometry and the surface of is marginal.
the tools, as well as the load control during compaction, are three of In opposite to the simple phenomenological models, the compaction
the most important influencing factors of the process. processes can as well be described by means of mechanical models as
Simulations are an important tool, next to experiments, to under- proposed in [8,9]. In these studies, yield surfaces for powders are de-
stand the details of a process and to optimize powder compaction. The rived analytically, leading to low computational costs compared to the
advantage of process optimization via simulation is the possibility to model in the current work. The shape of the yield surfaces changes as
perform extensive parameter studies without any investment costs for function of a parameter describing the cohesion strength. The model is
prototypes. The drawback of simulation is the simplification of the also capable of describing the anisotropic yield behavior of powder
after closed die compaction. Comparable results are obtained by the
⁎ Corresponding author. network model proposed in [10]. Compaction and sintering is also ana-
E-mail addresses: manfred.ulz@tugraz.at (M.H. Ulz), khinast@tugraz.at (J. Khinast). lyzed theoretically in the work of [11]. The drawback of analytical

https://doi.org/10.1016/j.powtec.2019.01.048
0032-5910/© 2019 Published by Elsevier B.V.
590 P. Loidolt et al. / Powder Technology 345 (2019) 589–600

models is the need of many simplifications as it is not possible to con- for orthotropic materials are wood, fabrics, fiber composite materials
sider rearrangement and deformation of particles within these models. and rolled sheet metals. If there exist an axis of rotational symmetry,
Due to the availability of high-performance computers the applica- then the material is called transverse isotropic. In this case two of the
tion of numerical methods became more common in recent years. In Young's moduli, shear moduli and Poisson's ratios are equal. In case of
the field of powder simulations, the discrete element method (DEM) an isotropic material there exists only one Young's modulus, one shear
and the finite element method (FEM) are used. In a DEM model the po- modulus and one Poisson's ratio.
sition of every particle is tracked. If two particles (theoretically) overlap, If the stress exceeds a certain threshold, then the powder deforms
a contact force is computed as a function of the overlap. The particles are plastically. To model yielding of powder in FEM the yield surface has
accelerated by the sum of all contact forces and the position of the par- to be provided in addition to the elastic properties. As soon as the stress
ticles is calculated based on the Newton's equations of motion. In this state reaches the yield surface the powder deforms plastically. A general
way rearrangement of particles is considered. Comprehensive studies form of the yield surface is given in Eq. (4). A good discussion about
which use DEM to obtain the yield properties of compacted powder standard requirements to yield surfaces is found in [18]. The Drucker-
are presented in [12–14]. With efficient codes, this method can be Prager/Cap model is a widely used form of the yield surface for powders
used to model up to many millions of particles [15,16]. Still, in most and is shown in Fig. 1 (see [19]). In the Drucker-Prager/Cap model the
cases the number of particles inside the process is too high to be cap- yield surface in stress space is divided into a shear failure surface Fs
tured by a DEM model. Furthermore, the deformation of particles can- and into a cap surface Fc. In both cases the von Mises equivalent stress
not be considered effectively in DEM. In contrast to DEM, the number q is computed as a function of the hydrostatic pressure p.
of particles is no issue in a FEM simulation since the powder is consid-
ered as a continuum. Therefore, individual particles are not considered, F ðσ Þ ¼ 0 ð4Þ
but the powder is assumed to have field properties which are distrib-
uted continuously, e.g., density, stress and strain. During a FEM simula- A suitable yield model will need two important ingredients for our
tion the stress inside the part is computed as a function of the strain. In study. Firstly, as the strength of the powder changes in a compaction
Eq. (1) the stress-strain relation is written in the Voigt matrix notation process, the yield surface has to grow during compaction (Fig. 1 right).
(symmetric tensors are denoted as vectors) where Cij is the 6 × 6 elas- Such a surface growth is commonly modeled by an isotropic hardening
ticity tensor. The inverse relation is shown in Eq. (2) where Sij is the law [18]. Hence, a huge amount of compression tests are necessary to
compliance tensor. determine the evolution of the yield surface. Secondly, the boundary
conditions of compaction may render the powder compact's material
σ i ¼ C ij ε j ð1Þ properties to be anisotropic. An arbitrary initial arrangement of parti-
cles, which are made of an isotropic material, will show different mate-
ε i ¼ Sij σ j ð2Þ rial properties in the three principal directions after closed die
compaction. Once again, the strength of the compact has to be tested
Since the elasticity tensor is symmetric there are only 21 indepen- in different loading directions to construct an anisotropic yield surface
dent elastic constants for a general elastic material (see for example that is not known a priori.
[17]). The number of elastic constants reduces in case of material sym- Several yield surfaces to describe anisotropic plastic deformations
metry. In Eq. (3) the compliance tensor is shown for a material with were developed by Hill [20]. These yield criteria are based on the isotro-
three perpendicular planes of symmetry (orthotropic material). pic von Mises yield criterion, are widely used for metals and composites,
2 3 2 32 3 and can be considered to be standard for FEM applications. We mention
ε 11 1 ν 12 ν 31 σ 11 exemplarily some modifications of this approach [21–23]. An alterna-
6 7 6 E1 − − 0 0 0 76 7
6 7 6 E1 E3 76 7 tive criterion for plastic anisotropy was presented by Barlat [24], who
6 7 6 1 ν 23 76 7
6 ε22 7 6 7 6 σ 22 7 introduced a criterion based on the weighted deviatoric stress tensor.
6 7 6 − 0 0 0 76 7
6 7 6 E2 E2 76 7 Variations of this work were describes by [25–28]. A recent book [29] re-
6 7 6 76 7
6 7 6 1 76 7 views the theory of pressure-sensitive plasticity which is important for
6 ε33 7 6 0 0 6
0 7 6 σ 33 7
7
6 7 6 E3 7 structured materials including voids. Recent studies of powder compac-
6 7¼6 7∙6 7 ð3Þ
6 7 6 1 76 7
6 7 6 0 0 7 6 7 tion reported experimental investigations of the strength anisotropy
6 2ε23 7 6 G23 7 6 σ 23 7
6 7 6 76 7 after closed die compaction in [30] and determination of the compact's
6 7 6 76 7
6 7 6 1 76 7 strength by simulation in [31].
6 2ε31 7 6 sym: 0 76 7
6 7 6 G31 7 6 σ 31 7 The objective of the current work is the investigation of the expected
6 7 6 76 7
4 5 4 1 54 5 anisotropic material properties after non-isostatic compaction. The goal
2ε12 G12 σ is to obtain reliable material properties (elastic and plastic) after com-
12
paction with different strain paths (i.e., different Dirichlet boundary
Here E1, E2 and E3 describe the Young's moduli, G12, G23 and G31 are conditions on a representative volume element, RVE, in compaction)
the shear moduli and ν12, ν23 and ν31 are the Poisson's ratios. Examples and different compact densities. Therefore, we employ a multi-particle

Fig. 1. left: Drucker-Prager/Cap model in the hydrostatic stress-equivalent stress plane; right: Representation of material hardening during compaction of the Drucker-Prager/Cap yield
surface.
P. Loidolt et al. / Powder Technology 345 (2019) 589–600 591

FEM (MPFEM [32]) model already introduced by us in [33] to study the theory is used, since only small strains are considered during elastic
elastic and plastic properties of a representative volume element RVE deformation and yielding of the RVE. This means the size of the RVE is
consisting of deformable particles. First the model is described. Then re- assumed to stay constant during testing. The Cauchy stress tensor is
sults on the elastic properties are given, followed by results for the yield defined as.
surfaces. Finally, a conclusion and outlook are provided. 2 3
F xx F xy F xz
6 Ax Ax Ax 7
2. Model 6 7
6F F yz 7
6 yx F yy 7
σ ¼6 7 ð5Þ
2.1. Representative volume element (RVE) of deformable particles 6 Ay Ay Ay 7
6 7
4 F zx F zy F zz 5
The present work is based on a model introduced by the authors [33] Az Az Az
where detailed information about the implementation is given and
compared to similar approaches in the literature. Extensive conver- The three force vectors Fx, Fy and Fz are taken from the three auxil-
gence studies for mesh resolution, the number of particles and the iary nodes and their components are used to compute the stress tensor.
solver settings can also be found there. The model is used to obtain ho- The cross sections of the RVE faces are denoted as Ax, Ay and Az. Simi-
mogenized quantities of a particulate system. Specifically, the homoge- larly, the strain tensor.
nized stress and strain inside a powder bed are determined. The stress 2    3
and the strain are needed to derive the elastic and plastic properties Δuxx 1 Δuxy Δuyx 1 Δuzx Δuxz
6 þ þ 7
for a continuum model of the powder. The model arranges a certain 6 lx 2 lx ly 2 lz lx 7
6    7
number of particles in a periodic RVE using the commercial multi- 6 1 Δu Δuyx Δuyy 1 Δuyz Δuzy 7
6 xy 7
purpose finite element package Abaqus 6.14. Due to the periodic bound- ε¼6 þ þ 7 ð6Þ
62 lx ly ly 2 ly lz 7
6 7
ary conditions an RVE of 50 monodisperse spherical particles as shown 6     7
4 1 Δuzx Δuxz 1 Δuyz Δuzy Δuzz 5
in Fig. 2 is sufficient to obtain representative averaged properties. To þ þ
eliminate the variation between different arrangements of particles, 2 lz lx 2 ly lz lz
three arrangements are simulated and averaged properties are deter-
mined. Besides, the particles in the multi-particle finite element can be computed based on components of the displacement vectors of
(MPFEM) model are discretized with standard three-dimensional solid the auxiliary nodes Δux, Δuy and Δuz. The dimensions of the RVE are
elements in Abaqus. The explicit solver is used during simulation and denoted as lx, ly and lz.
an appropriate mass scaling (mass scaling factor 10,000) is employed Each particle in the RVE is modeled as an elasto-plastic body which
to increase the time step size to 0.0005 s [33]. does not break during deformation. In reality this is the case for most
The periodic boundary conditions are a key factor in this model. The metallic materials. In this work the material properties of copper are
arrangement of the particles on the boundary is periodic, i.e., for each used. The yield stress is defined as a piecewise linear function of the
particle on one face of the RVE there exists a particle on the opposite plastic strain based on the experimental data in [34,35]. All material
face. The displacements of each pair of corresponding particles is con- properties are summarized in Table 1. If brittle materials would be con-
straint via linear constraint equation. The constraint equations include sidered, such as ceramics or some excipients in pharmaceutical indus-
the displacement of an auxiliary node. Three auxiliary nodes are used try, it would be necessary to use an appropriate fracture criterion to
to govern the relative displacements of the particles at two opposing model cracking of the particles. In this case, the MPFEM model is more
faces of the RVE. In addition to that, the constraint forces of all included complex as new surfaces are created during compaction and additional
constraints are summarized as reaction force at the auxiliary nodes. The material properties need to be considered.
displacements of auxiliary nodes are used to compute the strain of the The contact properties between particles are of importance, since
RVE, and the reaction force yields the stress of the RVE. During testing failure of powders often occurs at contacts between particles. To com-
of the mechanical properties of the powder the infinitesimal strain pute the contact stress a simple contact model including cohesion
stresses was implemented as VUINTERACTION subroutine in Abaqus.
The contact normal stress is computed as a function of the normal dis-
tance δN and the normal relative velocity δ_N of the slave node to the
master surface (see Table 2). The distance has positive values for pene-
trations and negative values for gaps. Similarly, the contact tangential
stress is computed as a function of the relative tangential displacement
δT and the tangential relative velocity δ_T . Three different cases are distin-
guished (see Table 2 and Fig. 3).
The total normal stress is a function of the repulsive stress σrep, the
contact cohesion stress σcoh and an additional damping stress. The

Table 1
Material properties of copper.

Property Value

Young's modulus 115 GPa


Poisson ratio 0.34
density 8920 kg/m3
σ/ MPa εplastic
150 0.00
250 0.06
300 0.30
350 1.00
400 2.50
450 5.00
Fig. 2. RVE of 50 deformable particles with periodic boundary conditions.
592 P. Loidolt et al. / Powder Technology 345 (2019) 589–600

Table 2
Contact model.

Normal direction Tangential direction

δN b δ0 σN = 0 σT = 0
δ0 b δN b 0 σrep = 0 σfric = 0
σ coh ¼ cmax ð1− δδN0 Þ
0 b δN σrep = k ∙ δN σcoh = cmax |k ∙ δT| b σrep ∙ μ : σfric = − k ∙ δT
jk∙δT jNσ rep ∙μ : σ fric ¼ −σ rep ∙μ∙ jδδTT j
total σ N ¼ σ rep −σ coh þ d∙δ_N σ T ¼ σ fric −0:01∙d∙δ_T

signs follow from the convention where pressure stress is counted as during compaction can therefore be written as shown in Eq. (8). Fur-
positive. The damping stress is a linear function of the relative normal thermore, the mean strain e is computed in a standard fashion from
speed and a damping coefficient d. The damping coefficient is chosen the dilatation (relative variation of the volume) as given in Eq. (9).
to damp contact oscillations and to simultaneously avoid substantial ad- The values of the principal strains for the seven different strain paths
ditional macroscopic stresses. Similar to the normal stress the total tan- are given as a fraction of the mean strain. In this way the ratios of the
gential stress is a function of the friction force and a low tangential three principal strains are fixed for a certain strain path irrespective of
damping stress. how dense the powder is compacted. In strain path case A (see Fig. 5)
The contact parameters used in the simulations are summarized in all three strains are equal, and therefore, this case describes isostatic
Table 3. Two different values of the contact stiffness are used for testing compaction. Isostatic compaction is relevant in powder metallurgy dur-
the yield properties and the elastic properties. A comparably high con- ing hot isostatic pressing of high quality products. In case B the powder
tact stiffness is used during testing the elastic properties to avoid artifi- is only compressed in x-direction and no strain is allowed perpendicular
cial softening due to too soft contacts. A lower contact stiffness is used to the x-direction. This case refers to closed die compaction and is the
during testing the yield properties, since it has no essential influence most often used strain path during compaction of powders. It is widely
on the yield strength, but a larger simulation time step can be used to used during tableting in the pharmaceutical industry and during pro-
speed up the simulations. The maximum cohesion strength cmax is var- duction of green bodies before sintering. In real compaction processes
ied in the range from 1 to 1000 MPa. The lower values describe the weak ideal homogenous deformation of the powder does not occur due to
bonding of particles after cold compaction while the high values exceed the friction between tools and powder, which affect the distribution of
the yield strength of the particle material and are therefore useful to de- the strain inside the powder (e.g., see [36]). Therefore, the strain path
scribe strong bonding of particles after sintering. The cohesion interac- is also not constant across the powder compact. We consider strain
tion length is given as a fraction of the particle diameter. The latter is path case C to case G to account for a possible inhomogeneous deforma-
chosen for simplicity to be 1 m. tion during compaction. Case C works similar to case B but in this case
the powder is stretched perpendicular to the x-direction. In cases D
2.2. Compaction simulation and E, the powder is compacted in y- and z-direction. While in case D
no straining takes place in x-direction, in case E a stretch is applied. Fi-
The MPFEM simulation procedure is divided in three steps. In Fig. 4 nally, three different strains are applied in cases F and G. The absolute
the evolution of the stress inside the powder is shown for simplicity in values for the strains are determined based on the initial packing den-
the 2D stress space although the real simulation is in 3D. In the first sity and the final packing density after compaction. The final packing
step the initial packing is compacted with strain- (deformation-)con- density was varied in the range from 0.65 to 0.95. For all cases the pow-
trolled tri-axial compression. After compaction a short dwell step is der was unloaded after compaction to obtain the stress free reference
used to allow for dissipation of possible particle oscillations. Then, the state which is tested in the last step. The dimension of the unloaded
load control of the RVE is shifted to stress (force) control and the load RVE is used to calculate the stresses and strains needed during the test-
is ramped down thoroughly to obtain a stress-free RVE. Starting from ing step.
this configuration the RVE is reloaded in different directions during
the testing step to determine the mechanical properties of the
Δli
compacted powder. ei ¼ − ð7Þ
L0
We introduce seven different strain paths for the compaction step
(see Fig. 5). The initially cubic RVE is strained along the x-, y- and z-
2 3
direction, i.e., the three different strains in direction of the coordinate e1
system are equal to the principal strains during compaction as no e ¼ 4 e2 5 ð8Þ
shear strains develop. The strain is computed based on the engineering e3
strain measure as shown in Eq. (7). The vector of the principal strains
e ¼ e1 þ e2 þ e3
ð9Þ
3

Table 3
Contact properties.

Property Value

Contact stiffness k (yield tests) 1.0∙1012 N/m


Contact stiffness k (elasticity tests) 1.0∙1014 N/m
Coefficient of friction μ 0.2
Maximum cohesion strength cmax [1, 10, 100, 300, 600, 1000] MPa
Cohesion interaction length δ0 0.01·dParticle
Damping coefficient d 500 MPa∙s/m
Fig. 3. Contact normal stress as function of the relative overlap (overlap/diameter).
P. Loidolt et al. / Powder Technology 345 (2019) 589–600 593

Fig. 4. Schematic evolution of the stress during three steps of the compaction simulation
and subsequent testing. The real simulation is in 3D.

2.3. Testing of the elastic properties

The elastic properties of the compacted powder are tested with our
model. For that the reference state of the RVE is externally loaded for
each of the six stress components in Eq. (3). The normal stresses are in- Fig. 6. Determination of the Young's modulus E1 based on linear regression of the stress-
vestigated for positive and negative signs to take the difference of ten- strain curve for small strains. Different cohesion strengths of the particle contacts are
sion and compression into account. The stress inside the RVE is considered as shown in the legend. Straight lines are a fit of the strain-stress relationship.
calculated based on Eq. (5), while Eq. (6) is used to determine the cor-
responding strain. We plot exemplarily the stress as a function of the
strain for uniaxial compression and tension in the direction of the first During loading the stress and the strain inside the powder are evaluated
principal axis in Fig. 6. Since all stress components, except of the loading based on Eq. (5) and (6), respectively. Yielding is defined based on a sca-
direction, are zero it is possible to make use of Eq. (3) to obtain the elas- lar plastic strain measure already proposed in [33] to describe the onset
tic constant relevant for this type of loading. In this way, all components of yielding for general loading cases (see Eq. (10)).
of the orthotropic elasticity model given in Eq. (3) can be obtained. rffiffiffiffiffiffiffiffiffiffiffiffiffi rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Table 4 presents the evaluation of the nine independent elastic con- 2 2
εpl ¼ ε : ε− ε :ε ð10Þ
stants based on the six different loading cases. As shown, each Poisson's 3 3 el el
ratio can be computed based on two different loading cases. Both cases
are used in our subsequent analysis and the average value of the Therefore, the yield point is found by evaluating the equivalent plas-
Poisson's ratio is reported. The stress-strain curves of the compacted tic strain εpl of Eq. (10) after probing the RVE in a certain direction. The
powders are generally no straight lines. Especially in case of low contact strain tensor ε is computed with Eq. (6) based on the displacements of
cohesion they are non-linear, as can be seen in Fig. 6. This means the the auxiliary nodes during testing. The elastic strain εel at the yield
elastic constants are strictly speaking no constants, but are a function point is found during unloading. To avoid cyclic loading and unloading
of the stress. In this work we aim for a description of the powder as a lin- for every single yield point, we define the unloading strain after com-
ear elastic material. Hence, the elastic constants are found by lineariza- paction as the elastic strain for all loading directions. The yield point
tion of the stress-strain curve for strains smaller than a defined for a certain load case is reached if the plastic equivalent strain exceeds
threshold strain which is chosen here to be 0.002. a defined threshold. Below this threshold of the plastic equivalent strain,
the deformation is assumed to be elastic. The threshold for the plastic
equivalent strain is chosen to be 0.002 in analogy to the offset yield
2.4. Testing of the yield properties point in uniaxial tensile testing, where a plastic strain of 0.2% is used
to define the yield strength. The corresponding yield stress tensor of
The uniaxial yield strength of an RVE of the compacted powder is this loading direction is evaluated according to Eq. (5) in a post-
tested by applying uniaxial loading starting from the reference state. processing step.

Fig. 5. The seven different strain paths during compaction. The continuous line indicates the state before compaction and the dashed line the state after compaction.
594 P. Loidolt et al. / Powder Technology 345 (2019) 589–600

Table 4
Evaluation of the elastic constants for different loading directions.

Loading direction Elastic modulus Poisson's ratio

±σ11 E1 ¼ σε1111 ν 12 ¼ − εε22 ; ν31 ¼ − εε33 E3


11 11 E1

±σ22 E2 ¼ σε2222 ν 23 ¼ − εε33 ; ν12 ¼ − εε11 E1


22 22 E2

±σ33 E3 ¼ σε3333 ν 31 ¼ − εε11 ; ν 23 ¼ − εε22 E2


33 33 E3
σ 23
σ23 G23 ¼ 2ε23
σ 31
σ31 G31 ¼ 2ε31
σ 12
σ12 G12 ¼ 2ε12

During determination of the yield surface of the compacted powder the strain path during compaction. The shear modulus can therefore be
a total of 82 yield points are tested. We aim for an even distribution of considered to be isotropic in the studied strain paths which are chosen
these points on the yield surface in stress space. To find a proper distri- to be practically important for compaction processes.
bution of the yield points the rough dimensions of the yield surface is The regression of the Poisson's ratios for different loading directions
estimated during 4 trial tests. In the 4 trial tests the yield point for the proved to be difficult. It was not possible to obtain a unified equation
loading direction during compaction, the yield point for isostatic com- which described all strain paths during compaction. Hence, we reduce
pression and their inverted loading directions are tested. The remaining ourselves to describe the averaged Poisson's ratio based on Eq. (13).
78 yield points are evenly distributed based on the 4 obtained yield The averaged Poisson's ratio is approximately 0.3 for a large range of
points. During testing the 78 yield points, the powder is first loaded to the relative density and the contact cohesion. Only in case of very
obtain a stress state in the center of the yield surface. Starting from weak contact cohesion the Poisson's ratio increases up to 0.5. In this
this stress state an additional load is applied into different directions case the powder has only low strength and therefore the deformation
to hit the yield surface. during testing is plastic rather than elastic. According to basic plasticity
theory of metals a Poisson's ratio of 0.5 is expected during plastic yield-
3. Results - elastic properties ing, if plastic incompressibility is assumed [37]. The difference of the
Poisson's ratios for different directions is especially distinct for low rel-
The elastic properties of the powder compacted in different strain ative density and low contact cohesion. There seem to be preferred di-
paths are summarized in Fig. 7. The plots show the Young's modulus rections for transverse strain which are correlated to the principal
E, the shear modulus G and the Poisson's ratio ν as a function of the max- strains during compaction as similar evolutions are observed for similar
imum contact cohesion cmax and the relative density ρrel of the powder strain paths, e.g., similar evolutions of the different Poisson's ratios are
after compaction. The simulation series of contact cohesion was con- found for strain path case B and case C since in both cases the powder
ducted for a constant relative density of 0.85 while the contact cohesion is only compacted in x-direction. However, there is no unified rule
was set to 100 MPa during the series of the relative densities. The points how to describe the Poisson's ratio based on the principal strains for
in the plots describe the average elastic properties of three different all different strain paths during compaction.
packings of 50 particles. The lines indicate the regression curves, also
c 0:559
given in Eqs. (11) to (13), which describe the simulated properties as Ei ¼3:04∙1011 ∙ðρrel −0:6Þ1:34 ∙atan
max

a function of the relative density, the maximum contact cohesion and  246
e0:899 e  ð11Þ
the main principal strains during compaction. ∙ 1−0:04∙std þ 0:571∙ð1:14−ρrel Þ∙atan i −1
The regression equation of the Young's modulus Ei for a certain load- e e
ing direction as shown in Eq. (11) can be divided into two parts. The first c 0:846  e1:35 
part describes the increase of the Young's modulus as a function of the max
Gij ¼ 9:34∙1010 ∙ðρrel −0:6Þ1:26 ∙atan ∙ 1−0:0343∙std
relative density and the contact cohesion. The influence of the contact 118 e
cohesion is described by means of the arc tangent since the observed ð12Þ
saturation behavior can be modeled in this manner. The second part  
of the equation includes the influence of the strain path during compac- 0:00038 0:1
νmean ¼ 0:305∙ð1−ρrel Þ2:19 þ atan ð13Þ
tion and is only important for non-isostatic compaction. The Young's cmax
modulus averaged over the three principal axes is a maximum for iso-
static compaction and decreases as a function of the standard deviation
of the three principal strains in case of non-isostatic compaction. The 4. Results - yield properties
Young's modulus is furthermore higher for directions of high compac-
tion strains and lower for directions of low or even negative strains as 4.1. Uniaxial strength
described by the arc tangent. In case of the important application of
closed-die compaction (case B), the Young's modulus is higher in the di- The evolution of the uniaxial yield strength as a function of the rela-
rection of the punch axis than perpendicular to the punch axis. This may tive density and the cohesion strength is shown in Fig. 8 for different
be crucial during unloading of the powder compact inside the die after strain paths during compaction. The tensile strength is plotted in posi-
compaction, since usually first the upper punch is removed before the tive direction of the ordinate while the compression strength is nega-
powder compact is ejected. This step-wise elastic unloading leads to tive. For each strain path two simulation series are performed. The left
stress states which may cause failures like the well-known capping of diagrams show the uniaxial strengths as a function of the cohesion
tablets after compaction. strength for a constant relative density of 0.85 after compaction while
The shear modulus is described in Eq. (12) as a function of the rela- the right diagrams show the uniaxial strengths as a function of the rel-
tive density, the contact cohesion and the principal strains during com- ative density for a cohesion strength of 100 MPa. The full lines indicate
paction. The shear modulus increases with increasing relative density the regression curves shown in Eq. (14) and (15).
and contact cohesion and decreases with increasing standard deviation Eq. (14) shows the regression of the tensile strength for a direction
of the strain components during compaction. In contrast to the Young's of principal strain during compaction as a function of the relative den-
modulus there is no big influence of the loading direction irrespective of sity, the cohesion strength and the compaction strain in the direction
P. Loidolt et al. / Powder Technology 345 (2019) 589–600 595

direction of the highest compaction strain. Hence, if isotropic me-


chanical properties are important then it is necessary to use isostatic
compaction.
The regression of the compression strength in Eq. (15) is com-
plex. The first part of the equation describes the averaged compres-
sion strength which increases as a function of the relative density
and the contact cohesion and decreases as a function of the stan-
dard deviation of the principal strains during compaction. The sec-
ond part describes the anisotropic behavior. In general, the
compression strength is higher for the directions of high compaction
strains. The difference between different directions increases for low
relative density and decreases for high relative densities. This can be
explained by contact surfaces between particles which are first
formed normal to the main compaction direction. In case of high
relative densities of the compacted powder the contact surfaces
are also formed parallel to the compaction direction and the anisot-
ropy of the compression strength diminishes. The evolution for both
uniaxial strengths as a function of the contact cohesion increase
strongly for low cohesion strengths and saturate for high cohesion
strengths. The denominator of the arc tangent indicates the onset
of saturation and is therefore referred to as the saturation cohesion
strength. The saturation cohesion strength describes the transition
between contact failure of the particle contacts in case of low cohe-
sion strength and plastic flow of the particle material in case of high
cohesion strengths. The saturation cohesion strength is remarkably
higher for tensile load compared to compression load. This behavior
can be attributed to the different failure mechanisms for the particle
material and the contacts between particles. The particle material
follows a von Mises yield criterion which is independent from the
loading direction while the contact failure is different for tension
and compression. During tension loading the contacts are pulled
apart facilitating contact failure while contacts are pressed together
during compression load impeding contact failure. This leads to a
transition from contact failure to particle material flow which
takes place at a lower contact cohesion strength in case of compres-
sion than in case of tension.

c 1:13  e 
max
σ ti ¼ 772∙ðρrel −0:6Þ1:44 ∙atan ∙ 1 þ 0:195∙atan i −1 ð14Þ
332 e
c 0:502  e0:33 
max
σ ci ¼−1403∙ðρrel −0:6Þ1:73 ∙atan ∙ 1−0:0647∙std
223 e
e  c 0:216
1:73 i max
−2157∙ðρrel −0:6Þ ∙ð1:03−ρrel Þ∙atan −1 ∙atan
e 223
ð15Þ

4.2. Yield surface

The simulation results for the yield points in the 3D stress space are
shown in Fig. 9 and Fig. 10. Fig. 9 shows the yield points after compac-
tion to different relative densities and a constant cohesion strength of
100 MPa for different strain paths. In Fig. 10 the relative density is
kept constant at 0.85 while different cohesion strengths are used. The
plots show 2D projections parallel to the Cartesian coordinate axes.
The coordinate axes show pressure stresses as positive, as common for
yield surfaces of powders (e.g., see Fig. 1). For each projection only the
yield points included in the convex hull are shown for clarity. The
Fig. 7. Elastic properties for 7 different compaction cases. The points are the simulation solid lines indicate the regression curves for the yield points, and there-
results and the curves are a regression fit of the simulated points. The results for contact fore, define the yield surface.
cohesion (left panels) were obtained for a constant relative density of ρrel = 0.85.
The mathematical description of the yield surface is performed in cy-
Results for density (right panels) were obtained for a contact cohesion of cmax = 100 MPa.
lindrical coordinates: r, φ and z (see Fig. 11 and Eq. (16)). The origin of
the cylindrical coordinate system coincides with the origin of the Carte-
under consideration. The tensile strength increases as a function of sian coordinate system. The components zi of the unit vector pointing in
the relative density and the contact cohesion. As it was seen for the z-direction of the cylindrical coordinates are given in Eq. (17), while
the Young's modulus also the tensile strength is highest for the the axis for φ = 0 is given in Eq. (18). Eq. (16) determines the radius for
596 P. Loidolt et al. / Powder Technology 345 (2019) 589–600

Fig. 8. Uniaxial strength as a function of relative density and cohesion strength. Left panels show dependence on cohesion strength for constant relative density of ρrel = 0.85. Right panels
show dependence on relative density for cohesion strength of cmax = 100 MPa.

each pair of φ and z as a product of three terms. The first two terms de- always points towards the direction of minimum strain since the yield
scribe a 2D cut of the yield surface in the r-z plane which has the shape surface is flattest in this direction.
of an ellipse (first term) modified by a linear equation (second term). rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 ffi
The 3D yield surface is obtained by rotating the described shape a2 −ðz−zm Þ2  
around the z-axis resulting in an ellipsoid. During rotation the surface r ðφ; zÞ ¼ ∙ðk∙ðz−zmax Þ þ 1Þ∙ l∙ sinðφÞ2 þ 1 ð16Þ
is stretched by a factor (l ∙ sin (φ)2 + 1) with l being a positive number, q
meaning the yield surface is flattest for φ = 0, π and thickest for φ ¼ π2 ;  e  
1
3π pffiffiffi þ 0:397∙atan i −1 ∙ð1−ρrel Þ0:615
2. e
3
zi ¼ rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
The orientation of the rotation axis of the yield surface is influ-
P3  1   2ffi ð17Þ
enced by the compaction conditions (see Eq. (17)). It is aligned pffiffi þ 0:397∙atan ei −1 ∙ð1−ρ Þ0:615
i¼1 3 e rel
with the [111]T-space diagonal of the Cartesian coordinate system
in case of isostatic compaction and rotated towards the direction of 8  e 
>
> þ 1:21−1:02∙atan i −0:753
high compaction strain in case of non-isostatic compaction. The ori- >
> e
> rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
entation of the rotation axis is therefore a measure of the anisotropy
>
>
> X3   2 ; ei ¼ minðeÞ
>
> ei
of the yield surface. As can be seen, the anisotropy diminishes for in- >
< i¼1
1:21−1:02∙atan e
−0:753
creasing relative density. This can, as already stated earlier, be attrib- φ0i ¼  e  ð18Þ
>
>
uted to the formation of contact surfaces which are first formed >
> − 1:21−1:02∙atan i
−0:753
>
> e
>
> rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2 ; ei ≠ minðeÞ
normal to the main compaction direction and formed parallel to >
> X3  
>
: 1:21−1:02∙atan ei −0:753
the compaction direction in case of high relative densities. The orien- i¼1 e
tation of the yield surface is not dependent on the contact cohesion
according to Eq. (17) and Fig. 10. The constants in Eq. (16) are elaborate functions to some degree
The orientation of the coordinate axis for φ = 0 is only influenced by of the relative density after compaction, the contact cohesion and the
the principal strains during compaction (see Eq. (18) and Fig. 10). It standard deviation of the strain ratios during compaction (see Eqs.
P. Loidolt et al. / Powder Technology 345 (2019) 589–600 597

Fig. 9. 2D representation of the yield points in the stress space as a function of the relative Fig. 10. 2D representation of the yield points in the stress space as a function of the contact
density after compaction for a constant cohesion strength of cmax = 100 MPa. The dot- cohesion strength for a constant relative density of ρrel = 0.85. The dot-dashed line
dashed line indicates the rotation axis of the yield surface, while the dashed line points indicates the rotation axis of the yield surface while the dashed line points towards the
towards the flattest side of the yield surface (φ = 0). flattest side of the yield surface (φ = 0).
598 P. Loidolt et al. / Powder Technology 345 (2019) 589–600

Although the initial packing of the particles is random and isotropic, a


clear anisotropy of the mechanical properties is observed after compac-
tion. The simulations show a direct correlation of the anisotropy of the
principal strains during compaction and the anisotropy of the elastic
and plastic properties of the compacted powder. In case of isostatic
compaction, the elastic and plastic properties are observed to be isotro-
pic. If the compaction strain in one direction differs from the others,
then also a different Young's modulus and yield strength is observed
in this direction. A higher compaction strain results in a higher Young's
modulus and yield strength, while the opposite is true for lower com-
paction strain. The properties of the directions with equal strain during
compaction are identical leading to a transverse isotropic elastic mate-
Fig. 11. Cylindrical coordinate system of the yield surface.
rial model. This is the case for the most often used closed-die compac-
tion. It is therefore an oversimplification to use an isotropic material
(19)–(25)). The constants zmax and zmin describe the maximal extension model for closed-die compaction, as it is widely done for tableting
of the yield surfaces along its rotation axis in positive and negative [19,36]. In case of three different compaction strains all three Young's
direction. The respective Eqs. (19) and (20) have the same structure ex- moduli and uniaxial yield strengths are different, leading to an
cept that zmin includes the cohesion strength as it describes the exten- orthotropic elasticity model of the compacted powder. The yield sur-
sion of the yield surface for tension load. The parameters zm and a face, described as a rotated surface, shows the anisotropy of the yield
describe the center and the semi-major axis of the ellipse shown in properties as well. The rotational axis is not aligned with the space diag-
the first term of Eq. (16). The ratio of the major and the minor axes of onal, but is rotated towards the direction of highest compaction strain.
the ellipse is given by the parameter q. The parameter k describes the Furthermore, the yield surface is compressed in the direction of the
slope of the linear equation of the second term in Eq. (16). The last pa- smallest strain during compaction. The anisotropy is not solely a func-
rameter l is used in the third term of Eq. (16) and describes the differ- tion of the strain path, but is also influenced by the relative density
ence of the ellipsoid's extension in r-direction. As l is zero in case of after compaction. The anisotropy is more important at low relative den-
isostatic compaction and increases for anisotropic compaction it is a pri- sities compared to high relative densities, since contact areas are partic-
mary measure for the anisotropy of the yield surface in addition to the ularly formed normal to the direction of the highest strain in case of low
orientation of the rotation axis described in Eq. (17). relative densities. This behavior was already reported in [31,38]. In
It would be fruitful to combine the regression equations of the ob- Fig. 12 (right) the experimental data of compacted spherical copper par-
served yield surface given in Eqs. (16) to (18) with the existing concepts ticles from [38] are compared to the simulated yield points of the pow-
of anisotropic plasticity [20,24] in the literature. However, we refrain der in the current work. The evolution of the yield surface as a function
from this step in the current work, as it was our goal to best fit the nu- of the relative density of the powder is in reasonable agreement. The
merically obtained yield surface with flexible analytical equations yield strength in our simulations is higher at low relative densities com-
which describe the anisotropy of the yield surface as a function of the pared to the reported experiments and lower in case of high relative
condition of the compacted powder (relative density after compaction, densities. This difference may be attributed to a different particle size
the contact cohesion strength and the strain path during compaction). distribution since we used monodispersed particles. Furthermore, the
Our description of the yield surface in cylindrical coordinate systems di- hardening behavior of the material in the experiment and the simula-
rectly follows from the visual inspection of the yield surface for different tion may be different since the material data for the simulation was
compaction conditions of the powder. taken from a different experimental study.
 The contact cohesion has a minor influence on the anisotropy of the
e0:86 
zmax ¼ 3356∙ðρrel −0:6Þ1:79−0:1∙stdðeÞ ∙ 1−0:16∙std compacted powder, but it has a major influence on the mechanical
e
ð19Þ
e properties of the compacted powder in general. In case of very low con-
 e0:86  c 1:54 tact cohesion the Young's modulus, the shear modulus and the uniaxial
zmin ¼ −2097∙ðρrel −0:6Þ1:79−0:1∙stdðeÞ ∙ 1−0:16∙std
e max
∙atan ð20Þ strength are negligibly small and the Poisson's ratio is close to 0.5. This
e 233
indicates a pure plastic behavior of the powder rather than an elastic-
zmax þ zmin plastic behavior. Only in case of a positive pressure noteworthy stresses
zm ¼ ð21Þ can be applied to the cohesionless powder as can be seen in Fig. 10. In
2
case of cohesion strengths which are in the range or even above the
zmax −zmin yield strength of the particle material a clear elastic-plastic behavior of
a¼ ð22Þ
2 the compacted powder can be observed. The powder can furthermore
  e0:558  take up substantial tensile load, leading to yield surfaces which are
q ¼ 2:28 þ 7:93∙ð1−ρrel Þ2:05 ∙ 1 þ 0:146∙std ð23Þ equally extended for positive and negative pressure. The same behavior
e is observed for the Fleck model [8] including a cohesion parameter that
  c  allows modeling the whole range from cohesionless to perfect bonding,
k ¼ 0:000447 þ 4:98∙10−6 ∙ðρrel −0:6Þ−2:02 ∙ 0:64−atan
max
ð24Þ see Fig. 12 (left). Our simulated yield points for no cohesion and very
176:1
high cohesion are compared to the Fleck model and the isostatically
e0:366 cold compacted copper powder from [39]. Obviously, the contact cohe-
l ¼ 0:154∙ð1 þ 1:52∙ð1:52−ρrel ÞÞ∙ð1−0:000256∙cohÞ∙std ð25Þ sion after cold compaction is small since the yield surface without cohe-
e
sion fits the experimental points quite well. In contrast to that the
experimental results for the Young's modulus of cold compacted pow-
5. Conclusion and outlook der from [40] are closer to the simulated Young's modulus with medium
to high cohesion (see Fig. 13). The Young's modulus seems to be
The elastic and plastic properties of compacted powders were inves- underestimated during simulation, especially for low cohesion strength.
tigated in this work during extensive simulation studies based on a This can be attributed to the procedure of how the elastic properties are
MPFEM model including 50 particles in a periodic RVE. The RVE was tested. During the test procedure the RVE is loaded uniaxially to a strain
compacted to different relative densities using different strain paths. of 0.002 which is identical to the equivalent plastic strain during
P. Loidolt et al. / Powder Technology 345 (2019) 589–600 599

Fig. 12. (left): Simulated yield points after isostatic compaction with no cohesion and high cohesion compared with the Fleck model (dashed line). The experimental data of compacted
copper powder from [39] is also shown. (right): Comparison of the yield points after closed die compaction to different relative densities (filled symbols) with the experimental data of [38]
(open symbols).

evaluation of the yield point. This approach works well for the powder RVE used in this work was determined via a convergence study in our
of high cohesion strength, which deforms elastically below a strain of earlier work [33] and was found for a relative density of 0.9 and a con-
0.002 (see Fig. 6). In contrast, the powder with low cohesion strength tact cohesion strength of 100 MPa. Therefore, this RVE size is strictly
immediately starts to deform plastically after loading. This leads to speaking only valid for this parameter setting. In the current study a
very low Young's moduli and to Poisson's ratios in the range of 0.5 wide range of relative densities and cohesion strengths was investi-
(see Fig. 7), although it should be close to 0 for low relative densities gated and it was not feasible to determine the minimum RVE size for
and low cohesion strength [19]. all combinations. Especially in case of low cohesion and low packing
This work clearly reveals the need of an alternative yield surface density a larger RVE might be necessary. Therefore, it is recommended
for powder compaction as the most often used Drucker-Prager/Cap for future work to perform a convergence study based on the model
model (see Fig. 1) is not able to describe the yield behavior of pow- parameters used. In fact, it is necessary to adapt the contact model to en-
ders after compaction using different strain paths. Based on the re- hance the physical interpretation of parameters in the modeled contact
gression equations shown in this work, it is possible to implement between particles. In particular a more advanced cohesion model could
an anisotropic yield surface and elasticity model for a macroscopic be used which must be calibrated based on experiments to model the
FEM model of a powder compaction process in a future work. Since contact behavior of a certain powder. In case of low cohesion strength
the relative density and the strain ratios during compaction are in- an alternative testing procedure has to be employed to determine the
cluded in the regression equation they are useful to determine the elastic properties. One possibility would be to consider the elastic defor-
distribution of the material properties of inhomogeneous compacted mation during unloading. To enable comparability to a real powder it
powder. This allows to model compaction of complex shapes of pow- will additionally be necessary to use a proper particle size distribution
der compacts and to model the deformation close to walls and cor- and non-spherical particles. As there are many brittle particle materials
ners. The variety of the contact cohesion strength can be used to it would be a valuable extension to include particle breakage into the
model different contact properties, e.g., a low contact cohesion can model. Furthermore, the flow rule which describes the direction of the
be used to model cold compaction of powders with weak cohesion flow during yielding of the compacted powder needs to be considered.
interaction and high contact cohesion can be used to model powders In case of solid metallic materials, it is often valid to assume plastic
after sintering. flow normal to the yield surface. This assumption is not necessarily ap-
Despite the considerable possibilities of the demonstrated model ad- plicable for compacted powders. Another different measure which can
ditional improvements are suggested for future work. The size of the also be of interest is the fracture energy of the compacted powder. The
fracture energy is an important quantity in fracture-mechanics analyses
which may be employed during brittle fracture of the compacted
powder.
Finally, the material models derived from our micromechanical
model have to be implemented in a suitable subroutine of a FEM code
(e.g., UMAT/VUMAT in Abaqus) to compute the stress distribution of
whole powder parts during compaction and processing. In this way pos-
sible failures of the compacted powder can be predicted and process
optimization can be performed.

Declaration of interest

None.

Funding

Fig. 13. The relative Young's modulus after isostatic compaction for different cohesion This research did not receive any specific grant from funding agen-
strength compared to the experimental data (circles) of copper powder from [40]. cies in the public, commercial, or not-for-profit sectors.
600 P. Loidolt et al. / Powder Technology 345 (2019) 589–600

Acknowledgements powders, Int. J. Solids Struct. 45 (2008) 3088–3106, https://doi.org/10.1016/j.


ijsolstr.2008.01.024.
[20] R. Hill, The Mathematical Theory of Plasticity, Oxford University Press, 1950.
The authors would like to thank Simon Rustige for the evaluation of [21] R.M. Caddell, R.S. Raghava, A.G. Atkins, A yield criterion for anisotropic and pressure
the yield surfaces and Daniela Fiedler for proof-reading of the manu- dependent solids such as oriented polymers, J. Mater. Sci. (1973), https://doi.org/10.
1007/BF00754900.
script. We also want to thank the reviewers for their careful analysis [22] V.S. Deshpande, N.A. Fleck, M.F. Ashby, Effective properties of the octet-truss lattice
and for their many suggestions. material, J. Mech. Phys. Solids. 49 (2001) 1747–1769, https://doi.org/10.1016/
S0022-5096(01)00010-2.
[23] Z. Xue, J.W. Hutchinson, Constitutive model for quasi-static deformation of metallic
References sandwich cores, Int. J. Numer. Methods Eng. 61 (2004) 2205–2238, https://doi.org/
10.1002/nme.1142.
[1] S. Inghelbrecht, J.P. Remon, Roller compaction and tableting of microcrystalline cel-
[24] F. Barlat, D.J. Lege, J.C. Brem, A six-component yield function for anisotropic mate-
lulose/drug mixtures, Int. J. Pharm. 161 (1998) 215–224, https://doi.org/10.1016/
rials, Int. J. Plast. (1991), https://doi.org/10.1016/0749-6419(91)90052-Z.
S0378-5173(97)00356-6.
[25] F. Bron, J. Besson, A yield function for anisotropic materials Application to aluminum
[2] S.C. Lee, K.T. Kim, Densification behavior of aluminum alloy powder under cold
alloys, Int. J. Plast. 20 (2004) 937–963, https://doi.org/10.1016/j.ijplas.2003.06.001.
compaction, Int. J. Mech. Sci. 44 (2002) 1295–1308.
[26] F. Barlat, J.W. Yoon, O. Cazacu, On linear transformations of stress tensors for the de-
[3] M. Trunec, K. Maca, Compaction and pressureless sintering of zirconia nanoparticles,
scription of plastic anisotropy, Int. J. Plast. 23 (2007) 876–896, https://doi.org/10.
J. Am. Ceram. Soc. 90 (2007) 2735–2740, https://doi.org/10.1111/j.1551-2916.2007.
1016/j.ijplas.2006.10.001.
01781.x.
[27] F. Yoshida, H. Hamasaki, T. Uemori, A user-friendly 3D yield function to describe an-
[4] W.R. Mitchell, L. Forny, T. Althaus, D. Dopfer, G. Niederreiter, S. Palzer, Compaction
isotropy of steel sheets, Int. J. Plast. 45 (2013) 119–139, https://doi.org/10.1016/j.
of food powders: the influence of material properties and process parameters on
ijplas.2013.01.010.
product structure, strength, and dissolution, Chem. Eng. Sci. 167 (2017) 29–41,
[28] J.W. Yoon, Y. Lou, J. Yoon, M.V. Glazoff, Asymmetric yield function based on the
https://doi.org/10.1016/j.ces.2017.03.056.
stress invariants for pressure sensitive metals, Int. J. Plast. 56 (2014) 184–202,
[5] R.W. Heckel, Density-pressure relationships in powder compaction, Trans. Metall.
https://doi.org/10.1016/j.ijplas.2013.11.008.
Soc. AIME 221 (1961) 671–675.
[29] H. Altenbach, A. Öchsner (Eds.), Plasticity of Pressure-Sensitive Materials, Springer
[6] K. Kawakita, K.-H. Lüdde, Some considerations on powder compression equations,
Verlag Heidelberg, 2014.
Powder Technol. 4 (1971) 61–68, https://doi.org/10.1016/0032-5910(71)80001-3.
[30] S. Galen, A. Zavaliangos, Strength anisotropy in cold compacted ductile and brittle
[7] E. Ryshkewitch, Compression strength of porous sintered alumina and zirconia, J.
powders, Acta Mater. 53 (2005) 4801–4815, https://doi.org/10.1016/j.actamat.
Am. Ceram. Soc. 36 (1953) 65–68, https://doi.org/10.1111/j.1151-2916.1953.
2005.06.023.
tb12837.x.
[31] B. Harthong, D. Imbault, P. Dorémus, The study of relations between loading history
[8] N.A. Fleck, On the cold compaction of powders, J. Mech. Phys. Solids. 43 (1995)
and yield surfaces in powder materials using discrete finite element simulations, J.
1409–1431, https://doi.org/10.1016/0022-5096(95)00039-L.
Mech. Phys. Solids. 60 (2012) 784–801, https://doi.org/10.1016/j.jmps.2011.11.009.
[9] N.A. Fleck, L.T. Kuhn, R.M. McMeeking, Yielding of metal powder bonded by isolated
[32] D.T. Gethin, R.W. Lewis, R.S. Ransing, A discrete deformable element approach for
contacts, J. Mech. Phys. Solids. 40 (1992) 1139–1162, https://doi.org/10.1016/0022-
the compaction of powder systems, Model. Simul. Mater. Sci. Eng. 11 (2002)
5096(92)90064-9.
101–114, https://doi.org/10.1088/0965-0393/11/1/308.
[10] P.R. Heyliger, R.M. McMeeking, Cold plastic compaction of powders by a network
[33] P. Loidolt, M.H. Ulz, J. Khinast, Modeling yield properties of compacted powder using
model, J. Mech. Phys. Solids. 49 (2001) 2031–2054, https://doi.org/10.1016/
a multi-particle finite element model with cohesive contacts, Powder Technol. 336
S0022-5096(01)00038-2.
(2018) 426–440, https://doi.org/10.1016/j.powtec.2018.06.018.
[11] E. Arzt, The influence of an increasing particle coordination on the densification of
[34] J.E. Flinn, D.P. Field, G.E. Korth, T.M. Lillo, J. Macheret, The flow stress behavior of
spherical polders, Acta Metall. 30 (1982) 1883–1890, https://doi.org/10.1016/
OFHC polycrystalline copper, Acta Mater. 49 (2001) 2065–2074, https://doi.org/
0001-6160(82)90028-1.
10.1016/S1359-6454(01)00102-1.
[12] P. Pizette, C.L. Martin, G. Delette, P. Sornay, F. Sans, Compaction of aggregated ce-
[35] M.H. Shih, C.Y. Yu, P.W. Kao, C.P. Chang, Microstructure and flow stress of copper de-
ramic powders: from contact laws to fracture and yield surfaces, Powder Technol.
formed to large plastic strains, Scr. Mater. 45 (2001) 793–799, https://doi.org/10.
198 (2010) 240–250, https://doi.org/10.1016/j.powtec.2009.11.013.
1016/S1359-6462(01)01098-3.
[13] C.L. Martin, Elasticity, fracture and yielding of cold compacted metal powders, J. Mech.
[36] I.C. Sinka, J.C. Cunningham, A. Zavaliangos, The effect of wall friction in the compac-
Phys. Solids. 52 (2004) 1691–1717, https://doi.org/10.1016/j.jmps.2004.03.004.
tion of pharmaceutical tablets with curved faces: a validation study of the Drucker-
[14] C.L. Martin, D. Bouvard, Isostatic compaction of bimodal powder mixtures and com-
Prager Cap model, Powder Technol. 133 (2003) 33–43, https://doi.org/10.1016/
posites, Int. J. Mech. Sci. 46 (2004) 907–927, https://doi.org/10.1016/j.ijmecsci.2004.
S0032-5910(03)00094-9.
05.012.
[37] C. Wu, Continuum Mechanics and Plasticity, Chapman & Hall/CRC, Boca Raton, USA,
[15] C.A. Radeke, B.J. Glasser, J.G. Khinast, Large-scale powder mixer simulations using
2005.
massively parallel GPUarchitectures, Chem. Eng. Sci. 65 (2010) 6435–6442,
[38] L.C.R. Schneider, a.C.F. Cocks, Experimental investigation of yield behaviour of metal
https://doi.org/10.1016/j.ces.2010.09.035.
powder compacts, Powder Metall. 45 (2002) 237–245, https://doi.org/10.1179/
[16] P. Loidolt, S. Madlmeir, J.G. Khinast, Mechanistic modeling of a capsule filling process,
003258902225006998.
Int. J. Pharm. 532 (2017) 47–54, https://doi.org/10.1016/j.ijpharm.2017.08.125.
[39] A.R. Akisanya, T. Street, The Yield Behaviour of Metal Powders, vol. 39, 1997
[17] P. Kelly, Anisotropic Elasticity, Solid Mech. Part I An Introd. to Solid Mech, 2014
1315–1324.
1–83, https://doi.org/10.1016/B978-0-12-374446-3.50015-7.
[40] P.C. Carnavas, N.W. Page, Elastic properties of compacted metal powders, J. Mater.
[18] J.C. Simo, T.J.R. Hughes, Computational Inelasticity, Springer Verlag, New York, 1998.
Sci. 33 (1998) 4647–4655, https://doi.org/10.1023/A:1004445527430.
[19] L.H. Han, J.A. Elliott, A.C. Bentham, A. Mills, G.E. Amidon, B.C. Hancock, A modified
drucker-prager cap model for die compaction simulation of pharmaceutical

Das könnte Ihnen auch gefallen