Sei sulla pagina 1di 16

G Model

PARTIC-636; No. of Pages 16

ARTICLE IN PRESS
Particuology xxx (2014) xxxxxx

Contents lists available at ScienceDirect

Particuology
journal homepage: www.elsevier.com/locate/partic

Numerical investigation into the inuence of the punch shape on the


mechanical behavior of pharmaceutical powders during compaction
Alexander Krok , Marin Peciar, Roman Fekete
Slovak University of Technology in Bratislava, Faculty of Mechanical Engineering, Bratislava, Slovak Republic

a r t i c l e

i n f o

Article history:
Received 9 September 2013
Received in revised form
21 November 2013
Accepted 3 December 2013
Keywords:
Powder
Die compaction
FEM

a b s t r a c t
During the production of pharmaceutical tablets using powder compaction, certain common problems
can occur, such as sticking, tearing, cutting, and lamination. In the past, the compressibility of the powder
was calculated only along the axis of the device; consequently, critical areas of the material throughout
the volume could not be identied. Therefore, nite element method (FEM) can be used to predict these
defects in conjunction with the use of an appropriate constitutive model. This article summarizes the
current research in the eld of powder compaction, describes the DruckerPrager Cap model calibration
procedure and its implementation in FEM, and also examines the mechanical behavior of powder during
compaction. In addition, the mechanical behavior of pharmaceutical powders in relation to changes in
friction at the wall of the system is examined, and the dependence of lubrication effect on the geometry
of the compaction space is also investigated. The inuence of friction on the compaction process for the
at-face, at-face radius edge, and standard convex tablets is examined while highlighting how the effects
of friction change depending on the shape of these tablets.
2014 Published by Elsevier B.V. on behalf of Chinese Society of Particuology and Institute of Process
Engineering, Chinese Academy of Sciences.

1. Introduction
Tablets are a notably popular form of medicine in the pharmaceutical industry. Tablets have numerous advantages over other
dosage forms, such as their low cost, long storage life, mechanical
and chemical stability, good heat and moisture resistance, accurate dosage, and ease of use for the patient. Tablets are generally
manufactured by compressing a dry powder mixtures; compression may be one of the most signicant operational processes
in the pharmaceutical industry. Through research, many experimental methods for analyzing the compression process have been
developed. Empirical and regression models (Guyot, Delacourte, &
Marie, 1986; Heckel, 1961; Jetzer, Leuenberger, & Sucker, 1983;
Leuenberger & Rohera, 1986; Rue & Rees, 1978; York, 1979) have
been used for a long time. Although these models provide numerical data that are obtained using the comparison method, they are
not usually appropriate for quantitative forecasting. In most cases,
these methods offer a calculation of compressibility only along the
axis of the equipment; therefore, identifying the critical areas of the
material throughout the entire volume is not possible. In addition,
these experimental methods are often unable to identify the physical nature of the entire compaction process. Later, the process was

Corresponding author. Tel.: +421 949431543; fax: +421 252962454.


E-mail address: alexander.krok@stuba.sk (A. Krok).

subjected to a more detailed analysis, and the given process was


simulated with the constructed models. These phenomenological
constituent models are based on experimental measurements and
the mechanics of particulate materials. Currently, these models can
be applied to FEM (nite element method) (Cunningham, Sinka, &
Zavaliangos, 2004; Diarra et al., 2012; Kadiri, Michrafy, & Dodds,
2005; Michrafy, Ringenbacher, & Tchoreloff, 2002; Shang, Sinka,
& Pan, 2012; Sinka, Cunningham, & Zavaliangos, 2003; Wu et al.,
2005; Zhang et al., 2010). These works describe the calibration procedure for the constituent model, its application in FEM, and the
investigation into the mechanical behavior of the powder during
compaction. While manufacturing these products, some common
faults, such as sticking, stretching, chipping, restricting or laminating of the product, may occur. Accordingly, FEM in conjunction
with a suitable constituent model is likely to be a very effective
instrument for predicting problems and defects during this process. Numerous publications (Han et al., 2008; Kadiri et al., 2005;
Sinha, Bharadwaj, Curtis, Hancock, & Wassgren, 2010; Sinka, Burch,
Tweed, & Cunningham, 2004; Wu et al., 2008) compared the experimental and calculated data from the FEM simulations, both for
pharmaceutical and cosmetic powders, revealing very good conformity. Microcrystalline cellulose (MCC) Avicel is one of the most
frequently tested powder materials.
Intensive research into powder compaction has revealed that
the elastic properties of a powder signicantly affect the appearance of cracks (Han et al., 2008; Kadiri & Michrafy, 2013;

1674-2001/$ see front matter 2014 Published by Elsevier B.V. on behalf of Chinese Society of Particuology and Institute of Process Engineering, Chinese Academy of Sciences.

http://dx.doi.org/10.1016/j.partic.2013.12.003

Please cite this article in press as: Krok, A., et al. Numerical investigation into the inuence of the punch shape on the mechanical
behavior of pharmaceutical powders during compaction. Particuology (2014), http://dx.doi.org/10.1016/j.partic.2013.12.003

G Model
PARTIC-636; No. of Pages 16

ARTICLE IN PRESS
A. Krok et al. / Particuology xxx (2014) xxxxxx

Wu et al., 2005). In the past, interesting work regarding the


distribution of density in die pressing was carried out by
Macleod and Marshall (1977), and Aydin, Briscoe, and Sanhtrk
(1996) through auto-radiography or Nebgen, Gross, Lehmann,
and Mller (1995) through nuclear magnetic resonance (NMR)
microscopy. Later, this density distribution was investigated using
the indentation-hardness mapping test (Sinka et al., 2003) or with
micro-identication (Han, Kim, Oh, & Lee, 1994; Kim, Choi, & Park,
2000). For example, in the articles (Djemai & Sinka, 2006; Farber,
Tardos, & Michaels, 2003; Sinka et al., 2004; Wu et al., 2005),
testing by X-ray tomography is described. The experimental measurements of pharmaceutical powders show non-linear behavior
during the unloading phase. The experiments show that non-linear
segments at the end of the compression curves can occur; and
these segments become even more apparent in a relatively uncompressed material. Some researchers (Aydin et al., 1996; Han et al.,
2008; Sandler, Baladi, & DiMaggio, 1976; Wu et al., 2005) consider the non-linear segments at the end of the curves to be the
result of dilation during compaction and therefore dependent on
wall friction. The authors of the following references (Cunningham
et al., 2004; Han et al., 2008; James, 1987; Kadiri & Michrafy, 2013;
Michrafy, Dodds, & Kadiri, 2004; Sinka et al., 2003; Wu et al., 2008)
indicated that the effects of wall friction can be partially reduced by
using a lubrication system. One method for retaining the quality of
the tablet is to change its shape. The distribution of the shear stress
depends on the speed of the punch. The shape of the punch has been
investigated by FEM (Eiliazadeh, Pitt, & Briscoe, 2004; Han et al.,
2008; Kadiri & Michrafy, 2013; Newton, Haririan, & Podczeck, 2000;
Wu et al., 2008). The previous researchers concluded that by changing the wall friction during the production of at-faced tablets, the
change in stress for the powder during compression is not significant. By changing the shape of the punch, the effects of the wall
friction may intensify, commonly causing tablet cracking during
compression. The authors also present photographic evidence of
the cracked tablets using X-ray tomography.
The presented article offers results contributing to an investigation into the mechanical behavior of pharmaceutical powders
relative to changes in the wall friction of the system and analyses
how the effect of lubrication can be observed as a function of the
geometry of the compression space. The inuence of friction on
the process of compaction is identied for at-face (FF), at-face
radius edge (FFRE) and standard convex (SC) tablets; the changes
in the effects of friction relative to the shape of these tablets are
also reported. The results contribute to the research in this eld
because to date, no previous simulations have been undertaken to
the extent that they have spanned in this study (eight different wall
friction coefcients for eight different geometry congurations).
The aim of this contribution was to address the real effect of lubrication use and focus on extreme states that have no signicance
from the practical point of view, but may provide a broader picture
of the behavior of the materials beyond the boundaries of practical use. This study was performed to verify whether reducing the
friction is a sensible strategy and whether any anomalies in the
distribution of the stress and in the density of the tablet appear.

2. DPC phenomenological model


2.1. Background
Various phenomenological models that describe geo-materials
are available (Chtourou, Gakwaya, & Guillot, 2002; Coube & Riedel,
2000; DiMaggio & Sandler, 1971; Drucker & Prager, 1952; Green,
1972; Gurson & McCabe, 1992; Schoeld & Wroth, 1968; Wood,
1991). In most cases, however, these models describe the behavior

of granular materials or soils, but only the static behavior of a


material without addressing the various states of the powder, such
as hardening, softening or consolidation. The phenomenological
Drucker Prager Cap model (DPC model) was originally developed
to predict the plastic deformation in soils during compaction.
Since the rst presentation by Drucker, Gibson, and Henkel (1957),
this plastic model has changed, becoming modied and expanded
(Cunningham et al., 2004; Diarra, Mazel, Busignies, & Tchoreloff,
2013; Frenning, 2007; Han et al., 2008; Michrafy et al., 2002; Sinka
et al., 2003, 2004; Watson & Wert, 1993). In conjunction with this
approach, and due to the increasing efciency of computer technology, DEM (discrete element method) methods are frequently
employed (Cleary, 2000; Coube, Cocks, & Wu, 2005; Cundall &
Strack, 1979; Korachkin, Gethin, Lewis, & Tweed, 2008; Sharma,
Saxena, & Woods, 1999; Wu, Cocks, Gilia, & Thompson, 2003; Zhu,
Zhou, Yang, & Yu, 2007).
Fig. 1 assumes that the behavior of the material is isotropic,
and the area of the DPC model is composed of three segments
(Chtourou, Gakwaya, et al., 2002; Hofstetter, Simo, & Taylor, 1993):
(i) the shear failure surface (Fs ), depicting the main shear ow of the
material during compaction; (ii) the cap surface Fc , depicting the
mechanism of non-elastic hardening during plastic compaction;
(iii) and the transition surface (Ft ) that joins the Fs and Fc . This segment was created to facilitate the numerical application of FEM. The
authors (Coube & Riedel, 2000; Cunningham et al., 2004; Jonsn &
Hggblad, 2005) forecast that the DPC areas depend on the density, and when describing the unloading process, utilize the law of
non-linear elasticity.
The shear failure surface (Fs ) in the DPC model provides the
criterion for achieving shear ow and depends on the cohesion and
the internal friction angle (dened according to the MohrCoulomb
hypothesis), as expressed by Eq. (1):
Fs (p, q) = q p tan d = 0,

(1)

where is the internal friction angle,


 d is the cohesion, p = (1/3)

(3/2)(S : S) is the Mises equivis the hydrostatic stress, and q =


alent stress. S is the tensor of deviator stress: S = + pI, is the
stress tensor, and I is the identity matrix. For the uniaxial cylindrical
die compaction test, p and q, are expressed as follows:
p=

1
(z + 2r ),
3

(2)

q = |z r |,

(3)

where z and r are the axial and radial stresses.


The cap surface (Fc ) represents an ellipse with a constant eccentricity in the pq plane, as expressed in Eq. (4):

Fc (p, q) =

(p pa )2 +

Rq
1 + / cos

2

R(d + pa tan ) = 0,
(4)

where R is the cap eccentricity (cap shape parameter), representing a material constant and determining the shape of the cap
surface. secures the constant ow between Fc and Fs and is designated as the coefcient for the transition curve. Furthermore, pa is
the evolution parameter that determines the compaction, depending on the volume of plastic deformation. The law of hardening
p
pb = f (v ) determines the relationship of the hydrostatic compression yield stress (pb ) and the corresponding volumetric plastic
p
strain v = ln(RD/RD0 ), where RD is the relative density and RD0
is the initial relative density of the non-compacted powder. During
our calculations and evaluations, we will work exclusively with the
relative density; this parameter reects the changing composition
of the powder during compaction more accurately. RD is the ratio

Please cite this article in press as: Krok, A., et al. Numerical investigation into the inuence of the punch shape on the mechanical
behavior of pharmaceutical powders during compaction. Particuology (2014), http://dx.doi.org/10.1016/j.partic.2013.12.003

G Model

ARTICLE IN PRESS

PARTIC-636; No. of Pages 16

A. Krok et al. / Particuology xxx (2014) xxxxxx

Fig. 1. An illustration of DruckerPrager Cap model.

of the bulk density (b ) of the material to the true density of the
particles of which the material is composed (T ).
The transition surface (Ft ) is given by Eq. (5) and the main shape
of the DPC area does not depend on the Ft area.

Ft (p, q) =

(p pa )2 + q 1
(d + pa tan ) = 0.

cos

2

(d + pa tan )

(5)

2.2. Calibration of DPC model


The most effective method for calibrating the DPC model is
through an experimental investigation of the compaction of particulate material using a three-axis test. This experiment is expensive
and requires a long-term and complicated procedure for evaluating the measured data (Koerner, 1971; Pavier & Doremus, 1999;
Rottmann, Coube, & Riedel, 2001; Zhang et al., 2010), precluding it
from industrial engineering applications. Other calibration methods have been developed to identify the material parameters of the
DPC model, such as an experimental investigation in a die compaction test equipped with a sensor for measuring the axial and
radial stress (Chtourou, Guillot, & Gakwaya, 2002; Han et al., 2008;
Shang et al., 2012). A diametrical compression test and a uniaxial
compression test are undertaken to measure the strength of the
product using this equipment. The situation in which it is possible for the powder to change its stress state throughout the entire
process is illustrated in Fig. 2. A short description of the situation
is included in Fig. 2.
During the rst phase, it is assumed that the material is compressed, undergoing plastic deformation to behave according to
the non-elastic mechanism of hardening. Depending on the loading
size on the material by the punch, the powder decreases in porosity while increasing its density. This phase is characterized by the
curve AB. During the second phase, the unloading occurs. During
this phase, the material exhibits its elastic properties and attempts
to regain its original state. In this case, the material is acting in
accordance with the elastic mechanism of unloading. This phase
is characterized by the curve BE. The curve BC expresses the rst

stage of unloading. This process continues up to the moment when


the stresses in the radian and axial directions are equal, specifically when the material attains its hydrostatic equivalent stress
(p > 0) and its von Mises equivalent stress is zero (q = 0). To revert
to its original state, the material expands, using some of the energy
gained during compression, and its stress decreases. The process
does not end here. The unloading continues, as expressed by the
curve CD, such that the material continues to reduce its stress up to
the moment when it achieves shear failure (curve Fs ). After reaching this stage, cracks usually appear in the structure of the material.
The nal part of the unloading process is the so-called dilation of
the material, meaning that the material reduces its stress to a small
extent but does not exceed the shear failure curve. This state is
expressed by the curve DE.

2.2.1. Cohesion and internal friction angle


To dene the shear failure surface, the cohesion (d) and the
internal friction angle () must be known; these values are obtained
experimentally (Brewin, Coube, Doremus, & Tweed, 2008; Fell &
Newton, 1970; Frocht, 1947; Timoshenko & Goodier, 1970). As
shown in Fig. 1, the uniaxial tension (1), shear (2), diametrical compression (3) and uniaxial compression (4) (Doremus, Toussaint, &
Alvain, 2001; Procopio, Zavaliangos, & Cunningham, 2003) are used
to construct the shear failure curve (Fs ). After plotting the results
of these measurements in the pq plane, a straight line emerges;
the slope of this line gives the internal friction angle (). The displacement of this line represents the cohesion (d). This procedure
is laborious, time-consuming and expensive. Accordingly, the procedure is partially modied and the number of measurements is
reduced. The diametrical compression test (3) is often carried out
on a Brazilian compactor, and the uniaxial compression (4) is tested
such that the shear failure surface retains its linear shape along the
total range.
Diametrical compression is a simple test in which two opposing forces (FD ) compress a tablet composed of the material under
investigation to reach various volume densities. In an idealized
example, the behavior is isotropic and linearly elastic, therefore
creating a diagonal load across the tablet. For practical reasons, the
diameter should be ve times bigger than the thickness. When the
thickness is larger, the stress state will only partially widen along

Please cite this article in press as: Krok, A., et al. Numerical investigation into the inuence of the punch shape on the mechanical
behavior of pharmaceutical powders during compaction. Particuology (2014), http://dx.doi.org/10.1016/j.partic.2013.12.003

G Model
PARTIC-636; No. of Pages 16

ARTICLE IN PRESS
A. Krok et al. / Particuology xxx (2014) xxxxxx

Fig. 2. Evolution of powder stress state during compaction: (a) axis stress ( z ) vs. axial strain (z ), (b) axial stress ( z ) vs. radial stress ( r ), and (c) stress state change on the
pq plane.

the plane, affecting the results by 30%. During the experiment, the
crack must expand to the center of the tablet. Subsequently, Eq. (6)
is used:
d =

2FD
,
Dh

(6)

where D is the diameter of the tablet and FD represents the break


force of diametrical compression and  d is diametrical tensile
strength. On the pq plane (pd , qd stress state for diametrical
compression), the boundary state is expressed as follows:
pd =

2
 ,
3 d

qd =

13d .

(7)

Uniaxial compression is a common test used to determine the


parameters of the material, including those in other industries.
A cylindrical object is pressed from both sides by FC . The relation h/D is very important; h represents the height of the tablet.
This test illustrates the strong inuence of slip forces. These forces
are created by pressure, resulting in a non-homogeneous stress
distribution. After a certain angle, a break-line is formed that continues throughout the entire body and appears in the nal product,
depending on the h/D ratio. From the measured forces, the axial
compression strength ( c ) is calculated:
c =

4FC
,
D2

(8)

where FC represents the uniaxial tensile strength at the yield point.


In the pq (pc , qc stress state for uniaxial compression) plane, the
boundary state will be expressed as follows:
1
pc = c ,
3

qc = c .

(9)

In the previous tests, the tested tablet was brought to the


breaking point. The force that broke this tablet can be calculated
according to the above-stated formulas for stress,  d and  c (i.e.
Eqs. (6) and (8)); these values are then plotted onto the p-q plane,
and thus pc , pd , qc , and qd (Eqs. (7) and (9)) are found. From this data,
the numerous points [pc , qc ], [pd , qd ] at which the tablet failed the
strength-tests can be calculated. A straight line is constructed, representing one given density of the tablet. If a tablet with different
geometric dimensions is used, the measured value of the forces
changes, as does the slope and intercept of the straight line. The
intercept of these lines represents the cohesion (d) and the slope
of this straight line represents the internal friction angle (). To
facilitate these calculations, these parameters can be analytically
deduced (Cunningham et al., 2004; Han et al., 2008; Shang, 2012)
and the resultant shape of these parameters is described by Eqs.
(10) and (11).

c d ( 13 2)
d=
,
(10)
c + 2d
= tan1

 3( + d) 
c
c

(11)

Please cite this article in press as: Krok, A., et al. Numerical investigation into the inuence of the punch shape on the mechanical
behavior of pharmaceutical powders during compaction. Particuology (2014), http://dx.doi.org/10.1016/j.partic.2013.12.003

G Model

ARTICLE IN PRESS

PARTIC-636; No. of Pages 16

A. Krok et al. / Particuology xxx (2014) xxxxxx

Table 1
Measurement data obtained from the diametrical and uniaxial tests.
mT (g)

h (mm)

RD

FD (N)

mT (g)

h (mm)

RD

1.391
1.366
1.398
1.366
1.375
1.398
1.418
1.4
1.396
1.387
1.412
1.415
1.419
1.396
1.392
1.436
1.432
1.414
1.4
1.444
1.443
1.459
1.418
1.453
1.408

2.986
2.934
2.982
2.922
2.976
2.739
2.778
2.743
2.729
2.708
2.598
2.6
2.61
2.584
2.557
2.423
2.409
2.385
2.353
2.423
2.294
2.343
2.249
2.331
2.238

0.675
0.675
0.679
0.677
0.67
0.737
0.738
0.738
0.739
0.74
0.784
0.785
0.785
0.779
0.785
0.853
0.855
0.853
0.856
0.857
0.903
0.894
0.894
0.895
0.903

68
61
69
65
63
144
144
148
136
136
244
246
229
236
223
431
421
431
427
442
651
687
656
652
658

18.124
18.103
18.308
18.33
18.206
19.698
19.287
18.571
17.954
18.404
18.081
18.25
17.828
18.078
17.945
18.029
18.358
18.546
18.656
18.8
18.982
19.236
19.162
19.29
19.662

30.54
30.46
30.84
30.86
30.78
30.78
30.04
28.8
27.92
28.6
26.4
26.78
26.06
26.5
26.24
24.42
24.98
25.16
25.24
25.4
24.44
25.1
25
25.24
25.68

0.652
0.653
0.652
0.653
0.65
0.703
0.704
0.707
0.705
0.706
0.75
0.747
0.749
0.747
0.749
0.807
0.803
0.805
0.808
0.809
0.848
0.837
0.837
0.835
0.836

FC (N)
4659
4787
4999
4827
4840
8192
8106
8157
8208
8336
11,869
11,640
11,860
11,608
11,823
17,756
17,061
17,314
17,393
17,706
23,943
20,901
20,854
20,470
20,812

Fig. 4. Dependence of cohesion (d) on the relative density (RD) of the tablet.
Fig. 3. Dependence of the internal friction angle () on the relative density (RD) of
the tablet.

A summary of the measured data from the diametrical and


uniaxial tests is depicted in Table 1. From the known mass (mT ),
height (h) and given the diameter D = 1.18 mm of the tablet, the
relative density of the tablet (RD) can be calculated. From the measured break force of diametrical compression (FD ) and break force
of uniaxial compression (FC ), the appropriate parameters can be
calculated.
Table 1 shows that the diametrical test has been carried out for
samples with a relative density in the range of RD 0.675; 0.903,
while the uniaxial test uses samples with a relative density in
the range of RD 0.652; 0.836. By using regression analysis, the
parameters  d and  c can be calculated for any selected value of
RD with good conformity (R2 > 98%); this step must be carried out
to synchronize both tests. By using Eqs. (10) and (11), the values
for d and are thus calculated. The dependence of and d on the
relative density of the tablet is illustrated in Figs. 3 and 4.
2.2.2. Cap shape and the hardening curve
To dene the cap surface, the eccentricity (R), the evolution
parameter (pa ), the hydrostatic yield stress (pb ), and the transition coefcient () must be known. is chosen as a small number

(0.01). The evolution parameter and the eccentricity are calculated


by measuring the die compression system.
In Fig. 5 the curves that determine the relationship between
 z and the volume plastic strain (v ) are measured. In Fig. 6 the

Fig. 5. Changes of axial stress ( z ) with the axial strain (v ) obtained experimentally
under various values of compaction stress (100300 MPa).

Please cite this article in press as: Krok, A., et al. Numerical investigation into the inuence of the punch shape on the mechanical
behavior of pharmaceutical powders during compaction. Particuology (2014), http://dx.doi.org/10.1016/j.partic.2013.12.003

G Model

ARTICLE IN PRESS

PARTIC-636; No. of Pages 16

A. Krok et al. / Particuology xxx (2014) xxxxxx

Fig. 7. The eccentricity (R) vs. the relative density of the tablet corresponding to the
ve values of compaction stress.

Fig. 6. Changes of axial stress ( z ) with radial stress ( r ) obtained experimentally


under various values of compaction stress (100300 MPa).

curves that determine the relationship between the axial and radial
stresses during the compaction process are measured. These gures show that ve curves were measured with varied compaction
forces and loading stresses from the punch (100300 MPa). To calculate pa and R, point B is assumed to represent the maximum axial
and radial stress during the process (Fig. 2). By using Eqs. (2) and
(3), these measured points are plotted onto the pq plane, providing pB and qB . During 2D simulations, the cap line is assumed to run
through point B, as does the derived cap line. At this point (B), the
following is valid:

Fc
p
ij = 
ij

= f d, , , pa , R = 0,

(12)

pB ,qB

Fc (pB , qB ) = f (d, , , pa , R) = 0.

(13)

The required parameters can be analytically derived


(Cunningham et al., 2004; Han et al., 2008; Shang, 2012) from Eqs.
(4), (12), and (13):

Fig. 8. The hydrostatic equivalent yield stress (pb ) as a function of the volume plastic
strain (v ) the hardening curve.

( z ) is proportional to the radial stress on the wall, i.e. z =


r ,
where
is the wall friction coefcient between the material and
the die wall. As mentioned in the introduction, prior experiments
showed that non-linear segments exist at the end of the unloading
curves (Aydin et al., 1996; Cunningham et al., 2004; Han et al., 2008;
Jonsn & Hggblad, 2005; Sandler et al., 1976; Wu et al., 2005) due
to dilation during the die compaction. Using this assumption has
led to an unrealistic simulation of decompression, during which
the plastic compression almost ceased (Wu et al., 2005). To avoid
this problem, the law of non-linear elasticity is used to describe the

pa =


R=

[3qB + 4d tan (1 + / cos ) ]


2

4[(1 + / cos ) ]

9q2B + 24dqB (1 + / cos ) tan + 8(3pB qB + 2q2B )[(1 + / cos ) tan ]


2

4[(1 + / cos ) ]

(14)

2(1 + / cos )
(pB pa ).
3qB

(15)

The hardening curve (pb = f(v )) is calculated using Eqs.


(14)(16); Eq. (16) is the result of the Eq. (4) when assuming q = 0
and p = pb . The graphical form of the hardening curve depends on
the change in eccentricity (R) and on the relative density of the
tablet (RD), as stated in Figs. 7 and 8.
pb = pa (1 + R tan ) + RD.

(16)

2.2.3. Youngs modulus and the Poissons ratio


The elastic properties of the powder may be characterized using
Youngs modulus (E) and the Poissons ratio (). These parameters are discovered through experimental measurements. One
advantage of this route is that the measuring equipment is a die
compression tester. To proceed with the measurements, numerous
assumptions must be accepted: (i) the radial stress is proportional
to the axial stress, that is, r = kz , where k represents the proportionality coefcient (if a sensor for  r is not available); (ii) the
gradient of the radial stress is negligible; (iii) the tangential stress

decompression. In the elastic portion, the elastic behavior of the


materials must be independent of the stress path to avoid hysteresis. The shear modulus K and the bulk modulus G may be expressed
as follows:
K+

4
 B zC
,
G = zB
3
z Cz

(17)

2G
qB
,
= B

pC
p
3K

(18)

E=

9GK
,
3K + G

(19)

=

3K 2G
,
2(3K + G)

(20)

where B and C express the position of the stress state of the material
(Fig. 2) with the axial strain (Bz and Cz ) and axial stress (zB and zC ).
Using these equations, K and G are calculated, and according to Eqs.
(19) and (20), Youngs modulus (E) and the Poissons ratio () can
be calculated. The dependence of the change in these parameters
on the relative density of the tablet is illustrated in Figs. 9 and 10.

Please cite this article in press as: Krok, A., et al. Numerical investigation into the inuence of the punch shape on the mechanical
behavior of pharmaceutical powders during compaction. Particuology (2014), http://dx.doi.org/10.1016/j.partic.2013.12.003

G Model

ARTICLE IN PRESS

PARTIC-636; No. of Pages 16

A. Krok et al. / Particuology xxx (2014) xxxxxx

Fig. 9. The Youngs modulus (E) as a function of the relative density (RD) of the
tablet for four values of compaction stress.

2.2.4. Wall friction coefcient


Previously, two methods were developed to measure wall friction with a shear measurement device (ASM Handbook, 1998; Bell,
Grygo, Duffy, & Puri, 1995; Carr & Walker, 1968; Jenike, 1964) (at
or circular), or by a die compactor (Aydin et al., 1996; Han et al.,
2008; Jonsn & Hggblad, 2005; Sandler et al., 1976; Wu et al.,
2005). Measuring wall friction using a shear device is a sufciently
widely investigated issue that will not be presented for any details
of its procedure. Apart from a change of the position (z) of the upper
and lower punch, the stresses ( T ,  B ) are measured with a radial
pressure sensor using the following relation (Cunningham et al.,
2004; Shang et al., 2012; Sinka et al., 2003; Wu et al., 2005):

D B
4h r

  z/H
T

B

ln

T
.
B

(21)

To calculate
, the dependence of a change in wall friction on the
height of the equipment can be observed. Beyond a certain position,
this parameter does not change, remaining equal throughout the
entire process. As observed in Fig. 11 for this parameter of FEM
modeling,
= 0.1 is used.
The role of lubricants during production processes is often a
sensitive issue. The negative effects of lubricants on the physicochemical properties of the tablet are often manifested as a defect
in the mechanical tensile strength and/or restrict the dissolution of
the tablet. Magnesium stearate (MgSt) is one of the most widely
used lubricants, and its usual concentration uctuates between
0.25% and 0.5% (w/w).
The creation of a suitable compaction mixture from which the
tablet is made is a blend of a lubricant (MgSt), an active material,
and an excipient material (e.g. MCC Avicel PH 102) (Bolhuis, Lerk,
Zijlstra, & De Boer, 1975; Hlzer & Sjgren, 1981; Otsuka, Yamane, &
Matsuda, 2004; Pingali et al., 2011; Shotton & Lewis, 1964; Yu et al.,
2013). The mixing can be carried out in mixers with varied geometries and degrees of mixing exibility. Mixers with a double wall
(Otsuka et al., 2004; Pingali et al., 2011), high-speed mixers (Otsuka

Fig. 10. The Poisson ratio () as a function of the relative density (RD) of the tablet
for four values of compaction stress.

Fig. 11. The change in the wall friction coefcient (


) with radial pressure during
compaction.

et al., 2004) or planetary, drum, and cubic mixers, which differ


from each other in terms of rotational speed, capacity or critical
mixing time (Bolhuis, De Jong, Van Kamp, & Dettmers, 1987; Celik,
2011), are used. According to the authors (Kushner, 2012; Porion,
Sommier, Faugre, & Evesque, 2004; Sommier et al., 2001), Turbula
mixtures have proven to be very suitable. In contrast to simple diffusion mixers (e.g. V-mixers and Bin mixers) that provide mixing
by rotation along a single axis, the Turbula mixer ensures rotation,
transmission, and inversion of the powder via Schatz geometry. The
authors who have engaged in research regarding the lubrication of
pharmaceutical powders have kept strictly to short mixing times,
(in the range of 220 min) depending on the type of mixer used,
while using only very small amounts of lubrication, (in the range
of 0.11%) (Celik, 2011).
Ragnarsson, Hlzer, and Sjgren (1979), and Soininen and
Kuusivuori (1980) have shown that by increasing the lubrication
content in the tablet mixture, as well as extending the mixing
time, a reduction in the tensile strength of the manufactured
tablet occurred; a more obvious decrease may be observed in crystalline materials. The decreasing tensile strength of the tablets
due to increased mixing times and increased amounts of lubricant
are caused by the creation of a lubrication lm. At the interface point, the particles bind together, and the reduced tensile

Fig. 12. Photograph of micro-crystalline cellulose taken by an electron microscope


(Shang et al., 2012).

Please cite this article in press as: Krok, A., et al. Numerical investigation into the inuence of the punch shape on the mechanical
behavior of pharmaceutical powders during compaction. Particuology (2014), http://dx.doi.org/10.1016/j.partic.2013.12.003

G Model
PARTIC-636; No. of Pages 16

ARTICLE IN PRESS
A. Krok et al. / Particuology xxx (2014) xxxxxx

Fig. 13. Geometry and boundary conditions of the FEM model for (a) FF tablet, (b) FFRE tablet, and (c) SC tablet.

Fig. 14. Distributions of the (a) von Mises stress (q), (b) relative density (RD) in the powder during compaction for an FF tablet. (For interpretation of the references to color
in text, the reader is referred to the web version of the article.)

Please cite this article in press as: Krok, A., et al. Numerical investigation into the inuence of the punch shape on the mechanical
behavior of pharmaceutical powders during compaction. Particuology (2014), http://dx.doi.org/10.1016/j.partic.2013.12.003

G Model
PARTIC-636; No. of Pages 16

ARTICLE IN PRESS
A. Krok et al. / Particuology xxx (2014) xxxxxx

strength can be attributed to the emergence of weaker bonds


between lubricantlubricant compared to the excipientexcipient
bond, where the binding capacity is stronger. During the mixing process, the shear forces are sufcient to tear off the layers
of the molecules from the magnesium stearate particles, forming a thin layer on the particles of the excipient material. Hlzer
and Sjgren (1981), and Johansson and Nicklasson (1986) presume
that the MgSt is later absorbed into the excipient material particles and is consequently distributed equally across the surface
of the substrate. Concurrently, the diffusion or penetration of the
solid particles plays a lesser role. After adding excess hydrophobic
stearate to the mixture or increasing the mixing time, the bonding properties and the dissolution speed decreases. Therefore, if
the lubricant is added to the tablet mixture and mixed, the distribution of the particles is of a different fraction. Alternatively, if the
lubricant is prone to deagglomeration and delamination, it appears
as a layer on the surface of the substrate.
The DPC model was successfully calibrated for the purposes of
this work. All the necessary parameters regarding compression and
decompression were calculated.
3. Modeling powder compaction by FEM

the roughness is equal along the entire interior wall of the die. The
compacted powder should be a homogenous and isotropic material.
In the initial simulations, the results given by the 3D model are
almost identical to the 2D (two-dimensional) axisymmetric model.
The difference in the results did not exceed 1%, and the deviation
occurred only in the calculation time. The 3D simulation of the die
compression of the powder was carried out on PC Intel Core i7-CPU
2.2 GHz, 4 GB RAM; the calculation time was three hours. For the 2D
axisymmertric simulation, the calculation time was 15 s. Therefore,
in this specic case, the 2D axisymmetric model was more suitable
(Diarra et al., 2013; Michrafy et al., 2011; Sinha et al., 2010).
After applying the calculation model to ABAQUS, the calculations could be completed. The results verify the measured data on
the die compactor. The amount of simulated powder in the die was
13.5 mm high for both the 2D axisymmetric model, and for the 3D
model; the diameter of the die was 5.25 mm. A mapped orphan,
non-adaptive mesh, was chosen, and the material thus happened
to be as a deformable continuum. The upper and lower punches,
as well as the die, were modeled as a solid body without any
potential deformation. Apart from the elastic properties of the
material, such as Youngs modulus and the Poissons ratio, the

3.1. Model material


At the beginning of this section, the measured data used during
the numerical FEM investigation was already published in (Shang
et al., 2012); the authors concluded that these data are of particular interest, and offer complete measurements that may also help
while investigating other tasks related to this issue.
Microcrystalline cellulose (MCC) Avicel PH 102 is one of the
important pharmaceutical materials; therefore, in this article it
is used as the model material. The particles of this material are
very porous and have a large surface area caused by the random
grouping of the micro-crystalline bers. The capillary mechanism
between the particles facilitates the transfer of hydrophilic uids
into the pressed material, inducing the rapid dissolution of the
tablets. MCCs great compaction potential is used to good effect;
this material may contain up to 70% of the active medical ingredient. Therefore, this tablet is designed to administer medications
requiring regulated release of the medicine, ensuring a prolonged,
delayed or pulsed release of the medicine into the living organism. These sophisticated medical forms are also used as sensitive
hydrogels that react to changes in temperature, pressure or pH.
MCC Avicel PH 102 has an average particle size of 180 m and moisture content below 5%. The bulk density of the powder is 300 kg/m3 ,
while its true density is 1590 kg/m3 . The structure of the MCC Avicel
PH 102 is illustrated in Fig. 12.

Fig. 15. Comparison of the experimental and calculated data for the axial stress ( z )
vs. the axial strain (z ) at ve different loading stresses.

3.2. 3D modeling of powder compaction


The creation of a 3D (three dimensional) model is necessary
only when the geometry is different at each cross-section or
the simulated material exhibits anisotropic behavior. One example is the roll compaction of powder: the walls of the rolls are
moving with the model. In this case, 3D modeling is required
(Cunningham, Winstead, & Zavaliangos, 2010; Michrafy, Diarra,
Dodds, & Michrafy, 2011; Muliadi, Litster, & Wassgren, 2013).
When modeling the die compaction of powders, if the shape
of the compacting geometry is cylindrical (a rotationally symmetric body), the geometry in the cross-section does not change. The
behavior of the powder during compaction will change only along
the vertical axis (i.e. height) of the apparatus. The behavior of compression does not depend on the angle of the cross-section because

Fig. 16. The change of the maximum powder stress on the pq plane during compaction for ve tablets with different relative densities.

Please cite this article in press as: Krok, A., et al. Numerical investigation into the inuence of the punch shape on the mechanical
behavior of pharmaceutical powders during compaction. Particuology (2014), http://dx.doi.org/10.1016/j.partic.2013.12.003

G Model
PARTIC-636; No. of Pages 16

10

ARTICLE IN PRESS
A. Krok et al. / Particuology xxx (2014) xxxxxx

Fig. 17. Distributions of the stresses for the modied FFRE tablets: (a) von Mises stress (q) and (b) shear stress (s12 ).

cap-plasticity model was chosen for which input parameters, such


as hardening curve, cohesion, internal friction angle, eccentricity,
and the transition coefcient from calibration were dened.
Master-slave contacts with nite sliding were chosen to model
the interaction between the wall of the apparatus and the powder,
while the slip interactions were expressed using the coefcient of
friction for the wall. The difference between creating a 2D axisymmetric or a 3D model was as follows: for the 3D task, the element
type was C3D8R (three-dimensional hexahedral element), while
in the 2D axisymmetric case, the element type was CAX4R (twodimensional axisymmetric quadrilateral element). The USDFL user
subroutine was developed to calculate the distribution of the relative density of the powder during the entire process.
As observed in Fig. 13, for the FFRE tablet, the height and width
of the tablet could be changed, similar to the FF tablet. The following
dimensions were retained: h = 13.5 mm and D = 5.25 mm. Changing
the shape of the radius at the corner of the tablet was also possible.
In this study the radius was set as 1, 2, 3, or 3.5 mm for FFRE tablet.

For the SC tablet, the height and width of the tablet also remained
the same. Besides h and D, additional parameters d2 and d1 were
chosen to dene the tablet geometry. It is assumed that d1 = d2 /2,
where d1 represents the distance between the punch and the center
of the tablet, and d2 , representing the distance between the punch
and the corner of the tablet, was set as 1, 1.5, 2, or 2.5.
4. Results and discussion
Fig. 14(a) shows the distribution of von Mises stress in the material cross-section, and Fig. 14(b) shows the distribution of relative
density in the powder throughout the entire process. These states
are marked with letters A through H. During the FEM simulation,
the rst tablet was an FF tablet, whose properties for the modeling process are as follows: cohesion d = 0.279, internal friction
angle = 70.56 , eccentricity R = 0.519, transit coefcient = 0.01,
Youngs modulus E = 12.039 GPa, Poissons ratio  = 0.148 and hardening curve pb = 2.6423e4.7119v . The initial yield surface size is

Please cite this article in press as: Krok, A., et al. Numerical investigation into the inuence of the punch shape on the mechanical
behavior of pharmaceutical powders during compaction. Particuology (2014), http://dx.doi.org/10.1016/j.partic.2013.12.003

G Model
PARTIC-636; No. of Pages 16

ARTICLE IN PRESS
A. Krok et al. / Particuology xxx (2014) xxxxxx

11

Fig. 18. Distributions of stresses for the modied SC tablets: (a) von Mises stress (q) and (b) shear stress (s12 ).

0.0005, and the ow stress ratio is zero. Compaction begins in


Fig. 14(a) dataset A, when the stress across the entire cross-section
is zero. In datasets B and C, the course of compaction can be
observed, and in dataset D, the material is compacted to its maximum value. Dataset E shows the course of unloading when the
compression stress from the punch no longer acts on the particulate material and elastic behavior occurs more intensively. The red
lines in the pictures reveal the shift away from the maximum level
of deformation in the material when the von Mises stress achieves
its lowest values. This state conforms with point C in the DPC model
in Fig. 2. In dataset G, the level of the von Mises stress increased
slightly. The radial stress is approximately equal to the axial stress.
In dataset H, the material reaches the end of its decompression.
Figs. 15 and 16 compare the FE analysis from the ABAQUS program with the results from the powder compaction experiment

for ve tablet samples with different RD. Depending on the loading pressure of the punch, some properties of the powder have to
be changed because relative density is bound to the parameters
of the DPC model. Therefore, these parameters cannot be selected
independently. For each selected RD, the correct cohesion (d), internal friction (), eccentricity (R), Youngs modulus (E) and Poissons
number () must be set according to the measured data from the
DPC model calibration. The geometry, boundary conditions of the
apparatus walls, and the wall friction (
) remain the same.
4.1. Inuence of punch shape on compaction
To demonstrate the inuence of the punch shape, the authors
began with the assumption that the properties of the powder, the
coefcient of friction for the wall (
= 0.1), and the punch speed

Please cite this article in press as: Krok, A., et al. Numerical investigation into the inuence of the punch shape on the mechanical
behavior of pharmaceutical powders during compaction. Particuology (2014), http://dx.doi.org/10.1016/j.partic.2013.12.003

G Model
PARTIC-636; No. of Pages 16

12

ARTICLE IN PRESS
A. Krok et al. / Particuology xxx (2014) xxxxxx

Fig. 19. Change of the powder stress on the pq plane with the change in the geometry of the model (d2 ) and the wall friction coefcient (
) for the SC tablets. (For
interpretation of the references to color in text, the reader is referred to the web
version of the article.)

Fig. 20. Change of the powder stress on the pq plane with the change in the geometry of the model (R) and the coefcient of wall stress (
) for the FFRE tablets. (For
interpretation of the references to color in text, the reader is referred to the web
version of the article.)

(v = 0.7 mm/s) remain the same as for the FF tablet (see the above
part of Section 4), while the actual change to the simulation model
consists only in a change in the geometry. The geometry of the
model and the boundary conditions are stated in Fig. 13.
As observed in Fig. 14(a), for the FF tablet, the distribution of
the von Mises stress is homogenous, and signicant changes can be
observed only in the corners of this model due to the wall friction. In

this case, the shear surfaces are created only in proximity to the wall
of the die because the wall presents the greatest resistance toward
the movement of the particles during compaction. The interaction
of the shear stress between the powder and the punch is negligible
for the FF tablets. Figs. 17 and 18 show that when changing the
shape of the compression space, the pressure distribution changes.
The interaction between the powder and the wall of the die, as well

Fig. 21. Dependence of the powder stress on the wall friction coefcient (
) for the FF tablets. The top row represents the distributions of von Mises stress and the bottom
row represents the distributions of shear stress in the FF tablet samples. (For interpretation of the references to color in text, the reader is referred to the web version of the
article.)

Please cite this article in press as: Krok, A., et al. Numerical investigation into the inuence of the punch shape on the mechanical
behavior of pharmaceutical powders during compaction. Particuology (2014), http://dx.doi.org/10.1016/j.partic.2013.12.003

G Model
PARTIC-636; No. of Pages 16

ARTICLE IN PRESS
A. Krok et al. / Particuology xxx (2014) xxxxxx

13

Fig. 22. Dependence of the powder stress on the wall friction coefcient (
) for the SC tablets. The top row represents the distributions of von Mises stress and the bottom
row represents the distributions of shear stress in the SC tablet samples.

as the interaction between the powder and the compression punch,


change the stresses more signicantly for the FFRE tablets.
This result is natural because, when increasing for an FFRE
tablet, the contact interaction is increased by the contribution of
interaction between the powder and the punches. The slope angle
of the punch surface is no longer favorable; therefore, the material
must exert part of the energy from compaction to overcome this
resistance. The result is a non-homogeneous distribution of the von
Mises stress in the compacted material, as well as a large shear that
also appears due to the radial elasticity of the material. Concurrently, capping and lamination faults can be caused by increasing
the local concentration of the shear stress. In Fig. 18, the distribution of the von Mises pressure for the SC tablets is shown. This
gure shows that the effects of the wall friction that originate from
the interaction with the top and bottom punches, against which the
material is pressed, are signicant for this shape. In addition, the
stress on the powder is inuenced by the shear friction from the
wall of the die, as well as the interaction with the top and bottom
punches.
4.2. Inuences of friction and punch shape on compaction
In this section, the impact of the coefcient of the wall friction
on the mechanical behavior of the powder was investigated. In this
case, FFRE and SC tablets were used. As observed in Figs. 19 and 20,
the results of this investigation are presented on the pq plane for
a clearer interpretation of the results. The procedure used to test
the inuence of the wall friction involved using eight different wall
friction coefcients (0.05
0.4) for each geometric model. This

range was selected by (Eiliazadeh et al., 2004; Han et al., 2008;


Kadiri & Michrafy, 2013; Wu et al., 2008) and was found to be
acceptable. For each calculation, the von Mises stress (q) was set
to a maximum value on the pq plane, along with the corresponding hydrostatic stress (p). In Figs. 19 and 20 four differently colored
lines can be observed. These lines express how the maximum q and
p values changed (Fig. 2, point B) with the wall friction coefcient
for the given shape of the model.
As observed in Fig. 21, for the FF tablets with a at punch, lubrication has no effect on the distribution of stress in the volume of the
tablet; after greasing the punch surface and the walls of the apparatus, the von Mises stress (q) changes only minimally (Han et al.,
2008; Wu et al., 2008). The effect of lubrication can be observed
in the lower right corner (blue area), that is, after an increase in
lubrication, the stress also increases due to the better transfer of
the force in the volume of the sample.
The inuence of lubrication appears more signicant after
changing the punch geometry. In Fig. 19 the reduction of stresses p
and q due to lubrication in the range of 0.05
0.4 for the geometry of the SC tablets (four full black lines marked with colored
points) can be observed. Therefore, by reducing the wall friction
coefcient, the total average stress of the SC tablet decreases. Concurrently, the sensitivity of the reduction of this stress depends on
the geometry of the punches. Fig. 22 shows that the shear friction is
primarily concentrated near the interface between the punch and
material; adding lubrication can change the shape of the emerging
shear areas. If the friction coefcient is decreased, the resistance
that the punch shows in relation to the movement of the material
is also reduced; in this case, a more intensive transfer of the stress

Please cite this article in press as: Krok, A., et al. Numerical investigation into the inuence of the punch shape on the mechanical
behavior of pharmaceutical powders during compaction. Particuology (2014), http://dx.doi.org/10.1016/j.partic.2013.12.003

G Model
PARTIC-636; No. of Pages 16

14

ARTICLE IN PRESS
A. Krok et al. / Particuology xxx (2014) xxxxxx

Fig. 23. Dependence of the powder stress on the wall friction coefcient (
) for the FFRE tablets. The top row represents the distributions of von Mises stress and the bottom
row represents the distributions of shear stress in the SC tablet samples. (For interpretation of the references to color in text, the reader is referred to the web version of the
article.)

among the particles occurs. Consequently, when


= 0.4, the von
Mises stress was concentrated near the wall of the compaction
space, while at
= 0.05 the emerging area was already distanced
from the wall in the direction of the internal volume of the tablet.
Therefore, lubrication yielded a tablet with lower stress than its
original value, but the average stress takes up a greater volume,
making the tablet more compact.
In Fig. 20, the change in the maximum p and q caused by the
lubrication for the geometry of the FFRE tablet (four full black lines
marked with colored dots) can be observed. If the friction is large
(
= 0.4), the mobility of the particles at the contact point with the
punch surface is restricted. Around the circumference of the wall,
shear areas are formed; during the nal analysis, the stresses q
are concentrated at this point (Fig. 23). If the wall friction begins
to decrease signicantly due to lubrication, the mobility of the
particles in contact with the punch surface increases. Because
the punch has a conical surface, the horizontal component of
the force begins to press the powder toward the center of the
sample (tablet) due to the disintegration of the forces acting
between the punch and the powder. In the range of 0.15
0.4,
the stresses (p and q) increase. After reducing the wall friction
coefcient to 0.05
0.15, this trend reverses, and the stress
begins to decrease, similar to the SC tablet. Consequently, the

stress is concentrated in the center of the tablet, and in the wall


surroundings, the stress is minimal. (See Fig. 23 for FFRE tablets
with a wall friction coefcient
= 0.05.)
The shape of the curves in Fig. 20 is interesting. When the friction
is high (
= 0.150.4), the sample reforms near the wall of the cylinder (see red areas in Fig. 23), while the core remains only slightly
compressed (Fig. 23). After a reduction in friction, the movement of
the particles in proximity to the surfaces of the compacting apparatus is freed, and the reformation is shifted toward the core of
the tablet, which is where the most stress is located. This change
may result in the change in stresses of  z and  r due to the reduced
stress. While  z dominated initially, the effect of  r began to appear
under reduced friction.
To measure the compaction of the powder in a die with a range of

(0.050.4) experimentally, a few minor alterations are necessary.


The principle for obtaining the shear friction coefcient has been
presented in Section 2.2.4. To achieve a higher wall friction coefcient (
= 0.40.5), the interior surface of the die must be roughened
or a different construction material that suits the required properties for the apparatus must be used. To acquire a lower wall friction
coefcient (
= 0.160.35), the experience of other authors (e.g.
Han et al., 2008; Michrafy et al., 2004; Wu et al., 2008) was useful.
Reaching such a low wall friction coefcient for tablets without

Please cite this article in press as: Krok, A., et al. Numerical investigation into the inuence of the punch shape on the mechanical
behavior of pharmaceutical powders during compaction. Particuology (2014), http://dx.doi.org/10.1016/j.partic.2013.12.003

G Model
PARTIC-636; No. of Pages 16

ARTICLE IN PRESS
A. Krok et al. / Particuology xxx (2014) xxxxxx

mechanical defects is difcult. In this case, our knowledge of


external lubrication should be used. External lubrication could be
achieved by creating a liquid lm on the surface of the die. However,
this state is hard to achieve in practice due to the high compaction
pressures and the small dimensions of the powder particles.
5. Conclusions
The reduction of wall friction should help the tablet retain its
strength and density while distributing these parameters homogeneously throughout the tablet. The results described in this report
indicate that this effect is closely connected with the geometry of
the tablet model. To retain the desired properties, the correct values for lubrication must be chosen, depending on the shape of the
tablet. Therefore, to produce a homogeneous tablet, neither too little (tablet with the greatest stress near the wall of the die), nor
too much lubrication, (when the stress is greatest in its center) is
appropriate. These simulations indicate that the optimal value for
the given material is
= 0.2 and that the appearance of signicant
shear areas for the FFRE and SC tablets might indicate the possibility
of increased damage to the tablets.
Acknowledgments
This contribution was supported by the Scientic Grant Agency
of the Ministry of Education of the Slovak Republic for the research
project, titled Basic Research of the Operations in the Mechanics of
Particulate Material No. 1/0652/13, and by the Slovak University of
Technology to support young researchers (Grant No. 1312)/2013.
References
ASM Handbook. (1998). Vol. 7: Powder metal technologies and applications. OH, USA:
ASM International Materials Park.
Aydin, I., Briscoe, B. J., & Sanhtrk, K. Y. (1996). The internal form of compacted
ceramic components: A comparison of a nite element modeling with experiment. Powder Technology, 89, 239254.
Bell, T. A., Grygo, R. J., Duffy, S. P., & Puri, V. M. (1995). Simplied methods of measuring powder cohesive strength. In Preprints of the 3rd European symposium
Storage and ow of particulate solids, PARTEC 95 Nrnberg, Germany, (pp. 7688).
Bolhuis, G. K., De Jong, S. W., Van Kamp, H. V., & Dettmers, H. (1987). The effect on
tablet crushing strength of magnesium stearate admixing in different types of
lab scale and production-scale mixers. Pharm Technology, 11(3), 3644.
Bolhuis, G. K., Lerk, C. F., Zijlstra, H. T., & De Boer, A. H. (1975). Film formation by
magnesium stearate during mixing and its effect on tabletting. Pharmacetisch
Weekblad, 110(16), 317325.
Brewin, P. R., Coube, O., Doremus, P., & Tweed, J. H. (2008). Modelling of powder die
compaction. New York: Springer.
Carr, J. F., & Walker, D. M. (1968). An annular shear cell for granular materials. Powder
Technology, 1, 369373.
Celik, M. (2011). Pharmaceutical powder compaction technology (2nd ed.). London:
Informa Healthcare.
Chtourou, H., Gakwaya, A., & Guillot, M. (2002). Modeling of the metal powder
compaction process using the cap model. Part II: Numerical implementation
and practical applications. International Journal of Solids and Structures, 39,
10771096.
Chtourou, H., Guillot, M., & Gakwaya, A. (2002). Modeling of the metal powder
compaction process using the cap model. Part I: Experimental material characterization and validation. International Journal of Solids and Structures, 39,
10591075.
Cleary, P. W. (2000). DEM simulation of industrial particle ows: Case study of
dragline excavators, mixing in tumblers and centrifugal mills. Powder Technology, 109, 83104.
Coube, O., Cocks, A. C. F., & Wu, C.-Y. (2005). Experimental and numerical study of
die lling, powder transfer and die compaction. Powder Metallurgy, 48, 6875.
Coube, O., & Riedel, H. (2000). Numerical simulation of metal powder die compaction
with special consideration of cracking. Powder Metallurgy, 43, 123131.
Cundall, P. A., & Strack, O. D. L. (1979). A discrete numerical model for granular
assemblies. Geotechniques, 29, 4765.
Cunningham, J. C., Sinka, I. C., & Zavaliangos, A. (2004). Analysis of tablet compaction.
I. Characterization of mechanical behaviour of powder and powder/tooling friction. Journal of Pharmaceutical Sciences, 93, 20222039.
Cunningham, J. C., Winstead, D., & Zavaliangos, A. (2010). Understanding variation in
roller compaction through nite element-based process. Computers & Chemical
Engineering, 34, 10581071.

15

Diarra, H., Mazel, V., Boillon, A., Rehault, L., Busignies, V., Bureau, S., et al. (2012).
Finite element method (FEM) modeling of the powder compaction of cosmetic
products: Comparison between simulated and experimental results. Powder
Technology, 224, 233240.
Diarra, H., Mazel, V., Busignies, V., & Tchoreloff, P. (2013). FEM simulation of the die
compaction of pharmaceutical products: Inuence of visco-elastic phenomena
and comparison with experiments. International Journal of Pharmaceutics, 453,
389394.
DiMaggio, F. L., & Sandler, I. S. (1971). Material model for granular soils. Journal of
the Engineering Mechanics Division, 97, 935950.
Djemai, A., & Sinka, I. C. (2006). NMR imaging of density distributions in tablets.
International Journal of Pharmaceutics, 319, 5562.
Doremus, P., Toussaint, F., & Alvain, O. (2001). Simple tests and standard procedure
for the characterisation of green compacted powder. In A. Zavaliangos, & A.
Laptev (Eds.), Recent developments in computer modeling of powder metallurgy
processes 1518 May, Kiev, Ukraine, (pp. 2941). IOS Press.
Drucker, D. C., Gibson, R. E., & Henkel, D. J. (1957). Soil mechanics and workhardening theories of plasticity. Transactions of the American Society of Civil
Engineers, 122, 338346.
Drucker, D. C., & Prager, W. (1952). Soil mechanics and plastic analysis or limit design.
Quarterly of Applied Mathematics, 10(2), 157165.
Eiliazadeh, B., Pitt, K., & Briscoe, B. (2004). Effects of punch geometry on powder
movement during pharmaceutical tabletting processes. International Journal of
Solids and Structures, 41, 59675977.
Farber, L., Tardos, G., & Michaels, J. N. (2003). Use of X-ray tomography to
study the porosity and morphology of granules. Powder Technology, 132,
5763.
Fell, J. T., & Newton, J. M. (1970). Determination of tablet strength by the diametral
compression test. Journal of Pharmaceutical Sciences, 59, 688691.
Frenning, G. (2007). Analysis of pharmaceutical powder compaction using multiplicative hyperelasto-plastic theory. Powder Technology, 172, 103112.
Frocht, M. M. (1947). Photoelasticity. New York: John Wiley and Sons.
Green, R. J. (1972). A plasticity theory for porous solids. International Journal of
Mechanical Sciences, 14, 215224.
Gurson, A. L., & McCabe, T. J. (1992). Experimental determination of yield functions
for compaction of blended powders. In Proceedings of MPIF/APMI world congress
on powder metallurgy and particulate materials San Francisco, USA.
Guyot, J. C., Delacourte, A., & Marie, B. (1986). Computer determination and comparison of the compression behaviour of powder mixtures. Drug Development
and Industrial Pharmacy, 12(11), 18691884.
Han, H. N., Kim, H. S., Oh, K. H., & Lee, D. N. (1994). Elastoplastic nite element
analysis for porous metals. Powder Metallurgy, 37, 140146.
Han, L. H., Elliott, J. A., Bentham, A. C., Mills, A., Amidon, G. E., & Hancock, B. C.
(2008). A modied DruckerPrager Cap model for die compaction simulation
of pharmaceutical powders. International Journal of Solids and Structures, 45,
30883106.
Heckel, R. W. (1961). Densitypressure relationships in powder compaction. Transactions of the Metallurgical Society of AIME, 221, 671675.
Hofstetter, G., Simo, J. C., & Taylor, R. L. (1993). A modied cap model: Closest point
solution algorithms. Computers & Structures, 46, 203214.
Hlzer, A. W., & Sjgren, J. (1981). Evaluation of some lubricants by the comparison
of friction coefcients and tablet properties. Acta Pharmaceutica Suecica, 18(3),
139148.
James, B. A. (1987). Die-wall lubrication for powder compacting: A feasible solution?
Powder Metallurgy, 30, 273.
Jenike, A. W. (1964). Storage and ow of solids. Salt Lake City, USA: Bulletin 123 of
Utah Engineering Experiment Station, University of Utah.
Jetzer, W., Leuenberger, H., & Sucker, H. (1983). The compressibility and compactibility of powder mixtures. Pharm. Technol., 7(11), 3348.
Johansson, M. E., & Nicklasson, M. (1986). Investigation of the lm formation of
magnesium stearate by applying a ow-through dissolution technique. Journal
of Pharmacy and Pharmacology, 38(1), 5154.
Jonsn, P., & Hggblad, H. A. (2005). Modelling and numerical investigation of the
residual stress state in a green metal powder body. Powder Technology, 155,
196208.
Kadiri, M. S., & Michrafy, A. (2013). The effect of punchs shape on die compaction
of pharmaceutical powders. Powder Technology, 239, 467477.
Kadiri, M. S., Michrafy, A., & Dodds, J. A. (2005). Pharmaceutical powders compaction:
Experimental and numerical analysis of the density distribution. Powder Technology, 157, 176182.
Kim, K. T., Choi, S. W., & Park, H. (2000). Densication behaviour of ceramic powder
under cold compaction. Journal of Engineering Materials and Technology, 122,
238244.
Koerner, R. M. (1971). Triaxial compaction of metal powders. Powder Metallurgy
International, 3, 186188.
Korachkin, D., Gethin, D. T., Lewis, R. W., & Tweed, J. H. (2008). Effect of die
lling on powder compaction. International Journal of Powder Metallurgy, 44,
2234.
Kushner, J. I. V. (2012). Incorporating Turbula mixers into a blending scale-up
model for evaluating the effect of magnesium stearate on tablet tensile strength
and bulk specic volume. International Journal of Pharmaceutics, 429(12),
111.
Leuenberger, H., & Rohera, B. D. (1986). Fundamentals of powder compression: 1.
The compactibility and compressibility of pharmaceutical powders. Pharmaceutical Research, 3(1), 1222.

Please cite this article in press as: Krok, A., et al. Numerical investigation into the inuence of the punch shape on the mechanical
behavior of pharmaceutical powders during compaction. Particuology (2014), http://dx.doi.org/10.1016/j.partic.2013.12.003

G Model
PARTIC-636; No. of Pages 16

16

ARTICLE IN PRESS
A. Krok et al. / Particuology xxx (2014) xxxxxx

Macleod, H. M., & Marshall, U. (1977). The determination of density distribution in


ceramic compacts using autoradiography. Powder Technology, 16, 107122.
Michrafy, A., Diarra, H., Dodds, J. A., & Michrafy, M. (2011). Experimental and
numerical analyses of homogeneity over strip width in roll compaction. Powder
Technology, 206, 154160.
Michrafy, A., Dodds, J. A., & Kadiri, M. S. (2004). Wall friction in the compaction of
pharmaceutical powders: Measurement and effect on the density distribution.
Powder Technology, 148, 5355.
Michrafy, A., Ringenbacher, D., & Tchoreloff, P. (2002). Modelling the compaction
behaviour of powders: Application to pharmaceutical powders. Powder Technology, 127, 257266.
Muliadi, A. R., Litster, J. D., & Wassgren, C. R. (2013). Validation of 3-D nite element
analysis for predicting the density distribution of roll compacted pharmaceutical
powder. Powder Technology, 237, 386399.
Nebgen, G., Gross, D., Lehmann, V., & Mller, F. (1995). H-NMR microscopy of tablets.
Journal of Pharmaceutical Sciences, 84, 283291.
Newton, J. M., Haririan, I., & Podczeck, F. (2000). The inuence of punch curvature
on the mechanical properties of compacted powders. Powder Technology, 107,
7983.
Otsuka, M., Yamane, I., & Matsuda, Y. (2004). Effects of lubricant mixing on compression properties of various kinds of direct compression excipients and physical
properties of the tablets. Advanced Powder Technology, 15(4), 477493.
Pavier, E., & Doremus, P. (1999). Triaxial characterisation of iron powder behaviour.
Powder Metallurgy, 42, 345352.
Pingali, K., Mendez, R., Lewis, D., Michniak-Kohn, B., Cuitino, A., & Muzzio, F. (2011).
Mixing order of glidant and lubricant Inuence on powder and tablet properties. International Journal of Pharmaceutics, 409(12), 269277.
Porion, P., Sommier, N., Faugre, A. M., & Evesque, P. (2004). Dynamics of size segregation and mixing of granular materials in a 3D-blender by NMR imaging
investigation. Powder Technology, 141(12), 5568.
Procopio, A. T., Zavaliangos, A., & Cunningham, J. C. (2003). Analysis of the diametrical
compression test and the applicability to plastically deforming materials. Journal
of Materials Science, 38, 36293639.
Ragnarsson, G., Hlzer, A. W., & Sjgren, J. (1979). The inuence of mixing time and
colloidal silica on the lubrication properties of magnesium stearate. International
Journal of Pharmaceutics, 3(23), 127131.
Rottmann, G., Coube, O., & Riedel, H. (2001). Comparison between triaxial results
and models prediction with special consideration of the anisotropy. In European
congress on powder metallurgy 2001 (Vol. 3) Shrewsbury, UK, (pp. 2937).
Rue, P. J., & Rees, J. E. (1978). Limitations of the Heckel relation for predicting powder
compaction mechanisms. Journal of Pharmacy and Pharmacology, 30, 642643.
Sandler, I. S., Baladi, G. Y., & DiMaggio, F. L. (1976). Generalized cap model for geological materials. Journal of the Geotechnical Engineering Division, 102, 683699.
Schoeld, A. N., & Wroth, C. P. (1968). Critical state soil mechanics. London: McGrawHill.
Shang, C. (2012). Modelling powder compaction and breakage of compacts (Doctoral
dissertation). UK: University of Leicester.
Shang, C., Sinka, I. C., & Pan, J. (2012). Constitutive model calibration for powder compaction using instrumented die testing. Experimental Mechanics, 52, 903916.

Sharma, V. M., Saxena, K. R., & Woods, R. D. (Eds.). (1999). Distinct element modelling
in geomechanics. Rotterdam: A. A Balkema.
Shotton, E., & Lewis, C. J. (1964). Some observations on the effect of lubrication on the
crushing strength of tablets. Journal of Pharmacy and Pharmacology, 16(Suppl.),
111T120T.
Sinha, T., Bharadwaj, R., Curtis, J. S., Hancock, B. C., & Wassgren, C. (2010). Finite element analysis of pharmaceutical tablet compaction using a density dependent
material plasticity model. Powder Technology, 202, 4654.
Sinka, I. C., Burch, S. F., Tweed, J. H., & Cunningham, J. C. (2004). Measurement of
density variations in tablets using X-ray computed tomography. International
Journal of Pharmaceutics, 271, 215224.
Sinka, I. C., Cunningham, J. C., & Zavaliangos, A. (2003). The effect of wall
friction in the compaction of pharmaceutical tablets with curved faces: A
validation study of the DruckerPrager Cap model. Powder Technology, 133,
3343.
Soininen, A., & Kuusivuori, P. (1980). Inuence of the length of mixing time on
bioavailability of capsule formulations containing magnesium stearate. Acta
Pharmaceutica Fennica, 89, 215222.
Sommier, N., Porion, P., Evesque, P., Leclerc, B., Tchoreloff, P., & Couarraze, G. (2001).
Magnetic resonance imaging investigation of the mixing-segregation process
in a pharmaceutical blender. International Journal of Pharmaceutics, 222(2),
243258.
Timoshenko, S. P., & Goodier, J. N. (1970). Theory of elasticity. New York: McGraw-hill.
Watson, T. J., & Wert, J. A. (1993). On the development of constitutive relations for
metallic powders. Metallurgical Transactions A, 24, 20712081.
Wood, D. M. (1991). Soil behaviour and critical state soil mechanics. Cambridge:
Cambridge University Press.
Wu, C. Y., Cocks, A. C. F., Gilia, T. G., & Thompson, D. A. (2003). Experimental and numerical investigations of powder transfer. Powder Technology, 138,
216228.
Wu, C. Y., Hancock, B. C., Mills, A., Benthama, A. C., Best, S. M., & Elliott,
J. A. (2008). Numerical and experimental investigation of capping mechanisms during pharmaceutical tablet compaction. Powder Technology, 181,
121129.
Wu, C. Y., Ruddy, O. M., Bentham, A. C., Hancock, B. C., Best, S. M., & Elliott, J. A.
(2005). Modelling the mechanical behaviour of pharmaceutical powders during
compaction. Powder Technology, 152, 107117.
York, P. (1979). A consideration of experimental variables in the analysis of powder
compaction behaviour. Journal of Pharmacy and Pharmacology, 31, 244246.
Yu, S., Adams, M., Gururajan, B., Reynolds, G., Roberts, R., & Wu, C. Y. (2013). The
effects of lubrication on roll compaction, ribbon milling and tabletting. Chemical
Engineering Science, 86, 918.
Zhang, B., Jain, M., Zhao, C., Bruhis, M., Lawcock, R., & Ly, K. (2010). Experimental calibration of density-dependent modied DruckerPrager/cap model
using an instrumented cubic die for powder compact. Powder Technology, 204,
2741.
Zhu, H. P., Zhou, Z. Y., Yang, R. Y., & Yu, A. B. (2007). Discrete particle simulation of
particulate systems: Theoretical development. Chemical Engineering Science, 62,
33783396.

Please cite this article in press as: Krok, A., et al. Numerical investigation into the inuence of the punch shape on the mechanical
behavior of pharmaceutical powders during compaction. Particuology (2014), http://dx.doi.org/10.1016/j.partic.2013.12.003

Potrebbero piacerti anche