Advances in Pharmaceutical Powder Processing
Advances in Pharmaceutical Powder Processing
Recent Advances in
Secondary Processing of
Pharmaceutical Powders
Edited by
Colin Hare
mdpi.com/journal/pharmaceutics
Recent Advances in Secondary
Processing of Pharmaceutical Powders
Recent Advances in Secondary
Processing of Pharmaceutical Powders
Editor
Colin Hare
Editorial Office
MDPI
St. Alban-Anlage 66
4052 Basel, Switzerland
This is a reprint of articles from the Special Issue published online in the open access journal
Pharmaceutics (ISSN 1999-4923) (available at: www.mdpi.com/journal/pharmaceutics/special
issues/pharm powders).
For citation purposes, cite each article independently as indicated on the article page online and as
indicated below:
Lastname, A.A.; Lastname, B.B. Article Title. Journal Name Year, Volume Number, Page Range.
© 2024 by the authors. Articles in this book are Open Access and distributed under the Creative
Commons Attribution (CC BY) license. The book as a whole is distributed by MDPI under the terms
and conditions of the Creative Commons Attribution-NonCommercial-NoDerivs (CC BY-NC-ND)
license.
Contents
Preface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii
Satyajeet Bhonsale, Lewis Scott, Mojtaba Ghadiri and Jan Van Impe
Numerical Simulation of Particle Dynamics in a Spiral Jet Mill via Coupled CFD-DEM
Reprinted from: Pharmaceutics 2021, 13, 937, doi:10.3390/pharmaceutics13070937 . . . . . . . . . 14
Danni Suhaidi, Yao-Da Dong, Paul Wynne, Karen P. Hapgood and David A. V. Morton
Bulk Flow Optimisation of Amorphous Solid Dispersion Excipient Powders through Surface
Modification
Reprinted from: Pharmaceutics 2023, 15, 1447, doi:10.3390/pharmaceutics15051447 . . . . . . . . 114
Niels Lasse Martin, Ann Kathrin Schomberg, Jan Henrik Finke, Tim Gyung-min Abraham,
Arno Kwade and Christoph Herrmann
Process Modeling and Simulation of Tableting—An Agent-Based Simulation Methodology for
Direct Compression
Reprinted from: Pharmaceutics 2023, 13, 996, doi:10.3390/pharmaceutics13070996 . . . . . . . . . 197
v
Owen Jones-Salkey, Zoe Chu, Andrew Ingram, and Christopher Windows-Yule
Reviewing the Impact of Powder Cohesion on Continuous Direct Compression (CDC)
Performance
Reprinted from: Pharmaceutics 2023, 15, 1587, doi:10.3390/pharmaceutics15061587 . . . . . . . . 221
vi
Preface
Pharmaceuticals are becoming increasingly important in modern society. In most cases, active
pharmaceutical ingredients (APIs) are formed through crystallisation but need to be combined with
excipients to form a tablet, or with carriers to form an oral suspension. The formation of tablets
requires several secondary processing steps, such as milling of the API and mixing with excipients
(often involving granulation) before finally being tabletted. Due to the differences in physical and
mechanical properties that individual APIs and excipients exhibit, these secondary processing steps
are often fraught with challenges. This Special Issue comprises 11 research articles and 1 review
which outline the recent advances in secondary processing of pharmaceutical powders, and aims
to enhance our understanding of powder behaviour and help us tackle these challenges. Through
careful experimentation and the application of sophisticated computational techniques, the authors
of these papers demonstrate advances in our understanding of many of the processing steps for
pharmaceutical products. With this Special Issue, I hope to engage a broad readership in order
to spark new ideas to further advance our knowledge of powder processing and to optimise
production.
Colin Hare
Editor
vii
pharmaceutics
Article
Experimental Elucidation of Templated Crystallization and
Secondary Processing of Peptides
Vivek Verma 1, *, Isha Bade 1 , Vikram Karde 1 and Jerry Y. Y. Heng 1,2, *
1 Department of Chemical Engineering, Imperial College London, London SW7 2AZ, UK;
[email protected] (I.B.); [email protected] (V.K.)
2 Institute for Molecular Science and Engineering, Imperial College London, London SW7 2AZ, UK
* Correspondence: [email protected] (V.V.); [email protected] (J.Y.Y.H.)
Abstract: The crystallization of peptides offers a sustainable and inexpensive alternative to the
purification process. In this study, diglycine was crystallised in porous silica, showing the porous
templates’ positive yet discriminating effect. The diglycine induction time was reduced by five-
fold and three-fold upon crystallising in the presence of silica with pore sizes of 6 nm and 10 nm,
respectively. The diglycine induction time had a direct relationship with the silica pore size. The
stable form (α-form) of diglycine was crystallised in the presence of porous silica, with the diglycine
crystals obtained associated with the silica particles. Further, we studied the mechanical properties of
diglycine tablets for their tabletability, compactability, and compressibility. The mechanical properties
of the diglycine tablets were similar to those of pure MCC, even with the presence of diglycine
crystals in the tablets. The diffusion studies of the tablets using the dialysis membrane presented an
extended release of diglycine through the dialysis membrane, confirming that the peptide crystal can
be used for oral formulation. Hence, the crystallization of peptides preserved their mechanical and
pharmacological properties. More data on different peptides can help us produce oral formulation
peptides faster than usual.
2
Pharmaceutics 2023, 15, 1288
for high-concentration doses without increasing the viscosity of the suspension as well
as the ability to reach a concentration greater than 200 mg/mL in the final dosage form,
leading to a higher bioavailability and lower dosage requirement [25]. Hence, peptide
crystallization offers great applicability for oral doses of the current peptide molecules on
the market.
At present, oral formulations are available for several hormones, including insulin,
vasopressin, somatostatin, calcitonin, parathyroid hormone (PTH), uroguanylin, thyroid
hormone-releasing hormone, and GLP1. These treatments can be classified into two
categories based on their intended action within the body—those that require oral ab-
sorption and those that require retention in the gastrointestinal tract [26,27]. All of these
formulations use one of the above-mentioned strategies; more specifically, well-known per-
meation enhancers such as sodium caprate (C10; also known as decanoic acid) and sodium
N-[8-(2-hydroxybenzoyl)amino] caprylate (SNAC; also known as salcaprozate sodium)
improve the transcellular and paracellular permeation [20]. However, there has not yet been
a report of an oral formulation of the peptide using just the peptide crystals on their own.
This is potentially due to the fact that no peptide drug is small enough to be compliant with
Lipinski’s rule of five (MW < 500 g/mol, H-bond donors < 5 and H-bond acceptors < 10,
LogP < 5, rotatable bonds < 10, and total polar surface area < 140 Å2 ) for predicting good
absorption and permeation [28]. Surprisingly, the peptides that have been identified as
having good oral activity and that could be potential candidates for oral formulations, in
addition to the one in the clinical trial, are mostly cyclic [29]. This suggests that further
research is needed to explore more cyclic peptides with good oral activity.
This work focuses on improving the crystallization of peptides using the templated
crystallization approach, already established by the authors [9,10,13,14]. The templates in-
teract with the crystallising solute through functional group complementarity, allowing the
sequestration of solute molecules on the surface of the templates with h-bond interactions
for a long enough time to achieve a fully grown crystal. This was recently complemented
with a molecular dynamics simulation, as mentioned earlier in the section [14]. In this
work, diglycine was crystallised in the presence of different pore sizes of silica particles
to examine the effect of template pore size on the crystallization time and rate. Diglycine
is made by combining the simplest amino acid, i.e., glycine, with itself though a peptide
bond, making it the smallest known peptide, which behaves similarly to a small molecule
due to the lack of degrees of freedom, defined unit cell, and well-defined intermolecular
contacts. The crystallised peptide was later blended with a widely used pharmaceutical
excipient, microcrystalline cellulose (MCC) [30], to obtain an oral formulation. The formu-
lation was later compressed to form tablets, which were studied to obtain the tabletability,
compressibility, and compactibility for the peptide tablets. This is the first study reporting
the tablet properties of a peptide formulation. The tablets were later studied to obtain the
permeability rate of different formulations.
3
Pharmaceutics 2023, 15, 1288
rated diglycine solution was heated to 5 ◦ C above the saturation temperature for complete
dissolution of diglycine. This solution was cooled to 32.7 ◦ C to induce crystallization at
relative supersaturations of 1.20, which is in the metastable zone width limit, enabling us
to capture the effect of templates on heterogenous nucleation of diglycine. The change in
concentration upon nucleation of diglycine in the absence and presence of porous silica
particles ((10% w/w loading) was captured using Mettler-Toledo ReactIR 15 system, an
in situ Fourier transform infrared (FTIR) probe. Each experiment was carried out at least
twice to ensure the reproducibility of the results. The induction time was accessed from
the desupersaturation curves using the tangent method that was previously used by the
authors [14].
4
Pharmaceutics 2023, 15, 1288
The ρtrue is the true density (g/cm3 ) calculated using the helium pycnometer men-
tioned in Section 2.4.3, while ρapp is the apparent density calculated by dividing tablet
weight by its volume.
2.5.2. Tabletability
The tabletability of a material is represented by the relationship between its tensile
strength and the compaction pressure applied. Tabletability is expressed by a linear relation
between tensile strength and compaction force, as given by Newton et al. [32] and presented
in Equation (2).
σt = C p P + b (2)
where P is the compaction pressure, Cp is the tabletability parameter, and b is a constant.
2.5.3. Compactability
An exponential relation between the tensile strength and porosity expresses the com-
pactability of a composite. Ryshkewitch–Duckworth proposed a mathematical equation to
understand compactability, as shown in Equation (3) [33].
where tablet tensile strength (MPa) is represented by σt and tablet tensile strength at zero
porosity (MPa) by σt0 . The tablet porosity is given by P, while b is an empirical constant
representing bonding capacity, in which stronger bonding between primary particles is
expressed by higher b value [34].
2.5.4. Compressibility
Compressibility profile shows the change in the tablet porosity with the increasing
compaction pressure. A tablet with low porosity is associated with the capping problem
and can result in slow tablet dissolution. Compressibility can be assessed by change in
tablet porosity with compaction pressure as expressed by Heckle’s model according to
Equation (4).
1
− ln ε = ln = kP + A (4)
1−D
The compressibility of a powder can be indicated by the values of the Heckle coefficient
(k) and its reciprocal giving yield pressure (Py ). In the Heckle’s model, tablet density,
relative tablet density, compression pressure, and intercept are represented by ε, D, P, and
A, respectively.
5
Pharmaceutics 2023, 15, 1288
100
90
80
% Desupersaturation
70
60
50
40
Homogeneous
30
6nm Silica
20 10nm silica
30nm silica
10
50nm Silica
0
0 25 50 75 100 125 150 175 200 225 250 275 300 325 350
Time (min)
Figure 1. Comparison of % desupersaturation curves of diglycine in the absence and presence of
porous silica at S = 1.20; volume = 40 mL; Tsat = 40 ◦ C; Tcry = 32.7 ◦ C.
The reduction in the induction time in the presence of porous silica has a direct relation
with the pore size. The smallest silica with a pore size of 6 nm exhibited a five-fold reduction
in the induction time compared to that of the homogeneous nucleation, as presented in
Table 1. This was followed by silica with pore sizes of 10 nm, 30 nm, and 50 nm, which had
3-fold, 2-fold, and 1.5-fold reductions, respectively.
Table 1. Average induction time of diglycine in the absence and presence of porous silica at S = 1.20;
volume = 40 mL; Tsat = 40 ◦ C; Tcry = 32.7 ◦ C.
This direct relation of pore size to the reduction in the induction time is potentially
due to the diglycine cluster size matching of pore sizes. DLS data suggest a hydrodynamic
radius of 0.94 nm for diglycine in water, while a unit cell of diglycine contains four diglycine
molecules [13], bringing the diglycine cluster size in the range of silica with a 6 nm pore
size. A pore size (length) of 6 nm is sufficient to sequester a critical size cluster of diglycine,
creating local supersaturation and resulting in the nucleation of diglycine in the pores.
Similar results were obtained by Shah et al. [36] when different-molecular-weight proteins
6
Pharmaceutics 2023, 15, 1288
were crystallised using the templates with engineering pores with an optimum size similar
to their hydrodynamic radius. Further, an antibody named anti-CD20 was crystallised
using the specifically designed porous silica template with a pore size in the range of the
molecular diameter of the antibody [37,38]. Silica with a 10 nm pore size also influenced
the diglycine induction time; 6 nm and 10 nm pores are significantly not different, and
hence are also able to crystallise diglycine and reduce the induction time. Due to the large
pore size of the other pores’ silica, the diglycine induction time was not affected much by
the other pores’ silica compared to the silica with a pore size of 6 nm.
Figure 2 presents a solid-state analysis of the isolated composite solids from the
desupersaturation experiments in the presence of different-pore-size silicas. The PXRD
graphs in Figure 2A show the crystallization of diglycine stable form (α-form) in the
presence of porous silica. There is a preferred orientation observed for the peaks at (100)
and (20–2) when crystallised in the presence of porous silica. This is potentially due to
the non-interaction between the functional groups (-COO, carboxyl group) present on
these surfaces with the silica hydroxyls. Figure 2B presents the SEM micrographs of the
isolated solids along with the bare silica and diglycine. The diglycine crystals obtained in
the presence of porous silica seem to be associated with the silica particles, as observed in
the SEM micrographs, suggesting the crystallization of diglycine either on the silica surface
or the pores. The particle size of the diglycine crystals obtained is in the range of 20–100 μm.
Figure 2. (A) Powder X-ray diffraction spectra of the diglycine–silica composite solids isolated upon
complete desupersaturation of diglycine in the presence of porous silica at S = 1.20, along with the
diglycine patterns; (B) Scanning electron microscopy images of diglycine, porous silica, and isolated
solids after the desupersaturation experiments in the presence of porous silica.
7
Pharmaceutics 2023, 15, 1288
has been extensively studied and reported in the literature, and can be easily compressed
with an excellent tensile strength. Hence, MCC was selected as the excipient to be used
in this study [41]. Due to its easy compressibility, the tensile strength of MCC is >2 MPa
at any compression force, while the tensile strength of the diglycine composite is >2 MPa
after 2 kN, irrespective of the silica pore size. Pharmaceutical tablets exhibiting a tensile
strength greater than 2 MPa are considered as passing the standard requirements for
manufacturability, quality, and biopharmaceutical performance [42]. The tabletability
parameter (Cp ) of the tablets can be obtained using the Newton equation [32]. The Cp
value of 1.61 exhibited by MCC is the highest (reported in Table 2) compared to that of the
diglycine composite, indicating tablets for the diglycine composite are less stable compared
to MCC. This is due to the presence of glycine crystals in the composite tablets.
Figure 3. (Top) Tabletability, (Middle) compactability, and (Bottom) compressibility profiles of the
25% loading blend of diglycine–silica–MCC composite tablets prepared along with MCC alone
(blue squares, PH101). Red circles, blue upwards triangles, green downwards triangles, and pink
rhombuses represent diglycine crystallised using silica with pore sizes of 6 nm, 10 nm, 30 nm, and
50 nm as templates, respectively, with n ≥ 3 (n is the number of experiments).
8
Pharmaceutics 2023, 15, 1288
Table 2. Summary of the mechanical parameters obtained from the curve fitting of tabletability,
compressibility, and compactability curves for diglycine–silica–MCC composites and bare MCC
tablets (Cp is the tabletability parameter, σt0 is the tablet tensile strength at zero porosity, b is an
empirical constant representing bonding capacity, k is the Heckle Coefficient, and Py is the yield
pressure).
Figure 3 (middle) presents a tensile strength vs. porosity graph, representing the
compactability of the tablets. Compactability is the ability of a powder to reduce porosity
under an applied pressure. The curve-fitting parameters obtained (reported in Table 2)
using the Ryshkewitch–Duckworth equation [43] were used to analyse the compactability
of the composite tablets. The tensile strength at zero porosity (σt0 ) is higher for all the
composite samples except for the silica composite with a 50 nm pore size compared to
that of MCC. This suggests that the porosity of the composite tablets is less than that
of MCC. The composite with silica with a 50 nm pore size could potentially have more
agglomerated samples, resulting in a high porosity in the tablets. Further, a higher value of
the constant ‘b’ represents a stronger bonding between the primary particles [34]. The value
of ‘b’ is higher for all the composite samples compared to that of MCC, suggesting that the
interparticle bonding is higher in the composite tablets than the MCC tablets, which is due
to the presence of silica in the composite tablets filling the interparticle distance, as also
observed in the SEM micrographs (Figure 2B) with the silica particles associated with the
diglycine crystals.
The compressibility of the powders was assessed using the Heckle equation. The
values for the Heckle coefficient, k, which were obtained from the slope of the Heckle plot,
and its inverse values, i.e., the yield pressure, (Py ), for the different composite systems
along with that of pure MCC are presented in Table 2. The high value of k and the low
value of Py suggests the good plasticity of MCC. On the contrary, the diglycine composite
exhibit an inverse behaviour to that of pure MCC, with low values for k and higher values
for Py , indicating poor plasticity and low compressibility properties. This is due to the
presence of diglycine crystals, which exhibit poor compressibility, leading to high porosity
in the tablets.
There were no differences observed in the flow properties of the formulations, and this
potentially was due to the presence of 75% w/w microcrystalline cellulose (MCC) in the
formulation mixture. This made the formulation mixture flow more like pure MCC. The
silica tested in the study were mostly identical from outside in terms of their shape, size,
and functionality, while the only difference was in their internal pore diameter. Due to the
identical nature of the silica tested, there was no obvious difference in the flow properties.
Moreover, variations in the particle size distribution of the crystallised diglycine will also
influence the flow properties of the formulation mixture. Particles with a larger crystal size
distribution are more likely to have poor flow properties due to the plate-shaped diglycine
crystals resisting a smooth flow, whereas this may not be the problem with crystals of a
smaller size distribution. However, this was not observed with the crystals in this work, as
all the crystalline particles obtained from the templated crystallization had a similar size
distribution, as can be seen in the SEM micrographs in Figure 2B. A similar phenomenon
was observed by a group member in the past, confirming that elongated needles were
more cohesive than hexagonal crystals, which is due to the combined effects of surface
9
Pharmaceutics 2023, 15, 1288
energy and surface area leading to poor flow behaviour due to the high cohesivity of the
crystals [44].
The diglycine–silica–MCC tablets were subjected to the diffusion test using the dialysis
tube to observe their diffusion rate from the dialysis membrane. Diglycine is readily
dissolved in water based on its solubility data, which have already been published in the
literature [31], and hence, performing the dissolution study is futile. In contrast, the low
intrinsic permeability of peptides is a bottleneck in the oral delivery of peptides. The oral
absorption of peptides in the body occurs through three possible pathways: [18]
i. paracellular, involving passing in between enterocyte cells, only allowing smaller
peptides with a low molecular weight;
ii. transcellular, involving passage through the enterocyte cell with the help of other cells;
iii. ligand-assisted transport, allowing transportation through the intestinal mucosal
membrane using a permeation enhancer.
The molecular weight, flexibility, and hydrophilic nature of smaller peptide-based
drugs provide them the advantage of paracellular passage in between cells [18]. This
phenomenon was later confirmed by Foger et al. [45] in a report that illustrated that the
peptide permeability increases as the molecular weight decreases, but only for peptides
with a molecular weight up to 1.4 kDa.
Figure 4 presents the diffusion of diglycine over a period through the permeation
membrane with a molecular weight cut-off of 6–8 kDa. The diffusion rate experiments
were performed using the diglycine composite tablets compressed at 2.5 kN, as the tablets
had the desired tensile strength of more than 2 MPa at this compression force. All the
tablets exhibited an extended-release profile, confirming their easy diffusion through the
permeation membrane with approximately a 50% diffusion achieved in the first hour and a
complete dissolution in 12–16 h. This steady diffusion suggests that permeation enhancers
are not required for small peptides, as mentioned by Klepach et al. [18].
Figure 4. %diffusion of diglycine from the diglycine, silica, and MCC composite tablet compressed at
2.5 kN. The diffusion medium was water; volume = 100 mL; and stirring = 100 rpm.
10
Pharmaceutics 2023, 15, 1288
potentially not only improve the downstream processing time, but also improve the particle
attributes, such as crystal size, shape, and form, also improving the mechanical properties
of the tablets. However, more work is needed in this area to prove the ability of templates
to crystallise larger peptides and formulate them in oral tablets to study their mechanical
properties.
4. Conclusions
In this study, diglycine, the simplest glycine homopeptide, was crystallised in the
presence of porous silica, presenting a template-assisted crystallization of peptides. The
induction time of diglycine was significantly reduced in the presence of silica with 6 nm
and 10 nm pore sizes due to the size complementarity in the cluster size of diglycine
and the pores. The induction time of diglycine had a direct relationship with the silica
pore size. The induction time increased with the increase in the silica pore size, but the
presence of silica had a positive and discriminating effect. Further, our PXRD graphs
suggested the crystallization of a stable α-form in the presence of porous silica, while all
the crystals were seen to be associated with the silica particles in our SEM micrographs.
The mechanical properties of the diglycine–silica–MCC tablets were accessed through
tabletability, compactability, and compressibility curves. The mechanical properties of the
tablets remained uninfluenced by the presence of diglycine–silica and presented similar
tablet properties as those of MCC alone. Lastly, our diffusion analysis of the diglycine
composite tablets presented an extended release with a constant flux of diglycine through
the membrane. The oral formulation of peptides is not limited to diglycine, and hence more
research is needed in the oral delivery of peptides.
Author Contributions: Conceptualization, V.V.; methodology, V.V., I.B. and V.K.; validation, V.V.,
I.B. and V.K.; formal analysis, V.V., I.B. and V.K.; investigation, V.V., I.B. and V.K.; resources, J.Y.Y.H.;
data curation, V.V., I.B. and V.K.; writing—original draft preparation, V.V.; writing—review and
editing, V.V., I.B. and J.Y.Y.H.; visualization, V.V.; supervision, J.Y.Y.H.; project administration, V.V.;
funding acquisition, J.Y.Y.H. and V.V. All authors have read and agreed to the published version of
the manuscript.
Funding: The results incorporated in this paper have received funding from the European Union’s
Horizon 2020 research and innovation program under the Marie Sklodowska-Curie grant, agreement
no. 101026339. The results were collected at the Centre for Rapid Online Analysis of Reactions at
Imperial College London (EPSRC grant nos. EP/R008825/1 and EP/V029037/1).
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: Not applicable.
Acknowledgments: The authors would like to thank Bed Deadman and Christopher S. Roberts for
their help with the crystallization experiments.
Conflicts of Interest: The authors declare no conflict of interest.
References
1. Wang, L.; Wang, N.; Zhang, W.; Cheng, X.; Yan, Z.; Shao, G.; Wang, X.; Wang, R.; Fu, C. Therapeutic peptides: Current applications
and future directions. Sig. Transduct. Target Ther. 2022, 7, 48. [CrossRef]
2. Muttenthaler, M.; King, G.F.; Adams, D.J.; Alewood, P.F. Trends in peptide drug discovery. Nat. Rev. Drug Discov. 2021, 20,
309–325. [CrossRef] [PubMed]
3. Craik, D.J.; Fairlie, D.P.; Liras, S.; Price, D. The Future of Peptide-based Drugs. Chem. Biol. Drug Des. 2013, 81, 136–147. [CrossRef]
[PubMed]
4. Ferrazzano, L.; Catani, M.; Cavazzini, A.; Martelli, G.; Corbisiero, D.; Cantelmi, P.; Fantoni, T.; Mattellone, A.; De Luca, C.; Felletti,
S.; et al. Sustainability in peptide chemistry: Current synthesis and purification technologies and future challenges. Green Chem.
2022, 24, 975–1020. [CrossRef]
5. Isidro-Llobet, A.; Kenworthy, M.N.; Mukherjee, S.; Kopach, M.E.; Wegner, K.; Gallou, F.; Smith, A.G.; Roschangar, F. Sustainability
Challenges in Peptide Synthesis and Purification: From R&D to Production. J. Org. Chem. 2019, 84, 4615–4628. [CrossRef]
11
Pharmaceutics 2023, 15, 1288
6. Roque, A.C.A.; Pina, A.S.; Azevedo, A.M.; Aires-Barros, R.; Jungbauer, A.; Di Profio, G.; Heng, J.Y.Y.; Haigh, J.; Ottens, M.
Anything but Conventional Chromatography Approaches in Bioseparation. Biotechnol. J. 2020, 15, 1900274. [CrossRef]
7. Guo, M.; Jones, M.J.; Goh, R.; Verma, V.; Guinn, E.; Heng, J.Y.Y. The Effect of Chain Length and Conformation on the Nucleation
of Glycine Homopeptides during the Crystallization Process. Cryst. Growth Des. 2023, 23, 1668–1675. [CrossRef]
8. Guo, M.; Rosbottom, I.; Zhou, L.; Yong, C.W.; Zhou, L.; Yin, Q.; Todorov, I.T.; Errington, E.; Heng, J.Y.Y. Triglycine (GGG) Adopts
a Polyproline II (pPII) Conformation in Its Hydrated Crystal Form: Revealing the Role of Water in Peptide Crystallization. J. Phys.
Chem. Lett. 2021, 12, 8416–8422. [CrossRef]
9. Link, F.J.; Heng, J.Y.Y. Enhancing the crystallisation of insulin using amino acids as soft-templates to control nucleation.
CrystEngComm 2021, 23, 3951–3960. [CrossRef]
10. Chen, W.; Park, S.J.; Kong, F.; Li, X.; Yang, H.; Heng, J.Y.Y. High Protein-Loading Silica Template for Heterogeneous Protein
Crystallization. Cryst. Growth Des. 2020, 20, 866–873. [CrossRef]
11. Li, X.; Heng, J.Y.Y. Protein crystallisation facilitated by silica particles to compensate for the adverse impact from protein
impurities. CrystEngComm 2021, 23, 8386–8391. [CrossRef]
12. Li, X.; Heng, J.Y.Y. The critical role of agitation in moving from preliminary screening results to reproducible batch protein
crystallisation. Chem. Eng. Res. Des. 2021, 173, 81–88. [CrossRef]
13. Verma, V.; Mitchell, H.; Guo, M.; Hodnett, B.K.; Heng, J.Y.Y. Studying the impact of the pre-exponential factor on templated
nucleation. Faraday Discuss. 2022, 235, 199–218. [CrossRef]
14. Verma, V.; Mitchell, H.; Errington, E.; Guo, M.; Heng, J.Y.Y. Templated Crystallization of Glycine Homopeptides: Experimental
and Computational Developments. Chem. Eng. Technol. 2023. [CrossRef]
15. Haddadzadegan, S.; Dorkoosh, F.; Bernkop-Schnürch, A. Oral delivery of therapeutic peptides and proteins: Technology
landscape of lipid-based nanocarriers. Adv. Drug Deliv. Rev. 2022, 182, 114097. [CrossRef]
16. Jenkins, K., II. Needle phobia: A psychological perspective. Br. J. Anaesth. 2014, 113, 4–6. [CrossRef]
17. Karsdal, M.A.; Byrjalsen, I.; Riis, B.J.; Christiansen, C. Optimizing bioavailability of oral administration of small peptides through
pharmacokinetic and pharmacodynamic parameters: The effect of water and timing of meal intake on oral delivery of Salmon
Calcitonin. BMC Clin. Pharmacol. 2008, 8, 5. [CrossRef]
18. Klepach, A.; Tran, H.; Ahmad Mohammed, F.; ElSayed, M.E.H. Characterization and impact of peptide physicochemical properties
on oral and subcutaneous delivery. Adv. Drug. Deliv. Rev. 2022, 186, 114322. [CrossRef]
19. Choonara, B.F.; Choonara, Y.E.; Kumar, P.; Bijukumar, D.; du Toit, L.C.; Pillay, V. A review of advanced oral drug delivery
technologies facilitating the protection and absorption of protein and peptide molecules. Biotechnol. Adv. 2014, 32, 1269–1282.
[CrossRef]
20. Maher, S.; Brayden, D.J. Formulation strategies to improve the efficacy of intestinal permeation enhancers. Adv. Drug Deliv. Rev.
2021, 177, 113925. [CrossRef]
21. Maher, S.; Ryan, B.; Duffy, A.; Brayden, D.J. Formulation strategies to improve oral peptide delivery. Pharm. Pat. Anal. 2014, 3,
313–336. [CrossRef]
22. Zhu, Q.; Chen, Z.; Paul, P.K.; Lu, Y.; Wu, W.; Qi, J. Oral delivery of proteins and peptides: Challenges, status quo and future
perspectives. Acta Pharm. Sin. B 2021, 11, 2416–2448. [CrossRef]
23. Anselmo, A.C.; Gokarn, Y.; Mitragotri, S. Non-invasive delivery strategies for biologics. Nat. Rev. Drug Discov. 2019, 18, 19–40.
[CrossRef]
24. Harrison, G.A. Insulin in Alcoholic Solution by the Mouth. Br. Med. J. 1923, 2, 1204. [CrossRef]
25. Basu, S.K.; Govardhan, C.P.; Jung, C.W.; Margolin, A.L. Protein crystals for the delivery of biopharmaceuticals. Expert Opin. Biol.
Ther. 2004, 4, 301–317. [CrossRef]
26. Aguirre, T.A.S.; Teijeiro-Osorio, D.; Rosa, M.; Coulter, I.S.; Alonso, M.J.; Brayden, D.J. Current status of selected oral peptide
technologies in advanced preclinical development and in clinical trials. Adv. Drug Deliv. Rev. 2016, 106, 223–241. [CrossRef]
27. Drucker, D.J. Advances in oral peptide therapeutics. Nat. Rev. Drug Discov. 2020, 19, 277–289. [CrossRef]
28. Lipinski, C.A.; Lombardo, F.; Dominy, B.W.; Feeney, P.J. Experimental and computational approaches to estimate solubility and
permeability in drug discovery and development settings. Adv. Drug Deliv. Rev. 2001, 46, 3–26. [CrossRef]
29. Brayden, D.J.; Hill, T.A.; Fairlie, D.P.; Maher, S.; Mrsny, R.J. Systemic delivery of peptides by the oral route: Formulation and
medicinal chemistry approaches. Adv. Drug Deliv. Rev. 2020, 157, 2–36. [CrossRef]
30. Verma, V.; Zeglinski, J.; Hudson, S.; Davern, P.; Hodnett, B.K. Dependence of Heterogeneous Nucleation on Hydrogen Bonding
Lifetime and Complementarity. Cryst. Growth Des. 2018, 18, 7158–7172. [CrossRef]
31. Guo, M.; Chang, Z.H.; Liang, E.; Mitchell, H.; Zhou, L.; Yin, Q.; Guinn, E.J.; Heng, J.Y.Y. The effect of chain length and side chains
on the solubility of peptides in water from 278.15 K to 313.15 K: A case study in glycine homopeptides and dipeptides. J. Mol. Liq.
2022, 352, 118681. [CrossRef]
32. Newton, J.M.; Rowley, G.; Fell, J.T.; Peacock, D.G.; Ridgway, K. Computer analysis of the relation between tablet strength and
compaction pressure. J. Pharm. Pharmacol. 1971, 23, 195S–201S. [CrossRef] [PubMed]
33. Ryshkewitch, E. Compression Strength of Porous Sintered Alumina and Zirconia. J. Am. Ceram. Soc. 1953, 36, 65–68. [CrossRef]
34. Steendam, R.; Lerk, C.F. Poly(dl-lactic acid) as a direct compression excipient in controlled release tablets: Part I. Compaction
behaviour and release characteristics of poly(dl-lactic acid) matrix tablets. Int. J. Pharm. 1998, 175, 33–46. [CrossRef]
12
Pharmaceutics 2023, 15, 1288
35. de Andrade, D.F.; Zuglianello, C.; Pohlmann, A.R.; Guterres, S.S.; Beck, R.C.R. Assessing the In Vitro Drug Release from
Lipid-Core Nanocapsules: A New Strategy Combining Dialysis Sac and a Continuous-Flow System. AAPS PharmSciTech 2015, 16,
1409–1417. [CrossRef]
36. Shah, U.V.; Williams, D.R.; Heng, J.Y.Y. Selective Crystallization of Proteins Using Engineered Nanonucleants. Cryst. Growth Des.
2012, 12, 1362–1369. [CrossRef]
37. Yang, H.; Belviso, B.D.; Li, X.; Chen, W.; Mastropietro, T.F.; Di Profio, G.; Caliandro, R.; Heng, J.Y.Y. Optimization of Vapor
Diffusion Conditions for Anti-CD20 Crystallization and Scale-Up to Meso Batch. Crystals 2019, 9, 230. [CrossRef]
38. Gerard, C.J.J.; Briuglia, M.L.; Rajoub, N.; Mastropietro, T.F.; Chen, W.; Heng, J.Y.Y.; Di Profio, G.; ter Horst, J.H. Template-Assisted
Crystallization Behavior in Stirred Solutions of the Monoclonal Antibody Anti-CD20: Probability Distributions of Induction
Times. Cryst. Growth Des. 2022, 22, 3637–3645. [CrossRef]
39. Michrafy, A.; Michrafy, M.; Kadiri, M.S.; Dodds, J.A. Predictions of tensile strength of binary tablets using linear and power law
mixing rules. Int. J. Pharm. 2007, 333, 118–126. [CrossRef]
40. Wu, C.-Y.; Best, S.M.; Bentham, A.C.; Hancock, B.C.; Bonfield, W. A simple predictive model for the tensile strength of binary
tablets. Eur. J. Pharm. Sci. 2005, 25, 331–336. [CrossRef]
41. Pudasaini, N.; Upadhyay, P.P.; Parker, C.R.; Hagen, S.U.; Bond, A.D.; Rantanen, J. Downstream Processability of Crystal
Habit-Modified Active Pharmaceutical Ingredient. Org. Process Res. Dev. 2017, 21, 571–577. [CrossRef]
42. Chen, H.; Aburub, A.; Sun, C.C. Direct Compression Tablet Containing 99% Active Ingredient&-A Tale of Spherical Crystallization.
J. Pharm. Sci. 2019, 108, 1396–1400. [CrossRef] [PubMed]
43. Barralet, J.E.; Gaunt, T.; Wright, A.J.; Gibson, I.R.; Knowles, J.C. Effect of porosity reduction by compaction on compressive
strength and microstructure of calcium phosphate cement. J. Biomed. Mater. Res. 2002, 63, 1–9. [CrossRef] [PubMed]
44. Shah, U.V.; Olusanmi, D.; Narang, A.S.; Hussain, M.A.; Gamble, J.F.; Tobyn, M.J.; Heng, J.Y.Y. Effect of crystal habits on the
surface energy and cohesion of crystalline powders. Int. J. Pharm. 2014, 472, 140–147. [CrossRef] [PubMed]
45. Föger, F.; Kopf, A.; Loretz, B.; Albrecht, K.; Bernkop-Schnürch, A. Correlation of in vitro and in vivo models for the oral absorption
of peptide drugs. Amino Acids 2008, 35, 233–241. [CrossRef] [PubMed]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual
author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to
people or property resulting from any ideas, methods, instructions or products referred to in the content.
13
pharmaceutics
Article
Numerical Simulation of Particle Dynamics in a Spiral Jet Mill
via Coupled CFD-DEM
Satyajeet Bhonsale 1 , Lewis Scott 2 , Mojtaba Ghadiri 2 and Jan Van Impe 1, *
Abstract: Spiral jet mills are ubiquitous in the pharmaceutical industry. Breakage and classification in
spiral jet mills occur due to complex interactions between the fluid and the solid phases. The study of
these interactions requires the use of computational fluid dynamics (CFD) for the fluid phase coupled
with discrete element models (DEM) for the particle phase. In this study, we investigate particle
dynamics in a 50-mm spiral jet mill through coupled CFD-DEM simulations. The simulations showed
that the fluid was significantly decelerated by the presence of the particles in the milling chamber.
Furthermore, we study the particle dynamics and collision statistics at two different operating
conditions and three different particle loadings. As expected, the particle velocity was affected by
both the particle loading and operating pressure. The particles moved slower at low pressures and
high loadings. We also found that particle–particle collisions outnumbered particle–wall collisions.
Keywords: spiral jet mills; discrete element models; computational fluid dynamics
Citation: Bhonsale, S.; Scott, L.;
Ghadiri, M.; Van Impe, J. Numerical
Simulation of Particle Dynamics in a
Spiral Jet Mill via Coupled CFD-DEM. 1. Introduction
Pharmaceutics 2021, 13, 937. https:// Particle size reduction is an important step in the design, development, and processing
doi.org/10.3390/pharmaceutics
of active pharmaceutical ingredients (API). Spiral jet mills are the preferred comminution
13070937
devices for ultra-fine grinding where particles less than 10 μm diameter are desired [1,2].
Since spiral jet mills were first patented in the 1930s [3], their design has remained relatively
Academic Editor: Anne Marie Healy
unchanged. Their design consists of a short cylindrical (or elliptical) milling chamber into
which high velocity gas is pushed through several nozzles (called the grinding nozzles),
Received: 19 May 2021
Accepted: 19 June 2021
which are at an angle to the mill perimeter. The gas jets entering through these nozzles
Published: 23 June 2021
create a vortex in the milling chamber.
Solid feed particles are fed to an injector, which delivers the feed to the vortex, wherein
Publisher’s Note: MDPI stays neutral
they are accelerated by the gas flow. The momentum gathered by the particles due to the
with regard to jurisdictional claims in
high velocity gas jets leads to high energy particle–particle and particle–wall collisions,
published maps and institutional affil- which cause breakage. The centrifugal forces in the vortex retain the coarse particles
iations. within the milling chamber. The centrifugal forces and the radial drag forces acting on the
particles in the vortex depend on the particle size (x). As the size decreases due to breakage,
centrifugal force (∼ x3 ) reduces faster than the radial drag forces (∼ x2 ).
When the radial drag force acting on a particle exceeds the centrifugal force, the par-
ticle is entrained out of the milling chamber via an outlet in the centre of the mill. Al-
Copyright: © 2021 by the authors.
Licensee MDPI, Basel, Switzerland.
though the energy consumption of spiral jet mills is relatively high, they provide various
This article is an open access article
advantages. Due to the absence of any moving parts and the self-classifying nature of the
distributed under the terms and mill, contamination can be completely avoided [4]. Moreover, the expansion of fluid from
conditions of the Creative Commons the jet into the grinding chamber leads to a cooling effect, which makes spiral jet mills
Attribution (CC BY) license (https:// attractive for heat sensitive materials [5].
creativecommons.org/licenses/by/ The popularity of the jet mills arises from the simplicity of its operation. In general,
4.0/). only three parameters are needed to control the operation of the jet mills: the injector nozzle
pressure (IP), the grinding nozzle pressure (GP), and the solid feed rate (FR). Of these,
the GP and FR have a significant impact on the milling performance, while the impact
of IP is not large [6]. The kinetic energy within the jet mill is directly related to the GP.
Higher kinetic energy leads to enhanced acceleration of the particles and hence higher
impact collisions.
A large number of studies assessing the influence of the GP (or in some cases the
gas flow rate) on milling performance arrived at the same conclusion: increasing the GP
leads to a decrease in the output median particle size [7–11]. As the IP has a negligible
impact on the mill performance, it is usually maintained at a pressure that is slightly
higher than the GP. Increasing the FR makes the output particle size coarser. Although the
frequency of collisions increases with increasing FR [12], the high particle concentration
within the mill causes the fluid energy in the mill to dissipate faster [13]. This leads to low
energy collisions.
A systematic mathematical description of the spiral jet mill is desirable for predictive
purposes. A validated model can reduce the need for extensive experimentation with
an expensive API. Moreover, a model-based process understanding is also a key element
of the Quality by Design paradigm adopted by the Food and Drugs Authority in the
United States. Despite the prevalence of spiral jet mills in the pharmaceutical industry,
modelling studies are relatively sparse. Some of these studies rely on the force and energy
balance approach [14–16], some on the population balance approach [17–19], and a few on
computational fluid dynamics [15,20,21].
The energy and force balances equate the two opposing forces acting on the particles:
the centrifugal force and the radial drag force. Based on the balance, a particle cut size that
depends on the ratio of tangential and radial particle velocities (called the spin number)
is derived. Under idealistic assumptions of Archimedes spiral flow, Tanaka [16] derived
expressions that defined the tangential and radial velocity in a mill as a function of the mill
parameters. Rodnianski et al. [15], on the other hand, used CFD simulations to obtain the
spin number and radial velocity. They described the spin number as a general function
of mill’s geometric and operational parameters. An important conclusion from the CFD
analysis of Rodnianski et al. [15] is that the gas flow rate does not affect the spin number.
MacDonald et al. [14] built upon the previous results and derived a cut size equation
by incorporating the energy balance. In all the above derivations, the particle tangential
velocity was assumed to be the same as the gas velocity.
Population balance modelling (PBM) has also been used to describe breakage is
described with empirical breakage distribution functions. As the fully described PBM is a
complex integro-differential set of equations, numerical methods are commonly utilized
to obtain a solution [22]. Gommeren et al. [17] presented a compartmentalized PBM
describing three zones within the grinding chamber: the comminution zone, central (feed)
zone, and the classifier zone. The model was then used to determine the residence time
distribution, hold up, and closed-loop control. However, no discussion on the estimation
of the parameters involved in the model was provided.
An highly empirical steady-state PBM was also described in Starkey et al. [18]. How-
ever, the algebraic equation set only considered six discrete size classes to describe the
particle size distribution of the product. A major drawback in the PBM approach is the
need for estimating the breakage parameters from experimental data. Although a variety
of approaches have been proposed for this inverse problem, they cannot be applied directly
to the spiral jet mill models. Bhonsale et al. [23] performed an identifiability analysis of
a discretized spiral jet mill PBM, and showed that the convolution between classification
and breakage in the jet mill led to non-identifiable parameters in the breakage kernels.
Computational fluid dynamics (CFD) relies on the numerical solution of the Navier–
Stokes equations to resolve the fluid flow field within the mill. The CFD simulations of
Kozawa et al. [20] showed that coarse particles near the upper wall could escape the mill
easily. Similarly, Rodnianski et al. [15] reported the invariability of the spin number with
gas mass flow rates. However, as the operation of the jet mill involves a complex interplay
15
Pharmaceutics 2021, 13, 937
between the fluid and particle phases, simulations solely via CFD can be misleading. When
the influence of particle phase cannot be ignored completely, a coupled CFD—Discrete
Element Method (DEM) approach needs to be adopted. The DEM approach solves the
Newtonian kinematic equations for individual particles to determine their trajectories.
Given the computational limitations, the CFD-DEM approach cannot be used for very fine
particles undergoing breakage. Thus, its application to modelling spiral jet mills requires
simplifying assumptions.
Han et al. [24] reported the influence of the feed rate, feed nozzle angle, and the gas
flow rate on the product particle size based on two dimensional CFD-DEM simulations.
Levy and Kalman [21] presented three-dimensional simulations of particle motion in an
industrial scale jet mill. Although particle breakage and particle–particle interactions are
completely ignored, the simulations provide interesting insights into the flow field in the
jet mill. Teng et al. [12] included particle–particle interactions but ignored particle breakage.
With simulations involving only 1000 particles, they reported the influence of the GP on
the particle velocity distribution. Along with an increase in the particle velocity, increasing
the GP also led to an increased width of the particle velocity distribution.
Moreover, particle–particle collisions were shown to be the primary cause of breakage.
By identifying the collision patterns, they concluded that the majority of collisions had a
much larger tangential component, which would lead to abrasion rather than fragmen-
tation. Brosh et al. [25] adopted the breakage model developed by Kalman et al. [26] to
incorporate comminution in the simulations. To avoid an excessive number of particles in
the simulation, particles that fell below 10 μm were removed from the simulation.
Bnà et al. [27] presented a thorough CFD-DEM simulation study. The cut size deter-
mined by their simulations is in good agreement with the previous studies by Dobson and
Rothwell [4]. They also highlighted the influence of the product hold up inside the mill.
They concluded that the fluid deceleration caused by the presence of the particle phase was
responsible for the classification efficiency of the mill. However, only a one-way coupling
between CFD-DEM was used. In such a coupling approach, the effect of the fluid phase on
the particulate phase was considered; however, the effect of the particulate phase on the
fluid was ignored.
Given the importance of hold up, Bnà et al. [27] recognized this limitation and under-
lined the need for a four-way coupling between CFD and DEM. Scott et al. [28] presented
such a simulation study using a four-way CFD-DEM coupling. Their simulation reported a
decrease in the tangential velocity component with increasing hold up. They also showed
that most energy was dissipated along the bed surface and in front of each jet.
Although the use of CFD-DEM models for predictive purposes is limited by computa-
tional restrictions, they provide useful insight into the particle dynamics of a mill. In this
paper, a coupled CFD-DEM simulation is used to analyse the particle dynamics in the
spiral jet mill. Unlike the coupling used by Bnà et al. [27], the coupling used here considers
both the influence of fluid on particles and particles on fluid. In the subsequent sections,
the CFD-DEM approach is described, followed by the simulation results and conclusions.
2. Numerical Methods
The mill geometry used for the simulations is based on the Hosokawa AS50 spiral jet
mill. However, the geometry is based on an in-house drawing made at the University of
Leeds and shown in Figure 1 [28,29]. The milling chamber is 50 mm in diameter and has
four jets angled at 50◦ from the radius. A special feature of the AS50 is its classifier design.
The milling gas spirals up into the classifier section where a vortex finder reverses the
flow direction. Figure 1a shows the CAD geometry of the mill used. Following Dogbe [29],
an annular manifold for gas distribution was included around the milling chamber as this
influences the fluid flow field within the chamber.
The numerical simulation of the process proceeds in three stages. The fluid field is
resolved by CFD (using ANSYS Fluent v19), the particle phase is resolved by DEM (using
EDEM 2019), and the coupling responsible for exchanging information on solid–fluid forces
16
Pharmaceutics 2021, 13, 937
is achieved by EDEM’s coupling tool. For the fluid field resolution, the gas is assumed to
behave ideally. The k-ω shear stress transport (SST) [30] was used as the turbulence model.
For its simplicity, the Morsi–Alexander correlation [31] was used to compute the fluid drag
on the particles.
There are several other drag laws available in the literature, and the choice influences
the results of the simulation. Although an evaluation of drag laws is out of the scope
of this paper, a few other studies have made the comparisons in the context of coupled
CFD-DEM [32–34]. The CFD simulations were carried out using the commercial software—
ANSYS Fluent v19 with the first-order upwind approach used to solve the individual
equations. The convergence tolerances were set at 10−4 for all the equations. All the
boundary conditions were set to pressure type boundaries. For the ‘feed hopper inlet’ and
the ‘mill outlet’, atmospheric conditions were assumed.
The ‘injector inlet’ and ‘milling inlet’ were set at the desired operational conditions.
Two operating conditions were considered: 1 bar IP and GP, and 3 bar IP and GP (based on
gauge pressure). A tetrahedral mesh was used with its size determined by the particle size
was used in the DEM simulation. The mesh size was constrained to be 40% larger than the
particle diameter [35]. A mesh convergence study was performed by recording the velocity
gradient across the milling chamber. Following Norouzi et al. [35], the time step for the
CFD simulations was set to 50 times the DEM time step (i.e., at 1 × 10−5 s).
Figure 1. The isometric view (a) of the CAD geometry used for CFD-DEM simulations of the jet mill
[28,29]. The red section in the top (b) and front (c) view of the geometry depicts the particle factory in
which the initial particle bed is generated. The green section in the front view (c) depicts the particle
factory through which particles are dynamically fed once the initial particle bed is dispersed.
For the DEM simulations, the particles were considered to be monosized perfect
spheres of 200 μm diameter. The particle properties used in the simulations are listed in
Table 1. The CFD-DEM coupling was handled by EDEM’s coupling tool, which is based on
the approach described by Tsuji et al. [36]. The mass of particles in the cell was decomposed
into a weighted volume so that the pressure calculation could be performed. The respective
velocity was then returned to EDEM to update the drag force acting on each individual
17
Pharmaceutics 2021, 13, 937
particle. The Hertz–Midlin model was used to model the contact forces. The integration
time step was fixed to 20% of the Rayleigh timestep. For the particle size and particle
properties used in the simulation, the time step used was 2 × 10−7 s.
To evaluate the effect of the hold up, three different particle loadings were consid-
ered: 10,000 particles (≈0.06 g), 40,000 particles (≈0.25 g), and 100,000 particles (≈0.63
g). In reality, the particles dropped into the feed hopper inlet are sucked into the milling
chamber by the injector gas flow. To avoid extensive simulation times, a particle bed was
pre-generated in a ring shape factory placed inside the milling chamber as depicted in red
in Figure 1b. Once the particle bed is dispersed by the flow and has reached a steady state,
10,000 particles are fed to the mill via the injector factory (coloured green in Figure 1c) at a
feed rate of 1.8 kg/h. It takes about 0.128 s to finish feeding the particles. The particle and
collision data were collected for 0.15 s from the start of the feeding phase.
3. Results
3.1. Fluid and Particle Dynamics
Figure 2 illustrates the velocity magnitude within the milling chamber at the mid plane.
These contours are plotted once the pre-generated particle bed is completely dispersed.
Although the fluid velocity reaches a magnitude of around 300 m/s, the contour plot
is clipped at 150 m/s to emphasize the lower velocity areas. The areas with low fluid
velocities are coloured blue, and high fluid velocities are coloured red. Once the pre-
generated particle bed is dispersed, particles form a bed on the mill periphery along which
they circulate. For all three loadings, the periphery of the mill along which the particles
circulate had the lowest velocity.
This particle bed is locally dispersed by the high velocity jet streams emitting from
the nozzles. As higher particle loadings lead to a thicker circulating particle bed, the low
velocity region around the wall increases in size. Moreover, the jets dispersing the particle
bed loose energy much faster when the particle bed is thick. Thus, the length of the jet
stream reduces with increasing particle load. This also leads to much lower fluid velocities
in the entire mill. This decrease in jet penetration length was also reported by Scott et al.
[28] for an even higher particle loading. It is evident that higher particle loadings lead to a
larger dampening of the fluid velocity. The velocity magnitude increases radially toward
the centre in all cases.
Figure 2. The velocity magnitude contours at the midplane for three particle loadings.
18
Pharmaceutics 2021, 13, 937
Snapshots of particle motion in the jet mill for the case with 100,000 particles and 3 bar
pressure are presented in Figure 3. At the time of 0.0 s, the pre-generated particle bed is
intact, and the CFD-DEM simulation is started. The particles start dispersing as the large
tangential component of the velocity accelerates them towards the wall. A steady state
is reached around 0.015 s, and a thick layer of particles is formed at the mill periphery,
which moves along the wall at a very low velocity. Meanwhile, the jet streams disperse
the particles when they approach the nozzles. Once the steady state has been achieved,
10,000 new particles are fed through a factory created in the injector nozzle at rate of
1.8 kg/h, taking roughly 0.128 s.
Figure 3. Particle motion in the jet mill for 100,000 particles and 3 bar pressure.
The evolution of the fluid field as the particles disperse is illustrated in Figure 4. As can
be deduced from the blue ring in the contour plot at time 0.00035 s, the pre-generated
particle bed slows the fluid notably. The particle bed, shown by the blue area, is pushed to
the wall as it is dispersed. Once the steady state is reached, the velocity profile does not
vary greatly. Even when new particles are fed, the velocity profile stays consistent. Thus, it
19
Pharmaceutics 2021, 13, 937
can be said that the particles being entrained into the milling chamber from the feeder do
not affect the fluid field in the mill. However, over time, as the particles build up in the mill,
the fluid behaviour will be affected. This is evident from Figure 2. The presence of a particle
bed circulating along the wall periphery has been reported in previous experimental and
numerical studies [6,12,28].
Figure 4. Evolution of fluid field in the jet mill along the mid plane for 100,000 particles with an
operating pressure of 3 bar.
Figure 5 displays heat maps of the particle velocities. The figure maps every particle
for 0.15 s from the time particle feeding starts. All the particles are coloured according to
their velocity at the given time. The prominence of the slow moving particle bed on the
mill periphery is evident for all three particle loadings. Similar to the observations from
20
Pharmaceutics 2021, 13, 937
the fluid fields, the increase in the thickness of this bed with particle loading is clear. In all
cases, particles with the highest velocity lie along the jet trajectory.
Figure 5. Particle velocity heat map for 0.15 s from the start of the feeding.
Figure 6a,b illustrate the effect of particle loading on the particle velocity distribution.
For both operating pressures, the particle velocity distribution shifts to the left with in-
creasing particle loading. The average and maximum particle velocities also decrease with
increasing particle loading. At low particle loadings, the mean free path (distance travelled
without colliding) of a particle is much larger. Thus, the particle can be accelerated over a
larger distance. Moreover, more fluid kinetic energy per unit particle mass is available for
the acceleration.
0.25 45
1 104
4 10 4 40
0.2 1 105
35
Particle Velocity (m/s)
30
Probability (-)
0.15 25
20
0.1
15
10
0.05
5
0
0
0 10 20 30 40 50 1 10 4 4 10 4 1 10 5
Particle Velocity (m/s) Particle Loading
(a) Particle velocity distribution for 3 (b) Box plot for 3 bar and three particle
bar and three particle loadings loadings
0.25
1 bar
3 bar 30
0.2
25
Particle Velocity (m/s)
Probability (-)
0.15 20
15
0.1
10
0.05 5
0
0
0 10 20 30 40 1 bar 3 bar
Particle Velocity (m/s) Operating Pressure
(c) Particle velocity distribution for (d) Box plot for 100,000 particles and
100,000 particles and two operating two operating pressures
pressures
Figure 6. Probability distribution and box plot for particle velocity with increasing particle loading
and two operating pressures
21
Pharmaceutics 2021, 13, 937
The effect of operating pressure can be discerned from Figure 6c,d wherein the particle
velocity distribution and box plot are depicted for 100,000 particles at the two operating
pressures. Increasing the operating pressure leads to a slight rightwards shift in the particle
velocity distribution while a much larger increase is noticed in the average and maximum
particle velocity. The increase in the velocity distribution is due to the much higher kinetic
energy provided by the fluid at higher operating pressures.
The motion of the particles fed into the jet mill via the injector is illustrated by the
streamlines plotted in Figure 7. The image on the left depicts the three particle loadings at
3 bar operating pressure, and the image on the right depicts the streamlines for the two
operating pressures with 100,000 particles. With higher particle loading, the particles are
ejected from the bed by the jet stream move more towards the centre of the mill. Similarly,
the particles travel closer to the classifier at higher operating pressure. In all cases, no
particles escape the mill via the classifier.
As Bnà et al. [27] indicated, particle classification is a function of the fluid deceleration
caused by the particle phase. At the current particle concentrations, very small particles
could be entrained. The cut size can be derived by a force balance between the centrifugal
and radial forces [4,14–16]. However, the particle size used in the current study is much
larger. Thus, no classification was observed from the simulations. Even with around
250,000 particles (some as small as 160 μm), Scott et al. [28] could not observe classification
in their four-way coupled simulations. This highlights the importance of the solid hold up
within the mill on the final particle size of the milled and classified product.
Figure 7. Streamlines of first five particles fed to the mill while describing their motion in the mill for
0.15 s after the feeding. Case (a) depicts the streamlines for three particle loadings at 3 bar operating
pressure, and case (b) depicts the streamlines at two operating conditions and particle loading of
100,000 particles.
22
Pharmaceutics 2021, 13, 937
the total number of collisions for those cases as a percentage of the maximum number of
collisions observed across all the case studies. In general, increased particle loading led to
an increase in the number of collisions.
In all the cases considered, particle–particle collisions were prevalent. At low particle
loading, a significant fraction of the collisions occurred between the particle and the wall.
The effect of the operating pressure on the number of collisions was prevalent at high
loadings. For 100,000 particles, 1 bar operating pressure led to only around 60% of the
collisions as 3 bar pressure. For 10,000 particles, increased pressure also led to a slight
increase in the fraction of particle–wall collisions.
Figure 8. The number of particle–particle (P–P) and particle–wall (P–W) collisions as a percentage of
the total collisions for three loadings and two operating pressures. The total number of collisions as a
normalized percent of the maximum collisions across all conditions. Collision data collected over
0.15 s after the start of particle feeding.
Figure 9 reports the number of collisions per particle for increasing particle loading
and the two operating pressures. Again, increased pressure and increased loading both led
to an increase in the number of collisions experienced by a particle.
50
1 bar
3 bar
40
#Collisions/Particle
30
20
10
0
1 104 4 104 1 105
Number of Particles
Figure 9. The total number of collisions per particle for the three loadings and two operating
pressures over 0.15 s after the start of the feeding phase.
The distribution of collision velocities is illustrated in Figure 10. For both particle–
particle and particle–wall collisions, increasing particle load leads to a reduction in the
impact velocities. This can be explained by the fact that, at lower particle concentrations,
23
Pharmaceutics 2021, 13, 937
each particle travels on a longer mean free path, thus, accelerating to higher velocities before
the collision. The high impact collisions at low loading and high operating pressure will
lead to a higher degree of breakage. This is corroborated by experimental studies that showed
that finer product was obtained at low feed rates and high operating pressures [7,8,11,12].
0.7 0.3
1 bar; 1 104 Particles 1 bar; 1 104 Particles
4
0.6 1 bar; 4 10 Particles 1 bar; 4 104 Particles
5
1 bar; 1 10 Particles
0.25 1 bar; 1 105 Particles
3 bar; 1 104 Particles 3 bar; 1 104 Particles
0.5 4
3 bar; 4 10 Particles 0.2 3 bar; 4 104 Particles
Probability (-)
Probability (-)
5
3 bar; 1 10 Particles 3 bar; 1 105 Particles
0.4
0.15
0.3
0.1
0.2
0.05
0.1
0 0
10-1 100 101 102 10-1 100 101 102
Collision Velocity (m/s) Collision Velocity (m/s)
(a) Particle Particle Collisons (b) Particle Wall Collision
Figure 10. Collision velocity distribution for particle–particle (P–P) and particle–wall (P–W) collision for all cases. Collision
data collected over 0.15 s after the start of particle feeding.
As the study by Bnà et al. [27] involved only a one-way coupling in which particles
mostly followed the fluid streamlines, they reported particle–wall collision velocities of
around 40–100 m/s. These were much higher than the collision velocities observed in
this study. Moreover, in contrast to the findings in this study, they reported no significant
effect of particle loading on the particle–wall collision velocity distribution. This shows the
importance of the fluid deceleration by particle phase for breakage as well as classification.
Figure 11 shows the scatter plots of the tangential component against the normal com-
ponent of the particle–particle and particle wall collision velocities. As already mentioned,
in all cases, the particle–particle collisions are prevalent. Increasing the particle loading
led to a decrease in the velocity magnitudes of both the components while increasing
the operating pressure led to an increase. At higher particle loadings, the particle–wall
collisions were characterized by a high tangential component, while the particle–particle
collisions had a much higher normal component.
This contradicts the findings of both Bnà et al. [27] and Teng et al. [12]. Both studies
reported a higher tangential component for all collisions. As mentioned, the study by
Bnà et al. [27] used only a one-way coupling scheme and, thus, ignored the fluid deceler-
ation by the particle phase. The study by Teng et al. [12], although a two-way coupling,
only considered 1000 particles. Such a low particle concentration cannot decelerate fluid to
the extent observed in the current study.
24
Pharmaceutics 2021, 13, 937
Figure 11. Scatter of the tangential component and normal component of the collision velocity for all
cases studied. Collision data collected over 0.15 s after the start of particle feeding.
4. Conclusions
We analysed the effect of particle loading and the operating pressure on the fluid
and particle dynamics in a spiral jet mill through coupled CFD-DEM simulations. Three
particle loadings and two operating pressures were considered in the study.
We found that the particles significantly decelerated the fluid. The particles dispersed
from the pre-generated bed formed a new bed along the mill periphery along which they
moved at very slow velocities. The reduction in the fluid velocities was at the maximum
in this area. The velocity profile increased monotonically along the radius from the wall
toward the mill centre. The particle loading and operating pressure had a much higher
impact on the tangential velocity than on the radial velocity. At increasing particle loading
and decreasing pressure, the particles experienced more radial forces, which led to the
entrainment of coarser particles.
The particle velocity also followed the same trend as the fluid velocity. The particles
in the bed moving along the mill periphery were ejected with force by the jet streams.
The highest particle velocities were observed directly in front of the jets. Higher operating
pressure and low loading led to higher particle velocities and, subsequently, higher impact
velocities. In all cases, substantially more particle–particle collisions were reported com-
pared with particle–wall collisions. Most high impact collisions occurred at the surface of
the circulating particle bed.
The results show that the particle loading had a profound effect on the fluid field,
which, in turn, influenced both the breakage and classification. The computational method
and the results presented provide a valuable tool-process optimisation for industrial appli-
cations of spiral jet mills.
Author Contributions: Conceptualization, S.B., L.S., M.G. and J.V.I.; Data curation, S.B.; Formal
analysis, S.B. and L.S.; Investigation, S.B. and L.S.; Methodology, S.B., L.S., M.G. and J.V.I.; Project
administration, M.G. and J.V.I.; Software, S.B., L.S. and M.G.; Supervision, M.G. and J.V.I.; Validation,
S.B., L.S., M.G. and J.V.I.; Visualization, S.B. and L.S.; Writing—original draft, S.B.; Writing—review
& editing, S.B., L.S., M.G. and J.V.I. All authors have read and agreed to the published version of
the manuscript.
25
Pharmaceutics 2021, 13, 937
References
1. Nakach, M.; Authelin, J.R.; Chamayou, A.; Dodds, J. Comparison of various milling technologies for grinding pharmaceutical
powders. Int. J. Miner. Process. 2004, 74, S173–S181. [CrossRef]
2. Parrott, E.L. Milling of Pharmaceutical Solids. J. Pharm. Sci. 1974, 63, 813–829. [CrossRef] [PubMed]
3. Andrews, N. Method and Apparatus for Providing Material in Finely Divided Form. U.S. Patent US002032827, 3 March 1936.
4. Dobson, B.; Rothwell, E. Particle size reduction in a fluid energy mill. Powder Technol. 1969, 3, 213–217. [CrossRef]
5. Liu, G. Hammer Milling and Jet Milling Fundamentals. Chem. Eng. Prog. 2017, 113, 48–54.
6. Teng, S.; Wang, P.; Zhu, L.; Young, M.W.; Gogos, C.G. Experimental and numerical analysis of a lab-scale fluid energy mill.
Powder Technol. 2009, 195, 31–39. [CrossRef]
7. Katz, A.; Kalman, H. Preliminary Experimental Analysis of a Spiral Jet Mill Performance. Part. Part. Syst. Charact. 2007,
24, 332–338. [CrossRef]
8. Midoux, N.; Hošek, P.; Pailleres, L.; Authelin, J. Micronization of pharmaceutical substances in a spiral jet mill. Powder Technol.
1999, 104, 113–120. [CrossRef]
9. Müller, F.; Polke, R.; Schädel, G. Spiral jet mills: Hold up and scale up. Int. J. Miner. Process. 1996, 44–45, 315–326. [CrossRef]
10. Ramanujam, M.; Venkateswarlu, D. Studies in fluid energy grinding. Powder Technol. 1969, 3, 92–101. [CrossRef]
11. Tuunila, R.; Nyström, L. Effects of grinding parameters on product fineness in jet mill grinding. Miner. Eng. 1998, 11, 1089–1094.
[CrossRef]
12. Teng, S.; Wang, P.; Zhang, Q.; Gogos, C. Analysis of Fluid Energy Mill by gas-solid two-phase flow simulation. Powder Technol.
2011, 208, 684–693. [CrossRef]
13. Luczak, B.; Müller, R.; Kessel, C.; Ulbricht, M.; Schultz, H.J. Visualization of flow conditions inside spiral jet mills with different
nozzle numbers—Analysis of unloaded and loaded mills and correlation with grinding performance. Powder Technol. 2019,
342, 108–117. [CrossRef]
14. MacDonald, R.; Rowe, D.; Martin, E.; Gorringe, L. The spiral jet mill cut size equation. Powder Technol. 2016, 299, 26–40. [CrossRef]
15. Rodnianski, V.; Krakauer, N.; Darwesh, K.; Levy, A.; Kalman, H.; Peyron, I.; Ricard, F. Aerodynamic classification in a spiral jet
mill. Powder Technol. 2013, 243, 110–119. [CrossRef]
16. Tanaka, T. Scale-Up Theory of Jet Mills on Basis of Comminution Kinetics. Ind. Eng. Chem. Process Des. Dev. 1972, 11, 238–241.
[CrossRef]
17. Gommeren, H.; Heitzmann, D.; Kramer, H.; Heiskanen, K.; Scarlett, B. Dynamic modeling of a closed loop jet mill. Int. J. Miner.
Process. 1996, 44-45, 497–506. [CrossRef]
18. Starkey, D.; Taylor, C.; Morgan, N.; Winston, K.; Svoronos, S.; Mecholsky, J.; Powers, K.; Iacocca, R. Modeling of continuous
self-classifying spiral jet mills part 1: Model structure and validation using mill experiments. AIChE J. 2014, 60, 4086–4095.
[CrossRef]
19. Starkey, D.; Taylor, C.; Siddabathuni, S.; Parikh, J.; Svoronos, S.; Mecholsky, J.; Powers, K.; Iacocca, R. Modeling of continuous self-
classifying spiral jet mills part 2: Powder-dependent parameters from characterization experiments. AIChE J. 2014, 60, 4096–4103.
[CrossRef]
20. Kozawa, K.; Seto, T.; Otani, Y. Development of a spiral-flow jet mill with improved classification performance. Adv. Powder
Technol. 2012, 23, 601–606. [CrossRef]
21. Levy, A.; Kalman, H. Numerical Study of Particle Motion in Jet Milling. Part. Sci. Technol. 2007, 25, 197–204. [CrossRef]
22. Bhonsale, S.S.; Telen, D.; Stokbroekx, B.; Impe, J.V. Comparison of numerical solution strategies for population balance model of
continuous cone mill. Powder Technol. 2019, 345, 739–749. [CrossRef]
23. Bhonsale, S.S.; Stokbroekx, B.; Van Impe, J. Assessment of the parameter identifiability of population balance models for air jet
mills. Comput. Chem. Eng. 2020, 143, 107056. [CrossRef]
24. Han, T.; Kalman, H.; Levy, A. DEM Simulation of Particle Comminutionin Jet Milling. Part. Sci. Technol. 2002, 20, 325–340.
[CrossRef]
25. Brosh, T.; Kalman, H.; Levy, A.; Peyron, I.; Ricard, F. DEM–CFD simulation of particle comminution in jet-mill. Powder Technol.
2014, 257, 104–112. [CrossRef]
26. Kalman, H.; Rodnianski, V.; Haim, M. A new method to implement comminution functions into DEM simulation of a size
reduction system due to particle–wall collisions. Granul. Matter 2009, 11, 253–266. [CrossRef]
26
Pharmaceutics 2021, 13, 937
27. Bnà, S.; Ponzini, R.; Cestari, M.; Cavazzoni, C.; Cottini, C.; Benassi, A. Investigation of particle dynamics and classification
mechanism in a spiral jet mill through computational fluid dynamics and discrete element methods. Powder Technol. 2020,
364, 746–773. [CrossRef]
28. Scott, L.; Borissova, A.; Burns, A.; Ghadiri, M. Influence of holdup on gas and particle flow patterns in a spiral jet mill. Powder
Technol. 2021, 377, 233–243. [CrossRef]
29. Dogbe, S. Predictive Milling of Active Pharmaceutical Ingredients and Excipients. Ph.D. Thesis, School of Chemical and Process
Engineering, The University of Leeds, Leeds, UK, 2016.
30. Menter, F.R. Two-equation eddy-viscosity turbulence models for engineering applications. AIAA J. 1994, 32, 1598–1605. [CrossRef]
31. Morsi, S.A.; Alexander, A.J. An investigation of particle trajectories in two-phase flow systems. J. Fluid Mech. 1972, 55, 193–208.
[CrossRef]
32. Agrawal, V.; Shinde, Y.; Shah, M.T.; Utikar, R.P.; Pareek, V.K.; Joshi, J.B. Effect of drag models on CFD–DEM predictions of
bubbling fluidized beds with Geldart D particles. Adv. Powder Technol. 2018, 29, 2658–2669. [CrossRef]
33. Marchelli, F.; Hou, Q.; Bosio, B.; Arato, E.; Yu, A. Comparison of different drag models in CFD-DEM simulations of spouted beds.
Powder Technol. 2020, 360, 1253–1270. [CrossRef]
34. Zhou, L.; Zhang, L.; Bai, L.; Shi, W.; Li, W.; Wang, C.; Agarwal, R. Experimental study and transient CFD/DEM simulation in a
fluidized bed based on different drag models. RSC Adv. 2017, 7, 12764–12774. [CrossRef]
35. Norouzi, H.R.; Zarghami, R.; Sotudeh-Gharebagh, R.; Mostoufi, N. Coupled CFD-DEM Modeling: Formulation, Implementation and
Application to Multiphase Flows; Wiley: Newark, NJ, USA, 2016.
36. Tsuji, Y.; Tanaka, T.; Ishida, T. Lagrangian numerical simulation of plug flow of cohesionless particles in a horizontal pipe. Powder
Technol. 1992, 71, 239–250. [CrossRef]
37. Dogbe, S.; Ghadiri, M.; Hassanpour, A.; Hare, C.; Wilson, D.; Storey, R.; Crosley, I. Fluid-particle energy transfer in spiral jet
milling. EPJ Web Conf. 2017, 140, 09040. [CrossRef]
38. Rodnianski, V.; Levy, A.; Kalman, H. A new method for simulation of comminution process in jet mills. Powder Technol. 2019,
343, 867–879. [CrossRef]
27
pharmaceutics
Article
Numerical Investigation of the Particle Dynamics in a
Rotorgranulator Depending on the Properties of the
Coating Liquid
Philipp Grohn 1, *, Stefan Heinrich 2 and Sergiy Antonyuk 1
Abstract: In the pharmaceutical industry, the coating of particles is a widely used technique to obtain
desired surface modifications of the final product, e.g., controlled release of the active agents. The
production of round, coated particles is particularly important, which is why fluidized bed rotor
granulators (FBRG) are often used for this process. In this work, Computational Fluid Dynamics
(CFD) coupled with the Discrete Element Method (DEM) is used to investigate the wet particle
dynamics, depending on the properties of the coating liquid in a FBRG. The DEM contact model was
extended by liquid bridge model to account for capillary and viscous forces during wet contact of
particles. The influence of the relative contact velocity on the maximum length of the liquid bridge is
also considered in the model. Five different cases were compared, in which the particles were initially
wetted, and the liquid loading as well as the surface tension and viscosity of the liquid were changed.
The results show that increasing viscosity leads to a denser particle bed and a significant decrease
in particle rotational velocities and particle motion in the poloidal plane of the FBRG. Reducing the
liquid loading and surface tension results in increased particle movement.
Keywords: CFD-DEM simulation; wet particles; capillary force; viscous force; fluidized bed
Citation: Grohn, P.; Heinrich, S.;
rotor granulator
Antonyuk, S. Numerical
Investigation of the Particle
Dynamics in a Rotorgranulator
Depending on the Properties of the
Coating Liquid. Pharmaceutics 2023,
1. Introduction
15, 469. https://doi.org/10.3390/ For various products in the pharmaceutical, chemical and food industries, the coating
pharmaceutics15020469 of particles is an important processing step in order to obtain desired surface modification
of the final product [1,2]. Numerous coating equipment exists for this purpose. The coating
Academic Editor: Colin Hare
devices can be distinguished according to their method of introducing kinetic energy into a
Received: 14 December 2022 particle bed, between a purely mechanical input (e.g., mixer, disc and drum granulators)
Revised: 20 January 2023 and a fluidization induced by the energy of the process gas flow (e.g., fluidized bed or
Accepted: 27 January 2023 spouted bed systems). Particularly in the pharmaceutical industry, fluidized bed rotor
Published: 31 January 2023 granulators (FBRG) are widely used to produce round coated pellets for oral drug delivery
with a narrow size distribution, high strength, smooth surface and high sphericity [3–7].
This is achieved by the special design of a FBRG. It consists of a rotating circular rotating
base plate and a stationary cylindrical wall. The fluidization gas flow passes through an
Copyright: © 2023 by the authors.
annular gap between the rotating plate and the cylindrical wall. This combination enables
Licensee MDPI, Basel, Switzerland.
the individual process steps of spheronization, coating and drying to be carried out in the
This article is an open access article
distributed under the terms and
same unit [8].
conditions of the Creative Commons
Although the technology of a FBRG is widely used, the particle dynamics are still not
Attribution (CC BY) license (https:// fully understood due to the complex micro mechanisms in the process. Several experimental
creativecommons.org/licenses/by/ studies can be found in the literature that describe some mechanisms during the granulation
4.0/). process [4,6,9–11]. However, the knowledge in this field is mainly empirical and all the
particle interactions are not yet fully understood. A detailed knowledge of the particle
motion on the micro level is required to better understand the coating process in the rotor
granulator. Numerical simulations are particularly suitable for this purpose [5,8]. The
widely used Euler–Lagrange approach can be applied to simulate the multiphase flow,
where Computational Fluid Dynamics (CFD) is coupled with the Discrete Element Method
(DEM) [12,13]. In CFD, the flow field of the gas in the process is calculated treating the fluid
phase as a continuum. For the DEM, the interactions of each particle are determined based
on contact models describing the physical properties of the particles, such as adhesion, and
their mechanical behavior under slow, fast and repeated loading. In two-way CFD-DEM
coupling, both the influence of the gas phase on the particle phase and the influence of the
particle phase on the gas phase are considered [13–17], while in one-way coupling, only
the influence of the gas phase on the particles is taken into account [5].
The particle dynamics in a rotor granulator were first investigated by Muguruma et al. [18]
numerically with DEM and experimentally with Particle Tracking Velocimetry (PTV). They
studied the influence of liquid on particle motion, but considered only capillary and not
viscous contact forces, and did not vary the properties of the liquid. Weis et al. [19,20]
used DEM simulations to obtain the particle dynamics and mixing behavior, as well as the
contact frequency of the particles in a spheronizer that, in contrast to a rotor granulator,
works without fluidization air and usually at higher rotation velocities of the structured
friction plate. In addition, they extended the DEM approach to consider particle rounding
during this process. Recently, Grohn et al. [21] investigated numerically the multiphase
flow of cylindrical particles in a FBRG with CFD-DEM simulations. A significant influence
of the particle shape on the particle dynamics was found. Neuwirth et al. [4,22] performed
an experimental study of the particle dynamics in a FBRG under dry and wet conditions
using magnetic particle tracking (MPT). The comparison with the CFD-DEM simulations
showed good agreement with the experiments for the dry case. However, the experiments
were performed with particles of 6-mm diameter, which are not representative for real
applications in FBRG. This particle size was required by the MPT measurement equipment
available at that time. In our last study [8], the dynamics of initially wetted particles in
the FBRG was investigated numerically by CFD-DEM simulations and experimentally
by an improved MPT measurement system. In the numerical simulations, the capillary
forces due to the presence of liquid on particles were considered based on the model
of Israelachvili [23] and the viscous forces were calculated according to the models of
Lian et al. [24] and Popov [25]. In addition, a new model was implemented to describe
the velocity-dependent rupture length of liquid bridges [8,26]. With the improved MPT
equipment, the particle dynamics of spherical particles with a minimum diameter of 2.8 mm
could be measured. It was possible to validate contact models used in simulations of dry
particles and particles wetted with water, and a good agreement was found.
Since in real applications, both the liquid spray rate and thus, the liquid loading of the
particles and the properties of the coating solution vary, the influence of these parameters
on particle dynamics and contact behavior in the FBRG are investigated in this work using
the model previously validated in [8]. On the one hand, the influence of the liquid loading
of 1 vol.-% and 5 vol.-% with water is investigated. On the other hand, the liquid properties
are varied three times at a constant liquid loading of 5 vol.-%. The basis of this liquid is
a coating solution frequently used in the pharmaceutical industry, consisting of distilled
water with 6 mass-% PHARMACOAT® 606 (hydroxypropyl methylcellulose, Shin-Etsu
Chemical Co., Ltd., Chiyoda-ku, Tokyo, Japan) [5]. This coating solution is characterized
by a reduced surface tension of 42.5 mN·m−1 compared to water and a strongly increased
viscosity of 61.9 mPa·s. The three other variants thus result from: a reduction of the surface
tension to the value of the coating solution while the viscosity of water remains unchanged,
the surface tension of water remains unchanged but the viscosity is increased to the value
of the coating solution, and both the surface tension and the viscosity are changed to the
values of the coating solution. To analyze the influence of the studied liquid parameters
on the particle dynamics, the distributions of solid volume fraction, tangential, poloidal
29
Pharmaceutics 2023, 15, 469
and rotating particle velocities are compared. For a deeper understanding of the process,
the particle contact phenomena, such as the resulting aggregate size, are investigated
with DEM.
2. Model Description
2.1. CFD Modeling
In the CFD, the gas flow field is calculated by solving the volume-averaged Navier–
Stokes equations [27]. For this purpose, the flow domain for the CFD simulation must first
be discretized by mesh cells. In order to take the influence of the particulate phase on the
gas flow into account, the volume fraction of the gas phase ε g in each mesh cell is included
in the volume-averaged Navier–Stokes equation. Therefore, the governing equations of the
mass and momentum conversation can be given as follows:
∂ ε g ρg →
+ ∇· ε g ρg u g = 0 , (1)
∂t
→
∂ ε g ρg u g → →
→ → →
+ ∇· ε g ρg u g u g = −ε g ∇ p + ∇· ε g τ g − S p + ε g ρg g , (2)
∂t
→ →
where, p represents the pressure, ρg describes the density, u g and τ g are the velocity and
the stress tensor of the gas phase, respectively. For the calculation of the volume fraction of
the gas phase ε g in each CFD mesh cell, the volume fraction xi of each particle volume Vp,i
within the cell were determined using the so-called sample points approach. The principle
of this method, where the volume of all particles z in a grid cell with the volume Vcell is
approximated by cubic sample volumes, was first presented by Hoomans et al. [28]. In the
used framework of CFDEM® coupling [29], the particle is divided into 29 non-overlapping
regions of equal volume, each with one sample point [30]. At each time step, the algorithm
checks which of the sample cubes are located in which mesh cell:
z
1
ε g = 1 − ∑ xi Vp,i . (3)
i =0
Vcell
To consider the interactions between the particulate phase and the gas phase, the
→
momentum balance is extended by the momentum sink term S p . The momentum sink
→
term can be determined from the drag forces F d,i of all particles n p in the mesh cell with
the volume Vcell :
→ np →
1
Vcell i∑
Sp = · F d,i . (4)
=0
Various gas–solid models can be found in the literature that describe the drag forces
acting on the particles in a fluidized bed [12,31]. As in our previous work [8], the drag forces
are calculated according to the widely used model of Di Felice [32], which describes the
entire porosity range and for particle Reynold numbers Rep,i up to 106 with a continuous
function. Here, the drag force counteracts the relative velocity of a particle in a fluid
→ →
( u g − u p ):
→ 1 → →
→ → 2− β
F d,i = CD,i Rep ρg πd2p,i u g − u p u g − u p ε g . (5)
8
The drag force is considerably influenced by the drag coefficient CD,i . This coefficient
is related to the Reynolds number of the particles, which takes into account the superficial
velocity differences between particles and the surrounding fluid [33,34]:
2
4.8
CD,i = 0.63 + , (6)
Rep,i
30
Pharmaceutics 2023, 15, 469
→ →
ε g ρg dp,i u g − u p
Rep,i = , (7)
ηf
where, ηf is the dynamic viscosity of the fluid. The influence of the particle concentration
in a mesh cell on the drag force in Equation (5) is modeled with a function [32]:
2
1.5 − log10 Rep,i
β = 3.7 − 0.65 exp − . (8)
2
k →
→
dω p,i →
Jp,i = ∑ Mt,ij + Mr,ij . (10)
dt j=0
→ →
The gravitational force F g,i and the sum of the contact forces F c,ij , which act on the
particle due to interactions with other particles j or the walls, are modeled to determine
→
the translational velocity of each particle v p,i with the mass mp,i . Similar to our latest
work [8], the particles are initially wetted. Therefore, viscous forces Fvis,ij and capillary
forces Fcap,ij act during a particle contact. Both forces are described in detail in the following
→
Section 2.2.2. The sum of the torques Mt,ij caused by the tangential forces acting on the
→
particle and the torques Mr,ij due to rolling friction if the particle rotates are calculated to
→
determine the angular velocity ω p,i of each particle with the moment of inertia Jp,i .
ηn = 2α m∗ kn δn1/4 , (13)
31
Pharmaceutics 2023, 15, 469
ηt = 2α m∗ kt δt1/4 , (14)
where, α represents a function of the restitution coefficient, and m∗ is the reduced mass of
the contact partners. A more detailed description can be found in Heinrich et al. [17] and
Salikov et al. [37].
−8πR∗ γcos(θ )
Fcap,pw = −1 , (16)
Vb
1+ 1 + πR∗ h2 − 1
where, γ represents the surface tension, θ is the wetting angle, Vb describes the volume of
the liquid bridge and h is the shortest distance between the particles or the particle and the
wall. R∗ represents the effective contact radius, which is expressed as:
ri rj
R∗ = , (17)
ri + rj
where, ri and rj are the radii of the two contact partners. The assumption is made that the
liquid on the particles forms a uniform thin film over the particle surface. During a wet
particle contact, a liquid bridge is formed between the contact partners in the rebound
phase. Shi and McCarthy’s [41] distribution model is used to determine the liquid volume
of these liquid bridges for particle–particle contacts. The distribution model ensures that
the liquid on the particle surface contributes to only one liquid bridge (Figure 1a). However,
this approach is only valid for monodisperse systems. The liquid volume Vb,i that particle
i contributes to the liquid bridge is then calculated as follows:
⎛ ⎞
Li ⎝ rj2
Vb,i = · 1 − 1 − 2 ⎠ , (18)
2 ri + rj
where, Li is the total liquid volume present on particle i. The contributed liquid volume
from particle j is determined in a similar manner:
⎛ ⎞
Lj ri2
Vb,j = ·⎝1 − 1 − 2 ⎠ . (19)
2 ri + rj
32
Pharmaceutics 2023, 15, 469
Figure 1. Determination of (a) the volume of liquid used to form the liquid bridge according to the
model of Shi and McCarthy [41] and (b) the volume of liquid passing from a wetted grid cell of the
wall to the liquid bridge.
Figure 1b shows the contact case between a particle and a wetted wall. The volume
of liquid contributed by the wall depends on the virtual liquid layer thickness hwall on
the surface grid cell of the wall geometry in contact. The layer thickness is calculated by
the liquid volume associated with the wetted wall grid cell divided by its surface area.
Often, the 2D surface grid cells of the geometry are of different sizes and often much
larger than the particle surfaces. Therefore, it is assumed that only the liquid in the region
corresponding to the projection area of the contacting particle needs to be considered. In
Figure 1b, this area is marked with a red circle. Thus, the amount of liquid in the wall grid
cell that contributes to the formation of the liquid bridge is expressed as:
Vb,j = hwall πri2 . (20)
The final volume of the liquid bridge is then the sum of both contributed liquid volumes:
Vb = Vb,i + Vb,j . (21)
As the particles rebound, the liquid bridge is stretched until it ruptures at a critical
distance between the contact partners. This critical distance, also called the maximum liquid
bridge length, is described by various models [38,39,42–45]. All models have in common
that they do not consider the significant influence of the impact velocity on the bridge
length, which was found in our recent experimental study [26]. In this work, three different
experimental setups were developed to investigate the maximum liquid bridge length in a
velocity range from 0.0001 s·m−1 to 4 s·m−1 for particle–particle as well as particle–wall
contact. Based on our experimental results, we extended the model of Mikami et al. [39] to
account for the strong influence of the impact velocity uim,ij on the maximum liquid bridge
length. For particle–wall contact, the maximum bridge length was expressed as:
2
lmax,pw = (0.95 + 0.22θ )Vb0.32 1 + Cpw uim,ij 3 , (22)
where, Cpw represents a constant parameter with was found in [26] to have a value
of 4.424 s·m−1 . For particle–particle contact, the maximum liquid bridge length was
calculated as:
2
lmax,pp = (0.99 + 0.62θ )Vb0.34 1 + Cpp uim,ij 3 , (23)
where, Cpw is a constant parameter with the value of 6.266 s·m−1 . The end of the contact
is indicated by the rupture of the liquid bridge. The volume of the liquid bridge is then
distributed evenly among the contact partners.
In addition to the capillary forces, viscous forces are also considered. They slow down
the contact velocity during the approach phase as well as the velocity of the rebound phase
after contact. Based on the Reynolds lubrication theory [46], Adams and Perchard [47]
33
Pharmaceutics 2023, 15, 469
developed a model to describe the viscous force in the normal direction, which is often
used in DEM simulations [24,41,48,49]. In the model, two particles are assumed to be in a
liquid layer and move with a relative velocity in the normal direction. The viscous forces
in normal direction Fvis,n between the particles can then be calculated by the Reynolds
lubrication equation:
6πηl R∗ 2 up,ij,n
Fvis,n = , (24)
h
where, ηl is the dynamic viscosity of the liquid, up,ij,n represents the relative velocity of
the colliding particles or particle with a wall in normal direction and h is the shortest
distance between the surfaces of the contact partners. Similar to DEM studies from other
authors [48,50,51], a minimum distance between the contact partners is set for the calcu-
lation of the viscous forces as it is physically limited by the roughness of the respective
surfaces. The viscous force in tangential direction is calculated according to the model of
Popov [25]. It describes the tangential force acting on a spherical particle moving along a
plate wetted with a liquid film:
∗ R∗
Fvis,t = 2πηf R up,ij,t ln 1 + , (25)
2h
where, up,ij,t describes the relative velocity of the colliding particles or particle with a wall
in tangential direction.
3. Simulation Setup
3.1. Geometry of the Fluidized Bed Rotor Granulator
In this study, a FBRG is investigated, whose dimensions are inspired by the commer-
cially used rotor granulator Rotor 300 (Glatt GmbH, Binzen, Germany). In Figure 2, the
geometry of the apparatus is shown. The diameter of the cylindrical process chamber is
295 mm; thus, the radius RFBRG is 147.5 mm (Figure 2b). FBRG has an unstructured rotating
plate with a diameter of 268 mm located in the middle of the apparatus. In addition, the
gas flows vertically into the particle bed via a two-millimeter-wide annular gap between
the rotor plate and the apparatus wall. Due to the small inflow surface of the annular
gap, the gas enters the process chamber with a much higher velocity than in conventional
fluidized beds. The direction of the air flow in the apparatus is shown by blue arrows and
the rotation of the plate is represented by green arrows. In real applications, an additional
nozzle is often placed above the plate to coat the particles, but in the simulated cases in this
study, only initially wetted particles are examined.
Figure 2. (a) Dimensions of the fluidized bed rotor granulator (blue arrows indicate the air flow and
the green arrows the direction of rotation of the plate) and (b) its STL 3D geometry for the simulations.
34
Pharmaceutics 2023, 15, 469
The particulate phase was calculated with DEM using the open-source software
LIGGGHTS® [55] and coupled with the CFD by the open-source software CFDEM® cou-
pling [29]. Similar to previous studies, round particles with a diameter of 2.8 mm consisting
of a ceramic core and a shell of polyvinyl butyral (PVB) were investigated [5,8,26,56]. Ini-
tially wetted particles with a total mass of 1 kg were generated above the rotor plate. It
was assumed that the liquid on the particles was evenly distributed on the particle surface
with a layer of equal thickness. Similar to our previous work [5,8], different setups were
used to obtain the particle properties needed for the contact model in DEM. With a free-fall
device [26,56], the restitution coefficient was determined. A Nanoindenter (Hysitron TI Pre-
mier, Bruker Corporation, Billerica, Massachusetts, USA) was used to measure the Young’s
modulus, and with a Texture Analyser® (TA.XTplus, Stable Micro Systems, Godalming,
United Kingdom), the static as well as the rolling friction coefficients were obtained [34].
The contact angle of water was determined with a camera setup and evaluated by a MAT-
LAB script [26]. The DEM time-step was 1 × 10−7 s. The initial liquid loading of the
particles in the bed is varied from 1 vol.-% to 5 vol.-% distilled water. In addition, the
surface tension and viscosity of the liquid are varied three times at 5 vol.-%. The basis is
a coating solution frequently used in the pharmaceutical industry, consisting of distilled
water with 6 mass-% PHARMACOAT® 606 (hydroxypropyl methylcellulose, Shin-Etsu
Chemical Co., Ltd., Chiyoda-ku, Tokyo, Japan) [8]. This coating solution is characterized
by a reduced surface tension of 42.5 mN·m−1 compared to water and a strongly increased
viscosity of 61.9 mPa·s. The three other cases thus result from a reduction in surface tension
with no change in the viscosity of the coating liquid, no change in the surface tension of
the coating liquid but an increase in viscosity, and both the change in surface tension and
viscosity to the values of the coating solution. The DEM parameters can be seen in Table 2
and the performed simulation cases are listed in Table 3.
35
Pharmaceutics 2023, 15, 469
Table 3. The five simulation cases with different liquid loading and properties at a fluidization flow
of 200 m3 ·h−1 , a rotation velocity of the rotor plate of 100 rpm and a bed mass of 1 kg.
4. Results
The particle dynamics and contact behavior in the fluidized bed rotor granulator
obtained with CFD-DEM simulations for the five cases with different liquid loading, as well
as different liquid viscosity and surface tension, are compared in the following sections.
36
Pharmaceutics 2023, 15, 469
the case of the increased viscosity at constant surface tension of water (Figure 3d) and is
therefore the densest. However, the particle concentration is highest at high axial positions.
The reason is that the particles adhere to the wall over time and, unlike in the other cases,
very rarely come off. Therefore, the concentration here increases over time. It can be clearly
seen that with high liquid loading, surface tension and viscosity, and thus, high liquid
bridge forces, the particle bed becomes denser.
Figure 3. Poloidal distribution of solid volume fraction for studied cases (Table 3): (a) liquid loading of
1 vol.-%, (b) liquid loading of 5 vol.-%, (c) surface tension of 42.5 mN·m−1 , (d) viscosity of 61.9 mPa·s,
and (e) surface tension of 42.5 mN·m−1 and viscosity of 61.9 mPa·s.
37
Pharmaceutics 2023, 15, 469
input from the rotor to the bed is lower and the mean velocity of the particles in the bed
decreases slightly. This leads to a small difference between the three cases.
Figure 4. Differential distributions of the absolute particle velocity for a liquid loading of 1 vol.-%, a
liquid loading of 5 vol.-%, a surface tension of 42.5 mN·m−1 , a viscosity of 61.9 mPa·s, and a surface
tension of 42.5 mN·m−1 and a viscosity of 61.9 mPa·s.
The increased viscosity of 61.9 mPa·s in the fourth case (curve (d)) again leads to a
lower slip, slightly increasing the mean particle velocity by 2.3%, compared to 5 vol.-%
water. It can be seen that the proportion of velocities between 0.5 m·s−1 and 0.6 m·s−1
increases significantly. In the last case (curve (e)), the positive effect that the increased
viscosity has on the slip of the particles on the rotation plate is counteracted by the negative
effect of the lower surface tension. As a result, the average particle velocity only changes
by less than 1%.
The tangential velocity distribution in the poloidal plane as a function of radial and
axial position at a different liquid loading and at different liquid properties is shown in
Figure 5. Although in all five cases, the acting liquid bridge forces differ, their velocity
profiles are quite similar. The highest tangential velocities can be seen in the region directly
above the rotating plate caused by transfer of momentum into the particle bed. Due to
liquid bridges and, therefore, acting adhesive forces, there is a significant reduction in
particle velocity near the wall in all cases investigated, as the particles repeatedly adhere
to the wall. Thus, the particles are strongly decelerated and have a tangential velocity
of 0 m·s−1 directly at the wall. It can be seen that with an increase in the liquid bridge
forces in cases (b), (d) and (e), the zone with low tangential velocities near to the stationary
apparatus wall decreases. The particles are more strongly connected to each other, which
improves the energy input through the rotor plate to the entire particle bed.
More pronounced differences between the five cases can be seen in the poloidal velocity
distribution (Figure 6). The poloidal velocity is the velocity component composed of the
z-velocity and the radial velocity [8]. It can be clearly seen that the poloidal velocities of
the particles are significantly lower than their tangential velocities. The direction of particle
motion in the poloidal plane is evident from the velocity vectors. Due to the rotation of
the plate and the axial acceleration above the annular gap caused by the fluidization air,
the particles obtain a circular movement in the poloidal plane of the particle bed. The
highest poloidal velocities can be observed near the wall directly above the annular gap
due to the high inflow velocity of the fluidization air. In addition, the particles have a high
poloidal velocity at the surface of the particle bed, where the particles fall down by gravity.
In contrast to the center of the particle bed, as well as near the apparatus wall, the particles
move very slowly. The particles in the two cases with increased viscosity (Figure 6d,e) have
the lowest poloidal velocities. The increased viscous forces, due to the higher viscosity of
the liquid, lead to higher energy dissipation, and as a result, the poloidal velocity of the
particles decreases. For the first and third case (Figure 7a,c), it can be seen that the poloidal
particle velocities are slightly higher compared to the case with 5 vol.-% water (Figure 7b).
38
Pharmaceutics 2023, 15, 469
Here, the lower capillary forces due to the smaller amount of liquid or the lower surface
tension are responsible for the reduced energy dissipation.
Figure 5. Poloidal distribution of tangential particle velocity at (a) liquid loading of 1 vol.-%,
(b) liquid loading of 5 vol.-%, (c) surface tension of 42.5 mN·m−1 , (d) viscosity of 61.9 mPa·s, and
(e) surface tension of 42.5 mN·m−1 and viscosity of 61.9 mPa·s.
Figure 6. Poloidal distribution of poloidal particle velocity at (a) liquid loading of 1 vol.-%, (b) liquid
loading of 5 vol.-%, (c) surface tension of 42.5 mN·m−1 , (d) viscosity of 61.9 mPa·s, and (e) surface
tension of 42.5 mN·m−1 and viscosity of 61.9 mPa·s.
39
Pharmaceutics 2023, 15, 469
Figure 7. Poloidal distribution of rotational particle velocity at (a) liquid loading of 1 vol.-%,
(b) liquid loading of 5 vol.-%, (c) surface tension of 42.5 mN·m−1 , (d) viscosity of 61.9 mPa·s, and
(e) surface tension of 42.5 mN·m−1 and viscosity of 61.9 mPa·s.
For a useful description of the particle dynamics, the particle rotation number (PRN)
and the radial movement proportion (RMP) can be calculated. Both parameters were
developed in our previous study [8]. The particle rotation number is defined as the number
of 360-degree rotations of the all particles around the central vertical axis of the process
chamber per second. If, in a later case, a nozzle is installed in the wall of the process
chamber to coat the particles, this key number describes how many times per second the
particles pass the wet zone of the nozzle. The second number RMP describes the proportion
of the radial velocity to the total velocity of the particles in the xy-plane. The higher this
value, the more the kinetic energy of the particles goes into their radial motion in the bed.
The values of the PRN and RMP for all simulated cases are given in Table 4. Due to the
increased slip at lower liquid loading and lower surface tension (cases (a) and (c), the PRN
decreases by 3.9% compared to the case (b) with 5 vol.-% of water. In contrast, it increases
by 14.5% with increased viscosity (cases (d) and (e). All five cases have a significantly lower
rotational speed than the rotor plate, which rotates at 1.67 s−1 . This is mainly due to the fact
that it is an unstructured plate, where the energy transfer is not as good as with structured
plates [57]. Looking at the RMP, it is clear that a reduction in liquid loading leads to an
increase in radial motion. This is even more pronounced for the case with reduced surface
tension. In both cases, the lower capillary forces compared to 5 vol.-% water lead to a
greater freedom of movement of the particles and thus to an increased poloidal velocity. As
already seen in Figure 6, the velocities are lower in the poloidal plane when the viscosity is
increased to 61.9 mPa·s. The reason is due to the significant increase in viscous forces, the
particles are slowed down more during contacts. As a consequence, the RMP also decreases
more significantly. A lower surface tension in case e) with 6 mass-% PHARMACOAT®
40
Pharmaceutics 2023, 15, 469
606 solution, and thus lower capillary forces, lead to reduced energy dissipation, which
again slightly increases the RMP compared to the fourth case.
Table 4. The particle rotation number and the radial movement proportion for the five
simulation cases.
41
Pharmaceutics 2023, 15, 469
Table 5. Contact rate and average number of contact partners for the five simulation cases.
Average Numbers of
Case Contact Rates/-
Contact Partners/-
(a) 1.0 vol.-% 2.14 3.23
(b) 5.0 vol.-% 2.96 3.45
(c) γ = 42.5 mN·m−1 2.21 3.34
(d) η = 61.9 mPa·s 2.56 3.76
(e) γ = 42.5 mN·m−1 , η = 61.9 mPa·s 2.24 3.43
For a more detailed analysis of the aggregates formed during the process, the differ-
ential distributions of the number of simultaneous contact partners for all five cases are
shown in Figure 8. While the distributions at 1 vol.-% water (Figure 8a) and a reduced
surface tension (Figure 8c) differ only slightly from the case with 5 vol.-% water, significant
differences can be seen between the two cases with increased viscosity. In the fourth case
(Figure 8d), the proportion of aggregates consisting of two or three particles decreases by
3.7%, while the proportion of aggregates consisting of more than seven particles increases
by 4.8%. In the last case studied (Figure 8e), the proportion of aggregates consisting of two
or three particles increases by 4.1%, and the proportion of aggregates of more than nine
particles increases slightly as well. In all cases, the large aggregates form in the upper region
of the particle bed, where the poloidal and tangential particle velocities are lowest. Since
the poloidal velocities are lowest in the cases with increased viscosity, the large aggregates
can exist here for the longest time before the particles separate from each other again due
to shear forces.
Figure 8. Differential distribution of the number of simultaneous contact partners for a (a) liquid
loading of 1 vol.-%, (b) liquid loading of 5 vol.-%, (c) surface tension of 42.5 mN·m−1 , (d) viscosity of
61.9 mPa·s, and (e) surface tension of 42.5 mN·m−1 and viscosity of 61.9 mPa·s.
42
Pharmaceutics 2023, 15, 469
Figure 9. Differential distribution of the contact velocities for the five simulation cases.
Table 6 summarizes all average contact velocities with variation of liquid loading and
liquid properties. As could already be seen from the differential distribution of the contact
velocities (Figure 9), with increasing coating liquid viscosity, the fraction of low contact
velocities grows. This can also be seen when looking at the average contact velocities. The
highest contact velocities occur during interactions of the particles with the rotating rotor
plate. In wet particle contacts, both contact forces due to viscoelastic deformation and
contact forces due to liquid bridges are existent. In the FBRG, however, the contacts in
which only a liquid bridge force acts predominate. Table 6 shows that the mean particle–
particle contact forces, mainly contributed by liquid bridge forces, increase in both normal
and tangential directions for the highly viscous cases (d) and (e). In the normal direction,
the contact force increases by 14.3% for the fourth case (d) and by 4.1% for the fifth case (e)
compared to the case (b) with 5 vol.-% water. When comparing the tangential force for
particle–particle contacts, an even more significant increase of 76.9% in the fourth case (d)
and 65.4% in the fifth case (e) is noticeable. The reason is that the viscous force increases
proportionally with the viscosity of the liquid (Equations (24) and (25)). Since the capillary
force is also proportionally dependent on the surface tension (Equations (15) and (16)), the
mean contact force decreases when the surface tension is reduced. The average particle–
particle contact forces in the normal and tangential directions in the case (a) with reduced
liquid loading and in the case (c) with lower surface tension are smaller than for 5 vol.-%
water (case (b)). The ratio of the normal to the tangential contact force changes significantly
for the last two cases (d) and (e) and decreases. This influence is most clearly seen for
the particle–rotor contacts. Thus, the high viscosity with simultaneously reduced surface
tension of the 6 mass-% PHARMACOAT® 606 solution (case (e)) leads to a ratio of normal
43
Pharmaceutics 2023, 15, 469
to tangential contact force of less than one. The explanation for this is the lower capillary
forces acting only in the normal direction, with a simultaneous sharp increase in the viscous
forces in the tangential direction. The results for case (a) with a reduced liquid loading
and case (c) with a reduced surface tension differ only slightly from the second case with
5 vol.-% water.
Table 6. Average contact velocities and forces for the five simulation cases in contacts between
particles (P–P), particles with the cylindrical wall (P–W) and particles with rotor plate (P–R).
5. Conclusions
In this work, the dynamics of wet particles in a fluidized bed rotor granulator was
investigated using CFD-DEM simulation. A liquid bridge model was implemented in DEM
to account for the acting physical adhesion mechanisms due to the capillary and viscous
forces. In general, the dynamics of the wet particles are affected by the rotation of the plate;
therefore, the particles are located near the apparatus wall. The particle concentration is
highest in the center of the particle bed and lowest directly above the annular gap. In
addition, in the region above the annular gap, due to the inflowing gas, the poloidal velocity
of the particles is highest. For the tangential and rotational velocities, the region with high
velocities is mainly directly above the rotation plate.
The following findings regarding the influence of liquid loading and liquid properties
on the particle dynamics and interactions were obtained:
• Increasing the viscosity to the value of a 6 mass-% PHARMACOAT® 606 coating
solution results in a denser particle bed. In addition, the particle rotation velocities
and the particle movement in the poloidal plane are reduced.
• A reduced liquid loading in the bed as well as a reduced surface tension of the coating
liquid lead to lower capillary forces, and thus, to increased particle movement.
• The fraction of high contact velocities increases at low liquid loading or low surface
tension, while it decreases at high viscosity. On the other hand, the average contact
force increases significantly with high viscosity.
• Based on the proportional dependence of capillary force on surface tension or viscosity
force on viscosity, it was found that an increase in viscosity leads to an increase in
aggregate size, whereas a reduction in surface tension results in a decrease.
44
Pharmaceutics 2023, 15, 469
References
1. Mörl, L.; Heinrich, S.; Peglow, M. Chapter 2 Fluidized bed spray granulation. In Granulation; Elsevier: Amsterdam, The
Netherlands, 2007; pp. 21–188, ISBN 9780444518712.
2. Tsotsas, E.; Mujumdar, A.S. Modern Drying Technology; Wiley: Hoboken, NJ, USA, 2011; Volume 3, ISBN 9783527315581.
3. Korakianiti, E.S.; Rekkas, D.M.; Dallas, P.P.; Choulis, N.H. Sequential optimization of a pelletization process in a fluid bed rotor
granulator. J. Drug Deliv. Sci. Technol. 2004, 14, 207–214. [CrossRef]
4. Neuwirth, J.; Antonyuk, S.; Heinrich, S.; Jacob, M. CFD–DEM study and direct measurement of the granular flow in a rotor
granulator. Chem. Eng. Sci. 2013, 86, 151–163. [CrossRef]
5. Grohn, P.; Lawall, M.; Oesau, T.; Heinrich, S.; Antonyuk, S. CFD-DEM Simulation of a Coating Process in a Fluidized Bed Rotor
Granulator. Processes 2020, 8, 1090. [CrossRef]
6. Langner, M.; Kitzmann, I.; Ruppert, A.-L.; Wittich, I.; Wolf, B. In-line particle size measurement and process influences on rotary
fluidized bed agglomeration. Powder Technol. 2020, 364, 673–679. [CrossRef]
7. Jacob, M. Chapter 9 Granulation equipment. In Granulation; Elsevier: Amsterdam, The Netherlands, 2007; pp. 417–476,
ISBN 9780444518712.
8. Grohn, P.; Oesau, T.; Heinrich, S.; Antonyuk, S. Investigation of the influence of wetting on the particle dynamics in a fluidized
bed rotor granulator by MPT measurements and CFD-DEM simulations. Powder Technol. 2022, 408, 117736. [CrossRef]
9. Vuppala, M.K.; Parikh, D.M.; Bhagat, H.R. Application of Powder-Layering Technology and Film Coating for Manufacture of
Sustained-Release Pellets Using a Rotary Fluid Bed Processor. Drug Dev. Ind. Pharm. 1997, 23, 687–694. [CrossRef]
10. Kristensen, J.; Schaefer, T.; Kleinebudde, P. Direct pelletization in a rotary processor controlled by torque measurements. II: Effects
of changes in the content of microcrystalline cellulose. AAPS PharmSci 2000, 2, 45. [CrossRef]
11. Gu, L.; Liew, C.V.; Heng, P.W.S. Wet spheronization by rotary processing—A multistage single-pot process for producing
spheroids. Drug Dev. Ind. Pharm. 2004, 30, 111–123. [CrossRef]
12. Deen, N.G.; van Sint Annaland, M.; van der Hoef, M.A.; Kuipers, J.A.M. Review of discrete particle modeling of fluidized beds.
Chem. Eng. Sci. 2007, 62, 28–44. [CrossRef]
13. Fries, L.; Dosta, M.; Antonyuk, S.; Heinrich, S.; Palzer, S. Moisture Distribution in Fluidized Beds with Liquid Injection. Chem.
Eng. Technol. 2011, 34, 1076–1084. [CrossRef]
14. Salikov, V.; Heinrich, S.; Antonyuk, S.; Sutkar, V.S.; Deen, N.G.; Kuipers, J.A.M. Investigations on the spouting stability in a
prismatic spouted bed and apparatus optimization. Adv. Powder Technol. 2015, 26, 718–733. [CrossRef]
15. Breuninger, P.; Weis, D.; Behrendt, I.; Grohn, P.; Krull, F.; Antonyuk, S. CFD–DEM simulation of fine particles in a spouted bed
apparatus with a Wurster tube. Particuology 2019, 42, 114–125. [CrossRef]
16. Sutkar, V.S.; Deen, N.G.; Salikov, V.; Antonyuk, S.; Heinrich, S.; Kuipers, J.A.M. Experimental and numerical investigations of a
pseudo-2D spout fluidized bed with draft plates. Powder Technol. 2015, 270, 537–547. [CrossRef]
17. Heinrich, S.; Dosta, M.; Antonyuk, S. Multiscale Analysis of a Coating Process in a Wurster Fluidized Bed Apparatus. In Mesoscale
Modeling in Chemical Engineering Part I; Elsevier: Amsterdam, The Netherlands, 2015; pp. 83–135, ISBN 9780128012475.
18. Muguruma, Y.; Tanaka, T.; Tsuji, Y. Numerical simulation of particulate flow with liquid bridge between particles (simulation of
centrifugal tumbling granulator). Powder Technol. 2000, 109, 49–57. [CrossRef]
19. Weis, D.; Evers, M.; Thommes, M.; Antonyuk, S. DEM simulation of the mixing behavior in a spheronization process. Chem. Eng.
Sci. 2018, 192, 803–815. [CrossRef]
20. Weis, D.; Grohn, P.; Evers, M.; Thommes, M.; García, E.; Antonyuk, S. Implementation of formation mechanisms in DEM
simulation of the spheronization process of pharmaceutical pellets. Powder Technol. 2021, 378, 667–679. [CrossRef]
21. Grohn, P.; Schaedler, L.; Atxutegi, A.; Heinrich, S.; Antonyuk, S. CFD-DEM Simulation of Superquadric Cylindrical Particles in a
Spouted Bed and a Rotor Granulator. Chem. Ing. Tech. 2023, 95, 244–255. [CrossRef]
45
Pharmaceutics 2023, 15, 469
22. Neuwirth, J. Charakterisierung und Diskrete-Partikel-Modellierung des Strömungs- und Dispersionsverhaltens im Rotorgranulator, 1st ed.;
Cuvillier Verlag: Göttingen, Germany, 2017; ISBN 9783736994768.
23. Israelachvili, J.N. Intermolecular and Surface Forces; Elsevier: Amsterdam, The Netherlands, 2011; ISBN 9780123919274.
24. Lian, G.; Thornton, C.; Adams, M.J. Discrete particle simulation of agglomerate impact coalescence. Chem. Eng. Sci. 1998, 53,
3381–3391. [CrossRef]
25. Popov, V.L. (Ed.) Contact Mechanics and Friction: Physical Principles and Applications, 1st ed.; Springer: Berlin, Germany, 2010;
ISBN 978-3-642-10802-0.
26. Grohn, P.; Oesau, T.; Heinrich, S.; Antonyuk, S. Investigation of the influence of impact velocity and liquid bridge volume on the
maximum liquid bridge length. Adv. Powder Technol. 2022, 33, 103630. [CrossRef]
27. Moukalled, F. The Finite Volume Method in Computational Fluid Dynamics: An Advanced Introduction with OpenFOAM® and Matlab,
1st ed.; Springer: Cham, Switzerland, 2016; ISBN 9783319168746.
28. Hoomans, B.P.B.; Kuipers, J.A.M.; Briels, W.J.; van Swaaij, W.P.M. Discrete particle simulation of bubble and slug formation in a
two-dimensional gas-fluidised bed: A hard-sphere approach. Chem. Eng. Sci. 1996, 51, 99–118. [CrossRef]
29. Goniva, C.; Kloss, C.; Deen, N.G.; Kuipers, J.A.M.; Pirker, S. Influence of rolling friction on single spout fluidized bed simulation.
Particuology 2012, 10, 582–591. [CrossRef]
30. Kanitz, M.; Grabe, J. The influence of the void fraction on the particle migration: A coupled computational fluid dynamics–discrete
element method study about drag force correlations. Int. J. Numer. Anal. Methods Geomech. 2021, 45, 45–63. [CrossRef]
31. Zhao, J.; Shan, T. Coupled CFD–DEM simulation of fluid–particle interaction in geomechanics. Powder Technol. 2013, 239, 248–258.
[CrossRef]
32. Di Felice, R. The voidage function for fluid-particle interaction systems. Int. J. Multiph. Flow 1994, 20, 153–159. [CrossRef]
33. Zhou, Z.Y.; Kuang, S.B.; Chu, K.W.; Yu, A.B. Discrete particle simulation of particle–fluid flow: Model formulations and their
applicability. J. Fluid Mech. 2010, 661, 482–510. [CrossRef]
34. Hesse, R.; Krull, F.; Antonyuk, S. Experimentally calibrated CFD-DEM study of air impairment during powder discharge for
varying hopper configurations. Powder Technol. 2020, 372, 404–419. [CrossRef]
35. Cundall, P.A.; Strack, O.D.L. A discrete numerical model for granular assemblies. Géotechnique 1979, 29, 47–65. [CrossRef]
36. Crowe, C.T.; Schwarzkopf, J.D.; Sommerfeld, M.; Tsuji, Y. Multiphase Flows with Droplets and Particles; CRC Press: Boca Raton, FL,
USA, 2011; ISBN 9780429106392.
37. Salikov, V.; Antonyuk, S.; Heinrich, S.; Sutkar, V.S.; Deen, N.G.; Kuipers, J.A.M. Characterization and CFD-DEM modelling of a
prismatic spouted bed. Powder Technol. 2015, 270, 622–636. [CrossRef]
38. Lian, G.; Thornton, C.; Adams, M.J. A Theoretical Study of the Liquid Bridge Forces between Two Rigid Spherical Bodies.
J. Colloid Interface Sci. 1993, 161, 138–147. [CrossRef]
39. Mikami, T.; Kamiya, H.; Horio, M. Numerical simulation of cohesive powder behavior in a fluidized bed. Chem. Eng. Sci. 1998,
53, 1927–1940. [CrossRef]
40. Rabinovich, Y.I.; Esayanur, M.S.; Moudgil, B.M. Capillary forces between two spheres with a fixed volume liquid bridge: Theory
and experiment. Langmuir 2005, 21, 10992–10997. [CrossRef] [PubMed]
41. Shi, D.; McCarthy, J.J. Numerical simulation of liquid transfer between particles. Powder Technol. 2008, 184, 64–75. [CrossRef]
42. Pitois, O.; Moucheront, P.; Chateau, X. Rupture energy of a pendular liquid bridge. Eur. Phys. J. B 2001, 23, 79–86. [CrossRef]
43. Antonyuk, S.; Heinrich, S.; Deen, N.; Kuipers, H. Influence of liquid layers on energy absorption during particle impact.
Particuology 2009, 7, 245–259. [CrossRef]
44. Gollwitzer, F.; Rehberg, I.; Kruelle, C.A.; Huang, K. Coefficient of restitution for wet particles. Phys. Rev. E Stat. Nonlin. Soft Matter
Phys. 2012, 86, 11303. [CrossRef]
45. Schmelzle, S.; Asylbekov, E.; Radel, B.; Nirschl, H. Modelling of partially wet particles in DEM simulations of a solid mixing
process. Powder Technol. 2018, 338, 354–364. [CrossRef]
46. Reynolds, O. IV. On the theory of lubrication and its application to Mr. Beauchamp tower’s experiments, including an experimental
determination of the viscosity of olive oil. Phil. Trans. R. Soc. 1886, 177, 157–234. [CrossRef]
47. Adams, M.J.; Perchard, V. The cohesive forces between particles with interstitial liquid. In Institution of Chemical Engineering
Symposium; Institution of Chemical Engineers: Rugby, UK, 1985; pp. 147–160.
48. Nase, S.T.; Vargas, W.L.; Abatan, A.A.; McCarthy, J.J. Discrete characterization tools for cohesive granular material. Powder Technol.
2001, 116, 214–223. [CrossRef]
49. Tang, T.; He, Y.; Tai, T.; Wen, D. DEM numerical investigation of wet particle flow behaviors in multiple-spout fluidized beds.
Chem. Eng. Sci. 2017, 172, 79–99. [CrossRef]
50. Anand, A.; Curtis, J.S.; Wassgren, C.R.; Hancock, B.C.; Ketterhagen, W.R. Predicting discharge dynamics of wet cohesive particles
from a rectangular hopper using the discrete element method (DEM). Chem. Eng. Sci. 2009, 64, 5268–5275. [CrossRef]
51. Washino, K.; Miyazaki, K.; Tsuji, T.; Tanaka, T. A new contact liquid dispersion model for discrete particle simulation. Chem. Eng.
Res. Des. 2016, 110, 123–130. [CrossRef]
52. Weller, H.G.; Tabor, G.; Jasak, H.; Fureby, C. A tensorial approach to computational continuum mechanics using object-oriented
techniques. Comput. Phys. 1998, 12, 620. [CrossRef]
53. Issa, R.I. Solution of the implicitly discretised fluid flow equations by operator-splitting. J. Comput. Phys. 1986, 62, 40–65.
[CrossRef]
46
Pharmaceutics 2023, 15, 469
54. Launder, B.E.; Spalding, D.B. The numerical computation of turbulent flows. Comput. Methods Appl. Mech. Eng. 1974, 3, 269–289.
[CrossRef]
55. Kloss, C.; Goniva, C.; Hager, A.; Amberger, S.; Pirker, S. Models, algorithms and validation for opensource DEM and CFD-DEM.
Prog. Comput. Fluid Dyn. 2012, 12, 140–152. [CrossRef]
56. Oesau, T.; Grohn, P.; Pietsch-Braune, S.; Antonyuk, S.; Heinrich, S. Novel approach for measurement of restitution coefficient by
magnetic particle tracking. Adv. Powder Technol. 2022, 33, 103362. [CrossRef]
57. Weis, D. Beschreibung des Sphäronisationsprozesses von Pharmazeutischen Pellets Mittels Numerischer Methoden. Ph.D. Thesis,
TU Kaiserslautern, Kaiserslautern, Germany, 2021.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual
author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to
people or property resulting from any ideas, methods, instructions or products referred to in the content.
47
pharmaceutics
Article
A Hybrid Model to Predict Formulation Dependent Granule
Growth in a Bi-Component Wet Granulation Process
Indu Muthancheri and Rohit Ramachandran *
Department of Chemical and Biochemical Engineering, Rutgers, The State University of New Jersey,
Piscataway, NJ 08854, USA; [email protected]
* Correspondence: [email protected]
Abstract: In this study, a hybrid modeling framework was developed for predicting size distribution
and content uniformity of granules in a bi-component wet granulation system with components of
differing hydrophobicities. Two bi-component formulations, (1) ibuprofen-USP and micro-crystalline
cellulose and (2) micronized acetaminophen and micro-crystalline cellulose, were used in this study.
First, a random forest method was used for predicting the probability of nucleation mechanism
(immersion and solid spread), depending upon the formulation hydrophobicity. The predicted
nucleation mechanism probability is used to determine the aggregation rate as well as the initial
particle distribution in the population balance model. The aggregation process was modeled as
Type-I: Sticking aggregation and Type-II: Deformation driven aggregation. In Type-I, the capillary
force dominant aggregation mechanism is represented by the particles sticking together without
deformation. In the case of Type-II, the particle deformation causes an increase in the contact area,
representing a viscous force dominant aggregation mechanism. The choice between Type-I and II
aggregation is determined based on the difference in nucleation mechanism that is predicted using the
Citation: Muthancheri, I.; random forest method. The model was optimized and validated using the granule content uniformity
Ramachandran, R. A Hybrid Model data and size distribution data obtained from the experimental studies. The proposed framework
to Predict Formulation Dependent predicted content non-uniform behavior for formulations that favored immersion nucleation and
Granule Growth in a Bi-Component
uniform behavior for formulations that favored solid-spreading nucleation.
Wet Granulation Process.
Pharmaceutics 2021, 13, 2063.
Keywords: wet granulation; multicomponent; population balance model; content uniformity
https://doi.org/10.3390/
pharmaceutics13122063
A significant number of modeling studies have been conducted for the prediction of a
single-component wet granulation process, and this includes both nucleation and growth
kinetics [6–9]. Although most of the wet granulation process involves more than one solid
component, there are not many models reported that consider a multi-component process.
Building a formulation-dependent constitutive equation requires the understanding of the
effects of primary particle properties such as maximum pore saturation, material yield
strength, cohesiveness, and primary particle size, on the final product granule attributes.
Matsoukas et al. [10,11] and Marshall Jr et al. [12] applied a multi-component PBM
for modeling a bi-component aggregation using a composition-dependent aggregation
kernel. In their work, two components, solute and solvent, are considered. The internal
coordinates of the granule or particle were determined by the total mass of the granule
and the mass of solute in it. In the Matsoukas et al. [10] model, the solute is soluble in
the solvent, and the bivariate distribution of number density is shown as the product
of the size distribution, with a Gaussian compositional distribution. In an insoluble
system, the Gaussian compositional distribution function that is used to represent the
relative concentration of one component to another may not be applicable. Matsoukas
et al. [11] represented a size-dependent aggregation kernel with a composition-dependent
multiplicative factor. They defined an adjustable interaction parameter that describes the
attraction or repulsion between the two components. The model was able to demonstrate
the extent of blending with positive and negative interaction parameters.
The granule growth (aggregation/coalescence), however, is expressed nearly always
as a semi-empirically based [4,13] kernel or as a function of the particle collision information
obtained from DEM modeling [14,15]. Some of the limitations of the existing coalescence
models for representing granule growth are that they neglect capillary force interaction and
use the static yield strength analyses of the powder bed to calculate the granule strength
during the collision. A multi-dimensional population balance [16] that accounts for size,
solid content, surface liquid, and deformability needs to be used to couple “aggregation”
and “layering” granule growth mechanisms.
Another mechanism to consider is the nucleation process at the start of the liquid
addition phase during granulation, that impacts the granule growth mechanism and the
final granule quality attributes. There are two important aspects of nucleation modeling:
the kinetics of nuclei formation and the physical attributes such as size, porosity, and con-
tent uniformity of the nuclei. In the case of immersion nucleation, kinetic models were
developed by Hounslow et al. [17], Hapgood et al. [18]. These kinetic models provide
the nuclei size distribution as model output and which can be used for predicting the
final granule size distribution [19]. The time scale of the nucleation process is relatively
faster than the rest of the granulation rate mechanisms [20], and thus, it is not necessary to
incorporate the dynamics of the nucleation process into the granulation modeling. How-
ever, the internal properties of nuclei such as size, deformability, surface liquid content,
and content uniformity affect the granulation growth kinetics, and the final granule quality
attributes [21,22]. Thus, it is important to have an experimental or modeling framework to
predict the properties of the initial nuclei to simulate granule properties other than size
distribution using PBM.
Physics-based models for simulating nuclei require complicated multi-phase simula-
tions. Washino et al. [23] presented a coupled DEM and constrained interpolation profile
(CIP) for simulating the nuclei during the wet granulation process. The effect of surface
tension on the liquid binder flow was modeled depending on the relative position of
the fluid interfaces to the solid particles, i.e., the model on the outside, inside, or on the
surface of the powder bed corresponds to a free surface, capillary action, and bed surface
wetting. Washino et al. [24] showed the CFD-DEM simulation of nuclei generation in
a dynamic powder bed. These studies consider a particle system with good wettability
or spreading coefficient [25]. In the case of powder mixture with hydrophobic powders,
the nuclei formed are solid-spread nuclei, and the physics-based approach developed for
immersion nuclei is not suitable for such systems. Due to the complexities of nucleation
49
Pharmaceutics 2021, 13, 2063
mechanisms that can be either an immersion or solid-spread nuclei based on the wettability
of the constituent material on the droplet vicinity during binder addition, the currently
available models fall short of representing the nucleation mechanism and nuclei property
in such systems.
Hybrid modeling has been demonstrated to have various advantages of improving
process understanding with the incorporation of empirically based statistical models with
mechanistic models [26–28]. A hybrid model consisting of both statistical correlations and
physics-based models is often used to simplify the computational efforts and incorporate
complex mechanisms into the model. Such models can overcome the disadvantages of
both purely data-driven and physics-based models [27]. A hybrid model such as PBM
with artificial neural network (ANN) was developed to substitute for the high-fidelity
PBM-DEM model by Barrasso et al. [29].
The objectives of this study are as follows:
1. Incorporate the nuclei particle characteristics in the population balance model based
on the classification model result from Muthancheri et al. [30].
2. Develop a composition-dependent PBM framework for bi-component wet granulation
process with a large binder droplet for predicting the granule quality attributes with
change in percentage formulation.
2. Model Development
2.1. Population Balance Model
The population balance equation as shown in Equations (1) and (3) are used in this
work to predict the particle size distribution, liquid distribution, and component distri-
bution or the content uniformity [31]. In this work, the liquid volume of the granules is
considered to be a lumped parameter under the assumption that all granules of the same
size with the same composition of solids and pore volume have the same average liquid
content. Such a reduced-order model was compared with higher-order models and was
reported to have a significant time saving without compromising much on accuracy in
previous studies [32].
∂F ∂ ds1 ∂ ds2 ∂ dp
+ F + F + F = nuc + agg + + brk (1)
∂t ∂s1 dt ∂s2 dt ∂p dt
∂Li dl
= nuc,li + agg,li + brk,li + F i (2)
∂t dt
∂Le dle
= nuc,le + agg,le + brk,le + F (3)
∂t dt
where F = F (s1 , s2 , p, t) is the number of particles with API volume s1 and excipient
volume s2 , pore volume p at time t. Le = Le (s1 , s2 , p, t) and Li = Li (s1 , s2 , p, t) is the
average external liquid volume and average internal liquid volume of particles with API
volume s1 and excipient volume s2 , pore volume p at time t, respectively. Table 1 represents
the dependent and independent variables. The rate mechanisms dsdt1 , dsdt2 , dl i dle
dt , dt , nuc , agg ,
and brk are detailed in the next section.
The relationship between the variables in Table 1 are summarized in the following
equations. The total granule volume v is obtained by Equation (4).
v = s1 + s2 + l e + p (4)
The surface area of the particles can be derived from the granule total volume as
shown in Equation (5)
1 2
a = π 3 (6v) 3 (5)
The porosity and content uniformity can be calculated using Equations (6) and (7),
respectively.
50
Pharmaceutics 2021, 13, 2063
p
= (6)
v
q = s1 /v (7)
Description Notation
Independent variables
API solid volume s1
Excipient solid volume s2
External liquid volume le
Internal liquid volume li
Pore volume p
Dependent variables
Total granule volume v
Surface area a
Porosity
Content uniformity q
Q̇spray × Δt
Nim = Pim (8)
vd
where Qspray is the volumetric spray rate of binder liquid and vd is the volume of a single
drop calculated from the nozzle opening. Pim is the probability of immersion nucleation
to happen for the given percentage composition of API powder bed (calculated from the
classification model from Muthancheri et al. [30]). All the immersion nuclei is assigned to
the first bin of same percentage composition as that of the powder bed and pore volume
close to the volume of single droplet.
The mass of API (Ms1 ) and excipient (Ms2 ) available is calculated at every time step
based on the mass balance of nuclei generated and excipient particle layering. The volume
of immersion nuclei (vim ) at the end of nucleation can be estimated from the equation
derived by Hounslow et al. [17] (Equation (9)).
1 − φcp
vim = vd 1 + (9)
φcp
1 − φcp
vim,s = vd (10)
φcp
51
Pharmaceutics 2021, 13, 2063
where φcp is the critical-packing liquid volume fraction, which is kept constant at 0.2 in this
study. vim,s is the solid volume in the nuclei. Total mass of solid component Mim,si , where
i = 1, 2 for API and excipient, respectively, utilized to form immersion nuclei at Δt can be
calculated as follows:
where mim,s is the solid mass in a single immersion nuclei and ρs is the weighted true
density of the solid components. f i is the fraction of solid component present in the powder
bed (i = 1, 2 for API and excipient, respectively).
Q̇spray × Δt
Nss = (1 − Pim ) (13)
vd
Assuming that in the solid-spread nucleation the liquid drop (diameter dd ) is sur-
rounded by hydrophobic API particles of diameter (d p ) (Figure 1), the approximate solid-
spread nuclei volume (vss ) and the volume of API particles in a solid-spread nuclei (vss,s1 )
can be calculated as follows:
π 3
vss = dd + 2d p (14)
6
π 3
vss,s1 = dd + 2d p − vd (15)
6
The total mass of API particles that form solid-spread nuclei at time Δt can be calcu-
lated as shown in Equation (17).
where mss,s is the solid mass in a single solid-spread nuclei and ρs1 is the weighted true
density of the API.
52
Pharmaceutics 2021, 13, 2063
dle Q̇spray × v
= (18)
dt rewetting ∑s1 ∑s2 ∑ p Fv
ds1
= k sg v2/3 (19)
dt sur f acegrowth
Planchette et al. [34] studied the transition of liquid marble onto solid surfaces. They
studied three mechanisms involved when solid-spread nuclei collide on a surface. The drop
extension of the solid-spread nuclei is related to the impact velocity as ( Dmax − D )/D
0.12 (We), where Dmax is the diameter of the disk shape the solid-spread nuclei takes
before rupture, D is the diameter of the solid-spread nuclei (=dd + 2d p ), and We is the
Weber number during collision. This equation gives the minimum size (= f ( Dmax )) of the
granule upon which, when the solid-spread nuclei collide, the impact results in the surface
growth of s1 particles, as shown in Figure 2.
It is modeled such that only the granules with surface wetness (le = 0) experience
layering. As a result, the increase in consolidation has a secondary effect on layering.
The depletion of excipient fines is given by Equation (21).
s1 s2 p
dms2 ds2
= ρs2 F ds ds2 dp (21)
dt 0 0 p0 dt 1
53
Pharmaceutics 2021, 13, 2063
surface area upon collision (Type I) and two granules of low strength, high deformability,
and higher capillary number are assumed to conserve pore volumes upon collision.
The solid volumes are conserved during the aggregation. The resulting aggregate
solid volume is a sum of the two colliding particle solid volume. The pore volume and
liquid volume undergo a different transformation rule. Depending on the deformability of
the colliding particles, the final granule pore volume is an interpolation of the two extremes
in Figure 3. The resulting pore volume can be written as,
3 3
a = ζ ( a A2 + a B2 ) + (1 − ζ )( a A + a B ), 0 < ζ < 1 (22)
3
a 2
p = √ − s1 − s2 − l e (23)
6 π
f orm dep
where R agg and R agg are the rates of formation of larger particles (Equation (26)) and rate
of depletion of smaller particles (Equation (27)), respectively.
s s2 p
f orm 1 1
R agg (s1 , s2 , p, t) = β(s1 − s1 , s2 − s2 , p − p , s1 , s2 , p )× (25)
2 0 0 p0
The rate of aggregation R agg,li and R agg,le in Equations (2) and (3) are the rates at
which the internal and external liquid volumes are transferred between particles due to
aggregation. Similar to the formation and depletion of what was discussed above, these
rates can be calculated as shown in Equations (28) and (29).
54
Pharmaceutics 2021, 13, 2063
s s2 p
1 1
R agg,li (s1 , s2 , p, t) = β(s1 − s1 , s2 − s2 , p − p , s1 , s2 , p )× (27)
2 0 0 p0
F ( s1 − s1 , s2 − s2 , p − p , t ) F ( s1 , s2 , p , t )
(li (s1 − s1 , s2 − s2 , p − p , t) + li (s1 , s2 , p , t) + le→i )ds1 ds2 dp
s s2,max −s2 pmax − p
1,max − s1
− Li (s1 , s2 , p, t)
0 0 p0
The transfer of external liquid volume to internal liquid volume due to aggrega-
tion (represented as le→i in Equations (28) and (29)) can be computed as discussed by
Braumann et al. [31].
1
2
l e →i = l e ( s1 − s1 , s2 − s2 , p − p , t ) l e ( s1 , s2 , p , t ) (29)
⎡ ⎤1
⎛ ⎞2 2
⎢ 3
v ( s1 − s1 , s2 − s2 , p − p , t ) − l e ( s1 − s1 , s2 − s2 , p − p , t ) ⎥
×⎢ − 1 − ⎝ ⎠ ⎥
⎣ 1 ⎦ (30)
3
v ( s1 − s1 , s2 − s2 , p − p , t ) + 3 v ( s1 , s2 , p , t )
⎡ ⎤ 12
⎛ ⎞2
⎢ v ( s1 , s2 , p , t ) − l e ( s1 , s2 , p , t )
3
⎥
×⎢ ⎝ ⎠ ⎥
⎣1 − 1 − ⎦
3
v ( s1 − s1 , s2 − s2 , p − p , t ) + v ( s1 , s2 , p , t )
3
The aggregation kernel β depends on the properties of the colliding particle A(s1 , s2 , p)
and B(s1 , s2 , p ) (as shown in Figure 3). β(A, B) = β 0 β∗ A, B. β 0 is independent of the
colliding particle properties and is an optimized parameter in this work. β∗ is the efficiency
of particle collision which can be determined based on the following model proposed by
Balakin et al. [35]. The model accounts for both capillary and viscous forces during particle
collision. The efficiency is determined as a ratio of the total work of forces within the liquid
bridge to the kinetic energy of the particle.
Wc + Wd
β∗ = Ψ (31)
Ek
where Wc and Wd are the work of the capillary and dissipative forces, respectively. Ek is the
kinetic energy calculated from the mean relative velocity (vr ), mass (m), and coefficient of
restitution (e, Equation (37)) of particle before collision, as shown in Equation (32). The rel-
ative velocity is calculated from the granular temperature (Θ), as given by Equation (33).
55
Pharmaceutics 2021, 13, 2063
1 2 2
Ek = me vr (32)
2
3√
vr = πΘ (33)
2
(5π/96)γ2 d2
Θ= 2 (34)
φp
12φ p 1 − 1−φ p
2
where φ p is the volume fraction of particle in the granulator, γ is the shear rate, and d is the
granule diameter. The work of dissipative force is calculated using the following equation.
3πμd˜2 ecoag vr h
Wd = ln (35)
4 ha
where d˜ is the harmonic mean diameter of the colliding two particles and μ is the viscosity
of the binder. h is the binder layer thickness calculated from the external liquid content
(le ) and h a represents the granule surface asperities. The resulting aggregated particle
coefficient of restitution (ecoag ) is represented as a function of coefficient of restitution of
the constituent material properties by Braumann et al. [31] by the following equations:
√
ecoag = eA eB (36)
∑ eα mα
ei = α , i ∈ {A,B} (37)
∑α mα
where α ∈ {s1 , s2 , p}, m is the mass of colliding granule and e is the ratio of rebound energy
to impact energy. It takes a value between 0 (totally plastic impact) and 1 (totally elastic
impact). e for pore is assumed to be 0.
The work of capillary force (Wc ) is experimentally determined from the regime map
analysis carried out Muthancheri and Ramachandran [22]. Figure 4 plots the capillary
number which is the ratio between capillary force and viscous force as a function of API
fraction. The equation determined from the experiment analysis is used to provide a
composition dependent work of capillary force in the model.
(a) (b)
Figure 4. Change in capillary numbers with increase in API fraction. (a) Ibuprofen and MCC101
formulation, (b) Acetaminophen (APAP) and MCC101 formulation.
2.2.7. Compaction
Compaction of granules occur during collision and result in porosity reduction of the
granules.
56
Pharmaceutics 2021, 13, 2063
where k con is the consolidation rate constant, U is the particle collision velocity, and min
is the minimum porosity. Two conditions are modeled in this study. A non-squeeze case:
if no internal liquid is transferred to the external surface. In this case, there is only pore
volume reduction due to consolidation (Equation (39)). The next scenario is a squeeze case.
Some liquid is transferred from internal to external liquid volume (Equation (40)). This
occurs if the porosity after consolidation is smaller than a critical porosity. In this scenario,
the pore volume is completely occupied by internal liquid volume (li = p).
dp 1
=− ( s + s2 + l e ) − v (39)
dt 1 − ( − Δ) 1
dp
= −((1 + ( − Δ))(s1 + s2 + le ) + ( − Δ)li − v) (40)
dt
57
Pharmaceutics 2021, 13, 2063
integration technique, which is popularly used to solve PBMs [5,9,36–38]. The numerical
stability of a PBM is complex due to the presence of multiple dimensions and the inherent
possibility of instability involved with the time-step of the integration. The integration
time-step was thus chosen, such that the rate of particles leaving a particular size class
(bin) is not higher than the number of particles in that size class at any time-step based on
the Courant–Friedrichs–Lewy (CFL) condition [8,39]. The partial derivatives with respect
to internal coordinate volume (u ∈ (s1 , s2 , p)) and time (t) were discretized using a non-
linear grid (ui = u × (4)i−1 ). Here, i represents the bin number in one dimension, and u
indicates the volume of particle in the smallest bin. The smallest particle size is 31.5 μm,
and the largest particle size is 6000 μm. There is a total of 20 bins or grid points. A cell
average technique discussed by Chaudhury et al. [7] is utilized in this study to distribute
the particles that are formed in the intermediate range of two bins, into the adjacent bins.
The computations were performed in MATLAB 2020a on an Intel(R) Core(TM) i7-8700 CPU
(3.20 GHz) with 16 GB RAM.
j
where Yi (t) is the value of granule property of interest in the ith perturbation of the jth
j
parameter and Y0 (t) is the base value for the jth parameter.
Figure 6 illustrates that the sensitivity of parameters on d10 , d50 , and d90 simulation. It
shows that growth parameters (k layer and k sg ) are much less sensitive than the aggregation
and consolidation parameters when the variables are perturbed ±20%. The average
diameter is found to be highly sensitivity toward the coefficient of restitution of API (es1 ).
The study shows a decrease in average diameter with an increase in es1 . A decrease in
es1 indicates that the API is very deformable, resulting in smaller average granule size.
The aggregation rate constant, β 0 , has a positive impact on the granule size, showing an
increase in the rate constant increasing the average granule size. The consolidation rate
equation has a negative term (Equations (39) and (40)), which means the increase in k con
results in a decrease in consolidation rate. In Figure 6c, it can be seen that a decrease in
consolidation rate to 20% results in larger granules.
Similarly, Figure 7 shows the effect of the adjustable parameters on the average
porosity and API content of the granules. The sensitivity of parameters to granule API
content is similar to that of the granule size. Aggregation rate constant, coefficient of
restitution, and consolidate rate were found to be most significant in impacting the granule
API content. A decrease in es1 results in a decrease in aggregation rate and thus results
in granules with less s1 or API content. Average porosity of granule is most impacted by
the consolidation rate. It can be seen that with increase in consolidation (or decrease in
consolidation rate constant) the average granule porosity decreases.
58
Pharmaceutics 2021, 13, 2063
(a) (b)
(c)
Figure 6. Effect of changes in adjustable parameter values on d10 , d50 , and d90 . Sensitivity to (a) d10 ,
(b) d50 , (c) d90 .
(a) (b)
Figure 7. Effect of changes in adjustable parameter values on porosity and API content. (a) Sensitivity
to average API content, (b) Sensitivity to granule average porosity
59
Pharmaceutics 2021, 13, 2063
parameter estimation, and two were used to validate the calibrated model. The optimized
values of the variables are provided in the Table 2.
60
Pharmaceutics 2021, 13, 2063
The content uniformity of the granules were evaluated using the demixing potential
(DP) introduced by Thiel and Nguyen [41] to quantify the distribution of a solid component
as a function of particle size. DP can be calculated using the following equation:
100
DP% =
x̄ ∑ w(x − x̄)2 (42)
where x is the API content in a particular size range, x̄ is the average API content, and w
is the weight fraction of granule in each size range. The quantity is similar to the relative
standard deviation used for non-uniformity in mixing by Oka et al. [42]. The larger the
value of de-mixing potential, the larger the extent of deviation from the mean of the API
across granule size classes. Figure 11 shows the ability of the presented hybrid-modeling
framework to predict the change in de-mixing potential with an increase in percentage API.
The model predicted a decrease in the extend of de-mixing with increase in percentage
API for ibuprofen formulation and an increase in the extent of de-mixing with increase in
percentage API for APAP formulation.
61
Pharmaceutics 2021, 13, 2063
Figure 11. Comparison between model prediction and experimentally obtained de-mixing potential.
Figure 12. Cumulative volume fraction prediction with change in percentage composition of API
and 0.6 liquid-to-solid ratio.
62
Pharmaceutics 2021, 13, 2063
(a) (b)
(c)
Figure 13. Granule size distribution with increase in granulation time at a varying degree of hy-
drophobic content (ibuprofen). (a) Granule size distribution with increase in granulation time at
40% hydrophobic component (ibuprofen), (b) granule size distribution with increase in granulation
time at 50% hydrophobic component (ibuprofen), and (c) granule size distribution with increase in
granulation time at 60% hydrophobic component (ibuprofen).
Figure 14. Model predicted granule size distribution with an increase in wet massing time (50% API
content).
63
Pharmaceutics 2021, 13, 2063
Figure 15. Model predicted granule API content with increase in wet massing time.
(a) (b)
Figure 16. Change in granule micro-structure with increase in API content. (a) Model predicted
average porosity, (b) Experimental envelop density.
64
Pharmaceutics 2021, 13, 2063
4. Conclusions
In this study, a hybrid model (Random Forest-PBM) is developed to describe the
bi-component high shear wet granulation process. The model incorporates immersion
and solid-spread nucleation based on the change in percentage API. The probability of
each nucleation mechanism to occur, for a given formulation, is obtained from the random
forest model. The probability is incorporated into the PBM framework such that the rate
equations are impacted by the availability of immersion and solid-spread nuclei. It was
found that the aggregation and consolidation rate are more sensitive to the granule critical
quality attribute predictions. The model predictions are qualitatively in agreement with
profiles obtained in the literature [22,43]. The discussed methodology and presented model
could be used for predicting various aspects of the granulation process and controlling
the transient behavior during the process. As an example, we have provided the average
particle size, porosity, liquid content, and demixing-potential of the granules with change
in API percentage. The developed model is an improvement to the existing mechanistic
modeling framework, such that it incorporates the effect of hydrophobicity to track the
granule critical quality attributes.
Author Contributions: Conceptualization, I.M. and R.R.; methodology, I.M.; validation, I.M.; formal
analysis, I.M.; investigation, I.M.; resources, R.R.; writing—original draft preparation, I.M.; writing—
review and editing, I.M. and R.R.; visualization, I.M.; supervision, R.R.; project administration, I.M.;
funding acquisition, R.R. All authors have read and agreed to the published version of the manuscript.
Funding: This research was funded by National Science Foundation CAREER program through
grant no: 1350152.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: Not applicable.
65
Pharmaceutics 2021, 13, 2063
Acknowledgments: The authors would like to acknowledge the Department of Chemical & Bio-
chemical Engineering, Rutgers University, for funding in the form of a teaching assistantship
to I. Muthancheri, as well as the National Science Foundation CAREER program through grant
no: 1350152.
Conflicts of Interest: The authors declare no conflict of interest.
Appendix A
Equation (A1) represents the breakage rate equation.
f orm dep
Rbrk (s1 , s2 , p, t) = Rbrk (s1 , s2 , p, t) − Rbrk (s1 , s2 , p, t) (A1)
f orm dep
where Rbrk and Rbrk are the rates of formation of smaller particles and rate of depletion
of larger particles.
s s2,max pmax
f orm 1,max
Rbrk (s1 , s2 , p, t) = b(s1 , s2 , p, s1 , s2 , p )Kbrk (s1 , s2 , p)× (A2)
s1 s2 p
where Kbrk is the breakage rate constant and b is the probability distribution function of
daughter particles. For the purposes of this study, a uniform probability distribution was
assumed for all possible daughter particles [32].
References
1. Suresh, P.; Sreedhar, I.; Vaidhiswaran, R.; Venugopal, A. A comprehensive review on process and engineering aspects of
pharmaceutical wet granulation. Chem. Eng. J. 2017, 328, 785–815. [CrossRef]
2. Gernaey, K.V.; Cervera-Padrell, A.E.; Woodley, J.M. A perspective on PSE in pharmaceutical process development and innovation.
Comput. Chem. Eng. 2012, 42, 15–29. [CrossRef]
3. Ramkrishna, D.; Singh, M.R. Population balance modeling: Current status and future prospects. Annu. Rev. Chem. Biomol. Eng.
2014, 5, 123–146. [CrossRef]
4. Liu, L.; Litster, J. Population balance modelling of granulation with a physically based coalescence kernel. Chem. Eng. Sci. 2002,
57, 2183–2191. [CrossRef]
5. Chaudhury, A.; Wu, H.; Khan, M.; Ramachandran, R. A mechanistic population balance model for granulation processes: Effect
of process and formulation parameters. Chem. Eng. Sci. 2014, 107, 76–92. [CrossRef]
6. Poon, J.M.H.; Immanuel, C.D.; Doyle, F.J., III; Litster, J.D. A three-dimensional population balance model of granulation with a
mechanistic representation of the nucleation and aggregation phenomena. Chem. Eng. Sci. 2008, 63, 1315–1329.
7. Chaudhury, A.; Kapadia, A.; Prakash, A.V.; Barrasso, D.; Ramachandran, R. An extended cell-average technique for a multi-
dimensional population balance of granulation describing aggregation and breakage. Adv. Powder Technol. 2013, 24, 962–971.
[CrossRef]
8. Ramachandran, R.; Barton, P.I. Effective parameter estimation within a multi-dimensional population balance model framework.
Chem. Eng. Sci. 2010, 65, 4884–4893. [CrossRef]
9. Barrasso, D.; Ramachandran, R. Multi-scale modeling of granulation processes: Bi-directional coupling of PBM with DEM via
collision frequencies. Chem. Eng. Res. Des. 2015, 93, 304–317. [CrossRef]
10. Matsoukas, T.; Lee, K.; Kim, T. Mixing of components in two-component aggregation. AIChE J. 2006, 52, 3088–3099. [CrossRef]
11. Matsoukas, T.; Kim, T.; Lee, K. Bicomponent aggregation with composition-dependent rates and the approach to well-mixed
state. Chem. Eng. Sci. 2009, 64, 787–799. [CrossRef]
12. Marshall, C.L., Jr.; Rajniak, P.; Matsoukas, T. Multi-component population balance modeling of granulation with continuous
addition of binder. Powder Technol. 2013, 236, 211–220. [CrossRef]
13. Kapur, P.; Kapur, P.; Fuerstenau, D. An auto-layering model for the granulation of iron ore fines. Int. J. Miner. Process. 1993,
39, 239–250. [CrossRef]
14. Ingram, G.D.; Cameron, I.T. Formulation and comparison of alternative multiscale models for drum granulation. Comput. Aided
Chem. Eng. 2005, 20, 481–486.
15. Barrasso, D.; Ramachandran, R. Qualitative assessment of a multi-scale, compartmental PBM-DEM model of a continuous
twin-screw wet granulation process. J. Pharm. Innov. 2016, 11, 231–249. [CrossRef]
16. Goodson, M.; Kraft, M.; Forrest, S.; Bridgwater, J. A multi-dimensional population balance model for agglomeration. In
Proceedings of the PARTEC 2004—International Congress for Particle Technology, Nuremburg, Germany, 16–18 March 2004.
66
Pharmaceutics 2021, 13, 2063
17. Hounslow, M.; Oullion, M.; Reynolds, G. Kinetic models for granule nucleation by the immersion mechanism. Powder Technol.
2009, 189, 177–189. [CrossRef]
18. Hapgood, K.P.; Litster, J.D.; Smith, R. Nucleation regime map for liquid bound granules. AIChE J. 2003, 49, 350–361.
19. Bellinghausen, S.; Gavi, E.; Jerke, L.; Ghosh, P.K.; Salman, A.D.; Litster, J.D. Nuclei size distribution modelling in wet granulation.
Chem. Eng. Sci. X 2019, 4, 100038. [CrossRef]
20. Immanuel, C.D.; Doyle, F.J., III. Mechanistic modelling of aggregation phenomena in population balances of granulation. IFAC
Proc. Vol. 2005, 38, 416–421. [CrossRef]
21. Farber, L.; Al-aaraj, D.K.; Smith, R.; Gentzler, M. Formation and internal microstructure of granules from wetting and non-wetting
efavirenz/iron oxide blends. Chem. Eng. Sci. 2020, 227, 115909. [CrossRef]
22. Muthancheri, I.; Ramachandran, R. Mechanistic understanding of granule growth behavior in bi-component wet granulation
processes with wettability differentials. Powder Technol. 2020, 367, 841–859. [CrossRef]
23. Washino, K.; Tan, H.; Hounslow, M.; Salman, A. Meso-scale coupling model of DEM and CIP for nucleation processes in wet
granulation. Chem. Eng. Sci. 2013, 86, 25–37. [CrossRef]
24. Washino, K.; Tan, H.; Hounslow, M.; Salman, A. A new capillary force model implemented in micro-scale CFD–DEM coupling
for wet granulation. Chem. Eng. Sci. 2013, 93, 197–205. [CrossRef]
25. Hapgood, K.; Farber, L.; Michaels, J. Agglomeration of hydrophobic powders via solid spreading nucleation. Powder Technol.
2009, 188, 248–254. [CrossRef]
26. Chen, Y.; Ierapetritou, M. A framework of hybrid model development with identification of plant-model mismatch. AIChE J.
2020, 66, e16996. [CrossRef]
27. Azarpour, A.; Borhani, T.N.; Alwi, S.R.W.; Manan, Z.A.; Mutalib, M.I.A. A generic hybrid model development for process
analysis of industrial fixed-bed catalytic reactors. Chem. Eng. Res. Des. 2017, 117, 149–167. [CrossRef]
28. Sampat, C.; Ramachandran, R. Identification of Granule Growth Regimes in High Shear Wet Granulation Processes Using a
Physics-Constrained Neural Network. Processes 2021, 9, 737. [CrossRef]
29. Barrasso, D.; Tamrakar, A.; Ramachandran, R. A reduced order PBM–ANN model of a multi-scale PBM–DEM description of a
wet granulation process. Chem. Eng. Sci. 2014, 119, 319–329. [CrossRef]
30. Muthancheri, I.; Oka, S.; Ramachandran, R. Analysis and prediction of nucleation mechanisms in a bi-component powder bed
with wettability differentials. Powder Technol. 2021, 390, 209–218. [CrossRef]
31. Braumann, A.; Goodson, M.J.; Kraft, M.; Mort, P.R. Modelling and validation of granulation with heterogeneous binder dispersion
and chemical reaction. Chem. Eng. Sci. 2007, 62, 4717–4728. [CrossRef]
32. Barrasso, D.; Ramachandran, R. A comparison of model order reduction techniques for a four-dimensional population balance
model describing multi-component wet granulation processes. Chem. Eng. Sci. 2012, 80, 380–392. [CrossRef]
33. Jeong, J.I.; Choi, M. A bimodal particle dynamics model considering coagulation, coalescence and surface growth, and its
application to the growth of titania aggregates. J. Colloid Interface Sci. 2005, 281, 351–359. [CrossRef]
34. Planchette, C.; Biance, A.; Lorenceau, E. Transition of liquid marble impacts onto solid surfaces. EPL (Europhys. Lett.) 2012,
97, 14003. [CrossRef]
35. Balakin, B.V.; Kutsenko, K.V.; Lavrukhin, A.A.; Kosinski, P. The collision efficiency of liquid bridge agglomeration. Chem. Eng.
Sci. 2015, 137, 590–600. [CrossRef]
36. Chaudhury, A.; Armenante, M.E.; Ramachandran, R. Compartment based population balance modeling of a high shear wet
granulation process using data analytics. Chem. Eng. Res. Des. 2015, 95, 211–228. [CrossRef]
37. Barrasso, D.; El Hagrasy, A.; Litster, J.D.; Ramachandran, R. Multi-dimensional population balance model development and
validation for a twin screw granulation process. Powder Technol. 2015, 270, 612–621. [CrossRef]
38. Muthancheri, I.; Chaturbedi, A.; Bétard, A.; Ramachandran, R. A compartment based population balance model for the prediction
of steady and induction granule growth behavior in high shear wet granulation. Adv. Powder Technol. 2021, 32, 2085–2096.
[CrossRef]
39. Courant, R.; Friedrichs, K.; Lewy, H. On the partial difference equations of mathematical physics. IBM J. Res. Dev. 1967,
11, 215–234. [CrossRef]
40. Gunantara, N. A review of multi-objective optimization: Methods and its applications. Cogent Eng. 2018, 5, 1502242. [CrossRef]
41. Thiel, W.; Nguyen, L. Fluidized bed granulation of an ordered powder mixture. J. Pharm. Pharmacol. 1982, 34, 692–699. [CrossRef]
[PubMed]
42. Oka, S.; Kašpar, O.; Tokárová, V.; Sowrirajan, K.; Wu, H.; Khan, M.; Muzzio, F.; Štěpánek, F.; Ramachandran, R. A quantitative
study of the effect of process parameters on key granule characteristics in a high shear wet granulation process involving a two
component pharmaceutical blend. Adv. Powder Technol. 2015, 26, 315–322. [CrossRef]
43. Chaudhury, A.; Ramachandran, R. Integrated population balance model development and validation of a granulation process.
Part. Sci. Technol. 2013, 31, 407–418. [CrossRef]
44. Ansari, M.A.; Stepanek, F. The effect of granule microstructure on dissolution rate. Powder Technol. 2008, 181, 104–114. [CrossRef]
67
pharmaceutics
Article
Conceptualisation of an Efficient Particle-Based Simulation
of a Twin-Screw Granulator
John P. Morrissey, Kevin J. Hanley and Jin Y. Ooi *
School of Engineering, Institute for Infrastructure & Environment, The University of Edinburgh,
Edinburgh EH9 3JL, UK; [email protected] (J.P.M.); [email protected] (K.J.H.)
* Correspondence: [email protected]
Abstract: Discrete Element Method (DEM) simulations have the potential to provide particle-scale
understanding of twin-screw granulators. This is difficult to obtain experimentally because of the
closed, tightly confined geometry. An essential prerequisite for successful DEM modelling of a twin-
screw granulator is making the simulations tractable, i.e., reducing the significant computational cost
while retaining the key physics. Four methods are evaluated in this paper to achieve this goal: (i)
develop reduced-scale periodic simulations to reduce the number of particles; (ii) further reduce this
number by scaling particle sizes appropriately; (iii) adopt an adhesive, elasto-plastic contact model
to capture the effect of the liquid binder rather than fluid coupling; (iv) identify the subset of model
parameters that are influential for calibration. All DEM simulations considered a GEA ConsiGma™ 1
twin-screw granulator with a 60◦ rearward configuration for kneading elements. Periodic simulations
yielded similar results to a full-scale simulation at significantly reduced computational cost. If the
level of cohesion in the contact model is calibrated using laboratory testing, valid results can be
obtained without fluid coupling. Friction between granules and the internal surfaces of the granulator
is a very influential parameter because the response of this system is dominated by interactions with
Citation: Morrissey, J.P.; Hanley, K.J.; the geometry.
Ooi, J.Y. Conceptualisation of an
Efficient Particle-Based Simulation of Keywords: powder agglomeration; Discrete Element Method; cohesion; wet granulation; twin-
a Twin-Screw Granulator. screw granulation
Pharmaceutics 2021, 13, 2136.
https://doi.org/10.3390/
pharmaceutics13122136
1. Introduction
Academic Editor: Colin Hare
Wet granulation is a process used to create larger stable agglomerates (granules)
from fine powders. This has many desirable outcomes such as improving flowability,
Received: 29 October 2021
Accepted: 6 December 2021
compactibility and homogeneity. Granulation is commonly employed in the food, phar-
Published: 12 December 2021
maceutical, detergent and fertilizer industries. Despite its widespread adoption, it is often
inefficiently operated [1,2], with high recycle ratios in continuous processes and high rejec-
Publisher’s Note: MDPI stays neutral
tion rates in batch processes [3]. Wet granulation is the most common type of granulation
with regard to jurisdictional claims in
and, in pharmaceutical applications, it is a critical step in tablet manufacturing that affects
published maps and institutional affil- the uniformity and compactibility of the final dosage form.
iations. Traditionally, batch granulation was the favoured granulation approach in the pharma-
ceutical industry due to the challenges associated with continuous processing such as the
changeover cost and inability to monitor product quality reliably. Since the introduction of
Quality by Design (QbD) and Process Analytical Technology (PAT) by the FDA in 2003 [4]
Copyright: © 2021 by the authors.
and due to recent advances in continuous monitoring, there has been a move from tradi-
Licensee MDPI, Basel, Switzerland.
tional batch towards continuous processing. Although tremendous efforts have been made
This article is an open access article
to gain scientific insight into the granulation process [5–12], a fundamental understanding
distributed under the terms and of wet granulation is still lacking due to the complexity of the mechanisms involved, i.e.,
conditions of the Creative Commons governing rate processes of wetting and nucleation, aggregation and consolidation (or
Attribution (CC BY) license (https:// consolidation and coalescence), breakage and attrition, and layering [1,9].
creativecommons.org/licenses/by/ Twin-screw granulation developed from early work on single- [13] and twin-screw
4.0/). extruders [14] in the late 1980s. It has become popular within the pharmaceutical industry
in the past decade due to the many advantages it offers over high shear and fluidised bed
granulators including limited or no scale-up requirement, lower space requirements, con-
tinuous operation with monitoring and higher throughput. Since a patent was awarded to
Ghebre-Sellassie et al. [15], there has been increased adoption of this method of granulation
in industry.
A twin-screw granulator (TSG) consists of three main components: the powder feed,
liquid addition mechanism and two intermeshed screws, which may be either co- or
counter-rotating, enclosed in a barrel. The powder is fed into the screw barrels at one end
and exits the open end of the barrel. This is in contrast to twin-screw extruders where
the granulated product is forced through a die or plate at the end of the barrel. Twin-
screw granulators are often described by their length-to-diameter (L/D) ratio [16] with
extruder lengths in the range of 12–50 L/D common in industry [17]. Screws are required
to tightly fit within the bore of the barrel to create a closely confined flow path for materials.
Researchers have reported that the confinement offered by the low clearance between the
screws and barrel ensures a similar shear history for all particles, which helps produce
more consistent granules than batch granulation processes [17,18]. Screws typically operate
part-filled, with solid fraction dependent on screw geometry, material feed rate and screw
speed [16]. The screws are modular in nature and comprise elements of different types to
give the required configuration.
In general, there are four element types: conveying elements, kneading elements,
chopping elements and comb mixer elements (not considered in this paper). Multiple
elements are combined to form a block; multiple alternating blocks form the complete
screw. Kneading blocks impart high mechanical energy to the wetted material, producing
high shear forces, compaction and distributive mixing. Further details on screw elements,
profiles and configuration can be found in the literature [16–22].
Depending on the angle of offset, kneading blocks can produce forwarding or revers-
ing flow [23,24]. Reversing kneading blocks force material back against the direction of
flow, which leads to areas of high pressure and compaction. Although this allows strong
granules to be formed, there is a high likelihood of blockages [25]. Chopping elements are
shortened kneading elements whose purpose is to break up oversized agglomerates at the
end of the screw.
Twin-screw granulation has been studied in much detail in recent years [16,17]. Sev-
eral experimental studies have investigated the effects of key process variables [26–37],
screw configurations [24,25,38–42] and formulation variables [26,36,40], and have resulted
in a regime map [43] of the twin-screw granulator. However, these experimental studies
cannot provide any insight into the micro-scale phenomena that govern granulation pro-
cesses. Particle-scale simulations have the potential to provide this insight. The Discrete
Element Method (DEM) has become the dominant particle-scale simulation tool in the
last 20 years. DEM has been applied to investigate various wet granulation processes,
e.g., Mishra et al. [44] and Liu et al. [45] investigated granulation in a rotating drum;
Goldschmidt et al. [46] and Kafui et al. [47,48] simulated fluidised bed granulation; Gantt
and Gatzke [49], Hassanpour et al. [50], Nakamura et al. [51], Watson et al. [52], Tamrakar
et al. [53] and Börner et al. [54] studied high shear granulation. Researchers have also ex-
plored granulation processes by coupling DEM and population balance modelling [55–59].
However, there are few prior DEM studies of twin-screw granulation. Dhenge et al. [29]
studied the effect of binder amount and binder viscosity at varying powder feed rates on
the granulation behaviour and final granule properties. Due to the large computational
cost, their simulations use conveying screws only within a 16 mm periodic domain. They
also do not include cohesion between particles in the simulations. Only a qualitative
visual comparison was made to the experimental results. Recently Zheng et al. [60] have
used DEM to study residence time distributions for dry, cohesionless, mono-sized elastic
spheres of various sizes for a fixed configuration of conveying elements with two blocks of
kneading elements at various operating RPMs. Zheng et al. [61] extended the study further
to include the effect of particle shape in the TSG. Spherical particles were observed to have
69
Pharmaceutics 2021, 13, 2136
slightly lower mean residence times than particles of other shapes, which were largely the
same, regardless of shape. The influence of the many DEM parameters on the residence
time has not been investigated in these studies, which only considered dry cohesionless
materials. Kumar et al. [62] used mono-disperse periodic DEM simulations to study liquid
distributions within the mixing zone of a GEA ConsiGma 25 TSG; however, this zone
was only one kneading element thick (8× particle diameters). While liquid transfer and
associated agglomeration could be observed, the limited domain means no longitudinal
travel or transfer from conveying elements were considered in this model, providing a
somewhat isolated overview of the dynamics within a TSG.
The objectives of this research study are threefold:
• To develop and verify computationally efficient reduced-scale DEM models that
closely match full-scale simulations of the entire granulator;
• To carry out a comprehensive sensitivity study to explore how operating parameters
and DEM model parameters affect the performance of a TSG, and thereby identify the
subset of influential parameters requiring calibration (noting that this study does not
attempt to perform a full calibration of the granular material against a real solid);
• To demonstrate that the effect of liquid binder can be captured using an adhesive,
elasto-plastic contact model.
In order to create a numerical model that is capable of producing numerical predictions
that can be validated against experimental results, it is necessary to first verify that the
conceptual model is operating in the correct manner, similar to its real-life counterpart.
As such, the operation of the TSG model is verified by comparing the particle dynamics,
flow characteristics and residence time distributions against previously reported sources.
Validation is concerned with establishing the predictive capability of the computational
model in relation to the real system. Computational results should be compared with
independent experimental results, and a computational model will be considered validated
once the discrepancy between the simulation and experimental result of interest falls
within pre-determined bounds. This discrepancy must be evaluated against the degree of
uncertainty in both the validation experiment and the multiscale model. Validation of the
model outputs against careful experimental measurements such as granule porosities and
granule size distributions is not part of this study.
70
Pharmaceutics 2021, 13, 2136
Property Value
Barrel Diameter, D [mm] 25
Channels [-] 2
Channel Depth [mm] 2.58
Centre line Distance [mm] 19.31
Barrel–Screw Clearance [mm] 0.440
Screw–Screw Clearance [mm] 0.175
Area of Double Intermeshed Barrels [mm2 ] 920.0
Free Barrel Area [mm2 ] 435.26
Free Barrel Fraction 0.4731
Property Value
Screw Radius—Outer, R [mm] 12.06
Screw Radius—Inner, r [mm] 6.90
Screw Diameter Ratio 1.75
Screw Transition Radius, R + r [mm] 18.96
Screw Length, L [mm] 500
Screw Lead, S [mm] 25
Screw Section Perimeter [mm] 59.56
Screw Section Area [mm2 ] 242.37
Screw Transition Angle 38.15◦
Screw Tip Angle 13.72◦
Figure 1. CAD model of ConsiGma 1 screw. Top row, from left to right: (i) Conveying element of
length 37.5 mm (ii) Kneading element (iii) 6 × 60 R block of kneading elements; Bottom row: (iv)
6 × 60 R block of kneading elements with adjoining conveying elements.
71
Pharmaceutics 2021, 13, 2136
The stainless-steel barrel and screws were modelled as rigid bodies in EDEM. Appro-
priate rotational dynamics were added for the chosen screw speed. The screw configuration
used in the simulations is that of a typical experimental setup, summarised in Figure 3,
which consists of two blocks of six kneading elements with a 60◦ offset in a reversing con-
figuration, separated by a 1.5D conveying element. Two chopping elements were located
at the screw exit to break up any oversized agglomerates that may have formed.
Figure 3. Typical experimental screw configuration, as used in the full-scale DEM simulations. CH represents a chopping
element, C1.5 represents a conveying element of length 1.5 times the screw lead and K denotes a kneading element.
Both cohesionless and cohesive systems were studied. Cohesionless systems are first
introduced to study the effect of parameters such as friction and restitution, in recogni-
tion of the fact that most DEM simulations of twin-screw granulators still adopt simple
cohesionless models. The subsequent inclusion of cohesion allows its influence to be
distinguished from the other parameters.
72
Pharmaceutics 2021, 13, 2136
Breakage of the primary particles was not considered in the model. However, agglom-
erates formed by cohesion were able to break and reform, similar to real agglomeration
behaviour. Liquid migration was not considered in the model to reduce the computational
cost. Values of key input parameters used for the cohesionless simulations are given in
Table 3. These parameters form the reference case for the sensitivity study presented later.
The chosen bulk density is that of a typical pharmaceutical solid: paracetamol. The remain-
ing parameters are chosen as typical median values that are representative of a cohesive
powder. The “standard deviation” of 0.05 in Table 3 defines the width of the normal
Gaussian distribution in particle diameter, about a mean value of 400 μm. This distribution
is truncated at an upper and lower limit (0.5x–1.5x), relative to the mean. The effect of
cohesion in the TSG is considered through a set of cohesion parameters that result in high
levels of cohesion (given later in Table 4). These parameters have not been calibrated
for any specific material or level of liquid binder and are simply to explore the effect of
cohesion in the system and its relative importance.
73
Pharmaceutics 2021, 13, 2136
Table 4. Variations from the reference case defined in Table 3, where χ, μs and μr apply to both
particle–particle and particle–geometry interactions.
Reference Variations
Coefficient of restitution, χ 0.5 0.1, 0.9
Static friction, μs 0.5 0.1
Rolling friction, μr 0.01 0.2
Relative Standard
0.05 0.2
Deviation
Cohesion Cohesionless High (γ = 15 J/m2 , f 0 = −0.001 N)
74
Pharmaceutics 2021, 13, 2136
artefacts of the discretistation of the CAD geometry. It is also important to ensure that
small, well-shaped meshes are used and to avoid any stretched triangles, especially in the
barrel, as this could cause some strange effects.
Finally, adding large amounts of cohesion to the system can cause many problems
such as creating excessive overlaps at collisions to forming a single agglomerate containing
all particles in the simulation. It should also be noted that even if the amount of cohesion
used is not excessive, it can still lead to the build-up of material in the kneading elements
that would cause jamming of the real granulator. In this numerical situation, rather than
the screws stopping, the particles will just be forced through the barrel walls.
75
Pharmaceutics 2021, 13, 2136
of the particles. This can make periodic sub-domain models longer than non-periodic
sub-domain models, which do not need to recycle particles. Periodic sub-domain models
will have a fixed number of particles, which means they cannot enforce a fixed flow rate,
whereas non-periodic sub-domain models utilise dynamic particle generation and can be
used to study the granular behaviour under fixed flow rates. Periodic sub-domain models
cannot be used to study the developing granule size distribution as formed granules would
be recycled through the initial flow boundary instead of virgin feed. In this type of study
some models with continuous generation would be required.
76
Pharmaceutics 2021, 13, 2136
Figure 7. Cont.
77
Pharmaceutics 2021, 13, 2136
The Residence Time Distribution (RTD) for the full TSG simulation was extracted,
which shows a mean total residence time of approximately 3 s. This is in the region of what
has been measured experimentally for a free-flowing material in similar
equipment [7,27,28,39]. RTDs for the full-size DEM simulation and the periodic simu-
lation in the same domain, over a 1 s period at steady state, are compared in Figure 8.
The periodic simulation is capturing the same behaviour as the full-size simulation. For the
total residence time, the periodic results include a spike at 0 s, which is an artefact due to
the re-circulation of particles through the periodic domain and the time they spend in the
buffer zone at each end. The agreement is also excellent for the RTDs by element type.
78
Pharmaceutics 2021, 13, 2136
Figure 8. Comparison of the Residence Time Distributions (RTDs) for the full model and
periodic simulation.
79
Pharmaceutics 2021, 13, 2136
the respective solid fractions and residence time values, whereas this is less clear for the
velocity magnitudes, which are similar across all kneading elements. The velocity magni-
tude, resulting from the longitudinal and in-plane movement, is, however, much higher
for kneading elements than for conveying elements due to the rotational path enforced by
the geometry and is, therefore, heavily influenced by the screw RPM.
Figure 9. Comparison between periodic and full-scale simulations for various measured quantities.
80
Pharmaceutics 2021, 13, 2136
Figure 10. Effect of particle size on solid fraction, velocity and residence time.
81
Pharmaceutics 2021, 13, 2136
Figure 11. Particle snapshots at steady state as the mean particle diameter is varied. The particle colour indicates longitudinal
velocity in the direction parallel to the screws; a negative velocity indicates the movement of particles towards the outlet of
the granulator.
The solid fraction results on conveying elements fall into three separate classes for
small (<500 μm), medium (500–1000 μm) and large (>1000 μm) particles. Large particles
lead to the lowest solid fraction on conveying elements due to the geometric constraints of
82
Pharmaceutics 2021, 13, 2136
the screw geometry: larger spheres cannot occupy the space around the screws as efficiently
as smaller spheres. The kneading block of six elements appears to act as two blocks of
three elements, with significantly different behaviours observed on the first three and the
last three kneading elements. In the last three elements, the observed solid fractions are
largely the same for all particle sizes, apart from the large 1500 μm particles. The first three
kneading elements show significant variation with the larger particles generally having a
higher solid fraction.
For the longitudinal (X) velocity of the particles along the barrel, the largest particles
have the lowest longitudinal velocity on the kneading elements, suggesting that these
particles experience more obstructions by geometrical constraints at the transitions between
elements, hampering their path through the granulator. All particle sizes follow the same
general trend where there is a significant velocity decrease at the transfer from conveying
to kneading elements. However, the magnitude of the retardation is heavily dependent
on particle size, with the larger particles reduced to the lowest velocity. The 1500 μm
particles do not follow the general trend of significantly increasing longitudinal velocity
as the particles progress through the kneading block, with the velocity remaining almost
constant until the last kneading element before an increase is noted. All other sizes show a
clearer velocity increase along the kneading block and have returned to the original screw
translational velocity by the time they have left the kneading block.
Figure 10 also shows a consistent trend of increasing velocity magnitude with increas-
ing size, with an almost constant value for each element type. Interestingly, there is a
large variation in the velocity magnitudes for differently sized particles in the conveying
elements, with larger particles having larger velocities. This appears to be a result of the
number of particles in the system. As the size of the particles increases, the number of
particles in the system decreases, which can lead to decreasing levels of particle–particle
collisions in the system. With only several large particles able to sit on a conveying element
at once, there are relatively few interparticle contacts, leading to lower energy dissipation.
This results in the larger particles being more active in the screw elements.
A similar trend of high velocity magnitude for larger particles is also observed in
the kneading elements. Within the kneading zone, the increased velocities can also be
attributed to the reduction in particle numbers. In the kneading zone, particle collisions
play a significant role in helping particles progress along the screw. Without the particle
collisions to help “nudge” particles onto the next element, the larger particles have a
tendency to remain circulating in-plane, being pushed around by the flat lobes of the
kneading elements. There is also a geometric constraint at play in the kneading zone,
with the larger particles more likely to collide with the next kneading element and remain
on the current element than smaller particles, which are less likely to be constrained in
this manner. These factors contribute to the reduced longitudinal velocity in the larger
particles, which is also reflected in the observed residence times that are highest for the
largest particles.
Pradhan et al. [77] found from experiments and geometrical analysis on static 3D
CAD models that particle breakage was related to the maximum void space on an element,
with 20% of particles on mixing elements broken at just 60% of the maximum size and
all particles suffering some breakage at less than the maximum size. This geometrical
breakage constraint was recently implemented in a new population balance model kernel
for particle breakage [78] and should be considered when choosing particle sizes for
simulations so as to not adversely affect the simulation results. The simulation results
in the current study suggest the existence of a critical ratio of particle diameter to the
screw void space (effectively the difference between the screw’s root and outer radii) for
unhindered particle flow that is significantly less than the breakage limit identified by
Pradhan et al. [77]. For the configuration used in this study, the maximum dimensions
within the void spaces are approximately 5.5 mm and 6.25 mm for conveying and kneading
elements, respectively, suggesting a diameter-to-void limit of approximately 10–20% for
unhindered particle flow. This limit may also be affected by both the offset angle and
83
Pharmaceutics 2021, 13, 2136
direction of the kneading/mixing elements, which could reduce the maximum dimensions
within the void spaces below the figures quoted here.
The significance of the particle size in relation to the screw–barrel clearance can also
be observed in Figure 10. This critical clearance is 440 μm (Tables 1 and 2) in this model,
which means approximately half of the particle sizes studied will not pass through this
outer region. With the exception of kneading element 1, the particle behaviour that is
dominated by collisions in the central intermeshing region, there is little difference between
the particle sizes that fit through this outer gap (200 μm, 400 μm) and those that are too
large (1500 μm particles excluded) to pass through. This suggests that the screw–barrel
clearance is not a significant concern when choosing particle size.
Finally, it is worth considering the qualitative particle dynamics when investigating
particle size effects. Figure 11 compares the snapshots at the same time instant for all
simulations. For the 1500 μm particles, there is no clear or obvious pattern in the particle
dynamics due to the low number of particles, with just under 1200 existing at this particle
size. However, a flow pattern starts to appear when the particle size is decreased to
1000 μm (approximately 4100 particles). At 800 μm and 600 μm, this “x-shaped” crossover
pattern becomes increasingly well defined. Further reductions in particle size do not
significantly enhance this flow pattern. This was checked for particles as small as 100 μm
(Figure 11g). The 400 μm reference case is a good compromise between computational
efficiency and capturing the particle dynamics of the system, and 400 μm has been selected
as the reference size for the study. Omitted from the comparison in Figure 10 were 100 μm
particles due to the large number of particles (approx. 4.1 M in total) and limited simulation
time limiting the amount of available data.
84
Pharmaceutics 2021, 13, 2136
hence, the observed residence times at the same mass flow rate. The average velocity
magnitude, which includes the in-plane particle movement, remains almost identical for
both distributions.
Figure 12. Effect of particle size distribution on solid fraction, velocity and residence time.
85
Pharmaceutics 2021, 13, 2136
Figure 13. Effect of coefficient of restitution on solid fraction, velocity and residence time.
However, increasing χ for both the particle–geometry contacts and the interparticle
contacts leads to a significant effect on the average solid fraction for both kneading and
conveying elements. Low particle–geometry damping leads to more chaotic behaviour
as less energy is dissipated. The increased chaotic behaviour appears to be preventing
particles from freely moving axially through the granulator, shown by the reduced axial
velocity for kneading elements in Figure 13, which in turn leads to higher residence times
for the kneading elements. While the longitudinal velocity is decreased when the particle–
geometry coefficient of restitution is increased, the opposite trend is seen in the average
velocity magnitude. This shows that there is significantly increased in-plane velocities,
which leads to the larger velocity magnitude. Provided that a sensible value is chosen,
the TSG model is relatively insensitive to the amount of damping between particles or
geometries, with the particle–geometry coefficient of restitution being the more influential.
This effect is likely to be relative to the screw speed, with higher screw speeds likely to
exaggerate the effect.
86
Pharmaceutics 2021, 13, 2136
used a relatively high value of μs = 0.5. Increasing μs beyond this value will not cause
very large changes as the effect of particle friction tends to saturate. This limited study
investigated the effect of friction by decreasing μs to 0.1 for both particle–particle (μs,pp )
and particle–geometry (μs,pg ) contacts, individually and in tandem. When the averaged
solid fraction for the four cases is studied (Figure 14), the results fall into two distinct
groups distinguished by the μs,pg values. A lower friction coefficient leads to significantly
lower solid fractions across all elements, particularly for the kneading elements. The trend
is less well defined for the first and second kneading elements due to the nearby transition
from conveying to kneading. The division into two groups defined by μs,pg is even more
apparent in the longitudinal velocity (also in Figure 14). Larger μs,pg leads to much lower
longitudinal velocities as the increased friction leads to increased shear along the barrel
surface, retarding flow. This also increases the residence time for the higher μs,pg . As
particles’ longitudinal velocity reduce and they spend longer on each element, the in-plane
velocity also increases due to the longer time spent being rotated by the element. This leads
to an increase in the velocity magnitude, especially for the kneading element at which
the longitudinal velocity is the lowest. In general, μs,pp has minimal influence in the TSG,
which appears to be a system dominated by geometry interactions and geometry dynamics.
Figure 14. Effect of static friction coefficient on solid fraction, velocity and residence time.
87
Pharmaceutics 2021, 13, 2136
Rolling friction is a commonly used DEM approach to incorporate the effect of particle
shape without the computational expense of simulating non-spherical particles. It works by
applying an additional torque to resist particle rolling. The reference case used a minimal
value of 0.01 as the rolling friction coefficient. In the same manner as for static friction,
the opposite extreme is explored for particle–particle (μr,pp ) and particle–geometry (μr,pg )
coefficients to assess the effect on the TSG model.
Figure 15 shows that rolling friction has a minimal effect on the average solid fraction
on most elements, except for the first and second kneading elements. More particles are
held up on these elements when either rolling friction coefficient increases. Increasing
rolling friction leads to reduced longitudinal velocities across all elements. Increasing μr,pp
and μr,pg has a similar effect, although the former seems to have a greater significance: the
longitudinal velocity reduced more for particle–particle contacts than for particle–geometry
contacts. The combination of both leads to the largest reduction in velocity. The average
residence time (Figure 15) is strongly linked to the average longitudinal velocity and
displays the same trends, with μr,pp having the most significant effect.
Figure 15. Effect of rolling friction coefficient on solid fraction, velocity and residence time.
88
Pharmaceutics 2021, 13, 2136
Figure 16. Effect of level of cohesion on solid fraction, velocity and residence time.
89
Pharmaceutics 2021, 13, 2136
Figure 17. Agglomerate formation in the non-periodic reduced domain model (direction of flow:
right to left).
90
Pharmaceutics 2021, 13, 2136
Capturing the formation of agglomerates within the TSG is a key aspect of the model,
and this is accomplished though the inclusion of the cohesive forces, which represent the
liquid binder in the real granulator. No changes in primary particle sizes occur in the
model; instead, primary particles agglomerate because of the cohesive forces.
4. Discussion
In this study, the reference case was the full-scale, cohesionless system, which had a
mass flow rate set at 14.4 kg/h. The initial mass of the periodic system was calculated from
the steady-state operation of the full-size model. The computationally efficient periodic
system is then used extensively to assess how the DEM input parameters would influence
the behaviour in the granulator. The results in Section 3 have considered how the particle
dynamics in the various elements were affected in terms of velocity, solid fraction and
residence time. However, due to the periodic system having a fixed mass, it is possible to
measure the change in mass flow rate on the conveying elements at the inlet and outlet of
the periodic domain. Any significant changes to the bulk behaviour will result in variations
in the observed mass flow rates and a difference from the full-scale model value, which
can be used to assess the magnitude of a parameter effect. The larger the difference is, the
more influential that parameter is.
To consider this effect and the sensitivity of the mass flow rate to the parameters, the
mass flow rates were calculated at sensors (MFS_1 and MFS_2) at each end of the periodic
system and plotted for each case in Figure 18. Steady state exists in the system when the
averaged mass flow rate is approximately equal for both the inlet and outlet sensor.
Figure 18. Observed changes in mass flow rate in the periodic system.
For a changing size distribution and low restitution, there is almost no change from
the expected 14.4–14.6 kg/h mass flow rate. The decrease in damping at χ = 0.9 leads to
slightly higher solid fractions being observed for the kneading elements, which had the
effect of slightly reducing the mass flow to 13.7–14.1 kg/h.
The reduction in static friction (μs ) led to reduced solid fractions, which manifested as
a significant increase in the mass flow rate of approximately 5 kg/h to just over 20 kg/h.
Increasing rolling friction, μr , led to an increase in solid fraction for some elements, which
led to a reduction in the mass flow rate to approximately 12–13.3 kg/h depending on the
combination of parameters used. The effect of cohesion is the most dramatic, with the
mass flow rates reduced by just over 10 kg/h for the cohesive case. This is due to the
significant agglomeration and build-up of solids on the kneading elements, which reduces
91
Pharmaceutics 2021, 13, 2136
the circulating mass in the periodic system. The dominant effect of cohesion explains why
it was considered separately from other parameters, allowing the relative importance of
each to be quantified without being outweighed by the more dominant parameters.
5. Conclusions
Discrete Element Method (DEM) simulations were used in this study of the particle
dynamics that occur in a twin-screw granulator comprising both conveying and kneading
elements. Both full-scale and reduced periodic models were used.
The DEM simulations for a cohesionless system provide a mean residence time—a key
characteristic of a TSG—that is in line with expectations when compared to experimental
results with various different granulators and cohesive materials. The residence time
increased, as expected, when an adhesive, elasto-plastic contact model was used to capture
the effect of liquid binder and the resulting agglomeration in the granulator.
The significant computational expense of the full-scale simulations was reduced
through (carefully defined) periodic simulations. Results show that the residence time
distributions across elements in a periodic simulation are almost identical to their coun-
terparts in the full-scale simulation. These reduced domain models allow the key particle
dynamics to be captured at a much lower computational cost, making larger studies more
feasible in the future. These models, combined with a DEM contact model that includes
cohesion, are an effective way to simulate and investigate the wet granulation process.
A sensitivity study showed that the level of cohesion is a key determinant of granule
size and must be carefully calibrated using experiments. An important finding is that the
size of the fundamental particles in the simulation may be larger than in reality while still
capturing the correct dynamics for the TSG. The results suggest the existence of a critical
ratio of particle diameter to the screw void space (effectively the difference between the
screw’s root and outer radii). Once the particle size is below a certain threshold, the velocity,
solid fraction and residence time are largely unaffected by particle size, with the exception
of the first kneading element where there is the transition from the conveying elements.
This is a highly chaotic region that appears to be heavily influenced by particle size. This
effect was also noted when a broader particle size distribution was used; the increased
numbers of larger particles lead to slightly increased solid fractions and residence times on
the first kneading element only. The effect of the screw–barrel clearance is also found to be
negligible with particles that are both smaller and larger than this critical size behaving the
same in most locations in the granulator.
The effect of restitution coefficient (damping) was relatively insignificant for materials
not considered to be elastic materials. The particle–geometry damping is more influential
than the particle–particle damping, which is a feature of a system where geometries are
moving at very high speeds, leading to high relative impact velocities. Similar trends were
seen for static friction: reducing the particle–geometry static friction leads to significant
changes in the solid fraction and residence times. Rolling friction has a lesser effect than
static friction, in general. Increased cohesion leads to increased solid fractions and residence
times on kneading elements. The results suggest that the TSG is a system that is domi-
nated by geometry interactions, with velocities very much controlled by the screw speed
and pitch.
Author Contributions: Conceptualization, J.P.M., K.J.H. and J.Y.O.; methodology, J.P.M., K.J.H. and
J.Y.O.; software, J.P.M.; validation, J.P.M., K.J.H. and J.Y.O.; formal analysis, J.P.M.; visualization,
J.P.M.; investigation, J.P.M.; data curation, J.P.M.; writing—original draft preparation, J.P.M.; writing—
review and editing, J.P.M., K.J.H. and J.Y.O.; supervision, J.Y.O.; project administration, J.Y.O.;
resources, J.Y.O.; funding acquisition, J.Y.O. All authors have read and agreed to the published
version of the manuscript.
Funding: This research was funded by the UK’s Centre for Process Innovation (CPI) as part of the
project “Models for Manufacturing of Particulate Products” (MMPP).
Institutional Review Board Statement: Not applicable.
92
Pharmaceutics 2021, 13, 2136
References
1. Cameron, I.T.; Wang, F.Y.; Immanuel, C.D.; Stepanek, F. Process Systems Modelling and Applications in Granulation: A Review.
Chem. Eng. Sci. 2005, 60, 3723–3750. [CrossRef]
2. Barrasso, D.; Ramachandran, R. Multi-Scale Modeling of Granulation Processes: Bi-Directional Coupling of PBM with DEM via
Collision Frequencies. Chem. Eng. Res. Des. 2015, 93, 304–317. [CrossRef]
3. Wang, F.Y.; Cameron, I.T. Review and Future Directions in the Modelling and Control of Continuous Drum Granulation. Powder
Technol. 2002, 124, 238–253. [CrossRef]
4. Department of Health and Human Services. Pharmaceutical CGMPs for the 21st Century—A Risk-Based Approach; FDA: Silver
Spring, MD, USA, 2004.
5. Litster, J.D. Scaleup of Wet Granulation Processes: Science Not Art. Powder Technol. 2003, 130, 35–40. [CrossRef]
6. Liu, L.; Litster, J.D.; Iveson, S.; Ennis, B. Coalescence of Deformable Granules in Wet Granulation Processes. AIChE J. 2000, 46,
529–539. [CrossRef]
7. Lee, K.T.; Ingram, A.; Rowson, N.A. Twin Screw Wet Granulation: The Study of a Continuous Twin Screw Granulator Using
Positron Emission Particle Tracking (PEPT) Technique. Eur. J. Pharm. Biopharm. 2012, 81, 666–673. [CrossRef]
8. Li, H.; Thompson, M.R.; O’Donnell, K.P. Understanding Wet Granulation in the Kneading Block of Twin Screw Extruders. Chem.
Eng. Sci. 2014, 113, 11–21. [CrossRef]
9. Iveson, S.M.; Litster, J.D.; Hapgood, K.P.; Ennis, B.J. Nucleation, Growth and Breakage Phenomena in Agitated Wet Granulation
Processes: A Review. Powder Technol. 2001, 117, 3–39. [CrossRef]
10. Kayrak-Talay, D.; Dale, S.; Wassgren, C.R.; Litster, J.D. Quality by Design for Wet Granulation in Pharmaceutical Processing:
Assessing Models for a Priori Design and Scaling. Powder Technol. 2013, 240, 7–18. [CrossRef]
11. El Hagrasy, A.S.; Litster, J.D. Granulation Rate Processes in the Kneading Elements of a Twin Screw Granulator. AIChE J. 2013, 59,
4100–4115. [CrossRef]
12. Hapgood, K.P.; Lveson, S.M.; Litster, J.D.; Liu, L.X. Granulation Rate Processes. In Handbook of Powder Technology; Salman, A.D.,
Hounslow, M.J., Seville, J.P.K., Eds.; Elsevier Science B.V.: Amsterdam, The Netherlands, 2007; Volume 11, pp. 897–977. ISBN
9780444518712.
13. Gamlen, M.J.; Eardley, C. Continuous Extrusion Using a Raker Perkins MP50 (Multipurpose) Extruder. Drug Dev. Ind. Pharm.
1986, 12, 1701–1713. [CrossRef]
14. Lindberg, N.-O.; Tufvesson, C.; Holm, P.; Olbjer, L. Extrusion of an Effervescent Granulation with a Twin Screw Extruder, Baker
Perkins MPF 50 D. Influence on Intragranular Porosity and Liquid Saturation. Drug Dev. Ind. Pharm. 1988, 14, 1791–1798.
[CrossRef]
15. Ghebre-Sellassie, I.; Mollan, M.J.; Pathak, N.; Lodaya, M.; Fessehaie, M. Continuous Production of Pharmaceutical Granulation
2000. U.S. Patent No. 6,499,984, 31 December 2002.
16. Seem, T.C.; Rowson, N.A.; Ingram, A.; Huang, Z.; Yu, S.; de Matas, M.; Gabbott, I.; Reynolds, G.K. Twin Screw Granulation—A
Literature Review. Powder Technol. 2015, 276, 89–102. [CrossRef]
17. Thompson, M.R. Twin Screw Granulation—Review of Current Progress. Drug Dev. Ind. Pharm. 2015, 41, 1223–1231. [CrossRef]
18. Keleb, E.I.; Vermeire, A.; Vervaet, C.; Remon, J.P. Extrusion Granulation and High Shear Granulation of Different Grades of
Lactose and Highly Dosed Drugs: A Comparative Study. Drug Dev. Ind. Pharm. 2004, 30, 679–691. [CrossRef]
19. Keleb, E.I.; Vermeire, A.; Vervaet, C.; Remon, J.P. Twin Screw Granulation as a Simple and Efficient Tool for Continuous Wet
Granulation. Int. J. Pharm. 2004, 273, 183–194. [CrossRef]
20. Rauwendaal, C. The Geometry of Self-Cleaning Twin-Screw Extruders. Adv. Polym. Technol. 1996, 15, 127–133. [CrossRef]
21. Rauwendaal, C. Twin Screw Extruders. In Polymer Extrusion; Carl Hanser Verlag GmbH & Co. KG: München, Germny, 2014; pp.
697–761.
93
Pharmaceutics 2021, 13, 2136
22. Rauwendaal, C. Extruder Screw Design. In Polymer Extrusion; Carl Hanser Verlag GmbH & Co. KG: München, Germny, 2014; pp.
509–652.
23. Barrasso, D.; el Hagrasy, A.; Litster, J.D.; Ramachandran, R. Multi-Dimensional Population Balance Model Development and
Validation for a Twin Screw Granulation Process. Powder Technol. 2015, 270, 612–621. [CrossRef]
24. Djuric, D. Continuous Granulation with a Twin-Screw Extruder; Cuvillier Verlag: Göttingen, Germany, 2008; ISBN 3867276439.
25. Djuric, D.; Kleinebudde, P. Impact of Screw Elements on Continuous Granulation with a Twin-Screw Extruder. J. Pharm. Sci. 2008,
97, 4934–4942. [CrossRef]
26. Dhenge, R.M.; Cartwright, J.J.; Hounslow, M.J.; Salman, A.D. Twin Screw Wet Granulation: Effects of Properties of Granulation
Liquid. Powder Technol. 2012, 229, 126–136. [CrossRef]
27. Dhenge, R.M.; Cartwright, J.J.; Doughty, D.G.; Hounslow, M.J.; Salman, A.D. Twin Screw Wet Granulation: Effect of Powder Feed
Rate. Adv. Powder Technol. 2011, 22, 162–166. [CrossRef]
28. Dhenge, R.M.; Fyles, R.S.; Cartwright, J.J.; Doughty, D.G.; Hounslow, M.J.; Salman, A.D. Twin Screw Wet Granulation: Granule
Properties. Chem. Eng. J. 2010, 164, 322–329. [CrossRef]
29. Dhenge, R.M.; Washino, K.; Cartwright, J.J.; Hounslow, M.J.; Salman, A.D. Twin Screw Granulation Using Conveying Screws:
Effects of Viscosity of Granulation Liquids and Flow of Powders. Powder Technol. 2013, 238, 77–90. [CrossRef]
30. Lute, S.v.; Dhenge, R.M.; Hounslow, M.J.; Salman, A.D. Twin Screw Granulation: Understanding the Mechanism of Granule
Formation along the Barrel Length. Chem. Eng. Res. Des. 2016, 110, 43–53. [CrossRef]
31. Lute, S.v.; Dhenge, R.M.; Salman, A.D. Twin Screw Granulation: An Investigation of the Effect of Barrel Fill Level. Pharmaceutics
2018, 10, 67. [CrossRef]
32. Lute, S.v.; Dhenge, R.M.; Salman, A.D. Twin Screw Granulation: Effects of Properties of Primary Powders. Pharmaceutics 2018,
10, 68. [CrossRef]
33. Vercruysse, J.; Córdoba Díaz, D.; Peeters, E.; Fonteyne, M.; Delaet, U.; van Assche, I.; de Beer, T.; Remon, J.P.; Vervaet, C.
Continuous Twin Screw Granulation: Influence of Process Variables on Granule and Tablet Quality. Eur. J. Pharm. Biopharm. 2012,
82, 205–211. [CrossRef] [PubMed]
34. Vercruysse, J.; Delaet, U.; van Assche, I.; Cappuyns, P.; Arata, F.; Caporicci, G.; de Beer, T.; Remon, J.P.; Vervaet, C. Stability
and Repeatability of a Continuous Twin Screw Granulation and Drying System. Eur. J. Pharm. Biopharm. 2013, 85, 1031–1038.
[CrossRef] [PubMed]
35. Fonteyne, M.; Vercruysse, J.; de Leersnyder, F.; Besseling, R.; Gerich, A.; Oostra, W.; Remon, J.P.; Vervaet, C.; de Beer, T. Blend
Uniformity Evaluation during Continuous Mixing in a Twin Screw Granulator by In-Line NIR Using a Moving F-Test. Anal.
Chim. Acta 2016, 935, 213–223. [CrossRef] [PubMed]
36. Kyttä, K.M.; Lakio, S.; Wikström, H.; Sulemanji, A.; Fransson, M.; Ketolainen, J.; Tajarobi, P. Comparison between Twin-Screw
and High-Shear Granulation—The Effect of Filler and Active Pharmaceutical Ingredient on the Granule and Tablet Properties.
Powder Technol. 2020, 376, 187–198. [CrossRef]
37. Ryckaert, A.; Stauffer, F.; Funke, A.; Djuric, D.; Vanhoorne, V.; Vervaet, C.; de Beer, T. Evaluation of Torque as an In-Process
Control for Granule Size during Twin-Screw Wet Granulation. Int. J. Pharm. 2021, 602, 120642. [CrossRef]
38. Vercruysse, J.; Toiviainen, M.; Fonteyne, M.; Helkimo, N.; Ketolainen, J.; Juuti, M.; Delaet, U.; van Assche, I.; Remon, J.P.;
Vervaet, C.; et al. Visualization and Understanding of the Granulation Liquid Mixing and Distribution during Continuous Twin
Screw Granulation Using NIR Chemical Imaging. Eur. J. Pharm. Biopharm. 2014, 86, 383–392. [CrossRef]
39. Kumar, A.; Vercruysse, J.; Toiviainen, M.; Panouillot, P.E.; Juuti, M.; Vanhoorne, V.; Vervaet, C.; Remon, J.P.; Gernaey, K.v.; de
Beer, T.; et al. Mixing and Transport during Pharmaceutical Twin-Screw Wet Granulation: Experimental Analysis via Chemical
Imaging. Eur. J. Pharm. Biopharm. 2014, 87, 279–289. [CrossRef]
40. el Hagrasy, A.S.; Hennenkamp, J.R.; Burke, M.D.; Cartwright, J.J.; Litster, J.D. Twin Screw Wet Granulation: Influence of
Formulation Parameters on Granule Properties and Growth Behavior. Powder Technol. 2013, 238, 108–115. [CrossRef]
41. Li, J.; Pradhan, S.U.; Wassgren, C.R. Granule Transformation in a Twin Screw Granulator: Effects of Conveying, Kneading, and
Distributive Mixing Elements. Powder Technol. 2019, 346, 363–372. [CrossRef]
42. Liu, H.; Ricart, B.; Stanton, C.; Smith-Goettler, B.; Verdi, L.; O’Connor, T.; Lee, S.; Yoon, S. Design Space Determination and
Process Optimization in At-Scale Continuous Twin Screw Wet Granulation. Comput. Chem. Eng. 2019, 125, 271–286. [CrossRef]
43. Da Tu, W.; Ingram, A.; Seville, J. Regime Map Development for Continuous Twin Screw Granulation. Chem. Eng. Sci. 2013, 87,
315–326. [CrossRef]
44. Mishra, B.K.; Thornton, C.; Bhimji, D. A Preliminary Numerical Investigation of Agglomeration in a Rotary Drum. Miner. Eng.
2002, 15, 27–33. [CrossRef]
45. Liu, P.Y.; Yang, R.Y.; Yu, A.B. Particle Scale Investigation of Flow and Mixing of Wet Particles in Rotating Drums. Chem. Eng. Sci.
2013, 86, 99–107. [CrossRef]
46. Goldschmidt, M.J.v.; Weijers, G.G.C.; Boerefijn, R.; Kuipers, J.A.M. Discrete Element Modelling of Fluidised Bed Spray Granulation.
Powder Technol. 2003, 138, 39–45. [CrossRef]
47. Kafui, D.K.; Thornton, C. Fully-3D DEM Simulation of Fluidised Bed Spray Granulation Using an Exploratory Surface Energy-
Based Spray Zone Concept. Powder Technol. 2008, 184, 177–188. [CrossRef]
48. Kafui, K.D.; Thornton, C.; Adams, M.J. Discrete Particle-Continuum Fluid Modelling of Gas-Solid Fluidised Beds. Chem. Eng. Sci.
2002, 57, 2395–2410. [CrossRef]
94
Pharmaceutics 2021, 13, 2136
49. Gantt, J.A.; Gatzke, E.P. High-Shear Granulation Modeling Using a Discrete Element Simulation Approach. Powder Technol. 2005,
156, 195–212. [CrossRef]
50. Hassanpour, A.; Pasha, M.; Susana, L.; Rahmanian, N.; Santomaso, A.C.; Ghadiri, M. Analysis of Seeded Granulation in High
Shear Granulators by Discrete Element Method. Powder Technol. 2013, 238, 50–55. [CrossRef]
51. Nakamura, H.; Fujii, H.; Watano, S. Scale-up of High Shear Mixer-Granulator Based on Discrete Element Analysis. Powder Technol.
2013, 236, 149–156. [CrossRef]
52. Watson, N.J.; Povey, M.J.W.; Reynolds, G.K.; Ding, Y.; Xu, B.H. Development of a Discrete Element Model with Moving Realistic
Geometry to Simulate Particle Motion in a Mi-Pro Granulator. Comput. Chem. Eng. 2016, 93, 234–247. [CrossRef]
53. Tamrakar, A.; Chen, S.W.; Ramachandran, R. A DEM Model-Based Study to Quantitatively Compare the Effect of Wet and Dry
Binder Addition in High-Shear Wet Granulation Processes. Chem. Eng. Res. Des. 2019, 142, 307–326. [CrossRef]
54. Börner, M.; Michaelis, M.; Siegmann, E.; Radeke, C.; Schmidt, U. Impact of Impeller Design on High-Shear Wet Granulation.
Powder Technol. 2016, 295, 261–271. [CrossRef]
55. Sen, M.; Barrasso, D.; Singh, R.; Ramachandran, R. A Multi-Scale Hybrid CFD-DEM-PBM Description of a Fluid-Bed Granulation
Process. Processes 2014, 2, 89–111. [CrossRef]
56. Barrasso, D.; Eppinger, T.; Pereira, F.E.; Aglave, R.; Debus, K.; Bermingham, S.K.; Ramachandran, R. A Multi-Scale, Mechanistic
Model of a Wet Granulation Process Using a Novel Bi-Directional PBM-DEM Coupling Algorithm. Chem. Eng. Sci. 2015, 123,
500–513. [CrossRef]
57. Barrasso, D.; Ramachandran, R. Qualitative Assessment of a Multi-Scale, Compartmental PBM-DEM Model of a Continuous
Twin-Screw Wet Granulation Process. J. Pharm. Innov. 2016, 11, 231–249. [CrossRef]
58. Kulju, T.; Paavola, M.; Spittka, H.; Keiski, R.L.; Juuso, E.; Leiviskä, K.; Muurinen, E. Modeling Continuous High-Shear Wet
Granulation with DEM-PB. Chem. Eng. Sci. 2016, 142, 190–200. [CrossRef]
59. Wang, L.G.; Morrissey, J.P.; Barrasso, D.; Slade, D.; Clifford, S.; Reynolds, G.; Ooi, J.Y.; Litster, J.D. Model Driven Design for Twin
Screw Granulation Using Mechanistic-Based Population Balance Model. Int. J. Pharm. 2021, 607, 120939. [CrossRef] [PubMed]
60. Zheng, C.; Zhang, L.; Govender, N.; Wu, C.Y. DEM Analysis of Residence Time Distribution during Twin Screw Granulation.
Powder Technol. 2021, 377, 924–938. [CrossRef]
61. Zheng, C.; Govender, N.; Zhang, L.; Wu, C.-Y. GPU-Enhanced DEM Analysis of Flow Behaviour of Irregularly Shaped Particles
in a Full-Scale Twin Screw Granulator. Particuology 2021, 61, 30–40. [CrossRef]
62. Kumar, A.; Radl, S.; Gernaey, K.v.; de Beer, T.; Nopens, I. Particle-Scale Modeling to Understand Liquid Distribution in Twin-Screw
Wet Granulation. Pharmaceutics 2021, 13, 928. [CrossRef]
63. Altair Engineering Inc. Altair EDEM, 2020.3 2021. Available online: https://www.altair.com/edem/(accessed on 7 July 2021).
64. Morrissey, J.P. Discrete Element Modelling of Iron Ore Fines to Include the Effects of Moisture and Fines. 2013. Available online:
https://era.ed.ac.uk/handle/1842/8270 (accessed on 7 July 2021).
65. Thakur, S.C.; Morrissey, J.P.; Sun, J.; Chen, J.F.; Ooi, J.Y. Micromechanical Analysis of Cohesive Granular Materials Using Discrete
Element Method with an Adhesive Elasto-Plastic Contact Model. Granul. Matter 2014, 16, 383–400. [CrossRef]
66. Jones, R. From Single Particle AFM Studies of Adhesion and Friction to Bulk Flow: Forging the Links. Granul. Matter 2003, 4,
191–204. [CrossRef]
67. Morrissey, J.P.; Ooi, J.Y.; Chen, J.F.F.; Tano, K.T.; Horrigmoe, G. Measurement and Prediction of Compression and Shear Behavior
of Wet Iron Ore Fines. In Proceedings of the 7th World Congress on Particle Technology (WCPT7), Beijing, China, 19–22 May
2014; p. 8.
68. Morrissey, J.P.; Ooi, J.Y.; Chen, J.F. Effect of Solid Cohesion and Friction on Silo Discharge. In Proceedings of the 7th World
Congress on Particle Technology (WCPT7), Beijing, China, 19–22 May 2014; p. 8.
69. Thakur, S.C.; Ahmadian, H.; Sun, J.; Ooi, J.Y. An Experimental and Numerical Study of Packing, Compression, and Caking
Behaviour of Detergent Powders. Particuology 2014, 12, 2–12. [CrossRef]
70. Thakur, S.C.; Ooi, J.Y.; Ahmadian, H. Scaling of Discrete Element Model Parameters for Cohesionless and Cohesive Solid. Powder
Technol. 2015, 293, 130–137. [CrossRef]
71. Pantaleev, S.; Yordanova, S.; Janda, A.; Marigo, M.; Ooi, J.Y. An Experimentally Validated DEM Study of Powder Mixing in a
Paddle Blade Mixer. Powder Technol. 2017, 311, 287–302. [CrossRef]
72. Janda, A.; Ooi, J.Y. DEM Modeling of Cone Penetration and Unconfined Compression in Cohesive Solids. Powder Technol. 2016,
293, 60–68. [CrossRef]
73. Härtl, J.; Ooi, J.Y. Experiments and Simulations of Direct Shear Tests: Porosity, Contact Friction and Bulk Friction. Granul. Matter
2008, 10, 263–271. [CrossRef]
74. Lommen, S.; Schott, D.; Lodewijks, G. DEM Speedup: Stiffness Effects on Behavior of Bulk Material. Particuology 2014, 12, 107–112.
[CrossRef]
75. Lee, K.T. Continuous Granulation of Pharmaceutical Powder Using a Twin Screw Granulator. 2013. Available online: https:
//etheses.bham.ac.uk/id/eprint/4002/ (accessed on 7 July 2021).
76. Chan Seem, T.; Rowson, N.A.; Gabbott, I.; de Matas, M.; Reynolds, G.K.; Ingram, A. Asymmetric Distribution in Twin Screw
Granulation. Eur. J. Pharm. Biopharm. 2015, 106, 50–58. [CrossRef]
95
Pharmaceutics 2021, 13, 2136
77. Pradhan, S.U.; Sen, M.; Li, J.; Litster, J.D.; Wassgren, C.R. Granule Breakage in Twin Screw Granulation: Effect of Material
Properties and Screw Element Geometry. Powder Technol. 2017, 315, 290–299. [CrossRef]
78. Wang, L.G.; Pradhan, S.U.; Wassgren, C.; Barrasso, D.; Slade, D.; Litster, J.D. A Breakage Kernel for Use in Population Balance
Modelling of Twin Screw Granulation. Powder Technol. 2020, 363, 525–540. [CrossRef]
96
pharmaceutics
Article
Impact of Powder Properties on the Rheological Behavior
of Excipients
Pauline H. M. Janssen 1, *, Sébastien Depaifve 2 , Aurélien Neveu 2 , Filip Francqui 2 and Bastiaan H. J. Dickhoff 1
Abstract: With the emergence of quality by design in the pharmaceutical industry, it becomes
imperative to gain a deeper mechanistic understanding of factors impacting the flow of a formulation
into tableting dies. Many flow characterization techniques are present, but so far only a few have
shown to mimic the die filling process successfully. One of the challenges in mimicking the die filling
process is the impact of rheological powder behavior as a result of differences in flow field in the
feeding frame. In the current study, the rheological behavior was investigated for a wide range of
excipients with a wide range of material properties. A new parameter for rheological behavior was
introduced, which is a measure for the change in dynamic cohesive index upon changes in flow field.
Particle size distribution was identified as a main contributing factor to the rheological behavior of
powders. The presence of fines between larger particles turned out to reduce the rheological index,
which the authors explain by improved particle separation at more dynamic flow fields. This study
also revealed that obtained insights on rheological behavior can be used to optimize agitator settings
in a tableting machine.
Citation: Janssen, P.H.M.; Depaifve,
S.; Neveu, A.; Francqui, F.; Dickhoff, Keywords: powder flow; die filling; powder rheology; dynamic cohesive index; excipients; tableting;
B.H.J. Impact of Powder Properties continuous manufacturing; quality by design
on the Rheological Behavior of
Excipients. Pharmaceutics 2021, 13,
1198. https://doi.org/10.3390/
pharmaceutics13081198 1. Introduction
Powders are used in many industries and in a broad range of processes and applica-
Academic Editor: Colin Hare
tions. Often understanding powder behavior is crucial to properly design the process and
equipment [1]. In a continuous manufacturing line for example, consistent and continuous
Received: 30 June 2021
Accepted: 2 August 2021
flow through the system is a critical requirement for finished product quality [2]. Flow is
Published: 4 August 2021
also relevant for manufacturing efficiency for batch processes. It determines for example
whether bins can be used or hand scooping is required, to what extend product is scraped
Publisher’s Note: MDPI stays neutral
at the beginning or end of a run, and the allowable production rate of products [3].
with regard to jurisdictional claims in
In pharmaceutical processing, insufficient flow can lead to product quality failures,
published maps and institutional affil- due to large weight or dosage variations [4]. Weight variations in the final dosage form can
iations. occur when the flowability of the final formulation limits the separation of small quantities
of powder from the larger mass of powder. Because of powder flow limitations in this step,
a slow-speed process that works well may not work at all when rates are increased [3]. The
production capacity of a tablet press is directly determined by the rotation frequency of the
Copyright: © 2021 by the authors.
die table and limited by powder flow into the dies [5,6].
Licensee MDPI, Basel, Switzerland.
Predicting how powder will flow into the dies is however not easy, as flowability
This article is an open access article
is a complex, multidimensional property. Flowability is the result of material physical
distributed under the terms and properties and the equipment used for handling, storing or processing the material [3].
conditions of the Creative Commons The forces that influence material flowability depend on the flow field and stress state as
Attribution (CC BY) license (https:// well as on external factors, like equipment design and material, temperature, and humidity.
creativecommons.org/licenses/by/ This also explains why flow behavior (and ranking) of powders can be different between
4.0/). different measurement methods and applications.
Many different characterization techniques to measure flow have been developed and
correlated to the powder flow behavior in different processing units [1,7]. Pharmaceutical
compendial flow characterization methods include angle of repose [8,9], compressibility
index [10,11], flow through an orifice [12], rotating drum [13,14], powder rheometers [15],
and several variants of the shear cell tester [16]. The different flow characterization methods
quantify powder flow differently, due to the differences in flow field and degree of stresses
applied during measurement [17]. Stavrou et al. [18] for example showed differences in
powder flow behavior when different stress levels were employed.
Although many different characterization techniques have been developed, care must
be taken to select the most suitable characterization technique for a specific approach [1].
Simple flow measurements tend to fail in the prediction of powder flow into tableting
dies, due to a lack of simulation of the right flow field and stress state. So far, only a few
characterization techniques have been developed with specific focus to mimic the die filling
process. Wu et al. [19] for example used transparent dies and moving feeding shoes to
study the powder flow of different metallurgical powder components in air and vacuum.
He showed that powder characteristics, shoe speed, and die geometry play an important
role in the die filling process. Mendez et al. [20] used a fixed frame and a moving die disc
system to examine the effect of blend composition, shoe properties and die parameters on
uniformity of die filling. Mehrotra et al. [21] simulated the die filling process for cohesive
materials. He concluded that cohesive powder take longer to fill dies and hence could be a
potential cause of tablet weight variability.
Described research has shown to predict the tablet die filling process from a die shoe
well. A common practical challenge of the described studies however, is that they all
studied the die filling process with simplified systems. The effect of different flow fields
in the feeding frame has not been considered. This factor can be very important in the
prediction of die filling, especially when powder fluidity is influenced by the presence of
paddle feeders, wipers, or agitator arms [3].
With the emergence of quality by design (QbD), it becomes imperative to gain a
deeper mechanistic understanding of how different materials respond to differences in flow
field [22]. The aim of this research is to investigate the effect of rheological behavior during
powder flow into tablet dies. A wide range of excipients with a broad range of material
properties was evaluated. A new parameter for rheological behavior is introduced, which is
a measure for the change in dynamic cohesive index upon changes in stress state and flow
field. The objective of this study is to identify which material properties do have an impact
on this rheological index parameter. In addition, the fidelity of the rheological index was
validated by correlating it to tableting performance in a rotary tablet press with agitators.
98
Pharmaceutics 2021, 13, 1198
Table 1. List of material characterization techniques, the material properties they measure and corresponding abbreviations
of the measured material properties
Characterization
Physical Property Abbreviation Range of Values Unit
Technique
Visual observation by scanning
Shape Shape - -
electron microscopy
10% cumulative undersize of
×10 3.0–77.6 μm
volumetric PSD
Particle size distribution (PSD)
50% cumulative undersize of
by laser diffraction ×50 18.3–243 μm
volumetric PSD
90% cumulative undersize of
×90 49.4–406 μm
volumetric PSD
Span of the volumetric PSD* Span 1.22–2.83 -
Karl fisher titration Total moisture content KF 0.1–5.8 %w/w
Thermogravimetric balance Loss on drying LOD 0.0–9.3 %w/w
Brunauer–Emmett–Teller
Specific surface area SSA 0.1–5.1 m2 /g
analysis with Krypton
Bulk density BD 0.35–0.79 g/mL
Graduated cylinder Tapped density TD 0.49–0.98 g/mL
Hausner ratio HR 1.15–1.60 -
Flow function coefficient at 4
Ring shear cell tester ffc 2.4–17.3 -
kPa pre-consolidation pressure
Initial charge density q0 −1.5–0.2 nC/g
Electric charge analyzer Final charge density qf −1.5–0.2 nC/g
Tribo-charging density
Δq −5.0–0.1 nC/g
variation
Gas pycnometer True density TrD 1.52–1.58 g/mL
Yield pressure at 0.01
PyS 78–229 MPa
Heckel testing by compaction mm/s—slow
simulation Yield pressure at 300
PyF 80–236 MPa
mm/s—fast
Strain Rate Sensitivity SRS 0–48 %
Rotating drum Rheological index RI −0.54–0.86 rpm−1
2.2.1. Shape
Scanning electron microscopy (SEM) images were recorded using a Phenom ProX
scanning electron microscope (Thermo Fischer Scientific, Waltham, MA, USA). Prior to the
measurements, samples were coated with a gold layer with a thickness of 4 nm. Images
were recorded at an acceleration voltage of 10 kV. The shape of particles is defined by
visual observation.
99
Pharmaceutics 2021, 13, 1198
Δq = qf − q0 (3)
100
Pharmaceutics 2021, 13, 1198
The strain rate sensitivity (SRS) is calculated by comparing the yield pressure at high
speed (PyF) and slow speed (PyS)
2.6. Tableting
For the formulations that are tableted, 99.5% w/w filler is blended with 0.5 % w/w
MgSt for 2 min in a Turbula blender T2 at 96 rpm. Blends are compressed on a RoTab rotary
tableting press at 25 rpm. 9 mm flat beveled punches (iHolland) are used and compaction
force is set to 10 kN. The filling depth is set to obtain tablets of 250 mg at 10 rpm agitator
(optfiller) speed. The agitator speed is increased from 10–45 rpm in steps of 5 rpm to get
different amounts of agitation without changing any further settings.
101
Pharmaceutics 2021, 13, 1198
Table 2. Physical properties such as type, shape and particle size distribution, density and flow properties for the set of
excipients that is used in this study
Bulk
×10 ×50 ×90 Hausner ffc @4
Grade Abbreviation Type Shape Span Density
(μm) (μm) (μm) Ratio (-) kPa (-)
(g/mL)
Lactopress®
LP anh Anhydrous lactose Shards 16.5 133 323 2.30 0.69 1.28 7.5
anhydrous
SuperTab® 21 AN 21 AN Anhydrous lactose Shards 24.1 180 387 2.02 0.72 1.27 7.7
SuperTab® 22 AN 22 AN Anhydrous lactose Shards 47.0 203 359 1.54 0.68 1.17 15
(Granulated)
SuperTab® 24 AN 24 AN Granular 37.0 121 298 2.15 0.54 1.25 13
anhydrous lactose
Lactose monohydrate
Pharmatose® 80 M 80 M Tomahawk 76.6 242 406 1.36 0.79 1.19 13
(sieved)
Lactose monohydrate
Pharmatose® 150 M 150 M Tomahawk/fines 7.4 68.4 189 2.66 0.72 1.36 3.8
(milled)
Lactose monohydrate
Pharmatose® 200 M 200 M Tomahawk/fines 4.4 37.7 111 2.83 0.62 1.58 3.7
(milled)
Lactose monohydrate
Pharmatose® 450 M 450 M Fines 3.0 18.3 49.4 2.54 0.50 1.60 2.4
(milled)
Modified lactose
SuperTab® 30 GR 30 GR Granular 38.3 126 297 2.05 0.63 1.24 17
monohydrate
® Modified lactose
SuperTab 11 SD 11 SD Spherical 44.0 119 223 1.51 0.63 1.19 17
monohydrate
® Modified lactose
SuperTab 14 SD 14 SD Spherical 47.7 124 227 1.44 0.62 1.15 14
monohydrate
® Modified lactose
SuperTab 50 ODT 50 ODT Spherical 30.9 106 199 1.58 0.71 1.17 13
monohydrate
Microcrystalline
Pharmacel® 101 MCC101 Spherical/Fibers 20.0 62.2 137 1.89 0.34 1.45 5.9
cellulose
® Microcrystalline
Pharmacel 102 MCC102 Spherical/Fibers 29.9 86.9 200 1.95 0.33 1.39 7.0
cellulose
Microcrystalline
cellulose,
Pharmacel® sMCC90 sMCC90 Spherical/Fibers 29.4 102 233 1.99 0.38 1.32 9.0
co-processed with
silicon dioxide
Primojel® PJ Superdisintegrant Spherical 21.1 42.3 72.6 1.22 0.79 1.21 12
Primellose® PL Superdisintegrant Fibers 24.4 54.4 114 1.65 0.55 1.35 7.4
The used set of materials covers a large variation in excipient type, particle shape,
particle size distribution, density, and powder flow parameters. Four grades (4) of an-
hydrous lactose are evaluated. SuperTab® 21 AN and Lactopress® anhydrous have a
relatively large proportion of fines, as indicated by the ×10 of 15 μm. SuperTab® 22 AN
contains a smaller proportion of fines and has improved flow properties according to the
Hausner ratio and the flow function coefficient. SuperTab® 24 AN has also improved flow
properties and a slightly different shape and lower density. Four grades of non-modified
lactose monohydrate are evaluated. Pharmatose® 80 M is a sieved tomahawk shaped
material, with a relatively large particle size. Pharmatose® 150 M, Pharmatose® 200 M
and Pharmatose® 450 M are milled materials. Particle size, density, and flow properties
of these grades decrease with increasing number. The four evaluated modified lactose
monohydrate grades consist of three spray dried grades with slightly different particle size
distribution and density, and one granulated lactose monohydrate. Flow parameters of
these grades all indicate very good flowability. More irregular shaped particles that are
evaluated are microcrystalline cellulose and Primellose® . Pharmacel® 101 and Pharmacel®
102 are two microcrystalline cellulose grades with spherical morphology consisting of fibers.
Particle size of Pharmacel® 101 is smaller than for Pharmacel® 102, which also explains
102
Pharmaceutics 2021, 13, 1198
the reduced flow properties. Pharmacel® sMCC90 is very similar to Pharmacel® 102, but
has been co-processed with 2% w/w silicon dioxide to increase the specific surface area
and improve the flow properties. Primojel® and Primellose® are two superdisintegrants
with small particle size and a spherical and fibrous morphology respectively. The good
flow indicated by the flow parameters of Primojel® , is in line with the expectation from the
spherical morphology.
Particle size distribution, particle shape and density are parameters that are known
to have an impact on the different forces acting on particles during powder flow. This
can be understood by looking at the different driving and drag forces that act on particles
and determine the powder flow. One of the driving forces for flowability is gravity.
Gravitational forces are higher for larger particles, and for particles with higher (true)
density [25]. Drag forces on the other hand, typically include adhesive and cohesive
forces, which produce a tendency for particles to stick to each other and to other surfaces.
Adhesive and cohesive forces are composed mainly of van der Waals forces, capillary
bridging, and electrostatics [26]. The magnitude of these forces depends on the nature of
the material and on the available surface. Adhesive and cohesive forces will be higher
for smaller or irregular shaped particles, as the available surface for these particles is
higher [23]. Irregular shaped particles also can have a negative impact on flowability due
to the increased risk for mechanical interlocking [27,28].
103
Pharmaceutics 2021, 13, 1198
index is around 30. Primojel® and Pharmacel® sMCC90 have a cohesive index below 20,
in line with the lowest Hausner ratio and highest ffc within its category. Also, over the
different product groups, the variance in cohesive index at 2 rpm can be explained for over
60% by linear models of both the Hausner ratio and the flow function coefficient at 4 kPa
pre-consolidation strength. The high overlap in powder ranking by these measurements is
explained by similarities in the stress state and flow field during these measurements.
Figure 1. Cohesive index values as function of Hausner ratio (left) and flow function coefficient
(right) at two different rotational speeds. The obtained cohesive index at low rotational speed (2 rpm)
correlated with both measurements, while no correlation is observed for the cohesive index at high
rotation speed (60 rpm).
104
Pharmaceutics 2021, 13, 1198
Figure 2. Cohesive index values as function of rotational speed for anhydrous lactose (top left),
lactose monohydrate (top right), modified lactose monohydrate (bottom left) and microcrystalline
cellulose/superdisintegrants (bottom right).
Most powders that are evaluated show shear thickening behavior. This indicates that
cohesiveness increases with rotational speed, which is associated with more intermittent
and irregular flow. This can be concern for the pharmaceutical industry as more cohesive
powders can lead to increased variability and mass flow excursions outside the acceptable
target range, as demonstrated by Allenspach et al. [37]. Notable is the shear thinning
behavior, which is observed for two grades of anhydrous lactose, two grades of lactose
monohydrate and microcrystalline cellulose. The shear thinning behavior can be explained
by aeration of the powder at higher rotational speeds, which increases the distance between
particles and thereby reduces the cohesive surface interactions [36].
The observed differences in cohesive index and rheological index for sieved (Pharmatose®
80 M) and milled lactose (Pharmatose® 150 M and Pharmatose® 200 M) are in line with findings
of Hickey et al. [38] for inhalation grades lactose. These authors reported that an apparent
contradiction is present in flow property parameters of milled and sieved powders. Static
powder flow measurements (bulk and tapped density, angle of repose) lead to the conclusion
that milled powders exhibit poor flow compared to sieved batches. Dynamic rotating drum
measurements indicated the reverse, where milled powder flows better than sieved powder.
To evaluate which material properties influence the shear thinning and thickening
behavior, a new parameter rheological index (RI) is introduced. The Rheological Index (RI)
105
Pharmaceutics 2021, 13, 1198
is defined as the linear slope of dynamic cohesive index as function of rotational speed
between 2 rpm and 60 rpm. Figure 3 shows the RI values for the investigated materials.
For Lactopress® anhydrous, SuperTab® 21 AN, Pharmatose® 150 M, Pharmatose® 200 M,
and Pharmacel® 101 the reduction in cohesiveness due to aeration seems to outweigh the
inertial effect, resulting in shear thinning behavior. For Pharmacel® 102, Primojel® , and
Primellose® the rheological index is close to zero.
Figure 3. Rheological index (RI) values for the different materials. A positive rheological index indicates shear thickening
behavior, while negative rheological index indicates shear thinning behavior. Different colours represent different types
of materials.
106
Pharmaceutics 2021, 13, 1198
Figure 4. PLS analyses of the dataset with a loading plot (top) and score plot (bottom).
107
Pharmaceutics 2021, 13, 1198
Figure 5. Variable Influence on Projection (VIP) plot indicating the relevance of terms for explaining the rheological index.
Error bars indicate the 95% confidence interval.
108
Pharmaceutics 2021, 13, 1198
at a fines content above 30% w/w was dominated by the behavior of fines. More recently,
Pillitery et al. [40] also found a tipping point in compaction of binary granular mixture
close to 30–35% w/w of fines. It is expected that, at this level of fines, the coarse particles
are completely embedded by the fines, and so flow behavior is governed by interparticle
forces between fines. Below this level, contacts between coarse particles dominate flow [41].
Figure 6. Cohesive index as function of rotational speed for sieved lactose with different amounts of fines added.
Figure 7. Rheological index parameter as function of the amount of fines added to the sieved lactose. The marked area at
10–35% w/w fines indicates shear thinning behavior.
The observed differences in cohesive index and rheological index for blends with
different amounts of fines, are in line with the earlier mentioned findings of Hickey [38] on
inhalation lactose. The authors explained that milled powder flows more readily because of
the presence of fines. Stretching the findings of Hickey et al. [38], the current studies shows
that the rheological behavior is not only different for milled and sieved materials. It reveals
109
Pharmaceutics 2021, 13, 1198
that the presence of a small amount of fines in a powder with larger particle size distribution
can have a positive effect on the dynamic flow properties and rheological behavior.
An explanation for the reduced cohesive index of blends with 10–35% w/w fines
at high rotational speeds is found in the flow field. In a more rapid flow regime, the
magnitude of forces acting on the particles are different. On the one hand, the inertial effect
is stronger, which could lead to increased cohesive index. On the other hand, friction by
cohesive and adhesive forces is typically reduced. This is due to the fact that rolling friction
is typically much smaller than static or sliding friction [34]. It is the authors hypothesis
that the presence of fines even further reduces the friction when particles are moving. This
is explained by a reduction of bridge formation in moving powder when small particles
are present between larger particles. This improves particle separation, resulting in more
readily powder flow. In this study, we find that blends with 10–35% w/w fines correspond
to optimal mixtures where the separation of large grains by interstitial fines improves the
rheological properties of the blends. Indeed, the cohesive index of the blends with 10–35%
w/w fines reduces with the rotational speed, which can be explained by an aeration of
the powders leading to a reduction of the cohesive forces acting on the grains. When the
fines content increases over 40% w/w, the flow behavior is dominated by the fine fraction
embedding the coarse particles, and a more cohesive powder is observed. The hypothesis
is that for blends of coarse and fine material a tipping point for rheological behavior exist
at the point where coarse particles are completely embedded by fines. This tipping point is
therefore expected to depend on the particle shape and particle size ratio of the fines and
the coarse particles.
Figure 8. Tableting results (n = 20) on average mass and mass variability (%RSD) of a shear thinning material (SuperTab®
21 AN) and a shear thickening material (SuperTab® 11 SD). The agitator rotational speed was increased from 10–45 rpm in
steps of 5 rpm.
110
Pharmaceutics 2021, 13, 1198
Differences in rheological behavior were also visible during flow into the dies in
the tableting process. Increased agitator speed resulted in higher mass tablets with less
variability for the formulation with SuperTab® 21 AN. This indicates that at low agitator
speed, the powder flow into the dies was incomplete and that improvement was obtained
by increased flow via adjustment of the flow field. These findings confirm that powders
with a negative rheological index exhibit a lower cohesion and better flowability with
a more dynamic process. This is especially visible for grades like SuperTab® 21 AN,
which have as a significant fraction of fines present and a large bridging propensity in
the quasi-static state [32]. For SuperTab® 11 SD on the other hand, increased agitator
speed did not result in higher mass tablets. A slight increase in mass RSD is observed for
this formulation, indicating more cohesive powder flow. This is in line with the positive
rheological index of SuperTab® 11 SD, in line with the sharp particle size distribution and
low bridging propensity in the quasi-static state [32]. These results show that rheological
index measurements in the lab can be a tool to set-up a tableting process.
4. Conclusions
In this study, the effect of rheological behavior during powder flow into tablet dies is
investigated. It was shown that a difference in flow field and stress state can have an effect
on the cohesive index and die filling of powders. This effect depends on the raw material
properties. A new parameter for rheological behavior (RI) was introduced, which is a
measure for the change in dynamic cohesive index upon changes in stress state and flow
field. Most powders evaluated show shear thickening behavior. However, shear thinning
is observed for some materials as well. Of all physical/chemical parameters tested, the
particle size distribution (×10) was shown to have the largest impact on the rheological
behavior of powders. At higher rotational speeds, addition of 5–35% w/w of fines to coarse
sieved lactose resulted in a decrease of cohesive index. This is explained by the presence
of fines between larger particles, that can improve particle separation thus reducing the
cohesive forces acting between grains. The fidelity of the rheological index was validated
by correlating it to tableting performance in a rotary tablet press with agitators. It was
shown that insights on the rheological index (RI) obtained by rotating drum experiments
can be used to optimize agitator settings in a tableting machine.
111
Pharmaceutics 2021, 13, 1198
References
1. Krantz, M.; Zhang, H.; Zhu, J. Characterization of Powder Flow: Static and Dynamic Testing. Powder Technol. 2009, 194, 239–245.
[CrossRef]
2. Pernenkil, L.; Cooney, C.L. A Review on the Continuous Blending of Powders. Chem. Eng. Sci. 2006, 61, 720–742. [CrossRef]
3. Prescott, J.K.; Barnum, R.A. On Powder Flowability. Pharm. Technol. 2000, 24, 60–85.
4. Blackshields, C.A.; Crean, A.M. Continuous Powder Feeding for Pharmaceutical Solid Dosage Form Manufacture: A Short
Review. Pharm. Dev. Technol. 2018, 23, 554–560. [CrossRef] [PubMed]
5. Puckhaber, D.; Eichler, S.; Kwade, A.; Finke, J.H. Impact of Particle and Equipment Properties on Residence Time Distribution of
Pharmaceutical Excipients in Rotary Tablet Presses. Pharmaceutics 2020, 12, 283. [CrossRef] [PubMed]
6. Peeters, E.; De Beer, T.; Vervaet, C.; Remon, J.P. Reduction of Tablet Weight Variability by Optimizing Paddle Speed in the Forced
Feeder of a High-Speed Rotary Tablet Press. Drug Dev. Ind. Pharm. 2015, 41, 530–539. [CrossRef]
7. Schwedes, J. Review on Testers for Measuring Flow Properties of Bulk Solids. Granul. Matter 2003, 5, 1–43. [CrossRef]
8. Zhou, Y.C.; Xu, B.H.; Yu, A.B.; Zulli, P. An Experimental and Numerical Study of the Angle of Repose of Coarse Spheres. Powder
Technol. 2002, 125, 45–54. [CrossRef]
9. Ileleji, K.E.; Zhou, B. The Angle of Repose of Bulk Corn Stover Particles. Powder Technol. 2008, 187, 110–118. [CrossRef]
10. Saker, A.; Cares-Pacheco, M.G.; Marchal, P.; Falk, V. Powders Flowability Assessment in Granular Compaction: What about the
Consistency of Hausner Ratio? Powder Technol. 2019, 354, 52–63. [CrossRef]
11. Saw, H.Y.; Davies, C.E.; Paterson, A.H.J.; Jones, J.R. Correlation between Powder Flow Properties Measured by Shear Testing and
Hausner Ratio. Procedia Eng. 2015, 102, 218–225. [CrossRef]
12. Zhou, X.; Nauka, E.; Narang, A.; Mao, C. Flow Function of Pharmaceutical Powders at Low-Stress Conditions Can Be Inferred
Using a Simple Flow-Through-Orifice Device. J. Pharm. Sci. 2020, 109, 2009–2017. [CrossRef]
13. Nalluri, V.R.; Kuentz, M. Flowability Characterisation of Drug-Excipient Blends Using a Novel Powder Avalanching Method. Eur.
J. Pharm. Biopharm. 2010, 74, 388–396. [CrossRef]
14. Pirard, S.L.; Lumay, G.; Vandewalle, N.; Pirard, J.P. Motion of Carbon Nanotubes in a Rotating Drum: The Dynamic Angle of
Repose and a Bed Behavior Diagram. Chem. Eng. J. 2009, 146, 143–147. [CrossRef]
15. Freeman, R. Measuring the Flow Properties of Consolidated, Conditioned and Aerated Powders-A Comparative Study Using a
Powder Rheometer and a Rotational Shear Cell. Powder Technol. 2007, 174, 25–33. [CrossRef]
16. Koynov, S.; Glasser, B.; Muzzio, F. Comparison of Three Rotational Shear Cell Testers: Powder Flowability and Bulk Density.
Powder Technol. 2015, 283, 103–112. [CrossRef]
17. Shah, R.B.; Tawakkul, M.A.; Khan, M.A. Comparative Evaluation of Flow for Pharmaceutical Powders and Granules. AAPS
PharmSciTech 2008, 9, 250–258. [CrossRef]
18. Stavrou, A.G.; Hare, C.; Hassanpour, A.; Wu, C.Y. Investigation of Powder Flowability at Low Stresses: Influence of Particle Size
and Size Distribution. Powder Technol. 2020, 364, 98–114. [CrossRef]
19. Wu, C.Y.; Dihoru, L.; Cocks, A.C.F. The Flow of Powder into Simple and Stepped Dies. Powder Technol. 2003, 134, 24–39. [CrossRef]
20. Mendez, R.; Muzzio, F.; Velazquez, C. Study of the Effects of Feed Frames on Powder Blend Properties during the Filling of Tablet
Press Dies. Powder Technol. 2010, 200, 105–116. [CrossRef]
21. Mehrotra, A.; Chaudhuri, B.; Faqih, A.; Tomassone, M.S.; Muzzio, F.J. A Modeling Approach for Understanding Effects of Powder
Flow Properties on Tablet Weight Variability. Powder Technol. 2009, 188, 295–300. [CrossRef]
22. Lee, W.B.; Widjaja, E.; Heng, P.W.S.; Chan, L.W. The Effect of Rotation Speed and Particle Size Distribution Variability on
Mixability: An Avalanche Rheological and Multivariate Image Analytical Approach. Int. J. Pharm. 2020, 579. [CrossRef]
23. Lumay, G.; Boschini, F.; Traina, K.; Bontempi, S.; Remy, J.C.; Cloots, R.; Vandewalle, N. Measuring the Flowing Properties of
Powders and Grains. Powder Technol. 2012, 224, 19–27. [CrossRef]
24. Boschini, F.; Delaval, V.; Traina, K.; Vandewalle, N.; Lumay, G. Linking Flowability and Granulometry of Lactose Powders. Int. J.
Pharm. 2015, 494, 312–320. [CrossRef] [PubMed]
25. Geldart, D. Types of Gas Fhidization. Powder Technol. 1973, 7, 285–292. [CrossRef]
26. Aulton, M.E.; Taylor, K.M.G. Powder flow. In Aulton’s Pharmaceutics E-Book: The Design and Manufacture of Medicines; Churchill
Livingstone (Elsevier): Edinburgh, Scotland, 2013; pp. 189–200. ISBN 978-0702042904.
27. Goh, H.P.; Heng, P.W.S.; Liew, C.V. Comparative Evaluation of Powder Flow Parameters with Reference to Particle Size and
Shape. Int. J. Pharm. 2018, 547, 133–141. [CrossRef]
28. Fu, X.; Huck, D.; Makein, L.; Armstrong, B.; Willen, U.; Freeman, T. Effect of Particle Shape and Size on Flow Properties of Lactose
Powders. Particuology 2012, 10, 203–208. [CrossRef]
29. Rajchenbach, J. Flow in Powders: From Discrete Avalanches to Continuous Regime. Phys. Rev. Lett. 1990, 65, 2221–2224.
[CrossRef]
112
Pharmaceutics 2021, 13, 1198
30. Castellanos, A.; Valverde, J.M.; Pérez, A.T.; Ramos, A.; Watson, P.K. Flow Regimes in Fine Cohesive Powders. Phys. Rev. Lett.
1999, 82, 1156–1159. [CrossRef]
31. Alexander, A.W.; Chaudhuri, B.; Faqih, A.M.; Muzzio, F.J.; Davies, C.; Tomassone, M.S. Avalanching Flow of Cohesive Powders.
Powder Technol. 2006, 164, 13–21. [CrossRef]
32. Neveu, A.; Janssen, P.; Lumay, G. The Flowability of Lactose Powders to Optimise Tableting Processes. ONdrugDelivery 2020,
2020, 58–62.
33. Francia, V.; Yahia, L.A.A.; Ocone, R.; Ozel, A. From Quasi-Static to Intermediate Regimes in Shear Cell Devices: Theory and
Characterisation. KONA Powder Part. J. 2021, 38, 3–25. [CrossRef]
34. Tardos, G.I.; McNamara, S.; Talu, I. Slow and Intermediate Flow of a Frictional Bulk Powder in the Couette Geometry. Powder
Technol. 2003, 131, 23–39. [CrossRef]
35. Amidon, G.E.; Secreast, P.J.; Mudie, D. Particle, Powder, and Compact Characterization. Dev. Solid Oral Dos. Forms 2009, 163–186.
[CrossRef]
36. Shi, H.; Lumay, G.; Luding, S. Stretching the Limits of Dynamic and Quasi-Static Flow Testing on Cohesive Limestone Powders.
Powder Technol. 2020, 367, 183–191. [CrossRef]
37. Allenspach, C.; Timmins, P.; Lumay, G.; Holman, J.; Minko, T. Loss-in-Weight Feeding, Powder Flow and Electrostatic Evaluation
for Direct Compression Hydroxypropyl Methylcellulose ( HPMC ) to Support Continuous Manufacturing. Int. J. Pharm. 2021,
596, 120259. [CrossRef] [PubMed]
38. Hickey, A.J.; Mansour, H.M.; Telko, M.J.; Xu, Z.; Smyth, H.D.C.; Mulder, T.; McLean, R.; Langridge, J.; Papadopoulos, D. Physical
Characterization of Component Particles Included in Dry Powder Inhalers. II. Dynamic Characteristics. J. Pharm. Sci. 2007, 96,
1302–1319. [CrossRef]
39. Molerus, O.; Nywlt, M. The Influence of the Fine Particle Content of the Flow Behaviour of Bulk Materials. Powder Technol. 1984,
37, 145–154. [CrossRef]
40. Pillitteri, S.; Opsomer, E.; Lumay, G.; Vandewalle, N. How Size Ratio and Segregation Affect the Packing of Binary Granular
Mixtures. Soft Matter 2020, 16, 9094–9100. [CrossRef]
41. Kojima, T.; Elliott, J.A. Incipient Flow Properties of Two-Component Fine Powder Systems and Their Relationships with Bulk
Density and Particle Contacts. Powder Technol. 2012, 228, 359–370. [CrossRef]
113
pharmaceutics
Article
Bulk Flow Optimisation of Amorphous Solid Dispersion
Excipient Powders through Surface Modification
Danni Suhaidi 1 , Yao-Da Dong 2 , Paul Wynne 3 , Karen P. Hapgood 4 and David A. V. Morton 1, *
1 School of Engineering, Deakin University, Waurn Ponds, VIC 3216, Australia; [email protected]
2 Drug Delivery, Disposition and Dynamics, Monash Institute of Pharmaceutical Sciences,
Monash University, Parkville, VIC 3052, Australia
3 Medicines Manufacturing Innovation Centre, Monash University, Clayton, VIC 3168, Australia
4 School of Engineering, Swinburne University, Hawthorn, VIC 3122, Australia
* Correspondence: [email protected]
Abstract: Particulate amorphous solid dispersions (ASDs) have been recognised for their poten-
tial to enhance the performance of various solid dose forms, especially oral bioavailability and
macromolecule stability. However, the inherent nature of spray-dried ASDs leads to their surface
cohesion/adhesion, including hygroscopicity, which hinders their bulk flow and affects their utility
and viability in terms of powder production, processing, and function. This study explores the
effectiveness of L-leucine (L-leu) coprocessing in modifying the particle surface of ASD-forming
materials. Various contrasting prototype coprocessed ASD excipients from both the food and pharma-
ceutical industries were examined for their effective coformulation with L-leu. The model/prototype
materials included maltodextrin, polyvinylpyrrolidone (PVP K10 and K90), trehalose, gum arabic,
and hydroxypropyl methylcellulose (HPMC E5LV and K100M). The spray-drying conditions were
set such that the particle size difference was minimised, so that it did not play a substantial role in
influencing powder cohesion. Scanning electron microscopy was used to evaluate the morphology of
each formulation. A combination of previously reported morphological progression typical of L-leu
surface modification and previously unreported physical characteristics was observed. The bulk
characteristics of these powders were assessed using a powder rheometer to evaluate their flowability
Citation: Suhaidi, D.; Dong, Y.-D.; under confined and unconfined stresses, flow rate sensitivities, and compactability. The data showed
Wynne, P.; Hapgood, K.P.; Morton, a general improvement in maltodextrin, PVP K10, trehalose and gum arabic flowability measures as
D.A.V. Bulk Flow Optimisation of L-leu concentrations increased. In contrast, PVP K90 and HPMC formulations experienced unique
Amorphous Solid Dispersion challenges that provided insight into the mechanistic behaviour of L-leu. Therefore, this study recom-
Excipient Powders through Surface mends further investigations into the interplay between L-leu and the physico-chemical properties of
Modification. Pharmaceutics 2023, 15, coformulated excipients in future amorphous powder design. This also revealed the need to enhance
1447. https://doi.org/10.3390/
bulk characterisation tools to unpack the multifactorial impact of L-leu surface modification.
pharmaceutics15051447
Academic Editor: Colin Hare Keywords: amorphous solid dispersions; spray drying; L-leucine; bulk powder characterization;
particle engineering; flowability enhancement; surface modification
Received: 31 March 2023
Revised: 4 May 2023
Accepted: 5 May 2023
Published: 9 May 2023
1. Introduction
In recent years, there has been substantial interest in the pharmaceutical particle
engineering research community for optimising spray-dried formulation powders and
Copyright: © 2023 by the authors. conditions to synthesise smaller-sized (≤10 μm) amorphous solid dispersions (ASDs) [1,2].
Licensee MDPI, Basel, Switzerland. Typically, ASD formulations are used to improve the stability of biomacromolecules [3]
This article is an open access article and dissolution characteristics of poorly water-soluble active pharmaceutical ingredients
distributed under the terms and
(APIs) [4,5]. Because of their small particle size and amorphous physical nature, these
conditions of the Creative Commons
spray-dried ASD powders have both a high surface-area-to-volume ratio and high surface
Attribution (CC BY) license (https://
energy, which increases the impact of cohesive interparticular forces, such as van der
creativecommons.org/licenses/by/
Waals interactions, on bulk behaviours [6]. Most industrially relevant ASD powders
4.0/).
115
Pharmaceutics 2023, 15, 1447
116
Pharmaceutics 2023, 15, 1447
a mass balance to within ±5% of the ideal target mass, and subsequently mixed with
400 mL of demineralized water (≤10 ppm total dissolved solids) using a magnetic stirrer
(800 rpm, 35 ◦ C). Spray drying was conducted using a Büchi B290 mini benchtop spray
dryer (Büchi Laboratory Equipment, Flawil, Switzerland) with a standard 0.5 mm two-fluid
nozzle. The standard operating conditions were as follows: Tinlet , 125 ◦ C; aspirator rate,
35 m3 /h; feedstock flow rate, 7.5 mL/min; and Toutlet , 76 ± 2 ◦ C. After spray drying, the
powders were quickly collected in a sealed container and stored away from direct sunlight
to minimise environmental exposure.
L-leucine Content
Formulation
Excipient Abbreviation Percentage
(Excipient/L-Leu Mixture)
(L-Leu wt% Dry Basis)
Maltodextrin MD MD/L-leu 0, 2.5, 5, 7.5, 10, 15, 20, 30%
Polyvinylpyrrolidone K10 PVP K10 PVP K10/L-leu 0, 2.5, 5, 7.5, 10, 15, 20, 30%
Polyvinylpyrrolidone K90 PVP K90 PVP K90/L-leu 0, 2.5, 5, 7.5, 10, 15, 20, 30%
Trehalose Trh Trh/L-leu 0, 2.5, 5, 7.5, 10, 15, 20, 30%
Gum Arabic GA GA/L-leu 0, 2.5, 5, 7.5, 10, 15, 20, 30%
Hydroxypropyl
HPMC E5LV HPMC E5LV/L-leu 0, 2.5, 5, 7.5, 10, 15, 20, 30%
methylcellulose E5LV
Hydroxypropyl
HPMC K100M HPMC K100M/L-leu 0, 2.5, 5, 7.5, 10, 15, 20, 30%
methylcellulose K100M
117
Pharmaceutics 2023, 15, 1447
τ = σ tan(η ) + C (2)
where τ denotes the shear stress, σ is the normal stress, η is the angle of internal friction
(AIF), and C is the cohesion force. For this analysis, the flow function coefficient (ffc), AIF
and Cohesion parameters were used to compare each formulation. The cohesion parameter
was derived by extrapolating the yield loci to identify the y-axis shear-stress intercept,
which represents the strength of a powder under zero confining stress [6]. The ffc represents
the ratio between the major principal stress (MPS) and unconfined yield strength (UYS).
The AIF was defined as the angle of the line drawn from the origin of the shear-stress vs.
normal-stress graph towards the preshear data point.
118
Pharmaceutics 2023, 15, 1447
divided by mass. Finally, the FRI is a measure of powder bed sensitivity to changes in shear
rate, defined as the ratio between the total energy recorded at a blade speed of 10 mm/s vs.
100 mm/s.
3. Results
3.1. Scanning Electron Microscopy
Selected representative electron microscopy images of the formulations are shown in
Figures 1–6. Morphologically, the maltodextrin/L-leu formulations displayed the most
visually observable changes with increasing leucine concentrations (Figure 1). At 0 wt%
L-leu, a substantial number of the spray dried powders appeared as typical collapsed
spheres which could be described as ‘blood cells’. The powders physically presented as
more agglomerated than those containing L-leu.
At L-leu concentrations of 5 wt%, some unusual structures appeared, notably a unique
‘cupcake’ morphology not seen in other formulations (Figure 1B). We propose that this
resulted from L-leu influencing the mechanical properties of the surface film of the drying
droplet. As the internal vapour pressure increased, a localised puncture appears to have
occurred, which likely led to an irregular structure. Starting from a 10 wt% concentration,
the powders began to exhibit a more heavily indented and wrinkled morphology, typically
associated with L-leu surface-modified powders. However, above 20 wt%, the wrinkled
morphology was accompanied by more spherical structures with apparent surface flaking.
PVP K10/L-leu formulations displayed degrees of morphological changes with in-
creasing L-leu, as previously reported (Figure 2). At 0 wt%, the powder exhibited a
smooth-dimpled morphology, which progressed to a more corrugated surface at 5 wt%.
At concentrations of ≥10 wt%, ‘collapsed’ morphologies were observed. As previously
theorised, above a certain concentration threshold, L-leu was able to form an outer shell
with sufficient mechanical and transport resistance to entrap the escaping water vapour.
This results in an increased internal vapour pressure, leading to expansion and eventual
structural failure, resulting in the observed collapsed morphology [22].
119
Pharmaceutics 2023, 15, 1447
ȱ
Figure 1. Representative SEM images of spray-dried maltodextrin/L-leu formulations: (A) MD/L-leu
(0%), (B) MD/L-leu (5%), (C) MD/L-leu (10%), (D) MD/L-leu (20%).
ȱ
Figure 2. Representative SEM images of spray dried PVP K10/L-leu formulations: (A) PVP K10/L-leu
(0%), (B) PVP K10/L-leu (5%), (C) PVP K10/L-leu (10%), (D) PVP K10/L-leu (20%).
120
Pharmaceutics 2023, 15, 1447
Figure 3. Representative SEM images of spray-dried PVP K90/L-leu formulations: (A) PVP K90/L-
leu (0%), (B) PVP K90/L-leu (5%), (C) PVP K90/L-leu (10%), (D) PVP K90/L-leu (20%).
121
Pharmaceutics 2023, 15, 1447
Figure 5. Representative SEM images of spray-dried gum arabic/L-leu formulations: (A) GA/L-leu
(0%), (B) GA/L-leu (5%), (C) GA/L-leu (10%), (D) GA/L-leu (20%).
Figure 6. Representative SEM images of spray dried HPMC E5LV/L-leu formulations: (A) HPMC
E5LV/L-leu (0%), (B) HPMC E5LV/L-leu (5%), (C) HPMC E5LV/L-leu (10%), (D) HPMC E5LV
/L-leu (20%).
122
Pharmaceutics 2023, 15, 1447
Unlike PVP K10, PVP K90 demonstrated previously unreported complications with
L-leu during spray drying (Figure 3). At concentrations of 2.5–10 wt%, spray drying
generated macroscale fibrous byproducts at the spray nozzle which severely affected the
yield (Figure 3B,C). In each case, long fibrous spindles of material reminiscent of silk
fibres formed within the drying chamber, preventing the recovery of most free powder.
Consequently, no powder was recovered at 2.5 wt% and 7.5 wt% L-leu concentrations,
preventing any bulk powder characterisation of these concentrations. However, no fibrous
byproducts were generated at concentrations of ≥15 wt%, with the collected powders
sharing corrugated morphologies similar to those of the PVP K10 formulations. Notably,
these fibrous formations did not occur during the initial feedstock formulation or in the
pure PVP K90 spray-drying runs.
Trehalose-based powders displayed a different morphological progression from the
other powders (Figure 4). SEM of individual particles did not display an apparent transition
from smooth dimpled spheres to corrugated structures. Instead, the powders maintained
a mostly lightly dimpled spherical shape from 0 wt% to 30 wt%. Gum arabic is a hetero-
geneous material that is composed of polysaccharides and glycoproteins. SEM images
showed that gum arabic exhibited a gradual morphological change as L-leu concentration
increased (Figure 5). At 0 wt%, it exhibited similar ‘blood-cell’ like particles, like pure
maltodextrin powder. Increasing the L-leu concentration resulted in the same wrinkled
corrugated morphologies associated with L-leu. However, unlike homogeneous excipients,
gum arabic required a higher L-leu concentration to confer the same morphological shifts.
Finally, electron microscopy showed that there were no discernible morphological
differences between spray-dried pure HPMC and those coprocessed with L-leu. At all L-leu
concentrations, both the HPMC E5LV and K100M (K100M in Appendix A as Figure A1)
powders presented themselves as smooth-dimpled powders (Figure 6).
123
Table 2. Particle size data of the spray-dried formulations. Data represent the mean ± SD (n = 3).
Formulation Code D10 (μm) D50 (μm) D90 (μm) Span Formulation Code D10 (μm) D50 (μm) D90 (μm) Span
MD/L-leu (0%) 3.17 ± 0.01 4.69 ± 0.03 8.79 ± 0.42 1.2 ± 0.15 GA/L-leu (0%) 3.78 ± 0.02 6.55 ± 0.04 11.61 ± 0.21 1.2 ± 0.09
MD/L-leu (2.5%) 3.55 ± 0.01 5.94 ± 0.01 11.07 ± 0.25 1.27 ± 0.09 GA/L-leu (2.5%) 3.74 ± 0.01 6.52 ± 0.02 11.54 ± 0.09 1.2 ± 0.04
MD/L-leu (5%) 3.14 ± 0.02 4.72 ± 0.01 8.72 ± 0.11 1.18 ± 0.04 GA/L-leu (5%) 3.67 ± 0.03 6.17 ± 0.07 10.76 ± 0.15 1.15 ± 0.08
MD/L-leu (7.5%) 3.49 ± 0.01 5.62 ± 0.07 10.2 ± 0.11 1.2 ± 0.06 GA/L-leu (7.5%) 3.78 ± 0.02 6.57 ± 0.06 11.27 ± 0.14 1.14 ± 0.08
MD/L-leu (10%) 3.52 ± 0.01 5.83 ± 0.05 10.46 ± 0.21 1.19 ± 0.09 GA/L-leu (10%) 3.48 ± 0.03 5.38 ± 0.06 9.55 ± 0.08 1.13 ± 0.05
Pharmaceutics 2023, 15, 1447
MD/L-leu (15%) 2.92 ± 0 4.49 ± 0.01 7.96 ± 0.13 1.12 ± 0.05 GA/L-leu (15%) 3.49 ± 0.01 5.45 ± 0.01 9.66 ± 0.06 1.13 ± 0.03
MD/L-leu (20%) 2.95 ± 0.01 4.54 ± 0.02 8.47 ± 0.26 1.21 ± 0.09 GA/L-leu (20%) 3.58 ± 0.01 6.28 ± 0.03 11.23 ± 0.13 1.22 ± 0.05
MD/L-leu (25%) 3.08 ± 0.04 4.71 ± 0.01 9.23 ± 0.04 1.31 ± 0.03 GA/L-leu (25%) 3.42 ± 0.01 5.4 ± 0.02 10.23 ± 0.08 1.26 ± 0.04
MD/L-leu (30%) 3.42 ± 0.02 5.78 ± 0.11 10.87 ± 0.22 1.29 ± 0.12 GA/L-leu (30%) 3.66 ± 0.02 6.63 ± 0.02 16.38 ± 0.79 1.92 ± 0.27
PVP K10/L-leu (0%) 3.69 ± 0.02 6.12 ± 0.06 10.38 ± 0.1 1.09 ± 0.06 HPMC E5LV/L-leu (0%) 4.05 ± 0.03 8.1 ± 0.14 17.77 ± 0.69 1.69 ± 0.29
PVP K10/L-leu (2.5%) 3.49 ± 0.02 5.48 ± 0.04 9.84 ± 0.05 1.16 ± 0.03 HPMC E5LV/L-leu (2.5%) 4 ± 0.02 7.95 ± 0.06 17.01 ± 0.54 1.63 ± 0.21
PVP K10/L-leu (5%) 3.19 ± 0 4.75 ± 0.02 8.88 ± 0.07 1.2 ± 0.03 HPMC E5LV/L-leu (5%) 4.18 ± 0.01 9.13 ± 0.1 19.74 ± 0.71 1.7 ± 0.27
PVP K10/L-leu (7.5%) 3.67 ± 0.03 6.13 ± 0.02 10.4 ± 0.06 1.1 ± 0.04 HPMC E5LV/L-leu (7.5%) 4.16 ± 0.06 9.09 ± 0.13 22.11 ± 0.94 1.98 ± 0.38
PVP K10/L-leu (10%) 3.41 ± 0.02 5.02 ± 0.05 8.98 ± 0.08 1.11 ± 0.05 HPMC E5LV/L-leu (10%) 3.93 ± 0.02 7.77 ± 0.07 17.64 ± 0.59 1.76 ± 0.23
PVP K10/L-leu (15%) 3.37 ± 0.01 5.01 ± 0.04 9.01 ± 0.06 1.12 ± 0.04 HPMC E5LV/L-leu (15%) 4.08 ± 0.01 8.33 ± 0.07 18.82 ± 0.76 1.77 ± 0.28
PVP K10/L-leu (20%) 3.42 ± 0.01 5.02 ± 0.04 9.04 ± 0.14 1.12 ± 0.06 HPMC E5LV/L-leu (20%) 3.95 ± 0.02 7.91 ± 0.1 17.43 ± 0.42 1.7 ± 0.18
PVP K10/L-leu (25%) 3.29 ± 0.01 4.75 ± 0.01 8.47 ± 0.12 1.09 ± 0.05 HPMC E5LV/L-leu (25%) 3.97 ± 0.01 7.84 ± 0.01 17.93 ± 0.29 1.78 ± 0.1
PVP K10/L-leu (30%) 3.39 ± 0.02 4.88 ± 0.02 8.75 ± 0.08 1.1 ± 0.04 HPMC E5LV/L-leu (30%) 3.89 ± 0.03 7.57 ± 0.12 17.66 ± 1.35 1.82 ± 0.5
Trh/L-leu (0%) 3.77 ± 0.09 6.21 ± 0.15 12.64 ± 0.54 1.43 ± 0.26 HPMC K100M/L-leu (0%) 4.29 ± 0.01 11.43 ± 0.15 26.4 ± 0.49 1.93 ± 0.22
Trh/L-leu (2.5%) HPMC K100M/L-leu (2.5%)
124
3.57 ± 0.04 6.06 ± 0.07 11.93 ± 0.48 1.38 ± 0.19 4.39 ± 0.02 11.01 ± 0.17 24.97 ± 0.48 1.87 ± 0.22
Trh/L-leu (5%) 3.15 ± 0.03 4.72 ± 0.04 9.71 ± 0.53 1.39 ± 0.2 HPMC K100M/L-leu (5%) 3.94 ± 0.01 8.71 ± 0.11 20.79 ± 0.33 1.93 ± 0.15
Trh/L-leu (7.5%) 3.13 ± 0.06 4.68 ± 0.04 8.99 ± 0.14 1.25 ± 0.08 HPMC K100M/L-leu (7.5%) 4.29 ± 0.02 10.75 ± 0.3 25.28 ± 1.34 1.95 ± 0.55
Trh/L-leu (10%) 3.14 ± 0.02 4.72 ± 0.01 9.82 ± 0.22 1.42 ± 0.08 HPMC K100M/L-leu (10%) 3.76 ± 0.02 8.38 ± 0.19 20.99 ± 0.95 2.06 ± 0.39
Trh/L-leu (15%) 3.07 ± 0.11 4.88 ± 0.36 13.33 ± 4.13 2.1 ± 1.53 HPMC K100M/L-leu (15%) 4.1 ± 0.02 9.82 ± 0.22 24.12 ± 0.81 2.04 ± 0.35
Trh/L-leu (20%) 3.36 ± 0.03 5.28 ± 0.03 11.11 ± 0.18 1.47 ± 0.08 HPMC K100M/L-leu (20%) 4.07 ± 0.03 8.99 ± 0.11 20.62 ± 1.3 1.84 ± 0.48
Trh/L-leu (25%) 3.64 ± 0.03 6.37 ± 0.06 12.49 ± 0.24 1.39 ± 0.11 HPMC K100M/L-leu (25%) 4.14 ± 0.02 9.49 ± 0.09 21.62 ± 0.47 1.84 ± 0.19
Trh/L-leu (30%) 3.76 ± 0.03 6.37 ± 0.05 11.07 ± 0.06 1.15 ± 0.04 HPMC K100M/L-leu (30%) 4.09 ± 0.02 9.08 ± 0.13 20.11 ± 0.22 1.76 ± 0.12
PVP K90/L-leu (0%) 3.96 ± 0.03 7.99 ± 0.07 16.3 ± 0.12 1.54 ± 0.07
PVP K90/L-leu (5%) 4.32 ± 0.04 8.93 ± 0.04 18.13 ± 0.31 1.55 ± 0.13
PVP K90/L-leu (10%) 4.06 ± 0.66 8.96 ± 0.62 18.41 ± 0.91 1.6 ± 0.73
PVP K90/L-leu (15%) 3.98 ± 0.01 8.19 ± 0.02 16.18 ± 0.15 1.49 ± 0.06
PVP K90/L-leu (20%) 3.98 ± 0.02 8 ± 0.04 15.72 ± 0.23 1.47 ± 0.09
PVP K90/L-leu (25%) 4.16 ± 0.02 8.89 ± 0.11 18.12 ± 0.53 1.57 ± 0.22
PVP K90/L-leu (30%) 4.09 ± 0.02 8.27 ± 0.05 16.41 ± 0.53 1.49 ± 0.2
Pharmaceutics 2023, 15, 1447
Figure 7. P-XRD diffractograms of various spray dried formulations: (A) MD/L-leu (0%), (B) MD/L-
leu (20%) (C) Trh/L-leu (0%), (D) Trh/L-leu (20%), (E) GA/L-leu (0%), (F) GA/L-leu (20%), (G) HPMC
E5LV/L-leu (0%), (H) HPMC E5LV/L-leu (20%), (I) Pure unspray-dried L-leu.
125
Pharmaceutics 2023, 15, 1447
Figure 8. Visual representation of formulation shear cell data from the Freeman FT4 powder rheome-
ter: (A) cohesion data, (B) flow function coefficient (ffc) data, and (C) angle of internal friction (AIF)
data. All data points are displayed as mean ± SD (n = 3).
126
Pharmaceutics 2023, 15, 1447
The relationship between cohesion and ffc values was previously reported by Wang et al.,
who showed that these parameters derived from low-cohesion powders with the same
preconsolidation stress had a statistically substantial inverse correlation with low-cohesion
powders [48]. The results of this study further support this relationship.
HPMC formulations could be classified under the ‘cohesive’ category, as the majority
of their ffc values remained below four. For maltodextrin, trehalose, and PVP K10, a L-leu
concentration of ≥5 wt% could be classified as easily flowing. Gum arabic only became
‘easy-flowing’ after a concentration of 15 wt%. No powders could be definitively identified
as ‘free-flowing” owing to the high degree of variability in the recorded data between the
different runs.
127
Pharmaceutics 2023, 15, 1447
128
Pharmaceutics 2023, 15, 1447
maximum flowability improvement through L-leu coprocessing remained the same for
trehalose and gum arabic powders, requiring concentrations of ≥5 wt% and ≥20 wt%, re-
spectively. These results highlight the differences in the behaviour of these surface-modified
powders under confined and unconfined flow conditions.
Figure 9. Visual representation of formulation stability data from the Freeman FT4 powder rheometer:
(A) BFE value derived from the standard protocol; (B) BFE value derived from the modified protocol;
(C) SE value derived from the standard protocol; (D) SE value derived from the modified protocol;
(E) MD/L-leu stability data from the standard protocol; (F) MD/L-leu stability data from the modified
protocol. All data points are displayed as mean ± SD (n = 3).
129
Pharmaceutics 2023, 15, 1447
Figure 10. Visual representation of formulation variable flow rate data from the Freeman FT4 powder
rheometer: (A) FRI values of each formulation tested; (B) MD/L-leu powder FRI data; (C) PVP
K10/L-leu powder FRI data; (D) PVP K90/L-leu powder FRI data; (E) Trh/L-leu powder FRI data;
and (F) GA/L-leu powder FRI data. All data points are displayed as mean ± SD (n = 3).
130
Pharmaceutics 2023, 15, 1447
L-leu coprocessing mostly did not adversely affect the flowrate sensitivity of the PVP
K10 and PVP K90 powders. Figure 10A shows that, at 5 wt%, the PVP K10 powders
experienced a substantial increase in the FRI values; however, this value plateaued at its
original level at ≥10 wt%. Furthermore, Figure 10C highlights PVP K10s with a generally
low flowrate sensitivity at most L-leu concentrations. A pattern repeated with PVP K90
powders, regardless of L-leu concentration, flowrate sensitivity, and FRI values, remained
consistent (Figure 10F). Conversely, gum arabic powder became increasingly more sensitive
to the flowrate as the L-leu concentration increased, reaching a high of 3.82 ± 0.27 at 20 wt%
(Figure 10A,F).
It was clear that the flowrate sensitivity was a bulk powder characteristic affected by L-
leu coprocessing. Further investigations are required to better understand the mechanistic
underpinnings of this phenomenon. This was coupled with another noteworthy behaviour:
during the stability test, formulations coprocessed with L-leu required several test cycles
before they reached a steady state (Figure 9E). We theorise that this was a consequence of
the powders needing to reach a steady state of compaction when exposed to downward
motion of the blade. Because the FRI value only took one total energy measurement at every
blade speed, it was considered that these powders did not reach a steady state. Therefore,
in response to these two observed powder behaviour issues, a modified stability test was
conducted to determine whether it could provide a better powder characterisation.
3.7. Permeability
Permeability testing combines both compaction and air entrapment to characterise
each formulation. It provided a relative assessment of the packing efficiency of each
powder under different normal-stress regimes, as packing controls the bed porosity and
thus the resistance to airflow. Figure 11 summarises the permeability test results for
several excipients of interest. It shows two types of bulk behaviour profiles. The first,
where the packing efficiency improved from L-leu coprocessing, was more noticeable
under higher normal-stress regimes. PVP K10, trehalose, and HPMC were classified
under this category. Except for 0 wt% and 2.5 wt%, PVP K10 formulations experienced a
narrow range of pressure drop across the bed (8.6 ± 1.7 mBar, 25 wt% to 12.3 ± 2.1 mBar,
10 wt%). However, as the normal pressure increased, substantial differentiation was
observed (Figure 11B). A similar pattern was also observed with the trehalose and HPMC
formulations; at low normal pressures (1 kPa), the powders had a narrow range of pressure
131
Pharmaceutics 2023, 15, 1447
drops between the different L-leu concentrations. At higher normal pressures (15 kPa),
the recorded pressure drops of each trehalose formulation were distinct from one another
(Figure 11C,E). Although relatively modest in its improvement of packing efficiency, data
from Figure 11E suggest that L-leu coprocessing had a positive impact on the performance
of HPMC powders.
Figure 11. Visual representation of the formulation permeability data from the Freeman FT4 powder
rheometer: (A) MD/L-leu powder permeability data, (B) PVP K10/L-leu powder permeability data,
(C) Trh/L-leu powder permeability data, (D) GA/L-leu powder permeability data, and (E) HPMC
E5LV/L-leu powder permeability data. All data points are displayed as mean ± SD (n = 3).
132
Pharmaceutics 2023, 15, 1447
The second category of powders had great differentiation between each L-leu concen-
tration, regardless of the applied normal stress. Both maltodextrin and gum arabic powders
displayed this behaviour (Figure 11A,D). At both low and high normal pressures, increas-
ing the L-leu concentration generally resulted in a higher pressure drop across the powder
bed. From these data, it could be inferred that the larger the increase in pressure drop as the
L-leu concentration increased, the greater the improvements conferred towards packing
efficiency. Therefore, PVP K10 and gum arabic benefitted the most from L-leu coprocessing.
4. Discussion
The recent focus of academic research on L-leu in the coformulation of powders has
been on understanding the mechanistic reasons behind its surface deposition and the
ongoing self-assembly and crystallisation processes. However, studies have not evaluated
its comparative effects when coprocessed into powders with different amorphous-forming
excipient groups, especially its effect on powder flow. Therefore, the primary goal of this
study was to investigate the effect of L-leu on the bulk flowability of excipients coprocessed
at varying levels.
Qualitative observation of the SEM images compiled during this study showed that
L-leu surface modification resulted in a wide range of different morphologies based on
the level of the coformulated excipient and excipient nature. Many morphologies have
been previously observed, while some of our observed morphologies relating to L-leu
levels have not been previously reported. The deleterious impact of a low concentration
(2.5–10 wt%) L-leu on PVP K90, forming fibrous spindle byproducts, has not been reported.
In addition, the formation of flaky spherical structures in maltodextrin does not conform to
the previously reported patterns of a more corrugated or collapsed morphology.
Another example of morphological differences is the comparison between PVP K10
and the trehalose-based formulations. The PVP K10 particles exhibited increased corru-
gation up to ≥10 wt% L-leu, where the collapsed morphologies became prominent. In
contrast, the trehalose powder retained a corrugated dimpled spherical morphology, re-
gardless of L-leu content. The morphological progression of both excipients had been
examined in past publications [22,49]. Vehring et al. theorised that spray-dried powder
morphology could be predicted using the Péclet number [18]. The Péclet number is defined
as the ratio between the droplet surface evaporation rate and the diffusivity of the excipient
material within the droplet. Given that spray-drying conditions remained constant for all
formulations, the only differences were the excipient physico-chemical properties and L-leu
concentrations. PVP K10 has a larger molecular structure than trehalose and is a known
film-forming material [50]. Because molecular size affects the diffusivity of a material,
PVP-based formulations would have a higher Péclet number than trehalose. Therefore,
we theorise that a combination of PVP K10 surface accumulation and a coherent L-leu
crystalline shell that forms at ≥10 wt% concentration offers stronger resistance, which
restricts the ability of water vapour to escape from the droplet core, leading to collapsed
morphologies. In contrast, trehalose is not classified as a film-forming agent. Hence, tre-
halose should diffuse more readily within the drying droplet, resulting in the formation of
more spherical particles.
Increasing the L-leu content substantially improved the bulk flowability performance
of the excipients tested, with the notable exception of HPMCs. Two mechanisms of action
have historically been used to account for flowability improvements: (a) surface corru-
gation, which results in decreased particle contact areas, and (b) a coherent L-leu shell,
which decreases the surface energy of spray-dried formulations. These mechanisms are not
mutually exclusive and can work in complement, but our results presented here indicated
that improvements across the different excipients were not simply explained by observed
surface corrugation, and that a coherent L-leu shell is a more appropriate explanation
for flowability improvements. For example, consider the case of trehalose and HPMC.
Increasing the concentration of L-leu did not result in observed morphological changes.
However, bulk characterisation data showed that coprocessing improved trehalose flow
133
Pharmaceutics 2023, 15, 1447
characteristics. In contrast, HPMC powders did not exhibit any substantial bulk flowability
with increasing L-leu concentration. Therefore, this study highlighted that the relation-
ship between flowability enhancement and morphological changes cannot be consistently
predicted or generalised.
Bulk characterisation data have previously shown that above a certain L-leu concen-
tration, flowability improvements reached an optimum and plateaued [22,47], which is
supported by our results. However, our results showed that the threshold concentration
differed between the excipients. Two flowability tests showed this progression: shear
cell and stability tests. Shear cell testing showed the behaviour of these powders in a
confined high-stress environment, reminiscent of what could be observed in a hopper feed
system [47]. PVP K10, maltodextrin, and trehalose exhibited this plateau behaviour in both
the cohesion and ffc data. Finally, the AIF data did not provide substantial insights into the
differences between formulations.
The stability test represented the behaviour of these powders in the case of an uncon-
fined powder flow. The BFE obtained from the standard stability test was insufficient to
differentiate the formulations. Unlike the shear cell, BFE did not differentiate any formula-
tion once L-leu had sufficiently minimised the interparticular cohesion (≤20 mJ). Therefore,
we conclude that SE provides a more representative measure of the flow behaviour of
a formulated powder under unconfined stress. As this test measured the blade leaving
the powder bed, it did not consider the energy required to compact the bed. The results
were normalised by sample mass, eliminating bulk density differences and giving a better
account of powder cohesion.
A comparison of both confined and unconfined data shows how these surface-modified
powders perform differently under different consolidation conditions. For example, shear
cell data showed that coprocessed maltodextrin achieved maximum flowability improve-
ment at a lower L-leu concentration than PVP K10 (≥5 wt% vs. ≥10 wt%). This trend was
reversed in the SE data, where PVP K10 performed better than maltodextrin, achieving an
improvement plateau at ≥5 wt% compared to ≥10 wt%.
In addition, the FRI data showed how altering L-leu level coprocessing altered the
flowrate sensitivity of the powder. This is attributed to the differences in packing behaviour
and efficiency of each formulation. At a lower speed, greater sensitivity is shown; therefore,
we propose that the slower application of force allowed more time and opportunity for
the materials to consolidate and form a bulk resistance towards the blade, increasing
the total energy measured. The FRI data showed that PVP K10 was not greatly affected
by alterations in blade speed, whereas maltodextrin and trehalose exhibited substantial
flowrate sensitivity as a function of blade speed.
To further explore this observation, a modified stability test was used, where the
blade speed was reduced from 100 mm/s to 60 mm/s. It showed better differentiation
for each formulation at a lower speed. For example, at a faster speed, trehalose displayed
the lowest BFE among all the excipients. However, in the lower speed regime, trehalose
formulations plateaued at a BFE higher than most other excipients. In addition, the stability
test was unable to differentiate maltodextrin formulations in the 10–30 wt% L-leu range,
whereas this modified test could differentiate the powders. We believe that this provides
an improved approach to enhance the characterisation and differentiation of these powders
based on their flow-rate sensitivities.
The permeability test provided further insights into the compactability and air en-
trapment efficiency of leucine coprocessed powders. The powders were evaluated under
a range of normal-stress regimes. The greatest difference between L-leu modifications
occurred in the higher-stress regime (15 kPa). In general, a higher L-leu concentration
resulted in an increase in the pressure drop across the powder bed. We attributed this to
the increased air resistance and entrapment to their surface state and morphologies, which
allowed them to pack more efficiently when subjected to an external force. All of these
bulk tests showed that there were multiple exceptions and variations in the generalised
preconception that L-leu improves bulk powder flowability.
134
Pharmaceutics 2023, 15, 1447
The excipients maltodextrin, PVP K10, and trehalose were homogeneous in nature
and exhibited substantial surface modification from L-leu coformulation. Gum arabic
coprocessed with L-leu was the only excipient investigated with a heterogeneous nature,
containing a mixture of polysaccharides and glycoproteins [51]. It displayed morphological
progression, such as maltodextrin, blood cell-like structures when purely spray dried, and
rugose morphologies with substantial (≥20 wt%) L-leu. There was also a gradual increase
in flowability at relatively higher L-leu concentrations. It is understood that other amino
acids have surface activity similar to that of L-leu; however, they do not display the same
degree of flowability improvements [17]. Since gum arabic contains a substantial proportion
of amino acids within its composition, it was speculated that this may lead to a degree
of interruption in the formation of the historically reported L-leu shell. This example of
gum arabic and L-leu formulations highlights the potential challenges and uncertainties in
outcomes when coprocessed with heterogeneous formulations, with one or more APIs and
other components, which may be expected in pharmaceutical and nutraceutical products.
Larger macromolecule excipients, such as PVP K90 and HPMCs, displayed contrasting
results when coprocessed. PVP K90 displayed severe incompatibility at low L-leu concen-
trations (2.5–10 wt%) with the formation of fibrous byproducts, which necessitates future
investigation. However, observing the morphologies of the free powders showed that they
exhibited the same morphologies as the PVP K10 powders. Furthermore, at ≥15 wt% L-leu,
PVP K90 formulations had bulk characteristics similar to those of PVP K10. This indicated
that the large difference in molecular size did not greatly impact the ability of L-leu to
modify the powders formed from these excipients.
The HPMC E5LV and K100M formulations displayed contrasting characteristics com-
pared to the other excipients. Morphologically, most bulk flowability tests showed that
L-leu had little to no effect on the particles formed. However, XRD analysis showed the
presence of the characteristic reported 6◦ and 19◦ 2θ peak, associated with a crystalline
L-leu surface shell. These peaks in the HPMC XRDs displayed a degree of broadening
compared with the other formulations. This may be consistent with the differences in
the L-leu formation and the alternative form on the surface. HPMC is a known viscosity-
modifying material [44]. Feng et al. reported that viscosity may play a role in hindering
the ability of L-leu to form a crystalline shell [16]. The HPMCs used were also the largest
molecules examined here and therefore would have a higher Péclet number relative to
L-leu. HPMC was also reported to have surface activity, where it was able to improve
the flowability of various ASDs in common with L-leu [52]. Therefore, a combination
of such factors may result in HPMC having a physico-chemical interaction with L-leu,
preventing it from adequately forming a well-ordered crystalline structure. Data from
this study showed the contextual interaction between L-leu’s mechanism of action and
the different physico-chemical characteristics of the coformulated excipients. Furthermore,
it highlighted how the different FT4 powder rheometer tests characterised the different
physical properties of each formulation.
This study demonstrated the impact of L-leu surface modification on the bulk flowa-
bility performance of excipients in spray-dried formulations, but it also raised questions
regarding the underlying mechanisms of L-leu surface crystallisation. In future work, to
gain a deeper understanding of these mechanisms, we propose that various spectroscopic
techniques, such as Fourier-transform infrared spectroscopy, X-ray photoelectron spec-
troscopy, and time-of-flight secondary ion mass spectrometry, can be used to investigate the
surface characteristics of the modified materials. Additional techniques, such as inverse gas
chromatography, can also provide insight into the surface characteristics of these surface
modified powders. Future studies may focus on gum arabic and HPMC formulations, as
their interaction with L-leu may offer insights into their mechanism of action through their
disruption. Future work should also incorporate model materials with additional studies
to examine aspects such as chemical stability and dissolution to further explore the impact
of L-leu surface modification on the behaviour of spray-dried formulations.
135
Pharmaceutics 2023, 15, 1447
5. Conclusions
Overall, this study showed the effects of L-leu coprocessing via spray drying on the
bulk powder characteristics of different excipient formulations. Effects, such as improved
flow and altered compaction, are likely to lead to better control of postprocessability. The
addition of L-leu substantially modified the performance of excipients, such as PVP K10,
maltodextrin, and trehalose, which may help in the design of potential ASDs. However, we
observed novel behaviours of other common excipients, such as PVP K90, HPMCs, and
gum arabic, which exhibited differing behaviours. This highlights that L-leu behaviour in
particle formation differs with the differing physico-chemical properties of the coformulated
excipients and will not be directly predictable for all materials. This work highlights
the need for a better understanding of its mechanistic behaviour in multicomponent
systems, especially with increasing complexity of the composition. This also demonstrates
the variability of results from different standard bulk characteristics of such formulated
powders and the importance of suitable test design and control in identifying the nature of
bulk character changes due to surface modification with a coexcipient, such as L-leu.
Author Contributions: Conceptualization, D.S., D.A.V.M. and K.P.H.; methodology, D.S., Y.-D.D., P.W.
and D.A.V.M.; validation, D.S. and D.A.V.M.; investigation, D.S. and Y.-D.D.; resources, D.S., D.A.V.M.,
P.W. and Y.-D.D.; data curation, D.S.; writing—original draft preparation, D.S.; writing—review and
editing, D.S. and D.A.V.M.; visualization, D.S.; supervision, D.A.V.M. and K.P.H.; project administration,
D.S., D.A.V.M. and K.P.H. All authors have read and agreed to the published version of the manuscript.
Funding: The authors would like to acknowledge the Australian Government for their support
through the Australian Government Research Training Program Scholarship.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: Not applicable.
Acknowledgments: Special mention to Reza Parvizi who was extraordinary in ensuring the smooth
safe operation in the laboratories. We would also like to thank Guy Stimpson and Amalia Thomas
from Freeman Technologies for their technical insight during methodology development. Addition-
ally, we would like to thank John Long for his insights into X-ray Diffraction. Finally, the author
would like to personally thank Wai Mun Wong for providing support in automating the data entry
process, saving a substantial amount of effort and sanity in the process.
Conflicts of Interest: The authors declare no conflict of interest.
Appendix A
Figure A1. Representative SEM images of spray dried HPMC K100M/L-leu formulations: (A) HPMC
K100M/L-leu (0%), (B) HPMC K100M/L-leu (5%), (C) HPMC K100M/L-leu (10%), (D) HPMC
K100M/L-leu (20%).
136
Pharmaceutics 2023, 15, 1447
References
1. Pourshahab, P.S.; Gilani, K.; Moazeni, E.; Eslahi, H.; Fazeli, M.R.; Jamalifar, H. Preparation and Characterization of Spray Dried
Inhalable Powders Containing Chitosan Nanoparticles for Pulmonary Delivery of Isoniazid. J. Microencapsul. 2011, 28, 605–613.
[CrossRef] [PubMed]
2. Mehanna, M.M.; Mohyeldin, S.M.; Elgindy, N.A. Rifampicin-Carbohydrate Spray-Dried Nanocomposite: A Futuristic Multipar-
ticulate Platform for Pulmonary Delivery. Int. J. Nanomed. 2019, 14, 9089–9112. [CrossRef] [PubMed]
3. Sou, T.; Forbes, R.T.; Gray, J.; Prankerd, R.J.; Kaminskas, L.M.; McIntosh, M.P.; Morton, D.A.V. Designing a Multi-Component
Spray-Dried Formulation Platform for Pulmonary Delivery of Biopharmaceuticals: The Use of Polyol, Disaccharide, Polysac-
charide and Synthetic Polymer to Modify Solid-State Properties for Glassy Stabilisation. Powder Technol. 2016, 287, 248–255.
[CrossRef]
4. Mangal, S.; Nie, H.; Xu, R.; Guo, R.; Cavallaro, A.; Zemlyanov, D.; Zhou, Q. (Tony) Physico-Chemical Properties, Aerosolization
and Dissolution of Co-Spray Dried Azithromycin Particles with L-Leucine for Inhalation. Pharm. Res. 2018, 35, 28. [CrossRef]
5. Shetty, N.; Park, H.; Zemlyanov, D.; Mangal, S.; Bhujbal, S.; Zhou, Q. (Tony) Influence of Excipients on Physical and Aerosolization
Stability of Spray Dried High-Dose Powder Formulations for Inhalation. Int. J. Pharm. 2018, 544, 222–234. [CrossRef]
6. Shi, H.; Mohanty, R.; Chakravarty, S.; Cabiscol, R.; Morgeneyer, M.; Zetzener, H.; Ooi, J.Y.; Kwade, A.; Luding, S.; Magnanimo, V.
Effect of Particle Size and Cohesion on Powder Yielding and Flow. KONA Powder Part. J. 2018, 2018, 226–250. [CrossRef]
7. Démuth, B.; Nagy, Z.K.; Balogh, A.; Vigh, T.; Marosi, G.; Verreck, G.; Van Assche, I.; Brewster, M.E. Downstream Processing of
Polymer-Based Amorphous Solid Dispersions to Generate Tablet Formulations. Int. J. Pharm. 2015, 486, 268–286. [CrossRef]
8. Tabor, D. Surface Forces and Surface Interactions; Academic Press, Inc.: New York, NY, USA, 1977.
9. Jain, S. Mechanical Properties of Powders. Bulk Solids Handl. 1988, 8, 615–624.
10. Singh, A.; Van den Mooter, G. Spray Drying Formulation of Amorphous Solid Dispersions. Adv. Drug Deliv. Rev. 2016, 100, 27–50.
[CrossRef]
11. Qu, L.; Zhou, Q.; Denman, J.A.; Stewart, P.J.; Hapgood, K.P.; Morton, D.A.V. Influence of Coating Material on the Flowability and
Dissolution of Dry-Coated Fine Ibuprofen Powders. Eur. J. Pharm. Sci. 2015, 78, 264–272. [CrossRef]
12. Ariyasu, A.; Hattori, Y.; Otsuka, M. Delay Effect of Magnesium Stearate on Tablet Dissolution in Acidic Medium. Int. J. Pharm.
2016, 511, 757–764. [CrossRef] [PubMed]
13. Seville, P.C.; Learoyd, T.P.; Li, H.Y.; Williamson, I.J.; Birchall, J.C. Amino Acid-Modified Spray-Dried Powders with Enhanced
Aerosolisation Properties for Pulmonary Drug Delivery. Powder Technol. 2007, 178, 40–50. [CrossRef]
14. Mah, P.T.; O’Connell, P.; Focaroli, S.; Lundy, R.; O’Mahony, T.F.; Hastedt, J.E.; Gitlin, I.; Oscarson, S.; Fahy, J.V.; Healy, A.M. The
Use of Hydrophobic Amino Acids in Protecting Spray Dried Trehalose Formulations against Moisture-Induced Changes. Eur. J.
Pharm. Biopharm. 2019, 144, 139–153. [CrossRef]
15. Alhajj, N.; O’Reilly, N.J.; Cathcart, H. Leucine as an Excipient in Spray Dried Powder for Inhalation. Drug Discov. Today 2021, 26,
2384–2396. [CrossRef] [PubMed]
16. Feng, A.L.; Boraey, M.A.; Gwin, M.A.; Finlay, P.R.; Kuehl, P.J.; Vehring, R. Mechanistic Models Facilitate Efficient Development of
Leucine Containing Microparticles for Pulmonary Drug Delivery. Int. J. Pharm. 2011, 409, 156–163. [CrossRef] [PubMed]
17. Sou, T.; Kaminskas, L.M.; Nguyen, T.H.; Carlberg, R.; McIntosh, M.P.; Morton, D.A.V. The Effect of Amino Acid Excipients on Mor-
phology and Solid-State Properties of Multi-Component Spray-Dried Formulations for Pulmonary Delivery of Biomacromolecules.
Eur. J. Pharm. Biopharm. 2013, 83, 234–243. [CrossRef]
18. Vehring, R. Pharmaceutical Particle Engineering via Spray Drying. Pharm. Res. 2008, 25, 999–1022. [CrossRef]
19. Lu, W.; Rades, T.; Rantanen, J.; Chan, H.K.; Yang, M. Amino Acids as Stabilizers for Spray-Dried Simvastatin Powder for
Inhalation. Int. J. Pharm. 2019, 572, 118724. [CrossRef]
20. Chang, R.Y.K.; Chan, H.K. Advancements in Particle Engineering for Inhalation Delivery of Small Molecules and Biotherapeutics.
Pharm. Res. 2022, 39, 3047–3061. [CrossRef]
21. Ordoubadi, M.; Shepard, K.B.; Wang, H.; Wang, Z.; Pluntze, A.M.; Churchman, J.P.; Vehring, R. On the Physical Stability of
Leucine-Containing Spray-Dried Powders for Respiratory Drug Delivery. Pharmaceutics 2023, 15, 435. [CrossRef]
22. Mangal, S.; Meiser, F.; Tan, G.; Gengenbach, T.; Denman, J.; Rowles, M.R.; Larson, I.; Morton, D.A.V. Relationship between Surface
Concentration of L-Leucine and Bulk Powder Properties in Spray Dried Formulations. Eur. J. Pharm. Biopharm. 2015, 94, 160–169.
[CrossRef] [PubMed]
23. Malamatari, M.; Somavarapu, S.; Kachrimanis, K.; Buckton, G.; Taylor, K.M.G. Preparation of Respirable Nanoparticle Agglom-
erates of the Low Melting and Ductile Drug Ibuprofen: Impact of Formulation Parameters. Powder Technol. 2017, 308, 123–134.
[CrossRef]
24. Sou, T.; Orlando, L.; McIntosh, M.P.; Kaminskas, L.M.; Morton, D.A.V. Investigating the Interactions of Amino Acid Components
on a Mannitol-Based Spray-Dried Powder Formulation for Pulmonary Delivery: A Design of Experiment Approach. Int. J. Pharm.
2011, 421, 220–229. [CrossRef]
25. Hofman, D.L.; van Buul, V.J.; Brouns, F.J.P.H. Nutrition, Health, and Regulatory Aspects of Digestible Maltodextrins. Crit. Rev.
Food Sci. Nutr. 2016, 56, 2091–2100. [CrossRef] [PubMed]
26. Sansone, F.; Mencherini, T.; Picerno, P.; D’Amore, M.; Aquino, R.P.; Lauro, M.R. Maltodextrin/Pectin Microparticles by Spray
Drying as Carrier for Nutraceutical Extracts. J. Food Eng. 2011, 105, 468–476. [CrossRef]
137
Pharmaceutics 2023, 15, 1447
27. Caliskan, G.; Nur Dirim, S. The Effects of the Different Drying Conditions and the Amounts of Maltodextrin Addition during
Spray Drying of Sumac Extract. Food Bioprod. Process. 2013, 91, 539–548. [CrossRef]
28. Ohtake, S.; Wang, Y.J. Trehalose: Current Use and Future Applications. J. Pharm. Sci. 2011, 100, 2020–2053. [CrossRef]
29. Maury, M.; Murphy, K.; Kumar, S.; Shi, L.; Lee, G. Effects of Process Variables on the Powder Yield of Spray-Dried Trehalose on a
Laboratory Spray-Dryer. Eur. J. Pharm. Biopharm. 2005, 59, 565–573. [CrossRef]
30. Focaroli, S.; Mah, P.T.; Hastedt, J.E.; Gitlin, I.; Oscarson, S.; Fahy, J.V.; Healy, A.M. A Design of Experiment (DoE) Approach to
Optimise Spray Drying Process Conditions for the Production of Trehalose/Leucine Formulations with Application in Pulmonary
Delivery. Int. J. Pharm. 2019, 562, 228–240. [CrossRef]
31. Sou, T.; Morton, D.A.V.; Williamson, M.; Meeusen, E.N.; Kaminskas, L.M.; McIntosh, M.P. Spray-Dried Influenza Antigen with
Trehalose and Leucine Produces an Aerosolizable Powder Vaccine Formulation That Induces Strong Systemic and Mucosal
Immunity after Pulmonary Administration. J. Aerosol Med. Pulm. Drug Deliv. 2015, 28, 361–371. [CrossRef]
32. Osman, M.E.; Williams, P.A.; Menzies, A.R.; Phillips, G. Characterization of Commercial Samples of Gum Arabic. J. Agric. Food
Chem. 1993, 41, 71–77. [CrossRef]
33. Ali, B.H.; Ziada, A.; Blunden, G. Biological Effects of Gum Arabic: A Review of Some Recent Research. Food Chem. Toxicol. 2009,
47, 1–8. [CrossRef] [PubMed]
34. Kurakula, M.; Rao, G.S.N.K. Pharmaceutical Assessment of Polyvinylpyrrolidone (PVP): As Excipient from Conventional to
Controlled Delivery Systems with a Spotlight on COVID-19 Inhibition. J. Drug Deliv. Sci. Technol. 2020, 60, 102046. [CrossRef]
[PubMed]
35. De Caro, V.; Murgia, D.; Seidita, F.; Bologna, E.; Alotta, G.; Zingales, M.; Campisi, G. Enhanced in Situ Availability of Apha-
nizomenon Flos-Aquae Constituents Entrapped in Buccal Films for the Treatment of Oxidative Stress-Related Oral Diseases:
Biomechanical Characterization and In Vitro/Ex Vivo Evaluation. Pharmaceutics 2019, 11, 35. [CrossRef] [PubMed]
36. Malik, S.; Kumar, A.; Ahuja, M. Synthesis of Gum Kondagogu-g-Poly(N-Vinyl-2-Pyrrolidone) and Its Evaluation as a Mucoadhe-
sive Polymer. Int. J. Biol. Macromol. 2012, 51, 756–762. [CrossRef] [PubMed]
37. Deshmukh, K.; Basheer Ahamed, M.; Deshmukh, R.R.; Khadheer Pasha, S.K.; Bhagat, P.R.; Chidambaram, K. Biopolymer Composites
with High Dielectric Performance: Interface Engineering; Elsevier Inc.: Amsterdam, The Netherlands, 2017; ISBN 9780081009741.
38. Chowhan, Z.T. Role of Binders in Moisture-induced Hardness Increase in Compressed Tablets and Its Effect on In Vitro
Disintegration and Dissolution. J. Pharm. Sci. 1980, 69, 1–4. [CrossRef] [PubMed]
39. Oh, C.M.; Heng, P.W.S.; Chan, L.W. A Study on the Impact of Hydroxypropyl Methylcellulose on the Viscosity of PEG Melt
Suspensions Using Surface Plots and Principal Component Analysis. AAPS PharmSciTech 2015, 16, 466–477. [CrossRef]
40. Ma, X.; Williams, R.O. Characterization of Amorphous Solid Dispersions: An Update. J. Drug Deliv. Sci. Technol. 2019, 50, 113–124.
[CrossRef]
41. Pandi, P.; Bulusu, R.; Kommineni, N.; Khan, W.; Singh, M. Amorphous Solid Dispersions: An Update for Preparation, Charac-
terization, Mechanism on Bioavailability, Stability, Regulatory Considerations and Marketed Products. Int. J. Pharm. 2020, 586,
119560. [CrossRef]
42. Freeman, R. Measuring the Flow Properties of Consolidated, Conditioned and Aerated Powders—A Comparative Study Using a
Powder Rheometer and a Rotational Shear Cell. Powder Technol. 2007, 174, 25–33. [CrossRef]
43. Freeman, R.E.; Cooke, J.R.; Schneider, L.C.R. Measuring Shear Properties and Normal Stresses Generated within a Rotational
Shear Cell for Consolidated and Non-Consolidated Powders. Powder Technol. 2009, 190, 65–69. [CrossRef]
44. Jin, C.; Wu, F.; Hong, Y.; Shen, L.; Lin, X.; Zhao, L.; Feng, Y. Updates on Applications of Low-Viscosity Grade Hydroxypropyl
Methylcellulose in Coprocessing for Improvement of Physical Properties of Pharmaceutical Powders. Carbohydr. Polym. 2023, 311,
120731. [CrossRef] [PubMed]
45. Ferdynand, M.S.; Nokhodchi, A. Co-Spraying of Carriers (Mannitol-Lactose) as a Method to Improve Aerosolization Performance
of Salbutamol Sulfate Dry Powder Inhaler. Drug Deliv. Transl. Res. 2020, 10, 1418–1427. [CrossRef] [PubMed]
46. Jiang, Z.; Bai, X. Effects of Polysaccharide Concentrations on the Formation and Physical Properties of Emulsion-Templated
Oleogels. Molecules 2022, 27, 5391. [CrossRef]
47. Schulze, D. Powders and Bulk Solids; Springer: Berlin/Heidelberg, Germany, 2007; Volume 13, ISBN 978-3-540-73767-4.
48. Wang, Y.; Koynov, S.; Glasser, B.J.; Muzzio, F.J. A Method to Analyze Shear Cell Data of Powders Measured under Different Initial
Consolidation Stresses. Powder Technol. 2016, 294, 105–112. [CrossRef]
49. Ordoubadi, M.; Gregson, F.K.A.; Wang, H.; Nicholas, M.; Gracin, S.; Lechuga-Ballesteros, D.; Reid, J.P.; Finlay, W.H.; Vehring, R.
On the Particle Formation of Leucine in Spray Drying of Inhalable Microparticles. Int. J. Pharm. 2021, 592, 120102. [CrossRef]
50. Franco, P.; De Marco, I. The Use of Poly(N-Vinyl Pyrrolidone) in the Delivery of Drugs: A Review. Polymers 2020, 12, 1114.
[CrossRef]
138
Pharmaceutics 2023, 15, 1447
51. Williams, P.A.; Phillips, G.O. Gum Arabic. In Handbook of Hydrocolloids; Elsevier: Amsterdam, The Netherlands, 2021; pp. 627–652.
52. Lin, X.; Chyi, C.W.; Ruan, K.F.; Feng, Y.; Heng, P.W.S. Development of Potential Novel Cushioning Agents for the Compaction
of Coated Multi-Particulates by Co-Processing Micronized Lactose with Polymers. Eur. J. Pharm. Biopharm. 2011, 79, 406–415.
[CrossRef]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual
author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to
people or property resulting from any ideas, methods, instructions or products referred to in the content.
139
pharmaceutics
Article
Effect of Mixer Type on Particle Coating by Magnesium
Stearate for Friction and Adhesion Modification
Wei Pin Goh, Ana Montoya Sanavia and Mojtaba Ghadiri *
Faculty of Engineering and Physical Sciences, University of Leeds, Leeds LS2 9JT, UK;
[email protected] (W.P.G.); [email protected] (A.M.S.)
* Correspondence: [email protected]
Abstract: Glidants and lubricants are often used to modify interparticle friction and adhesion
in order to improve powder characteristics, such as flowability and compactability. Magnesium
stearate (MgSt) powder is widely used as a lubricant. Shear straining causes MgSt particles to break,
delaminate, and adhere to the surfaces of the host particles. In this work, a comparison is made
of the effect of three mixer types on the lubricating role of MgSt particles. The flow behaviour
of α-lactose monohydrate, coated with MgSt at different mass percentages of 0.2, 0.5, 1, and 5 is
characterised. The mixing and coating process is carried out by dry blending using Turbula, ProCepT,
and Mechanofusion. Measures have been taken to operate under equivalent mixing conditions,
as reported in the literature. The flow resistance of the coated samples is measured using the FT4
rheometer. The results indicate that the flow characteristics of the processed powders are remarkably
similar in the cases of samples treated by Turbula and Mechanofusion, despite extreme conditions
of shear strain rate. The least flow resistance of samples is observed in the case of samples treated
by the ProCepT mixer. High-velocity collisions of particles round off the sharp corners and edges,
Citation: Goh, W.P.; making them less resistant to flow. The optimal percentage of magnesium stearate is found to be
Montoya Sanavia, A.; Ghadiri, M. approximately 1% by weight for all mixer types, as the addition of higher amounts of lubricant does
Effect of Mixer Type on Particle not further improve the flowability of the material.
Coating by Magnesium Stearate for
Friction and Adhesion Modification. Keywords: mixing; coating; magnesium stearate; lubricant; Turbula; ProCepT; Mechanofusion;
Pharmaceutics 2021, 13, 1211. flowability; rheometry; flow energy
https://doi.org/10.3390/
10.3390/pharmaceutics13081211
conclusion as that of Castellanos [14]. They used the image analysis technique to quantify
the SAC from SEM micrographs of zeolite-coated silanised glass beads, and found that
the optimal SAC for maximum flow improvement to be approximately 20%. Beyond this
optimal value, the flowabilility decreased, as the system was then dominated by guest–
guest rather than host–guest particle interactions. However, the work of Jallo et al. [16],
on the coating of APIs with nano-silica particles, found an optimal SAC value of 293%
instead, implying a multilayer coating gives rise to better flow performance. Sato et al. [17]
coated sugar granules with MgSt powder in a Cyclomix mixer and reported that the guest
particles were initially covered discretely, and then formed a film over the surfaces of the
host particles. In addition, attrition was induced by excessive operating conditions, which
resulted in the worsening of dry coating performance.
Through coating α-lactose monohydrate (α-LM) crystals with MgSt using the Mechan-
ofusion device, Zhou et al. [18] found that the optimum SAC to be 64.5%, estimated from
X-ray photoelectron spectroscopy and time-of-flight of the secondary ion mass spectrometry
(Tof-SIMS). However, MgSt delaminates upon shearing [19–21] and the optimal amount
would depend on the mixer type and shearing conditions. Moreover, Hussain et al. [19]
propose that a Langmuir-type coating is first achieved with MgSt, followed by a patchy
coverage in powder form. Therefore, the characterisation of SAC of MgSt on lactose crystals
has yet to be satisfactorily established.
The preparation of a powder formulation is usually carried out on a small scale in a
laboratory mixer, and extension to a large-scale operation is very challenging. Therefore, a
‘recipe’ for mixing equivalency is highly desirable. The attention here is on batch mixing,
for which the literature is vast, addressing numerous aspects. These include the evaluation
of the performance of various mixer types, assessment of mixing time and operational
scale, operating conditions, the optimal amount of flow aids, and the development of
models [5,12,22,23]. Horibe et al. [24] recently assessed the effect of mixing time and the
concentration of flow aids on the flow properties of the end product. They found that the
mixing time has a higher correlation to the flow properties than the concentration of flow
aids, signifying the importance of the process and mixing conditions in improving the flow
behaviour of a particulate system. As quoted from Bridgwater’s review paper [9], process
design and operation are largely based on judgement rather than science. Finding the
optimal mixing conditions for new materials is often a time-consuming and challenging
task. The quality of a mixture depends on the mixer selected. Asachi et al. [25] critically
evaluated current techniques for the evaluation of powder mixing. Of special interest here
is the work of Barling et al. [26] who carried out an experimental evaluation of various
mixer types, and presented a methodology for establishing an ‘equivalent extent’ of batch
mixing amongst various mixer types at different scales. They used fine iron oxide powder
as a tracer, and evaluated its spreading over the surfaces of α-lactose monohydrate in
several batch mixing devices by monitoring the changes in the colour intensity and hue
through a colour measurement methodology. The powder blend changed colour as a result
of mixing and coating. By quantifying changes in the intensity and hue of the colour of the
mixture, the extent of mixing and coating of iron oxide was determined, based on which
a mixing ‘equivalency’ was proposed for different mixer types. It was found that under
specific operating conditions, a mixing equivalency could be established among different
coating devices, namely Turbula T2F (figure-eight tumbler type), Key International KG5
(high shear mixer, two sizes, 1 L and 5 L), Hosokawa AMS-MINI Mechanofusion (extreme
shear rate), T.K. Fielder high shear mixers (TRV25 and PMA65), and Diosna P100 (high
shear mixer type).
In this work, the above equivalency criterion is followed and applied to the coating
of α-LM by MgSt powder. In particular, samples treated by different mixers, but with
equivalent mixing conditions, are prepared and analysed for the effect of MgSt on the flow
behaviour of α-LM. Three types of mixers, Turbula, Mechanofusion, and ProCepT high
shear mixer are used. The last one is close in geometry and operation to the T.K. Fielder
TRV25, 1 L high shear mixer. The equivalent mixing conditions, estimated based on the
141
Pharmaceutics 2021, 13, 1211
work of Barling et al. [26], are given in Table 1. The flowability of the coated samples,
prepared at different MgSt compositions and mixer types, is then evaluated based on the
mechanical work expended to penetrate a rotating impeller into a confined powder bed
using the Freeman Technology FT4 rheometer. This will also enable the sensitivity to the
shear strain rate to be quantified. The outcome provides a guideline on the amount of MgSt
to be used for its effect on the flowability.
Table 1. Mixing speed and time for achieving mixing equivalency adapted from Barling et al. [26],
Elsevier, 2015.
142
Pharmaceutics 2021, 13, 1211
143
Pharmaceutics 2021, 13, 1211
Figure 1. Flow chart of 11 standard FT4 tests (C: Conditioning; TN (RPM)): Test number N at
specific RPM).
144
Pharmaceutics 2021, 13, 1211
Figure 2. Specific downward flow energy of α-lactose monohydrate (α-LM) for twelve samples,
eight samples as received and unwashed, and four samples washed with isopropyl alcohol, showing
sample to sample variation. Data are for eleven consecutively repeated tests for each sample, showing
test stability and strain rate sensitivity, the first eight at 100 RPM and the last three at 70, 40, and
10 RPM.
145
Pharmaceutics 2021, 13, 1211
Figure 3. Specific flow energies of α-lactose monohydrate treated with different percentages of
magnesium stearate in the three mixers. (a) Downward test; (b) Upward test. Each data point
represents the mean specific flow energy of the triplicate FT4 measurements taken from Test 1. The
error bar represents the standard deviation. Note: Error bars that cannot be seen are smaller than
the symbols.
146
Pharmaceutics 2021, 13, 1211
Feed Turbula
Mechanofusion ProCepT
In terms of SUFE, the trend of the results of samples treated with MgSt by Turbula
and Mechanofusion almost overlap. ProCepT-treated samples generally have much better
flow performance compared to Turbula- and Mechanofusion-treated samples at the same
MgSt wt%.
147
Pharmaceutics 2021, 13, 1211
Table 2. Particle size distribution of α-lactose monohydrate particles, treated by Turbula, Pro-
CepT, and Mechanofusion with 0.2 wt% MgSt under equivalent mixing conditions according to
Barling et al. [26].
% Mass
Sieve Size, μm Mechanofusion ProCepT Turbula
Mean SD Mean SD Mean SD
<150 0.13 0.03 0.19 0.10 0.08 0.05
150–180 0.18 0.18 0.15 0.00 0.19 0.11
180–200 0.62 0.33 1.11 0.24 0.41 0.15
>200 99.06 0.29 98.55 0.29 99.32 0.23
4. Conclusions
The influence of mixer type on the coating of α-LM crystals with a small quantity of
MgSt powder has been assessed. Three different types of mixers, i.e., Turbula, ProCepT,
and Mechanofusion have been used. The assessment method is based on the effect of
MgSt coating on flowability, as measured by the mechanical work expended by penetrating
a rotating impeller into a powder bed by using an FT4 rheometer. The flowability of
α-LM crystals is affected by the presence of debris, i.e., fine particles of the same material,
adhered to their surfaces, causing a wide sample to sample variation. Washing the α-LM
crystals using a liquid carrier, which does not dissolve them, reduces the variation as
well as the magnitude of the expended work significantly, implying improved flowability.
The addition of MgSt to α-LM particles further improves the flowability for powders
coated in all of the three mixers under equivalent mixing conditions, described in the work
of Barling et al. [26]. There exists a threshold at ~1 wt% MgSt above which the further
addition of MgSt does not improve the bulk flow performance of α-LM. Samples coated
by both Turbula and Mechanofusion mixers show remarkably similar flow performance,
despite extreme shear strain rate. At the same MgSt composition, the ProCepT-treated
α-LM samples consistently perform better in terms of flow behaviour. This is attributed to
collisional impacts in this mixer, rounding off sharp corners and edges. Mixing equivalency
is best achieved between the Turbula and Mechanofusion mixers.
A breakage analysis due to mixing is also performed, and the results suggest that α-LM
samples coated using Turbula experience the least damage. The particle size distribution of
the treated samples reveals that attrition and surface abrasion create the dominant breakage
mode, particularly in the case of Turbula and Mechanofusion mixers, with their breakage
extent being found to be in the range of approximately 0.1%.
148
Pharmaceutics 2021, 13, 1211
References
1. Chen, Y.; Yang, J.; Dave, R.N.; Pfeffer, R. Fluidization of coated group C powders. AIChE J. 2008, 54, 104–121. [CrossRef]
2. Quintanilla, M.A.S.; Castellanos, A.; Valverde, J.M. Correlation between bulk stresses and interparticle contact forces in fine
powders. Phys. Rev. E 2001, 64, 031301. [CrossRef] [PubMed]
3. Espin, M.J.; Ebri, J.M.P.; Valverde, J.M. Tensile strength and compressibility of fine CaCO3 powders. Effect of nanosilica addition.
Chem. Eng. J. 2019, 378, 122166. [CrossRef]
4. Faqih, A.M.N.; Mehrotra, A.; Hammond, S.V.; Muzzio, F.J. Effect of moisture and magnesium stearate concentration on flow
properties of cohesive granular materials. Int. J. Pharm. 2007, 336, 338–345. [CrossRef]
5. Ouabbas, Y.; Chamayou, A.; Galet, L.; Baron, M.; Thomas, G.; Grosseau, P.; Guilhot, B. Surface modification of silica particles by
dry coating: Characterization and powder ageing. Powder Technol. 2009, 190, 200–209. [CrossRef]
6. Chattoraj, S.; Shi, L.; Sun, C.C. Profoundly improving flow properties of a cohesive cellulose powder by surface coating with
nano-silica through comilling. J. Pharm. Sci. 2011, 100, 4943–4952. [CrossRef] [PubMed]
7. Kunnath, K.; Huang, Z.; Chen, L.; Zheng, K.; Davé, R. Improved properties of fine active pharmaceutical ingredient powder
blends and tablets at high drug loading via dry particle coating. Int. J. Pharm. 2018, 543, 288–299. [CrossRef]
8. Muzzio, F.J.; Shinbrot, T.; Glasser, B.J. Powder technology in the pharmaceutical industry: The need to catch up fast. Powder
Technol. 2002, 124, 1–7. [CrossRef]
9. Bridgwater, J. Mixing of powders and granular materials by mechanical means—A perspective. Particuology 2012, 10, 397–427.
[CrossRef]
10. Mullarney, M.P.; Beach, L.E.; Dave, R.N.; Langdon, B.A.; Polizzi, M.; Blackwood, D.O. Applying dry powder coatings to
pharmaceutical powders using a comil for improving powder flow and bulk density. Powder Technol. 2011, 212, 397–402.
[CrossRef]
11. Pingali, K.; Mendez, R.; Lewis, D.; Michniak-Kohn, B.; Cuitino, A.; Muzzio, F. Evaluation of strain-induced hydrophobicity
of pharmaceutical blends and its effect on drug release rate under multiple compression conditions. Drug Dev. Ind. Pharm.
2011, 37, 428–435. [CrossRef] [PubMed]
12. Sato, A.; Serris, E.; Grosseau, P.; Thomas, G.; Chamayou, A.; Galet, L.; Baron, M. Effect of operating conditions on dry particle
coating in a high shear mixer. Powder Technol. 2012, 229, 97–103. [CrossRef]
149
Pharmaceutics 2021, 13, 1211
13. Conesa, C.; Saleh, K.; Thomas, A.; Guigon, P.; Guillot, N. Characterization of flow properties of powder coatings used in the
automotive industry. KONA Powder Part. J. 2004, 22, 94–106. [CrossRef]
14. Castellanos, A. The relationship between attractive interparticle forces and bulk behaviour in dry and uncharged fine powders.
Adv. Phys. 2005, 54, 263–376. [CrossRef]
15. Fulchini, F.; Zafar, U.; Hare, C.; Ghadiri, M.; Tantawy, H.; Ahmadian, H.; Poletto, M. Relationship between surface area coverage
of flow-aids and flowability of cohesive particles. Powder Technol. 2017, 322, 417–427. [CrossRef]
16. Jallo, L.J.; Ghoroi, C.; Gurumurthy, L.; Patel, U.; Dave, R.N. Improvement of flow and bulk density of pharmaceutical powders
using surface modification. Int. J. Pharm. 2012, 423, 213–225. [CrossRef] [PubMed]
17. Sato, A.; Serris, E.; Grosseau, P.; Thomas, G.; Galet, L.; Chamayou, A.; Baron, M. Experiment and simulation of dry particle
coating. Chem. Eng. Sci. 2013, 86, 164–172. [CrossRef]
18. Zhou, Q.; Qu, L.; Gengenbach, T.; Denman, J.A.; Larson, I.; Stewart, P.J.; Morton, D.A. Investigation of the extent of surface
coating via mechanofusion with varying additive levels and the influences on bulk powder flow properties. Int. J. Pharm.
2011, 413, 36–43. [CrossRef]
19. Hussain, M.S.H.; York, P.; Timmins, P. A study of the formation of magnesium stearate film on sodium chloride using energy-
dispersive X-ray analysis. Int. J. Pharm. 1988, 42, 89–95. [CrossRef]
20. Marwaha, S.B.; Rubinstein, M.H. Structure-lubricity evaluation of magnesium stearate. Int. J. Pharm. 1988, 43, 249–255. [CrossRef]
21. Shah, A.C.; Mlodozeniec, A.R. Mechanism of surface lubrication: Influence of duration of lubricant-excipient mixing on processing
characteristics of powders properties of compressed tablets. J. Pharm. Sci. 1977, 66, 1377–1382. [CrossRef]
22. Portillo, P.M.; Muzzio, F.J.; Ierapetritou, M.G. Characterizing powder mixing processes utilizing compartment models. Int. J.
Pharm. 2006, 320, 14–22. [CrossRef] [PubMed]
23. Siraj, M.S.; Radl, S.; Glasser, B.J.; Khinast, J.G. Effect of blade angle and particle size on powder mixing performance in a
rectangular box. Powder Technol. 2011, 211, 100–113. [CrossRef]
24. Horibe, M.; Sonoda, R.; Watano, S. Scale-Up of lubricant mixing process by using V-Type blender based on discrete element
method. Chem. Pharm. Bull. 2018, 66, 548–553. [CrossRef]
25. Asachi, M.; Nourafkan, E.; Hassanpour, A. A review of current techniques for the evaluation of powder mixing. Adv. Powder
Technol. 2018, 29, 1525–1549. [CrossRef]
26. Barling, D.; Morton, D.A.V.; Hapgood, K. Pharmaceutical dry powder blending and scale-up: Maintaining equivalent mixing
conditions using a coloured tracer powder. Powder Technol. 2015, 270, 461–469. [CrossRef]
27. Ho, R.; Muresan, A.S.; Hebbink, G.A.; Heng, J.Y.Y. Influence of fines on the surface energy heterogeneity of lactose for pulmonary
drug delivery. Int. J. Pharm. 2010, 388, 88–94. [CrossRef] [PubMed]
28. Bonakdar, T.; Ghadiri, M. Analysis of pin milling of pharmaceutical materials. Int. J. Pharm. 2018, 552, 394–400. [CrossRef]
29. Jones, M.D.; Price, R. The Influence of fine excipient particles on the performance of carrier-based dry powder inhalation
formulations. Pharm. Res. 2006, 23, 1665–1674. [CrossRef]
30. Majd, F.; Nickerson, T.A. Effect of Alcohols on Lactose Solubility. J. Dairy Sci. 1976, 59, 1025–1032. [CrossRef]
31. Freeman Technology. Stability & Variable Flow Rate Method. In FT4 Rheometer User Manual; Freeman Technology: Gloucester,
UK, 2007.
32. Madian, A.; Leturia, M.; Ablitzer, C.; Matheron, P.; Bernard-Granger, G.; Saleh, K. Impact of fine particles on the rheological
properties of uranium dioxide powders. Nucl. Eng. Technol. 2020, 52, 1714–1723. [CrossRef]
33. Kurz, H.P.; Munz, G. The influence of particle size distribution on the flow properties of limestone powders. Powder Technol.
1975, 11, 37–40. [CrossRef]
34. Liu, Y.; Lu, H.; Guo, X.; Gong, X.; Sun, X.; Zhang, Z. The influence of fine particles on bulk and flow behavior of pulverized coal.
Powder Technol. 2016, 303, 212–227. [CrossRef]
150
pharmaceutics
Article
Discrete Element Method Evaluation of Triboelectric Charging
Due to Powder Handling in the Capsule of a DPI
Francesca Orsola Alfano *, Alberto Di Renzo * and Francesco Paolo Di Maio
Abstract: The generation and accumulation of an electrostatic charge from handling pharmaceutical
powders is a well-known phenomenon, given the insulating nature of most APIs (Active Pharmaceu-
tical Ingredients) and excipients. In capsule-based DPIs (Dry Powder Inhalers), the formulation is
stored in a gelatine capsule placed in the inhaler just before inhalation. The action of capsule filling,
as well as tumbling or vibration effects during the capsule life cycle, implies a consistent amount
of particle–particle and particle–wall contacts. A significant contact-induced electrostatic charging
can then take place, potentially affecting the inhaler’s efficiency. DEM (Discrete Element Method)
simulations were performed on a carrier-based DPI formulation (salbutamol–lactose) to evaluate
such effects. After performing a comparison with the experimental data on a carrier-only system
under similar conditions, a detailed analysis was conducted on two carrier–API configurations with
different API loadings per carrier particle. The charge acquired by the two solid phases was tracked
in both the initial particle settling and the capsule shaking process. Alternating positive–negative
charging was observed. Particle charging was then investigated in relation to the collision statistics,
tracking the particle–particle and particle–wall events for the carrier and API. Finally, an analysis of
the relative importance of electrostatic, cohesive/adhesive, and inertial forces allowed the importance
of each term in determining the trajectory of the powder particles to be estimated.
lactose particles (a widely used carrier and excipient) depends on the chosen inhaler [8],
on the material of the containing capsule [12], and even on their manufacturing process;
for example, milled lactose tends to charge positively, while sieved lactose tends to charge
negatively [8]. Salbutamol sulphate (a commonly used API in inhalation formulations)
tends to acquire a positive charge if it is amorphous and a negative charge if it is in its
crystalline form [11]. Peart [13] reports that salbutamol charges positively when in contact
with PVC, while lactose charges negatively. However, when salbutamol particles detach
from lactose, the opposite polarity is measured for the two materials (salbutamol’s specific
charge is about −3000 nC/g, and lactose’s specific charge is about 100 nC/g).
When strong size polydispersity is present, the charge polarity also depends on
particle size, with smaller particles that charge negatively and larger particles that charge
positively [11,14–16]. This effect cannot be predicted using the triboelectric series, which is
usually employed to rank the electronegativity of powder materials [1,2]. The concentration
of the API plays a role as well, with the electrostatic charge of drug–carrier mixtures usually
decreasing by increasing the concentration of the API [10,17,18].
Electrostatics also plays a major role in the design and use of hard capsules for capsule-
based DPI. Chow et al. [19] found that mechanical vibration such as tapping induced
significant static charge on lactose stored in a gelatine capsule. Hoe et al. [10,17] hy-
pothesized that the surface charge on the capsule might be high enough to ionize the
surrounding air. Understanding such a variety of observations requires a careful analysis
of the links between microscopic charge transfer processes and the macroscopic mani-
festations. Particle-scale information is an important ingredient that is accessible only
through simulation.
In recent years, Discrete Element Method (DEM) modeling has emerged as a powerful
simulation tool for studying the behavior of powders in inhalation devices [20]. Cou-
pled DEM-CFD (Computational Fluid Dynamics) studies showed the detailed motion
of particles from the initial dispersion in air through to the mouthpiece [21–25]. In the
last decade, model formulations have been introduced to extend the particle-scale contact
tracking capability of DEM with surface–surface charge transfer and physical electrostatic
interaction models (see, e.g., [26–28]), allowing triboelectric charging phenomena at the
particle scale to be studied [27,29–33]. Naik et al. [34] studied the triboelectrification of
binary mixtures of drug and excipient in a blender. They found that particle–particle inter-
actions enhance the electrostatic interaction between the drug and excipient and decrease
the overall charge transfer between particles and walls. The overall charging process that
resulted was mitigated by this effect, leading to a lower total charge than that acquired by
single components, and differences in excipient concentration, in some cases, caused charge
polarity reversal. The importance of particle–particle contact charging was also highlighted
by Chowdhury et al. [33]. Zhu et al. [35] studied the contact electrification effect of selected
API agglomerates in the Turbuhaler® , finding a reduction in inhaler efficiency due to the
triboelectrification of powders. Specific numerical studies on the motion of particles in
the capsule of capsule-based DPIs are available in the literature [36–39], but the effects of
charges generated upon contact on the release of powder from the capsule itself are not
taken into account. Most of previous works in DPI applications focus on the influence
of electrostatic interactions between previously charged particles, rather than the charge
buildup process, and the charging of capsule walls is typically neglected.
The aim of the present study is to apply extended DEM simulations to evaluate
triboelectric charging dynamics of a lactose–salbutamol binary mixture in the chargeable
capsule of a DPI. The capsule is vibrated to simulate the routine handling operations to
which a filled capsule is subjected during its life cycle, with the aim to assess the extent
to which such movements give rise to electrostatic charge accumulation. The simulation
tool offers the possibility to relate individual contact events and local charge transfer to
the macroscopic influence that such a charge exerts on the material dynamics. The model
formulation, simulation setup, and material parameter are presented first. After validation,
152
Pharmaceutics 2023, 15, 1762
where γ is the surface energy, a is the radius of the contact area, vn is the normal velocity,
δn is the normal overlap, ηnH is the normal damping coefficient (related to the restitution
coefficient; see, e.g., [41]), and Eeq and Req are the equivalent Young modulus and radius of
the two contacting particles (i and j) [42].
The normal overlap, δn , is related to the radius of the contact area as follows:
"
a2 4πγa
δn = − (4)
Req Eeq
As mentioned above, according to the JKR theory, during the detachment phase,
the contact remains active at negative overlaps between the spheres (as actual surfaces
are elongated shapes that are still in contact) until a threshold overlap is reached. The
maximum attractive force, usually referred to as the pull-off force, occurs at a negative
overlap and is given by the following:
Fpull −o f f = 3πγReq (5)
The tangential contribution to the contact is considered following the no-slip solution
of Mindlin and Deresiewicz [43] for the frictional–elastic part and a velocity-dependent
dissipation term similar to the normal direction. The tangential contact force is calculated
as follows: 1 1
Fc,ij = − μs Fc,ij , 8Geq Req δn δt + ηt δn vt
t n 2 H 4
(6)
where μs is the static friction coefficient, Geq is the equivalent shear modulus, δt is the
tangential overlap, ηtH is the tangential damping coefficient, and vt is the tangential velocity.
153
Pharmaceutics 2023, 15, 1762
In the rotational motion, the contact torque results from the action of the tangential
contact force. The rolling friction torque is calculated according to the Constant Directional
Torque model [44], introducing the rolling friction coefficient, μr , as material parameter.
All the models presented above are described in more detail in Alfano et al. [39,45].
The DEM simulations were carried out using an in-house customized version of the
open-source code MFIX (NETL MFS, Department of Energy (Morgantown, WV, USA), ver-
sion 18.1.5 [46]. Johnson-Kendall-Roberts (JKR) model for the cohesive force and constant
directional torque (CDT) model for the rolling friction were implemented in the original
version of the code (see [45] for more details). Moreover, a different approach for wall
contacts [47] was preferred to the standard MFIX formulation. Special precautions have
also been taken to prevent very fine particles from being unrealistically pushed out of
the domain by the carrier particles, crossing the domain boundary. In terms of hardware
resources, an own-managed cluster was extensively utilized by running parallel tests on
up to 32 cores.
where z p is the cutoff distance for the particle–particle charge transfer, and it is set to
260 nm [27].
154
Pharmaceutics 2023, 15, 1762
䌚㻌
㻌
䌚㻌
(a) (b)
Figure 1. (a) Schematics of contact charging mechanism. (b) Graphical representation of particle–
particle electrostatic forces evaluated within a cutoff distance.
Once each particle has its own charge, the second ingredient of the model, the electrostatic
force between two charged particles (Figure 1b), is calculated according to Coulomb’s law:
→e 1 qi q j
F i,j = 2
n̂i,j (10)
4πε 0 ri,j
where n̂i,j is the unit vector defined along the direction connecting the two particles’ centers,
and ri,j is the distance between these centers. The cutoff distance for the calculation of the
electrostatic interactions is set to 1.2 times the sum of the two particles’ radii.
Coulombic interactions between charged particles and the walls of the capsule were
considered according to the method of mirror charges [50]:
→e 1 q2i
F i,s = n̂ (11)
4πε 0 (2ri,s )2 i,s
where ri,s is the distance between the wall surface and the center of the particle, and n̂i,s is the
unit vector perpendicular to the surface and passing through the particle center. It is useful
to note that while particle–particle interactions may be attractive or repulsive depending
on the charge polarity, particle–wall electrostatic interactions are always attractive.
155
Pharmaceutics 2023, 15, 1762
The work functions were calculated from molecular orbital calculation (MOPAC) by
Naik et al. [9] for the lactose and by Zellnitz et al. [53] for salbutamol. The work function
of the gelatine capsule was not found in the literature. Pinto et al. [12] observed that
the charge acquired by the capsules in contact with stainless steel is about 60% of the
charge acquired by the capsules in contact with PVC. Since the work function of PVC is
5.33 eV [34] and that of stainless steel is 5.05 eV [54], and since the charge is proportional to
the difference between work functions (Equation (9)), the value Φ = 4.60 eV was estimated.
By looking at the work functions, salbutamol is expected to become negatively charged,
while the capsule is expected to become positively charged. The behavior of the carrier
particles is less predictable: lactose will acquire a positive charge after contact with the API
and a negative charge if in contact with the capsule.
The cohesion properties are reported in Table 2 for each material pair [52,55].
A B
Sample total mass 25 mg 1 mg
Total no. of particles in sample 123,858 117,233
No. of API particles 92,883 115,825
No. of carrier particles 30,975 1408
API-to-carrier ratio (w/w) 1:3332 1:124
API loading (w/w) 0.03% 0.80%
Total surface area (cm2 ) 9.80 0.52
(a) (b)
Figure 2. Initial location of powder with the two powder loadings: (a) 25 mg and (b) 1 mg.
Configuration A (Figure 2a) has a total mass of 25 mg. It consists of about 31k carrier
particles and 93k API particles. Configuration B (Figure 2b) has a total mass of 1 mg and is
made of about 1500 carrier particles and 117k of API particles. The total number of particles
is similar, but the mixing ratio between the two solid phases differs consistently: 1:3332
(0.03%, w/w) for Configuration A, and 1:124 (0.80%, w/w) for Configuration B.
156
Pharmaceutics 2023, 15, 1762
The approach for the dry powder coating of the carrier with API is described in Alfano
et al. [22]. The difference in the coating degree between Configuration A and B can be
noticed in Figure 3, which shows two coated carrier particles.
(a) (b)
Figure 3. API-coated carrier particles in the two configurations considered: (a) 25 mg, 1:3332 (w/w);
and (b) 1 mg, 1:124 (w/w).
The coated carrier particles are initially placed in a regular cubic arrangement, which
was visualized in Figure 2. The initial random condition is obtained by letting them
settle under gravity. Then, the capsule is subject to a periodic oscillatory translation to
reproduce a tapping motion. The shaking frequency is 120 Hz, and the shaking amplitude
is 0.8 mm. The shaking direction changes during the simulations, so that the capsule
shakes alternatively in both horizontal and vertical directions (x and y, respectively). In
implementation, rather than computing the actual motion of the capsule, the motion is
tracked in the frame of reference of the capsule by adding the equivalent corresponding
fictitious forces on the particles. More details of this implementation in MFIX can be found
in Alfano et al. [39].
157
Pharmaceutics 2023, 15, 1762
no case, however, did the abrupt initial change and subsequent less pronounced increase
turn out to be predictable. Overall, the agreement can be judged to be sufficiently good to
proceed with the more complex carrier–API cases.
(a) (b)
Figure 4. (a) Initial charge distribution and (b) sample charge as a function of the number of taps. The
line is the result of the simulations, and the points are experimental data reported by Chow et al. [19].
(a) (b)
Figure 5. Powder configuration after gravity settling: (a) 25 mg (after 40 ms) and (b) 1 mg (after
200 ms). Carrier particles are shown in gray, and API particles are in purple.
Figure 6 shows the total charge acquired by the samples following the deposition
simulations. The API and carrier curves are shown individually, as well as the total charge
curve. Note that the charges are evaluated at very short time increments, and the symbols
on the plot lines are only for reference. Contrary to the simulation with only lactose
particles (Figure 4b), which became negatively charged, in this case, the carrier acquires a
positive charge, while the salbutamol acquires a negative charge. Figure 6a shows some
oscillations in the total charge of the carrier: first, it increases, then it decreases, and then it
increases again, probably depending on the instantaneous ratio between carrier–wall and
carrier–API interactions. Similar fluctuations can be observed in Figure 6b, which shows
158
Pharmaceutics 2023, 15, 1762
the result with the 1 mg sample and a different mixing ratio, but in a less pronounced
way, suggesting that, in this case, the greater quantity of active principle and the reduced
number of carrier particles promote API–carrier interactions at the expense of carrier–wall
interactions. In both cases, the total net charge is negative and has a greater magnitude for
the larger sample. A plateau in the contact charging process is observed more markedly in
the 25 mg case.
(a) (b)
Figure 6. Total charge acquired by (a) 25 mg and (b) 1 mg of powder after gravity settling in
the capsule. The lines show instantaneous evaluation of the charges, and the symbols are only
for references.
(a) (b)
Figure 7. Charge-to-surface ratio (CTS) acquired during gravity settling by (a) carrier and (b) API for
the two simulations (1 mg and 25 mg). The lines show the instantaneous evaluation of the charges,
and the symbols are only for references.
In Table 4, a summary of the net and specific charges after gravity settling is reported.
The 25 mg test shows a higher net charge (four times higher), but the specific charge is
higher for the 1 mg test, whether it is expressed as the CTS (charge-to-surface ratio) or as
the CTM (charge-to-mass ratio). The numerical value of the CTS is equal to the CTM for
API particles, since the mass/surface ratio is 1 g/m2 .
159
Pharmaceutics 2023, 15, 1762
(a) (b)
Figure 8. Total net charge acquired by (a) 25 mg and (b) 1 mg of powder after capsule shaking. The
lines show instantaneous evaluation of the charges, and the symbols are only for references.
160
Pharmaceutics 2023, 15, 1762
(a) (b)
Figure 9. Specific charge expressed as charge-to-mass ratio (CTM) during capsule shaking for (a) API
and (b) carrier particles. The lines show instantaneous evaluation of the charges, and the symbols are
only for references.
(a) (b)
Figure 10. Specific charge expressed as charge-to-surface ratio (CTS) during capsule shaking for
(a) API and (b) carrier particles. The lines show instantaneous evaluation of the charges, and the
symbols are only for references.
Observing the evolution of carrier particles’ specific charge (Figures 9b and 10b), the
curve for the 1 mg sample is of particular interest. An initial positive charge buildup is
observed, followed by a decrease in the charge and a subsequent markedly linear trend,
until it reaches a specific charge value of almost −130 nC/g (or −3000 nC/m2 ). The specific
charge for the 25 mg sample is more than one order of magnitude lower.
Figure 11 shows the individual charge distribution for API (top) and carrier (bottom)
particles at different times: 50 ms, 200 ms, and 400 ms. The charge distribution of API
particles becomes wider, going from 50 ms to 200 ms, slightly narrowing again at the end
of the simulation (Figure 11c, top). The mean value shifts to the left, i.e., with an increase in
charge with negative polarity.
The evolution of carrier particles’ charge distribution is different for the 1 mg and
25 mg samples. With the 25 mg sample, more and more carrier particles acquire a positive
charge over time, while the opposite tendency is observed for the 1 mg sample: after 50 ms,
most of the particles carry a positive charge (Figure 11a, bottom), while no carrier particle
is charged positively at the end of the simulation (Figure 11c, bottom).
161
Pharmaceutics 2023, 15, 1762
Figure 11. API (top) and carrier (bottom) charge distribution at different times: (a) 50 ms, (b) 200 ms,
and (c) 400 ms.
Figure 12 shows a colored map of the surface charge density (σ) observed in the capsule
after shaking in the x-direction for 60 ms, as estimated according to Equation (9). The charge
is positive, as the wall has the lowest work function. Comparable charge density values are
observed between the two different loadings. However, a wider and more scattered area
is associated to the 1 mg sample case, probably due to the more chaotic movement of API
particles, while the 25 mg case is characterized by higher values localized near the main
impact locations (normal to the shaking direction). The charge density estimated for the
capsule is about 10 times lower than the CTS of carrier particles in the 1 mg configuration
(see Figure 10b), while the opposite trend is observed in the 25 mg simulation. On the other
hand, in both configurations, the CTS calculated for API particles is significantly higher than
the capsule surface charge density (see Figure 10a).
Figure 12. Capsule charge density due to charge exchange with the bottom of the capsule after
shaking in the x-direction for 60 ms.
162
Pharmaceutics 2023, 15, 1762
Figure 13a shows the percentage of API particles in contact with the walls of the
capsule as a function of time. In terms of the fraction of the total, the 1 mg case shows a
higher value, but the time variation is not significant in both cases. Figure 13b shows the
evolution of the carrier-to-API coordination number (CN), i.e., the number of API particles
in contact with a carrier particle averaged over all carrier particles. Interestingly, the CN
remains almost constant in the carrier-rich 25 mg simulation, while in the API-rich 1 mg
simulation, the carrier CN goes from about 100 to 4 within the first 100 ms, meaning that
collisions are strong enough to determine that detachment of most of the API particles
from the carrier. This can be the explanation of the shift in the carrier charge distribution,
in which the mean value goes from a positive value at 0.05 s (Figure 11a, bottom) to a
progressively negative value (Figure 11b,c, bottom), as the carrier–API contact is the event
that can give rise to a positively charged carrier particle.
(a) (b)
Figure 13. Collision statistics: (a) percentage of API particles adhered to the walls and (b) mean
coordination number (CN) for carrier particles.
Figure 14a shows the instantaneous number of collision events during the last part of
the simulation (from 300 to 400 ms); the data for particle–particle (PP) and particle–wall (PW)
contacts are presented separately. In the 25 mg simulation, more than 104 particle–particle
contacts are recorded, while the number of particle–wall collision events does not exceed 1000.
The contact events in the 1 mg simulation stand in between, with more particle–wall contacts
than particle–particle contacts, most probably due to the lower total solid mass. Figure 14b
shows the average normal impact velocity; in this case, the velocity values are lower for
the 25 mg simulations. Particle–wall contacts tend to occur with a higher impact velocity in
both configurations.
The contact statistics data are completed in Figure 15, which shows how the col-
lision events are distributed among the various phases involved. The majority of the
collisions in the 25 mg simulations are low-energy carrier–carrier collisions, during
which the charge exchanged is minimal since particles are made of the same material.
The large number of carrier–carrier collisions could also be the reason why settling
occurs much faster than in the 1 mg simulation, since all of these collisions dissipate the
initial kinetic energy of the particles.
163
Pharmaceutics 2023, 15, 1762
(a) (b)
Figure 14. Collision statistics: (a) instantaneous number of contacts and (b) average impact
velocity for the two simulations. Data for particle–particle (PP) and particle–wall (PW) contacts
are presented separately.
(a)
(b)
Figure 15. (a) Percentage of API–API (A-A), API–carrier (A-C) and carrier–carrier (C-C) collisions
with respect to the total particle–particle contacts for the two simulations. (b) Percentage of API–
wall (A-W) and carrier–wall (C-W) collisions with respect to the total particle–wall contacts for the
two simulations.
By equating the weight force of the particle with the particle–wall electrostatic force
calculated according to Equation (11), it is possible to estimate the charge magnitude
necessary for the particle to remain attached to the wall, assuming for simplicity that the
electrostatic attractive force acts in the opposite direction to the gravitational force:
!
ε 0 ρRg
|qW | = 16πR 2
(12)
3
164
Pharmaceutics 2023, 15, 1762
Figure 16 shows |qW | as a function of particle diameter. The final charge magnitude
distribution is also reported in the form of a scatter plot for the two systems investigated.
Most API particles at the end of the simulation possess a charge level that is higher than
|qW |, while no carrier particle exceeds such a value (all points are below the blue curve).
The results suggest that during capsule handling and shaking, a fraction of API particles
are likely to end up retained inside the capsule, on its internal walls, due to the charge
buildup, while this is much less likely to happen for a carrier particle.
Figure 16. Charge at which P-W electrostatic force is equal to weight as a function of particle diameter
(blue line). A scatter plot of the final charge magnitude is reported in the plot for 1 mg and 25 mg
simulations, separately for API (5 μm diameter) and carrier particles (100 μm diameter).
It is useful to note that other cohesive/adhesive forces and inertial fictitious forces are
at play (see, e.g., Table 2), so their relative contribution should be compared. An evaluation
in terms of the maximum force values is discussed below.
The maximum electrostatic force was calculated with the Coulomb law, considering
the maximum particle charge magnitude in both simulations, which is 7.7 fC for API
particles and 170 fC for carrier particles. For the calculation of the distance of the Coulomb
force, it is assumed that the two bodies are in contact; therefore, for example, in the case of
a p-P-P interaction, the sum of the particle radii was considered.
To estimate the maximum cohesive and adhesive forces, the reference values of the
pull-off force in the JKR model are considered, i.e., the values reported in Table 2.
The fictitious force associated with the shaking motion follows a sinusoidal trend in
time. Its maximum value is given by the following:
Fshk = 4π 2 A f 2 × m p (13)
where A is the shaking amplitude, f is the shaking frequency, and mp is the mass of the
particle. In the present study, this force reaches about 50 times the weight of the particle.
Figure 17 shows a comparison between the estimated maximum force contribution for
carrier and API particles. As expected from the externally imposed motion, the fictitious
force associated with shaking primarily determines the motion of the carrier particles, as it
is higher than any other contributions. Electrostatic forces are lower than cohesive forces,
with the API–carrier electrostatic interaction being the lowest (and lower than weight). On
the other hand, the adhesive API–carrier force contribution is dominant in the case of API
particles, the fictitious force appears to be irrelevant compared to the other forces at play,
and the electrostatic forces are relevant.
It is interesting to note that the electrostatic API–API force is higher than the cohesive
JKR reference force, suggesting that the repulsive Coulomb force between like-charged
particles can overcome the attractive van der Waals interactions. In reality, with sufficient
charge difference, even like-charged particles can be attracted to one another due to a
mutual polarization of surface charge [57,58]. This is not accounted for in simple Coulombic
interactions, as modeled here, and more sophisticated approaches would be necessary.
165
Pharmaceutics 2023, 15, 1762
Figure 17. Comparison between estimated maximum force contributions for API and carrier particles.
AW = API–wall; CW = carrier–wall; AA = API–API; AC = API–carrier; CC = carrier–carrier.
4. Conclusions
An extended DEM approach was applied to evaluate triboelectric charging of a lactose–
salbutamol binary mixture in the gelatine capsule of a DPI. Two powder configurations
were considered, with different loaded doses and mixing ratios. The selected triboelectric
charging model and parameters were validated with the experimental data available in
the literature. Unlike in previous works, dynamic charging and interparticle electrostatic
interactions of carrier and API powders were detailed investigated, including the charge
buildup on the capsule walls. The detailed analysis of the flow and collision behavior
during tapping allowed us to elucidate the mechanisms leading to the final charge polarity
of all three materials, a condition not predictable a priori.
Gravity settling simulations were performed to estimate the charge buildup due to
operations such as capsule filling and storage. The total net charge after settling was found
to be negative, despite the opposite charge polarity acquired by the carrier (positive) and
API (negative). Then, the capsule was subjected to a vibrating motion to simulate routine
handling operations and check whether such movements can give rise to a significant
electrostatic charge. A consistently higher charge buildup was measured with the vibrating
motion compared to the gravity settling simulations, with an overall negative charge also
for carrier particles. The net charge was higher for the carrier-rich 25 mg formulation,
but the specific charge was higher for the API-rich 1 mg formulation. The two different
configurations also show a different charge distribution, with a more pronounced bipolar
charging tendency in the case of the 25 mg dose. To interpret the charging dynamics, a
detailed study of contact statistics was performed. The mechanical shaking movement
promoted API detachment from the carrier, especially in the 1 mg configuration (with higher
API dosage). With the 25 mg configuration, a high number of low-energy carrier-carrier
collisions were recorded which tended to dissipate the kinetic energy of the particles. On
the other hand, API–API and API–wall collisions were prevalent in the 1 mg configuration.
Finally, a comparison of the forces at play revealed that electrostatic forces are relevant
for API particles and might play a major role in cohesion and adhesion phenomena. The
present work lays the foundation for new developments in the relations between particle-
scale charge transfer and buildup and the macroscopic manifestations, particularly in
relation to DPI performances.
166
Pharmaceutics 2023, 15, 1762
Author Contributions: Conceptualization, F.O.A., A.D.R. and F.P.D.M.; methodology, F.O.A., A.D.R.
and F.P.D.M.; software, F.P.D.M.; validation, F.O.A.; formal analysis, F.O.A.; investigation, F.O.A.;
resources, A.D.R. and F.P.D.M.; data curation, F.O.A.; writing—original draft preparation, F.O.A.;
writing—review and editing, A.D.R. and F.P.D.M..; visualization, F.O.A.; supervision, A.D.R. and
F.P.D.M.; project administration, A.D.R. and F.P.D.M.; funding acquisition, A.D.R. and F.P.D.M. All
authors have read and agreed to the published version of the manuscript.
Funding: The present research was conducted with the support of the ICSC National Centre
for High Performance Computing, Big Data and Quantum Computing, funded by European
Union—NextGenerationEU, through Piano Nazionale di Ripresa e Resilienza (PNRR), Missione 4,
Componente 2, Investimento 1.4 (grant number CN00000013 and CUP H23C22000360005), which
is gratefully acknowledged.
Data Availability Statement: The data presented in this study are available from the corresponding
authors upon request.
Conflicts of Interest: The authors declare no conflict of interest.
References
1. Wong, J.; Chan, H.K.; Kwok, P.C.L. Electrostatics in Pharmaceutical Aerosols for Inhalation. Ther. Deliv. 2013, 4, 981–1002.
[CrossRef]
2. Matsusaka, S.; Maruyama, H.; Matsuyama, T.; Ghadiri, M. Triboelectric Charging of Powders: A Review. Chem. Eng. Sci. 2010, 65,
5781–5807. [CrossRef]
3. Šupuk, E.; Zarrebini, A.; Reddy, J.P.; Hughes, H.; Leane, M.M.; Tobyn, M.J.; Timmins, P.; Ghadiri, M. Tribo-Electrification of Active
Pharmaceutical Ingredients and Excipients. Powder Technol. 2012, 217, 427–434. [CrossRef]
4. Šupuk, E.; Hassanpour, A.; Ahmadian, H.; Ghadiri, M.; Matsuyama, T. Tribo-Electrification and Associated Segregation of
Pharmaceutical Bulk Powders. KONA Powder Part. J. 2011, 29, 208–223. [CrossRef]
5. Lacks, D.J.; Shinbrot, T. Long-Standing and Unresolved Issues in Triboelectric Charging. Nat. Rev. Chem. 2019, 3, 465–476.
[CrossRef]
6. Murtomaa, M.; Laine, E. Electrostatic Measurements on Lactose-Glucose Mixtures. J. Electrostat. 2000, 48, 155–162. [CrossRef]
7. Hoe, S.; Young, P.M.; Traini, D. A Review of Electrostatic Measurement Techniques for Aerosol Drug Delivery to the Lung:
Implications in Aerosol Particle Deposition. J. Adhes Sci. Technol. 2011, 25, 385–405. [CrossRef]
8. Telko, M.J.; Kujanpää, J.; Hickey, A.J. Investigation of Triboelectric Charging in Dry Powder Inhalers Using Electrical Low Pressure
Impactor (ELPITM ). Int. J. Pharm. 2007, 336, 352–360. [CrossRef]
9. Naik, S.; Sarkar, S.; Hancock, B.; Rowland, M.; Abramov, Y.; Yu, W.; Chaudhuri, B. An Experimental and Numerical Modeling
Study of Tribocharging in Pharmaceutical Granular Mixtures. Powder Technol. 2016, 297, 211–219. [CrossRef]
10. Hoe, S.; Traini, D.; Chan, H.K.; Young, P.M. The Contribution of Different Formulation Components on the Aerosol Charge in
Carrier-Based Dry Powder Inhaler Systems. Pharm. Res. 2010, 27, 1325–1336. [CrossRef] [PubMed]
11. Wong, J.; Kwok, P.C.L.; Noakes, T.; Fathi, A.; Dehghani, F.; Chan, H.K. Effect of Crystallinity on Electrostatic Charging in Dry
Powder Inhaler Formulations. Pharm. Res. 2014, 31, 1656–1664. [CrossRef] [PubMed]
12. Pinto, J.T.; Wutscher, T.; Stankovic-Brandl, M.; Zellnitz, S.; Biserni, S.; Mercandelli, A.; Kobler, M.; Buttini, F.; Andrade, L.;
Daza, V.; et al. Evaluation of the Physico-Mechanical Properties and Electrostatic Charging Behavior of Different Capsule Types
for Inhalation Under Distinct Environmental Conditions. AAPS PharmSciTech 2020, 21, 128. [CrossRef]
13. Peart, J. Powder Electrostatics: Theory, Techniques and Applications. KONA Powder Part. J. 2001, 19, 34–45. [CrossRef]
14. Mukherjee, R.; Gupta, V.; Naik, S.; Sarkar, S.; Sharma, V.; Peri, P.; Chaudhuri, B. Effects of Particle Size on the Triboelectrification
Phenomenon in Pharmaceutical Excipients: Experiments and Multi-Scale Modeling. Asian J. Pharm. Sci. 2016, 11, 603–617.
[CrossRef]
15. Lacks, D.J.; Sankaran, R.M. Triboelectric Charging in Single-Component Particle Systems. Part. Sci. Technol. 2016, 34, 55–62.
[CrossRef]
16. Chowdhury, F.; Elchamaa, B.; Ray, M.; Sowinski, A.; Passalacqua, A.; Mehrani, P. Apparatus Design for Measuring Electrostatic
Charge Transfer Due to Particle-Particle Collisions. Powder Technol. 2020, 361, 860–866. [CrossRef]
17. Hoe, S.; Young, P.M.; Traini, D. Dynamic electrostatic charge of lactose-salbutamol sulphate powder blends dispersed from a
Cyclohaler® . Drug Dev. Ind. Pharm. 2011, 37, 1365–1375. [CrossRef]
18. Kaialy, W. A Review of Factors Affecting Electrostatic Charging of Pharmaceuticals and Adhesive Mixtures for Inhalation. Int. J.
Pharm. 2016, 503, 262–276. [CrossRef] [PubMed]
19. Chow, K.T.; Zhu, K.; Tan, R.B.H.; Heng, P.W.S. Investigation of Electrostatic Behavior of a Lactose Carrier for Dry Powder Inhalers.
Pharm. Res. 2008, 25, 2822–2834. [CrossRef]
20. Capecelatro, J.; Longest, W.; Boerman, C.; Sulaiman, M.; Sundaresan, S. Recent Developments in the Computational Simulation of
Dry Powder Inhalers. Adv. Drug Deliv. Rev. 2022, 188, 114461. [CrossRef] [PubMed]
167
Pharmaceutics 2023, 15, 1762
21. Ponzini, R.; Da Vià, R.; Bnà, S.; Cottini, C.; Benassi, A. Coupled CFD-DEM Model for Dry Powder Inhalers Simulation: Validation
and Sensitivity Analysis for the Main Model Parameters. Powder Technol. 2021, 385, 199–226. [CrossRef]
22. Alfano, F.O.; Benassi, A.; Gaspari, R.; Di Renzo, A.; Di Maio, F.P. Full-Scale DEM Simulation of Coupled Fluid and Dry-Coated
Particle Flow in Swirl-Based Dry Powder Inhalers. Ind. Eng. Chem. Res. 2021, 60, 15310–15326. [CrossRef]
23. Benque, B.; Khinast, J.G. Carrier Particle Emission and Dispersion in Transient CFD-DEM Simulations of a Capsule-Based DPI.
Eur. J. Pharm. Sci. 2022, 168, 106073. [CrossRef] [PubMed]
24. Sulaiman, M.; Liu, X.; Sundaresan, S. Effects of Dose Loading Conditions and Device Geometry on the Transport and Aerosoliza-
tion in Dry Powder Inhalers: A Simulation Study. Int. J. Pharm. 2021, 610, 121219. [CrossRef] [PubMed]
25. Alfano, F.O.; Di Maio, F.P.; Di Renzo, A. Deagglomeration of Selected High-Load API-Carrier Particles in Swirl-Based Dry Powder
Inhalers. Powder Technol. 2022, 408, 117800. [CrossRef]
26. Korevaar, M.W.; Padding, J.T.; Van der Hoef, M.A.; Kuipers, J.A.M. Integrated DEM-CFD Modeling of the Contact Charging of
Pneumatically Conveyed Powders. Powder Technol. 2014, 258, 144–156. [CrossRef]
27. Pei, C.; Wu, C.Y.; England, D.; Byard, S.; Berchtold, H.; Adams, M. Numerical Analysis of Contact Electrification Using DEM-CFD.
Powder Technol. 2013, 248, 34–43. [CrossRef]
28. Kolehmainen, J.; Ozel, A.; Boyce, C.M.; Sundaresan, S. Triboelectric Charging of Monodisperse Particles in Fluidized Beds. AIChE
J. 2017, 63, 1872–1891. [CrossRef]
29. Hu, J.; Pei, C.; Zhang, L.; Liang, C.; Wu, C.Y. Numerical Analysis of Frictional Charging and Electrostatic Interaction of Particles.
AIChE J. 2022, 68, e17444. [CrossRef]
30. Zafar, U.; Alfano, F.; Ghadiri, M. Evaluation of a New Dispersion Technique for Assessing Triboelectric Charging of Powders. Int.
J. Pharm. 2018, 543, 151–159. [CrossRef]
31. Alfano, F.O.; Di Renzo, A.; Di Maio, F.P.; Ghadiri, M. Computational Analysis of Triboelectrification Due to Aerodynamic Powder
Dispersion. Powder Technol. 2021, 382, 491–504. [CrossRef]
32. Mukherjee, R.; Sansare, S.; Nagarajan, V.; Chaudhuri, B. Discrete Element Modeling (DEM) Based Investigation of Tribocharging
in the Pharmaceutical Powders during Hopper Discharge. Int. J. Pharm. 2021, 596, 120284. [CrossRef] [PubMed]
33. Chowdhury, F.; Ray, M.; Passalacqua, A.; Mehrani, P.; Sowinski, A. Evaluating the Electrostatic Charge Transfer Model for
Particle-Particle Interactions. J. Electrostat. 2021, 112, 103603. [CrossRef]
34. Naik, S.; Hancock, B.; Abramov, Y.; Yu, W.; Rowland, M.; Huang, Z.; Chaudhuri, B. Quantification of Tribocharging of Phar-
maceutical Powders in V-Blenders: Experiments, Multiscale Modeling, and Simulations. J. Pharm. Sci. 2016, 105, 1467–1477.
[CrossRef]
35. Zhu, Q.; Gou, D.; Li, L.; Chan, H.K.; Yang, R. Numerical Investigation of Powder Dispersion Mechanisms in Turbuhaler and the
Contact Electrification Effect. Adv. Powder Technol. 2022, 33, 103839. [CrossRef]
36. Almeida, L.C.; Bharadwaj, R.; Eliahu, A.; Wassgren, C.R.; Nagapudi, K.; Muliadi, A.R. Capsule-Based Dry Powder Inhaler
Evaluation Using CFD-DEM Simulations and next Generation Impactor Data. Eur. J. Pharm. Sci. 2022, 175, 106226. [CrossRef]
37. Mitani, R.; Ohsaki, S.; Nakamura, H.; Watano, S. Numerical Study on Particle Adhesion in Dry Powder Inhaler Device. Chem.
Pharm. Bull 2020, 68, 726–736. [CrossRef]
38. Benque, B.; Khinast, J.G. Understanding the Motion of Hard-Shell Capsules in Dry Powder Inhalers. Int. J. Pharm. 2019,
567, 118481. [CrossRef]
39. Alfano, F.O.; Sommerfeld, M.; Di Maio, F.P.; Di Renzo, A. DEM Analysis of Powder Deaggregation and Discharge from the
Capsule of a Carrier-Based Dry Powder Inhaler. Adv. Powder Technol. 2022, 33, 103853. [CrossRef]
40. Johnson, K.L.; Kendall, K.; Roberts, A.D. Surface Energy and the Contact of Elastic Solids. Proc. R. Soc. Lond. A 1971, 324, 301–313.
[CrossRef]
41. Di Renzo, A.; Napolitano, E.S.; Di Maio, F.P. Coarse-Grain Dem Modelling in Fluidized Bed Simulation: A Review. Processes 2021,
9, 279. [CrossRef]
42. Di Renzo, A.; Di Maio, F.P. Comparison of Contact-Force Models for the Simulation of Collisions in DEM-Based Granular Flow
Codes. Chem. Eng. Sci. 2004, 59, 525–541. [CrossRef]
43. Mindlin, R.D.; Deresiewicz, H. Elastic Spheres in Contact under Varying Oblique Forces. J. Appl. Mech. 1953, 20, 327–344.
[CrossRef]
44. Ai, J.; Chen, J.-F.; Rotter, J.M.; Ooi, J.Y. Assessment of Rolling Resistance Models in Discrete Element Simulations. Powder Technol.
2011, 206, 269–282. [CrossRef]
45. Alfano, F.O.; Di, R.; Gaspari, A.; Benassi, R.; Di Maio, A.; Alfano, F.O.; Di Renzo, A.; Gaspari, R.; Benassi, A.; Paolo, F.; et al.
Modelling Deaggregation Due to Normal Carrier-Wall Collision in Dry Powder Inhalers. Processes 2022, 10, 1661. [CrossRef]
46. Garg, R.; Galvin, J.; Li, T.; Pannala, S. Documentation of Open-Source MFIX–DEM Software for Gas-Solids Flows. 2012. Available
online: https://mfix.netl.doe.gov/documentation/dem_doc_2012-1.pdf (accessed on 1 February 2023).
47. Santasusana, M.; Irazábal, J.; Oñate, E.; Carbonell, J.M. The Double Hierarchy Method. A Parallel 3D Contact Method for the
Interaction of Spherical Particles with Rigid FE Boundaries Using the DEM. Comput. Part Mech. 2016, 3, 407–428. [CrossRef]
48. Matsusaka, S.; Ghadiri, M.; Masuda, H. Electrification of an Elastic Sphere by Repeated Impacts on a Metal Plate. J. Phys. D Appl.
Phys. 2000, 33, 2311–2319. [CrossRef]
49. Matsuyama, T.; Yamamoto, H. Charge Relaxation Process Dominates Contact Charging of a Particle in Atmospheric Conditions.
J. Phys. D Appl. Phys. 1995, 28, 2418–2423. [CrossRef]
168
Pharmaceutics 2023, 15, 1762
50. Feynman, R.P.; Leighton, R.B.; Sands, M. The Feynman Lectures in Physics, Mainly Electromagnetism and Matter; The New Millennium,
Ed.; Basic Books: New York, NY, USA, 2011; Volume 2, ISBN 978-0-465-07998-8.
51. van Wachem, B.; Thalberg, K.; Remmelgas, J.; Niklasson-Björn, I. Simulation of Dry Powder Inhalers: Combining Micro-Scale,
Meso-Scale and Macro-Scale Modeling. AIChE J. 2017, 63, 501–516. [CrossRef]
52. Nguyen, D.; Remmelgas, J.; Björn, I.N.; van Wachem, B.; Thalberg, K. Towards Quantitative Prediction of the Performance of Dry
Powder Inhalers by Multi-Scale Simulations and Experiments. Int. J. Pharm. 2018, 547, 31–43. [CrossRef]
53. Zellnitz, S.; Pinto, J.T.; Brunsteiner, M.; Schroettner, H.; Khinast, J.; Paudel, A. Tribo-Charging Behaviour of Inhalable Mannitol
Blends with Salbutamol Sulphate. Pharm. Res. 2019, 36, 80. [CrossRef]
54. Trigwell, S.; Grable, N.; Yurteri, C.U.; Sharma, R.; Mazumder, M.K. Effects of Surface Properties on the Tribocharging Characteris-
tics of Polymer Powder as Applied to Industrial Processes. IEEE Trans. Ind. Appl. 2003, 39, 79–86. [CrossRef]
55. Ibrahim, T.H.; Burk, T.R.; Etzler, F.M.; Neuman, R.D. Direct Adhesion Measurements of Pharmaceutical Particles to Gelatin
Capsule Surfaces. J. Adhes Sci. Technol. 2000, 14, 1225–1242. [CrossRef]
56. Coates, M.S.; Fletcher, D.F.; Chan, H.-K.; Raper, J.A. The Role of Capsule on the Performance of a Dry Powder Inhaler Using
Computational and Experimental Analyses. Pharm. Res. 2005, 22, 923–932. [CrossRef] [PubMed]
57. Lindgren, E.B.; Chan, H.K.; Stace, A.J.; Besley, E. Progress in the Theory of Electrostatic Interactions between Charged Particles.
Phys. Chem. Chem. Phys. 2016, 18, 5883–5895. [CrossRef] [PubMed]
58. Qin, J.; Li, J.; Lee, V.; Jaeger, H.; de Pablo, J.J.; Freed, K.F. A Theory of Interactions between Polarizable Dielectric Spheres. J. Colloid
Interface Sci. 2016, 469, 237–241. [CrossRef] [PubMed]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual
author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to
people or property resulting from any ideas, methods, instructions or products referred to in the content.
169
pharmaceutics
Article
Impact of Vertical Blender Unit Parameters on Subsequent
Process Parameters and Tablet Properties in a Continuous
Direct Compression Line
Marius J. Kreiser 1,2 , Christoph Wabel 1 and Karl G. Wagner 2, *
1 Product and Process Development, Pfizer Manufacturing Deutschland GmbH, 79108 Freiburg, Germany;
marius.kreiser@pfizer.com (M.J.K.); christoph.wabel@pfizer.com (C.W.)
2 Department of Pharmaceutical Technology and Biopharmaceutics, University of Bonn, 53121 Bonn, Germany
* Correspondence: [email protected]
Abstract: The continuous manufacturing of solid oral-dosage forms represents an emerging technol-
ogy among the pharmaceutical industry, where several process steps are combined in one production
line. As all mixture components, including the lubricant (magnesium stearate), are passing simultane-
ously through one blender, an impact on the subsequent process steps and critical product properties,
such as content uniformity and tablet tensile strength, is to be expected. A design of experiment (DoE)
was performed to investigate the impact of the blender variables hold-up mass (HUM), impeller
speed (IMP) and throughput (THR) on the mixing step and the subsequent continuous manufacturing
process steps. Significant impacts on the mixing parameters (exit valve opening width (EV), exit valve
opening width standard deviation (EV SD), torque of lower impeller (TL ), torque of lower impeller
SD (TL SD), HUM SD and blend potency SD), material attributes of the blend (conditioned bulk
density (CBD), flow rate index (FRI) and particle size (d10 values)), tableting parameters (fill depth
Citation: Kreiser, M.J.; Wabel, C.; (FD), bottom main compression height (BCH) and ejection force (EF)) and tablet properties (tablet
Wagner, K.G. Impact of Vertical thickness (TT), tablet weight (TW) and tensile strength (TS)) could be found. Furthermore, relations
Blender Unit Parameters on between these process parameters were evaluated to define which process states were caused by
Subsequent Process Parameters and which input variables. For example, the mixing parameters were mainly impacted by impeller speed,
Tablet Properties in a Continuous and material attributes, FD and TS were mainly influenced by variations in total blade passes (TBP).
Direct Compression Line. The current work presents a rational methodology to minimize process variability based on the main
Pharmaceutics 2022, 14, 278.
blender variables hold-up mass, impeller speed and throughput. Moreover, the results facilitated a
https://doi.org/10.3390/
knowledge-based optimization of the process parameters for optimum product properties.
pharmaceutics14020278
Academic Editor: Colin Hare Keywords: continuous manufacturing; continuous mixing technology; vertical blender; direct compression;
lubrication; material characterization
Received: 9 December 2021
Accepted: 20 January 2022
Published: 25 January 2022
Figure 1. Overview of the direct compression line used for this trial.
Basically, for each raw material, the powder is transferred from a polyethylene bag via
a top-up valve into an agitated hopper, where co-rotating screws supply the powder by the
loss in weight (LiW) principle at a composition related feed rate. The continuous process
demands a periodical refill of the hopper triggered by a defined refill level, performed by a
rotating volumetric refill device with flexible volume inserts [2,3].
Since feeding is the first step within a continuous process, it is consequently one of the
first critical control elements besides the material attributes. Accurate feeding is substantial
for the quality of a continuous process to avoid deviations regarding the quality of blend
and content uniformity of the tablets [4–8].
To provide low variability in feed rate, the optimal feeder design and the corresponding
parameter settings, such as refill level, top-up volume, screw pitch, feed-factor array (gov-
erning dosing in volumetric mode during e.g., refill) and gearbox type (Figure 2), should be
individually adjusted based on composition, throughput and powder attributes [3,4,7–13].
Several feeders supply each raw material separately, and the powder falls through
the conical-shaped inlet hopper into the vertical continuous mixer. It is composed of two
regions: the upper delumping region and the lower mixing region (Figure 3). In both, the
impellers can be adjusted independently regarding speed, direction and vertical position,
i.e., the gap between impeller and conical sieve. In the delumping region, a downstream
sieve (d = 2.1 mm) is set to delump possible agglomerates. The powder leaves the upper
region and arrives in the conical mixing region, where a second impeller is mounted. The
whole setup of the CMT is attached to load cells, which monitor the weight of the powder
within the mixer. This hold-up mass (HUM) is defined in the recipe and determines the
mass, will always be mixed in the CMT continuously throughout the process.
171
Pharmaceutics 2022, 14, 278
Figure 3. Overview of the CMT. View of the upper impeller is obstructed by the sieve.
172
Pharmaceutics 2022, 14, 278
Other papers focus on a horizontal continuous mixer, where the HUM is considered a
function of flow rate and impeller speed and cannot be set individually [14,15]. In contrast,
the HUM in a vertical continuous mixer remains constant, and various shear rates (impeller
speeds) can be applied despite a constant residence-time distribution (RTD) [8]. As in this
case the continuous DC line includes only one mixing step for all mixture components,
including the lubricant. An impact especially on lubricant-sensitive mixtures, as well as on
the blend uniformity of the mixture and, subsequently, content uniformity of the tablets,
can be expected [16–18].
The exit valve is located at the bottom of the CMT. By means of a proportional–integral–
derivative (PID) control loop, the exit valve opening width is adjusted automatically based
on the current HUM value in order to keep the mass of the CMT constant. The controlled
exit valve ensures that the same amount of mass entering the CMT will simultaneously
leave the CMT (massin = massout ). Feed fluctuations of each feeder and the respective
variability in the mass flow can be balanced that way. Smaller exit valve opening widths
are recommended so that newly entering raw materials can be properly mixed together
with the blend that is already present in the blender. Otherwise, unmixed or poorly mixed
material can pass by and leave the CMT without being blended, causing content-uniformity
variability [8].
The mean residence time (MRT, Equation (1)) of a particle can be calculated based on
the overall throughput and the HUM. It reflects the mixing period of that particle within
the CMT [8].
HUM [kg] min
MRT [min] = ∗ 60 (1)
THR
kg h
h
The total blade passes (TBP, Equation (2)) reveals how often the impeller, on average,
will pass a particle and show the intensity of the shear transmitted to the powder. With
an increasing number of revolutions and respectively increased shear, a lubricant, such
as magnesium stearate (MgSt), can be introduced more homogeneously into the blend
or even filmed onto the particles, potentially resulting in decreasing tensile strength of
tablets. Therefore, particular attention is paid to the single mixing step in the CMT, where
the lubricant will be mixed right from the start, together with the remaining raw ma-
terials, potentially resulting in a narrow process window between a homogeneous and
an over-lubricated blend. Hence, it is required to set a suitable combination for HUM
and IMP to ensure that TS and disintegration, as well as dissolution time, are within
specification [19–25]. Thresholds regarding HUM and IMP are, besides the mass balance
model (MBM), part of the control strategy of the CMT. If the process values exceed the
specific limits, an alarm occurs and the process stops. Furthermore, variations in HUM and
IMP could also impact the exit valve opening width and, therefore, the mixing quality.
After the powder exits the CMT, it travels through the feed chute into the feed frame,
where powder will be held up and be fed into the tablet press. Position sensors in the
feed chute measure the filling levels. Using an internal feedback loop, we can control the
turret speed of the tablet press according to the filling levels, thus preventing powder from
backing up or the tablet press from running empty. An increasing feed-chute level results in
an increased turret speed of the tablet press, i.e., increased powder demand, and vice versa.
The NIR probe in the feed frame is the first chemometric measurement in the process.
It is therefore important to understand the impact of upstream settings and process states
on the conformity of potency, as predicted by the NIR model.
The NIR probe measures a defined volume of the powder. The corresponding spectra
are used to predict the API (active pharmaceutical ingredient) content. If inhomogeneity
of the blend or variability in the upstream process units occurs, it can consequently be
detected with NIR and is seen as a disturbance in the blend potency measured by NIR
inside the feed frame [26].
173
Pharmaceutics 2022, 14, 278
Depending on the chosen control strategy, the impacted tablets can be diverted into
the waste channel if the signals exceed the specification limits. As soon the signals are
within specification limits again, the diverter switches back to the good product channel
after a defined lead-lag time [27–29].
The tablet press was running with a control mode enabled, wherein the tablet weight
control is based on the pre-compression displacement and the fill depth is adjusted accord-
ingly. The bottom main compression height controls the thickness and compression force
and, therefore, the crushing strength of the tablets. At the end of the tablet press, the tablets
can be directed into the good channel, diverted into the waste channel or directed to the
combi-tester, where at-line measurements regarding tablet properties can be performed in
containment (see Section 2.7).
This paper assesses to what extent HUM, IMP and THR impact the downstream
process of a direct compression mixture. It focuses on correlations and coherences and
evaluates the predictability of process parameters based on the CMT settings, especially
since the lubricant and all other formulation constituents are mixed simultaneously in
one single mixing step. As a model formulation, Saccharin Monohydrate was used as
API surrogate.
Table 1. Overview of considered responses, where HUM, IMP and THR were adjusted as
input variables.
Responses
Mixing parameters TL,
EV
Blend potency as predicted by the NIR model
Material attributes of the blend FRI
Particle size (d10 )
CBD
Tableting parameters FD
BCH
EF
TS
Tablet properties TT
TW
174
Pharmaceutics 2022, 14, 278
Table 2. DoE settings, where phase 7, 9 and 11 are the replicates of the center point.
During the transition phase and the compression-force profile, the tablet press was
operated in manual mode, without using the combi-tester, to analyze tablet properties. In
manual mode, samples were taken and weighed manually to select the correct fill depth.
During each steady-state phase, manual mode was switched to automatic mode, in which
the NIR probe was active. For each steady state phase, 275 MPa compression pressure was
set; a tablet sample was taken in the middle of the steady state phase, using the combi-tester;
and a powder sample was withdrawn at the end of each steady state phase by opening the
sampling port underneath the feed frame and collecting approximately 300 g of powder.
175
Pharmaceutics 2022, 14, 278
ρtapped − ρbulk
Carr Index = ∗ 100 (4)
ρtapped
Basically, at higher flow rates, less energy is required, since the entrained air acts as
a lubricant. At lower flow rates, the powder in front of the blades is more likely to be
consolidated, due to the absence of entrained air; therefore, the interlocking of particles
is more probable. Consequently, higher FRI values are common for cohesive powders. In
this study, FRI values < 1 are shown; they are typical for powders or blends containing
lubricants. The conditioned bulk density (CBD) was measured after the initial conditioning
cycle and the split of the powder, where agglomerates and air inclusions could be evened
to ensure reproducible measurements [11,30,32–34].
176
Pharmaceutics 2022, 14, 278
2.7. Tableting
A MODUL™ P tablet press (GEA Pharma Systems, Courtoy™, Halle, Belgium) was im-
plemented at the end of the continuous manufacturing line. Mode 2 (Courtoy dual-control
force method) was selected, where the tablet weight control is based on pre-compression
displacement measurements adjusting the fill depth, accordingly [40].
Round convex tablets with an 11 mm diameter and 1.12 mm cup height were manu-
factured. During steady state, a target compression pressure of 275 MPa was set.
The target tablet weight of 600 mg, tablet crushing strength and tablet thickness
were tested periodically in the middle of each steady state, using the at-line combi-tester
(Kraemer Elektronik GmbH, Darmstadt, Germany).
The feed chute level was controlled to a constant level at 40%, and the paddle speed
remained constant at 45 and 40 rpm. Turret speed set-points and speed tolerances of
the tablet press were adapted to the respective mass throughput (11 rpm ± 2.2 rpm;
21 rpm ± 4.2 rpm and 32 rpm ± 6.4 rpm).
Tensile Strength
The tensile strength of the convex round tablets was calculated based on the following
equation [41]:
−1
10Ps t t W
Tensile Strength = 2.84 − 0.126 + 3.15 + 0.01 (6)
πD2 D W D
where Ps = tablet core crushing strength, D = tablet core diameter, t = tablet core thickness
and W = cylinder length. Tablet-crushing strength was measured by using the combi-tester,
which is directly connected to the continuous manufacturing line.
177
Pharmaceutics 2022, 14, 278
2.9. Software
2.9.1. MODDE
The DoE was designed by using MODDE Pro 12.1. A multiple linear regression
(MLR) model was used to evaluate the significance of the input factors on the responses.
Furthermore, MODDE was used to obtain model equations to predict the responses.
2.9.2. Osi Pi
A considerable benefit of the PCMM is the implementation of OsiPi (OsiSoft, San
Leandro, CA, USA), which enables access to all essential process values. All data generated
by the PCMM are continuously monitored and stored by using OsiPi.
Pi Vision is a web-based tool wherein process data can be visualized in real time. Since
the process data are stored in the PI Server, PiVision also can visualize previous batches if
process states need to be evaluated retrospectively. For this trial, all process-related data
were gathered by using PiDataLink, which is an Add-In to Excel (Microsoft Corporation,
Redmond, Washington, USA) that enables data to be imported from the PI Server.
178
Pharmaceutics 2022, 14, 278
the same extent of deflection. With regard to HUM SD values, all three input factors and
HUM*IMP were significant.
Figure 4. Process overview of input factors (green, left side) and observed responses (blue, right side).
(96'a>PP@
(9a>PP@
7/a>1P@
7+5
7+5
7+5
,03
,03
,03
+80
+80
+80
7+5 7+5
7+5 7+5
7+5 7+5
7+5 ,03
7+5 ,03
7+5 ,03
,03 ,03
,03 ,03
,03 ,03
7+5 +80
7+5 +80
7+5 +80
+80 ,03
+80 ,03
+80 ,03
+80 +80
+80 +80
+80 +80
%OHQG3RWHQF\6'a>@
+806'a>NJ@
7/6'a>1P@
7+5
7+5
7+5
,03
,03
,03
+80
+80
+80
7+5 7+5
7+5 7+5
7+5 7+5
,03 ,03
,03 ,03
,03 ,03
7+5 ,03
7+5 ,03
7+5 ,03
7+5 +80
7+5 +80
7+5 +80
+80 ,03
+80 ,03
+80 ,03
+80 +80
+80 +80
+80 +80
Figure 5. Coefficients plot of the impact of input variables on responses regarding the blending unit
and uniformity of the blend. The 95% confidence interval is displayed as an error bar.
As shown in Table 6, Q2 and R2 imply, that exit valve opening width (+SD) and
torque (+SD) can be considered good models. As the variabilities of the responses were
179
Pharmaceutics 2022, 14, 278
not linearly distributed a logarithmic data transformation was conducted (as shown in the
corresponding chapters and Supplementary Figures S7 and S11).
Further details regarding fit statistics and model equations are shown in Supplemen-
tary Section A, “Summary of Fit: Mixing Parameter”.
Figure 6. Coefficients plot of model terms regarding material attributes of the blends. The 95%
confidence interval is displayed as an error bar.
Table 7 shows the fit statistics after removing non-significant model terms, whereby
models regarding CBD, FRI and d10 can be considered good models. For further details,
see Supplementary Section B, “Summary of Fit: Material Attributes of the Blend”.
180
Pharmaceutics 2022, 14, 278
lower impeller speed result in lower TBP and, therefore, in lower lubrication, leading to
lower powder densities, higher required fill depths and higher ejection forces.
%&+>PP@
)'>PP@
()>N1@
,03
,03
,03
7+5
7+5
7+5
+80
+80
+80
,03 ,03
,03 ,03
,03 ,03
7+5 ,03
7+5 ,03
7+5 ,03
7+5 7+5
7+5 7+5
7+5 7+5
+80 ,03
+80 ,03
+80 ,03
7+5 +80
7+5 +80
7+5 +80
+80 +80
+80 +80
+80 +80
Figure 7. Coefficients plot of model terms regarding tablet press parameters. The 95% confidence
interval is displayed as an error bar.
Furthermore, the fill depth and ejection force share the same deflection of the three
significant model terms, namely THR, IMP and THR*THR.
Table 8 shows the fit statistics after removing non-significant model terms. All three
parameters show high values regarding Q2 and R2 . For further information regarding
fit statistics and model equations, see Supplementary Section C, “Summary of Fit: Tablet
Press Parameters”.
181
Pharmaceutics 2022, 14, 278
7603Da>03D@
7:03Da>PJ@
7703Da>PP@
7+5
7+5
7+5
,03
,03
,03
+80
+80
+80
7+5 7+5
7+5 7+5
7+5 7+5
,03 ,03
,03 ,03
7+5 ,03
7+5 ,03
7+5 ,03
7+5 +80
7+5 +80
7+5 +80
+80 ,03
+80 ,03
+80 ,03
+80 +80
+80 +80
7603D6'>03D@
7:03D6'>PJ@
7703D6'>PP@
7+5
7+5
7+5
,03
,03
,03
+80
+80
+80
7+5 7+5
,03 ,03
7+5 7+5
,03 ,03
7+5 7+5
,03 ,03
7+5 ,03
7+5 ,03
7+5 ,03
7+5 +80
7+5 +80
7+5 +80
+80 ,03
+80 ,03
+80 ,03
+80 +80
+80 +80
+80 +80
Figure 8. Model terms regarding tensile strength (TS), tablet weight (TW), tablet thickness (TT) and
corresponding standard deviation. The 95% confidence interval is displayed as an error bar.
According to Table 9, tensile strength, tablet weight and tablet thickness can be con-
sidered good models. Again, as the variabilities of the responses were not linearly dis-
tributed, a logarithmic data transformation was conducted. Corresponding figures, model
equations and fit statistics are shown in Supplementary Section D, “Summary of Fit:
Tablet Properties”.
182
Pharmaceutics 2022, 14, 278
reason for an increasing EV, while variations in HUM seemingly did not impact the exit
valve (0.042 p = 0.874). Furthermore, a contour plot is used to demonstrate the significance
of both model terms throughput and impeller speed (Figure 10). To determine suitable
CMT settings based on this plot, small exit valve opening widths (<5 mm) are preferable,
which is in line with the findings of Toson [8]. Additionally, data regarding blend potency
SD confirmed the maximum of 5 mm opening value of the exit valve (further details below).
Figure 9. Qualitative overview of process parameter connections and correlations. Input factors
are marked in dark green (thick borders), confounding input parameters are marked in light green
and the considered response parameters are shown in light orange. The color/shape of the borders
classifies the responses into mixing parameters (orange line, rounded corners), material attributes
of the blend (purple, striped background), tableting parameters (blue, dotted borders) and tablet
properties (red, thin borders). Compression pressure (green) is considered an independent input
factor of the tablet press.
Figure 10. Contour plot of exit valve opening width in dependence of THR, HUM and IMP.
183
Pharmaceutics 2022, 14, 278
Regarding EV SD, Table 6 reveals that impeller speed was the only significant model
term (Q2 = 0.822 R2 = 0.933). Furthermore, Figure 11c shows the EV standard deviation as a
function of the EV opening width (0.785 p = 0.0002). This correlation leads to the conclusion
that higher EV values increased the risk of a fluctuating opening width, impacting the
variability of the blend potency values (0.952 p < 0.0001) and subsequently affecting content
uniformity of the tablets. A correlation matrix with downstream parameters concerning
the EV is shown in Figure 12.
(a) (b)
(d)
(c)
Figure 11. (a) Exit valve opening width vs. throughput (kg/h) in relation to varying impeller speeds.
HUM [g]
(b) EV in dependence of impeller speed. (c) EV SD vs. EV. (d) EV as function of kg
IMP [rpm ]∗THR
2 2
h
HUM SD
HUM is an essential variable in MRT and TBP (Equations (1) and (2)), and this is
why it is crucial to choose suitable blender parameters to maintain a consistent process.
Accordingly, the fluctuation in HUM led to variabilities in the MRT and TBP.
184
Pharmaceutics 2022, 14, 278
Figure 13a shows the HUM SD as a function of impeller speed (0.514 p = 0.035). It
reveals that the HUM standard deviations were not directly impacted by throughput.
However, throughput is a significant model term, since comparatively low HUM standard
deviations were obtained at low throughputs. On the other hand, higher impeller speeds
tended to result in a larger span of HUM SD, and this could be caused by an unfavorable
powder bed shape, due to higher centrifugal forces, as described by Toson et al. [8].
(a) (b)
Figure 13. (a) HUM standard deviation as a function of impeller speed. (b) Dependencies between
SD in EV and HUM.
Figure 13b shows that the previously mentioned EV SD correlated with HUM SD
(0.929 p < 0.001). That could be traced back to the PID control loop between HUM and
EV, where EV is a function of HUM process values in order to maintain massin = massout .
Therefore, if variability could be observed in the HUM, then it occurred in EV, as well.
To avoid those fluctuations, we can rely on the previous section, where impeller speed is
185
Pharmaceutics 2022, 14, 278
(a) (b)
(c)
Figure 14. (a) Torque of the lower impeller as a function of the sum of HUM and IMP. (b) Correlation
between variability in torque and exit valve opening width. (c) Impact of impeller speed on the
torque values.
Figure 14b demonstrates the linearity between TL SD and EV SD (0.906 p < 0.0001).
The correlation between these standard deviations is based on the impact of impeller speed
(IMP—TL SD: 0.874 p < 0.0001), wherein the higher impeller rotation resulted in higher
variabilities in both parameters (Figures 5 and 14c).
Since standard deviations in both the torque and exit valve were strongly correlated, it
is recommended to only focus on the EV values if monitoring is required.
186
Pharmaceutics 2022, 14, 278
Blend Potency SD
Reflecting previously described process parameters, the correlations between blend
potency SD and EV (0.843 p < 0.0001), EV SD (0.952 p < 0.0001), HUM SD (0.817 p < 0.0001),
TL SD (0.965 p < 0.0001) and IMP (0.753 p = 0.0005) could be observed (Figure 12).
Higher exit valve opening widths implicate that the powder bed was not entirely
closed at the bottom of the CMT and that particles newly entering the CMT could exit
unmixed [8]. Consequently, blend potency SDs and, therefore, blend inhomogeneities
could be explained by insufficient mixing based on the structure of the powder bed within
the blend. Figure 15a shows the blend potency standard deviation as a function of impeller
speed, wherein all values at 650 rpm were above 2.5%. This observation could also be
confirmed by using Figure 5, wherein IMP was the significant model term. Thus, to reduce
blend potency SDs and, therefore, improve blend homogeneity and content uniformity of
the tablets, reduction of the impeller speed is again proposed.
(a) (b)
Figure 15. Blend Potency SD as a function of (a) impeller speed and (b) exit valve opening width.
Figure 15b shows that all blend potency values obtained at exit valve opening widths
below 10 mm were smaller than 2.5%. To minimize the risk of a higher blend potency SD,
the presented results confirm maximum EV values below 5 mm.
Furthermore, independent of the blender variables, a potential risk for blend potency
inhomogeneity could be adhesion of API at the walls due to electrostatic charging of
particles [44].
Powder Density
With more impeller revolutions, more cavities of particles and granules can be filled
and a layer around the particles can be built. On one hand, that increases the weight without
increasing the volume and on the other hand, it is reducing particle-particle frictions due
to the reduced friction of magnesium stearate filmed particles. Particles can now arrange
more compactly, increasing the powder density [45,46].
187
Pharmaceutics 2022, 14, 278
Figure 16a demonstrates an exponential relationship between TBP and CBD asymptot-
g
ically reaching a value of 0.598 mL at 1560 TBP. At extreme values, such as 3120 revolutions,
powder density will not increase any further and a maximum seemed to be reached, which
led to the conclusion that increasing TBP only affected the material up to a certain limit.
(a) (b)
(c)
Figure 16. (a) Conditioned bulk density (CBD) (g/mL), (b) flow rate index (FRI) and (c) particle size
(d10 ) (μm) as a function of total blade passes.
Considering Figure 6 and the equation for TBP (Equation (2)), the significant model
terms of the DoE revealed the same information, where higher values in HUM and impeller
speed increased CBD, and higher throughputs decreased the powder density (Q2 = 0.735
and R2 = 0.850).
Particle Size
The description of density changes based on TBP also applies to the particle size (d10 )
(Figure 16c). At a high TBP, more magnesium stearate adhered to the particles, leading to a
lower amount of the remaining free MgSt particles within the blend, and thus increasing
the d10 values (0.836 p < 0.0001). As a reference, a blend without magnesium stearate was
mixed by using a Turbula blender (Willy A. Bachofen AG, Muttenz, Switzerland), where a
d10 value of 38.38 μm was obtained (Figure 16c).
188
Pharmaceutics 2022, 14, 278
Therefore, the appearance of smaller particle sizes in the blend could be traced back to
MgSt. As seen at 3120 revolutions, the d10 value was similar to the blend without MgSt,
implicating that the fine fraction of MgSt was almost completely attached to the remaining
raw materials at the higher TBP. Moreover, particle-size changes due to destruction of
particles could be ruled out. In this case, the d10 values would have decreased with a
higher shear.
Regarding the DoE results in Figure 6, a good model for d10 values could be obtained
(R2 = 0.842 and Q2 = 0.587). Particle sizes of raw materials and blends are shown in
Supplementary Table S34.
Ejection Force
The ejection force is the required force to eject the tablet from the die and depends on
the friction between the tablet and the die walls. Consequently, the reduction in ejection
force is mainly influenced by the lubrication of the powder [50]. Usually, high ejection forces
are accompanied by tableting problems and may cause damages to the tooling [51,52].
Regarding this dataset, the model-terms throughput and impeller speed shared the
same deflections as for the fill depth (Figure 7); that means, it is indicated that a higher
TBP results in a higher lubrication and lower ejection forces. However, although a strong
correlation between ejection force and TBP was expected, only a correlation between ejection
force and tablet-weight variability could be found (0.787 p = 0.0002). Nevertheless, a robust
model regarding ejection force could be obtained by an MLR analysis (Q2 = 0.892 and
R2 = 0.944). For further explanation regarding TBP and ejection force, see Supplementary
Section G, “Ejection Force”.
189
Pharmaceutics 2022, 14, 278
(a)
(b)
'>P@
(c)
Figure 17. (a) Fill depth as function of total blade passes compared to bulk density. (b) Linearity
between fill depth and d10 values. (c) Impact of compressibility on fill depth.
190
Pharmaceutics 2022, 14, 278
Tensile Strength
Figure 18 demonstrates the tensile strength (TS) as a function of TBP (−0.704 p = 0.002),
wherein a higher TBP resulted in lower tensile strength at the same compression pressure
(275 MPa), due to increased lubrication efficiency. According to the DoE results in Figure 8,
the significant model terms corresponded to the TBP Equation (2), where a higher through-
put, lower HUM and lower impeller speed result in a lower TBP and, therefore, in higher
tensile strengths of the tablet. If previous process states need to be optimized by adapting
CMT parameters, a similar TBP should be maintained to ensure the correct TS.
Again, after 1560 revolutions, a plateau was reached, and no further reduction in
tensile strength could be noticed with the increasing TBP.
Compression-Force Profile
A compression-force profile was conducted by using 118, 157, 169, 236 and 275 MPa
compression pressure. During phase 16, no compression-force profile could be performed,
because HUM increased from 0.8 to ~1.1 kg and the exit valve opened up to 45 mm, without
any chance of decreasing. Thus, a consistent process flow could not be reached, and the
correct setting of the FD and compression pressure was not possible.
Figure 19 includes the TS as a function of the corresponding compression pressure
and TBP. Figure 19a demonstrates the profiles of each phase as a function of compression
pressure, wherein the lowest TBP showed the highest values. Figure 19b reflects the
TS as a function of TBP, where higher compression pressure led to profiles with higher
values. Table 10 shows the fit statistics regarding tensile strengths obtained during the
compression-force profiles.
Table 10. Overview of fit statistics regarding tensile strengths obtained during compression-
force profiles.
191
Pharmaceutics 2022, 14, 278
(a)
(b)
Figure 19. (a) Overview of all phases (TBP) regarding compression pressure and tensile strength.
(b) Overview of all compression pressures and the corresponding tensile strength based on the
lubrication (TBP).
4. Sweet Spot
By using MODDE, it is possible to detect a sweet spot where several criteria are met.
For this paper, exit valve opening width (1–5 mm), blend potency SD (0–3%), tensile strength
(2–3 MPa) and tablet-weight variability (0–2.5 mg) are considered critical parameters. In
brackets, the favorable process values are shown. Figure 20 shows a visualization of a
combination of input variables (throughput, hold-up mass and impeller speed) in which
all criteria are met (light green). At an impeller speed of 650 rpm, no sweet spot could be
achieved. With reducing impeller speeds, sweet spots at low throughputs are possible at 425
and 200 rpm. At 200 rpm, sweet spots could be achieved at low throughputs independently
of HUM. Since it is preferred to run the process with higher throughputs, an optimal setting
for this formulation can be observed at a combination of high throughputs and high HUM
values at 200 rpm impeller speed.
192
Pharmaceutics 2022, 14, 278
Figure 20. Sweet spot (light green) reveals the combination of the DoE input variables in which the
criteria are met. The color of the borders indicate which criterion is not met anymore. Black borders =
TW SD, red borders = TS and orange boarders = EV.
5. Conclusions
This paper showed the evaluation of the downstream process states based on through-
put, hold-up mass and impeller speed in a continuous direct compression line, including a
single blending step, in a vertical blender (CMT). For all settings in the performed DoE, the
same composition and compounds were used, so that the initial material attributes and
lubrication sensitivity remained constant.
In this study, the model terms of the process states based on the CMT parameters were
evaluated by means of a MLR analysis. Corresponding fit statistics are shown in Table 11.
Responses Q2 R2 Adjusted R2
Exit valve opening width 0.860 0.905 0.883
Exit valve opening width SD 0.822 0.933 0.893
Torque of lower impeller 0.851 0.916 0.896
Torque of lower impeller SD 0.882 0.949 0.933
Conditioned bulk density 0.735 0.850 0.816
Flow rate index 0.800 0.896 0.848
Fill depth 0.873 0.941 0.914
Bottom main compression height 0.774 0.928 0.885
Ejection force 0.892 0.944 0.931
Tablet thickness 0.718 0.953 0.917
Tablet weight 0.642 0.904 0.847
Tensile strength 0.907 0.976 0.958
193
Pharmaceutics 2022, 14, 278
independent of HUM. To run the process as fast as possible, high throughput, high HUM
and 200 rpm IMP are required to fulfill the target criteria and, therefore, represent the
optimal setting for this formulation.
Abbreviations
194
Pharmaceutics 2022, 14, 278
References
1. Blackwood, D.O.; Bonnassieux, A.; Cogoni, G. Continuous direct compression using portable continuous miniature modular &
manufacturing (PCM&M). In Chemical Engineering in the Pharmaceutical Industry; John Wiley & Sons, Ltd.: Hoboken, NJ, USA,
2019; pp. 547–560. [CrossRef]
2. Hsiao, W.-K.; Hörmann, T.R.; Toson, P.; Paudel, A.; Ghiotti, P.; Stauffer, F.; Bauer, F.; Lakio, S.; Behrend, O.; Maurer, R.; et al.
Feeding of particle-based materials in continuous solid dosage manufacturing: A material science perspective. Drug Discov. Today
2020, 25, 800–806. [CrossRef] [PubMed]
3. Nowak, S.; K-Tron, C. Three Ways to Improve Continuous Loss-in-Weight Feeding Accuracy; CSC Publishing: St Paul, MN, USA, 2016.
4. Engisch, W.E.; Muzzio, F.J. Feedrate deviations caused by hopper refill of loss-in-weight feeders. Powder Technol. 2015, 283,
389–400. [CrossRef]
5. Gao, Y.; Muzzio, F.; Ierapetritou, M. Characterization of feeder effects on continuous solid mixing using fourier series analysis.
AIChE J. 2011, 57, 1144–1153. [CrossRef]
6. Hanson, J. Control of a system of loss-in-weight feeders for drug product continuous manufacturing. Powder Technol. 2018, 331,
236–243. [CrossRef]
7. Tahir, F.; Palmer, J.; Khoo, J.; Holman, J.; Yadav, I.K.; Reynolds, G.; Meehan, E.; Mitchell, A.; Bajwa, G. Development of feed factor
prediction models for loss-in-weight powder feeders. Powder Technol. 2019, 364, 1025–1038. [CrossRef]
8. Toson, P.; Siegmann, E.; Trogrlic, M.; Kureck, H.; Khinast, J.; Jajcevic, D.; Doshi, P.; Blackwood, D.; Bonnassieux, A.;
Daugherity, P.D.; et al. Detailed modeling and process design of an advanced continuous powder mixer. Int. J. Pharm. 2018, 552,
288–300. [CrossRef]
9. Engisch, W.E.; Muzzio, F.J. Loss-in-Weight Feeding Trials Case Study: Pharmaceutical Formulation. J. Pharm. Innov. 2015, 10,
56–75. [CrossRef]
10. Engisch, W.E.; Muzzio, F.J. Method for characterization of loss-in-weight feeder equipment. Powder Technol. 2012, 228, 395–403.
[CrossRef]
11. Escotet-Espinoza, M.S.; Moghtadernejad, S.; Scicolone, J.; Wang, Y.; Pereira, G.; Schäfer, E.; Vigh, T.; Klingeleers, D.;
Ierapetritou, M.; Muzzio, F.J. Using a material property library to find surrogate materials for pharmaceutical process
development. Powder Technol. 2018, 339, 659–676. [CrossRef]
12. Hopkins, M. Loss in Weight Feeder Systems. Meas. Control. 2006, 39, 237–240. [CrossRef]
13. Wang, Y.; Li, T.; Muzzio, F.J.; Glasser, B.J. Predicting feeder performance based on material flow properties. Powder Technol. 2017,
308, 135–148. [CrossRef]
14. Gao, Y.; Vanarase, A.; Muzzio, F.; Ierapetritou, M. Characterizing continuous powder mixing using residence time distribution.
Chem. Eng. Sci. 2011, 66, 417–425. [CrossRef]
15. Marikh, K.; Berthiaux, H.; Gatumel, C.; Mizonov, V.; Barantseva, E. Influence of stirrer type on mixture homogeneity in continuous
powder mixing: A model case and a pharmaceutical case. Chem. Eng. Res. Des. 2008, 86, 1027–1037. [CrossRef]
16. Lee, K.T.; Kimber, J.A.; Cogoni, G.; Brandon, J.K.; Wilsdon, D.; Verrier, H.M.; Grieb, S.; Blackwood, D.O.; Jain, A.C.; Doshi, P.
Continuous Mixing Technology: Characterization of a Vertical Mixer Using Residence Time Distribution. J. Pharm. Sci. 2021, 110,
2694–2702. [CrossRef]
17. Mehrotra, A.; Llusa, M.; Faqih, A.; Levin, M.; Muzzio, F.J. Influence of shear intensity and total shear on properties of blends and
tablets of lactose and cellulose lubricated with magnesium stearate. Int. J. Pharm. 2007, 336, 284–291. [CrossRef]
18. Swaminathan, V.; Kildsig, D.O. Effect of Magnesium Stearate on the Content Uniformity of Active Ingredient in Pharmaceutical
Powder Mixtures. Aaps Pharmscitech 2002, 3, 27. [CrossRef]
19. Johansson, M.E. Granular magnesium stearate as a lubricant in tablet formulations. Int. J. Pharm. 1984, 21, 307–315. [CrossRef]
20. Ketterhagen, W.R.; Mullarney, M.P.; Kresevic, J.; Blackwood, D. Computational approaches to predict the effect of shear during
processing of lubricated pharmaceutical blends. Powder Technol. 2018, 335, 427–439. [CrossRef]
21. Kushner, J. Incorporating Turbula mixers into a blending scale-up model for evaluating the effect of magnesium stearate on tablet
tensile strength and bulk specific volume. Int. J. Pharm. 2012, 429, 1–11. [CrossRef]
22. Kushner, J.; Moore, F. Scale-up model describing the impact of lubrication on tablet tensile strength. Int. J. Pharm. 2010, 399, 19–30.
[CrossRef]
23. Kushner, J.; Schlack, H. Commercial scale validation of a process scale-up model for lubricant blending of pharmaceutical
powders. Int. J. Pharm. 2014, 475, 147–155. [CrossRef] [PubMed]
24. Portillo, P.M.; Ierapetritou, M.G.; Muzzio, F.J. Characterization of continuous convective powder mixing processes. Powder
Technol. 2008, 182, 368–378. [CrossRef]
25. Wang, J.; Wen, H.; Desai, D. Lubrication in tablet formulations. Eur. J. Pharm. Biopharm. 2010, 75, 1–15. [CrossRef] [PubMed]
195
Pharmaceutics 2022, 14, 278
26. Vanarase, A.U.; Alcalà, M.; Rozo, J.I.J.; Muzzio, F.J.; Romañach, R.J. Real-time monitoring of drug concentration in a continuous
powder mixing process using NIR spectroscopy. Chem. Eng. Sci. 2010, 65, 5728–5733. [CrossRef]
27. De Leersnyder, F.; Peeters, E.; Djalabi, H.; Vanhoorne, V.; Van Snick, B.; Hong, K.; Hammond, S.; Liu, A.Y.; Ziemons, E.;
Vervaet, C.; et al. Development and validation of an in-line NIR spectroscopic method for continuous blend potency determination
in the feed frame of a tablet press. J. Pharm. Biomed. Anal. 2018, 151, 274–283. [CrossRef] [PubMed]
28. Peeters, M.; Peeters, E.; Van Hauwermeiren, D.; Cogoni, G.; Liu, Y.; De Beer, T. Determination and understanding of lead-lag
between in-line NIR tablet press feed frame and off-line NIR tablet measurements. Int. J. Pharm. 2021, 611, 121328. [CrossRef]
29. Van Hauwermeiren, D.; Peeters, M.; Peeters, E.; Cogoni, G.; Yang, L.A.; De Beer, T. Development of a tablet press feed frame lead
lag determination model using in-line and off-line NIR measurements. Int. J. Pharm. 2021, 612, 121284. [CrossRef]
30. Madian, A.; Leturia, M.; Ablitzer, C.; Matheron, P.; Bernard-Granger, G.; Saleh, K. Impact of fine particles on the rheological
properties of uranium dioxide powders. Nucl. Eng. Technol. 2020, 52, 1714–1723. [CrossRef]
31. Freemann Technology. Instruction Documents: W7013 Stability and Variable Flow Rate; Freemann Technology: Tewkesbury, UK, 2007.
32. Freeman, R. Measuring the flow properties of consolidated, conditioned and aerated powders—A comparative study using a
powder rheometer and a rotational shear cell. Powder Technol. 2007, 174, 25–33. [CrossRef]
33. Freemann Technology. Instruction Documents: W7012 Variable Flow Rate; Freemann Technology: Tewkesbury, UK, 2007.
34. Freemann Technology. Instruction Documents: W7030 Basic Flowability Energy; Freemann Technology: Tewkesbury, UK, 2007.
35. Freemann Technology. Instruction Documents: W7008 Compressibility; Freemann Technology: Tewkesbury, UK, 2007.
36. Freemann Technology. Instruction Documents: W7018 Shear Cell; Freemann Technology: Tewkesbury, UK, 2007.
37. Freemann Technology. Instruction Documents: W7050 1ml Shear Cell; Freemann Technology: Tewkesbury, UK, 2007.
38. Freeman, R.E.; Cooke, J.R.; Schneider, L.C.R. Measuring shear properties and normal stresses generated within a rotational shear
cell for consolidated and non-consolidated powders. Powder Technol. 2009, 190, 65–69. [CrossRef]
39. Wang, Y.; Koynov, S.; Glasser, B.J.; Muzzio, F.J. A method to analyze shear cell data of powders measured under different initial
consolidation stresses. Powder Technol. 2016, 294, 105–112. [CrossRef]
40. Peeters, E. Investigation of the Tableting Process in Continuous Production: Influence of Feeding and Extended Dwell Time during
Compression on Dependent Process Variables and Tablet Properties; Ghent Univeristy: Ghent, Belgium, 2014.
41. Pitt, K.G.; Newton, J.M.; Stanley, P. Tensile fracture of doubly-convex cylindrical discs under diametral loading. J. Mater. Sci. 1988,
23, 2723–2728. [CrossRef]
42. Mukaka, M.M. Statistics corner: A guide to appropriate use of correlation coefficient in medical research. Malawi Med. J. 2012, 24,
69–71. [PubMed]
43. Knight, P.; Seville, J.; Wellm, A.; Instone, T. Prediction of impeller torque in high shear powder mixers. Chem. Eng. Sci. 2001, 56,
4457–4471. [CrossRef]
44. Moghtadernejad, S.; Escotet-Espinoza, M.S.; Oka, S.; Singh, R.; Liu, Z.; Román-Ospino, A.D.; Li, T.; Razavi, S.; Panikar, S.;
Scicolone, J.; et al. A Training on: Continuous Manufacturing (Direct Compaction) of Solid Dose Pharmaceutical Products. J.
Pharm. Innov. 2018, 13, 155–187. [CrossRef]
45. Morin, G.; Briens, L. The Effect of Lubricants on Powder Flowability for Pharmaceutical Application. AAPS PharmSciTech 2013,
14, 1158–1168. [CrossRef]
46. Razavi, S.M.; Gonzalez, M.; Cuitiño, A.M. Quantification of lubrication and particle size distribution effects on tensile strength
and stiffness of tablets. Powder Technol. 2018, 336, 360–374. [CrossRef]
47. Goh, H.P.; Heng, P.W.S.; Liew, C.V. Comparative evaluation of powder flow parameters with reference to particle size and shape.
Int. J. Pharm. 2018, 547, 133–141. [CrossRef]
48. Xie, X.; Puri, V.M. Uniformity of Powder Die Filling Using a Feed Shoe: A Review. Part. Sci. Technol. 2006, 24, 411–426. [CrossRef]
49. Osorio, J.G.; Muzzio, F.J. Effects of powder flow properties on capsule filling weight uniformity. Drug Dev. Ind. Pharm. 2012, 39,
1464–1475. [CrossRef]
50. Uzondu, B.; Leung, L.Y.; Mao, C.; Yang, C.-Y. A mechanistic study on tablet ejection force and its sensitivity to lubrication for
pharmaceutical powders. Int. J. Pharm. 2018, 543, 234–244. [CrossRef] [PubMed]
51. Anuar, M.; Briscoe, B. The elastic relaxation of starch tablets during ejection. Powder Technol. 2009, 195, 96–104. [CrossRef]
52. Dun, J.; Osei-Yeboah, F.; Boulas, P.; Lin, Y.; Sun, C.C. A systematic evaluation of poloxamers as tablet lubricants. Int. J. Pharm.
2020, 576, 118994. [CrossRef] [PubMed]
196
pharmaceutics
Article
Process Modeling and Simulation of Tableting—An Agent-Based
Simulation Methodology for Direct Compression
Niels Lasse Martin 1,2, *,† , Ann Kathrin Schomberg 2,3, *,† , Jan Henrik Finke 2,3 , Tim Gyung-min Abraham 1,2 ,
Arno Kwade 2,3 and Christoph Herrmann 1,2
1 Institute of Machine Tools and Production Technology (IWF), Technische Universität Braunschweig,
Langer Kamp 19b, 38106 Braunschweig, Germany; [email protected] (T.A.); [email protected] (C.H.)
2 Center of Pharmaceutical Engineering (PVZ), Technische Universität Braunschweig, Franz-Liszt-Str. 35A,
38106 Braunschweig, Germany; jan.fi[email protected] (J.H.F.); [email protected] (A.K.)
3 Institute for Particle Technology (iPAT), Technische Universität Braunschweig, Volkmaroder Str. 5,
38104 Braunschweig, Germany
* Correspondence: [email protected] (N.L.M.); [email protected] (A.K.S.); Tel.: +49-531-391-7693 (N.L.M.)
† These authors contributed equally to this work.
Abstract: In pharmaceutical manufacturing, the utmost aim is reliably producing high quality
products. Simulation approaches allow virtual experiments of processes in the planning phase and
the implementation of digital twins in operation. The industrial processing of active pharmaceutical
ingredients (APIs) into tablets requires the combination of discrete and continuous sub-processes
with complex interdependencies regarding the material structures and characteristics. The API and
excipients are mixed, granulated if required, and subsequently tableted. Thereby, the structure as
well as the properties of the intermediate and final product are influenced by the raw materials,
Citation: Martin, N.L.; Schomberg,
the parametrized processes and environmental conditions, which are subject to certain fluctuations.
A.K.; Finke, J.H.; Abraham, T.;
In this study, for the first time, an agent-based simulation model is presented, which enables the
Kwade, A.; Herrmann, C. Process
Modeling and Simulation of
prediction, tracking, and tracing of resulting structures and properties of the intermediates of an
Tableting—An Agent-Based industrial tableting process. Therefore, the methodology for the identification and development of
Simulation Methodology for Direct product and process agents in an agent-based simulation is shown. Implemented physical models
Compression. Pharmaceutics 2021, 13, describe the impact of process parameters on material structures. The tablet production with a pilot
996. https://doi.org/10.3390/ scale rotary press is experimentally characterized to provide calibration and validation data. Finally,
pharmaceutics13070996 the simulation results, predicting the final structures, are compared to the experimental data.
Academic Editors: Colin Hare and Keywords: agent-based modeling and simulation; process modeling and simulation; tableting;
Anne Marie Healy product structures and characteristics
methodologies forecasting the final product quality or even feed-forward oriented control
strategies for process parameter adaption are mostly missing. Hence, this gap often causes
batch loss and imposes high cost.
Dynamic simulation approaches allow for the forecasting of processes and their
process chains and, when combined with real world data, data-driven, knowledge-based
control of these processes [3]. In the planning phase, forecasting the resulting product
structures and characteristics of the planned process chain is possible. In the operation
phase and based on proper simulation approaches, deviations can be compensated for by
adjusting the parameters of upcoming process steps, adapting treatment times, or repeating
or adding further steps. Furthermore, simulation approaches allow the analysis of complex
interdependencies and are therefore useful to generate process knowledge. Four typical
simulation paradigms can be distinguished in dynamic simulation modeling [4] and are
shown in Figure 1. Depending on the abstraction level of the model and the change of
process parameters over time (continuous or discrete), different paradigms are more likely
to be chosen. For example, models with a high degree of detail and a continuous parameter
change are likely to be modelled using the dynamic systems paradigm.
Discrete event simulations (DES) are widely used for modeling complex networks such
as Petri nets, which do not require continuous parameter changes but are very helpful for
the determination of key performance indicators of process chains. System dynamics (SD)
models describe the system behavior with few details of the entities on a high abstraction
level, and are often used for complex social or political systems [5].
Figure 1. Typical dynamic simulation approaches for the tableting process mapped on the simulation
modeling paradigms (following [4–6]).
For simulating a tableting process or sub-process, the common approach is the use of
dynamic systems simulations such as flowsheet simulations (FSS). Furthermore, discrete
element methods (DEM) are used, which can further be classified depending on their level
of detail. This scale can reach from cellular automata to combined finite-discrete element
simulation as illustrated in Figure 1. A detailed description for those methods can be
found in [6]. Simulations built upon the finite element method (FEM) are used to model
particle deformation behavior under external stresses and the propagation of stresses
within particles. Cellular automata approaches model the interaction with neighboring
particle groups rather than the particles themselves, while combined finite-discrete element
approaches model particle deformation, interaction with other particles, form and size
in detail. FSS are mainly used to describe the change of properties of continuous phases
(whereas solids are also considered continuous phases) and are well established in the
process industry. Similar to the DES models, networks of entities and transitions are formed,
representing process chains and aggregates, for example. The entities of a FSS network,
so called nodes, are used to calculate output streams of given input streams according
198
Pharmaceutics 2021, 13, 996
to physical circumstances. Due to the more complex description of solids, the use of FSS
models for solid processes has only recently been investigated [7–9].
Those dynamic systems simulation approaches have already proven their value in
simulating production processes and chains or detailed process behaviors, respectively.
However, considering the different abstraction levels of the approaches, FSS lack the ability
to determine distributed product properties and their influence on the process. In contrast
to that, discrete element simulations lack the ability to simulate whole process chains due
to high computational efforts [10]. In past research, there have been efforts to adapt FSS
especially for particulate processes in order to model detailed product characteristics to
overcome this gap. This resulted in a modular open source system covering diverse process
units by using multidimensional distributed product structures [11]. Additionally, reduced
order models allow the integration of originally highly detailed models in FSS approaches
at low computational expense [12].
In comparison to the common dynamic systems modeling approaches, agent-based
(AB) models allow both the ability to represent process chain behaviors and the ability
to determine heterogenic product properties as well as their interdependencies with the
processes and the consideration of different detail levels [4,13]. Therefore, AB models could
be used to simulate intermediate product changes and the effects of those changes on the
process and vice versa. Furthermore, AB models are able to represent cellular automata [13].
This enables the simulation of interrelations between neighboring materials and processes,
a typical discrete element simulation characteristic [6]. Therefore, AB simulations helps to
gain a better understanding of process chains, sub-processes as well as product properties,
and represents a practical alternative to the existing simulation approaches. Especially for
considering not only single processes, but the complete process chain of a pharmaceutical
production, AB approaches have advantages in traceability and the transfer of product
characteristics for specific entities.
The overall aim in this research is to evaluate processes and entire process chains in the
planning phase and to support process and process chain improvements during operation
with respect to the quality of pharmaceutical products using simulations. AB models are
capable of integrating the advantages of the common dynamic systems approaches. Conse-
quently, an AB model on the tableting process in a rotary press is introduced, providing
the basis for further research. The process step of tableting is chosen to demonstrate the
applicability of the approach, which is, however, a generic approach that can be applied
virtually on other process steps in the process chain. It highlights the necessity of careful
process description and analysis (Section 2), setup of an AB model framework for a rotary
press (Section 3), model identification, development and integration into this simulation
approach, and its application and comparison to experimental results (Section 4), demon-
strating its capability of determining product characteristics from parameter settings as
well as tracing discrete product entities and material structures.
199
Pharmaceutics 2021, 13, 996
Figure 2. Direct compression process chain with exemplary (intermediate) product structures.
In addition, the applicability of direct compression is limited due to the high demands
on the formulation regarding good flowability and low segregation tendencies. Depending
on the properties and structure of API particles, the basic process chain of direct compres-
sion needs necessarily to be extended by different possible process steps, which are for
example different granulation processes or subsequent tablet coating. Accordingly, the
combined simulation of all processes for direct compression and the inclusion of granu-
lation and coating will be the subject of future research. This research treats the direct
compression process chain with a special focus on the tableting process.
200
Pharmaceutics 2021, 13, 996
201
Pharmaceutics 2021, 13, 996
tableting model, they used a Kawakita equation for modeling the final tablet porosity and
the final tablet hardness is modelled as a function of the tablet relative density according to
Kuentz and Leuenberger [19,20].
Rogers et al. extend the work of Boukouvala et al. [2,20], enabling the simulation of
further CQAs such as the tensile strength [21], while Singh et al. even describe a control
strategy for a multi-purpose continuous processing of pharmaceutical processes using the
FSS model of Boukouvala et al. [9,20]. The dynamic FSS modeling approaches allow a
good understanding and overview over the process chains and their parameters. Those
parameters can even be controlled. However, the crucial components of FSS are the models
for the individual processes and sub-processes. Those processes can be modeled to a certain
degree of detail using semi-empirical or physical models. Comprehensive physical models
that need microscale simulation, such as DEM, are not considered within the dynamic
FSS, and therefore the determination of parameters for single products or product groups
remains vague and could be improved [8].
The attributes of an agent can be static or dynamic, and each agent has a behavior that
might change depending on rules adapting the behaviors. Such rules can be, for example,
physical models such as the Heckel equation or they can depend on the state of the agent.
The states are connected via transitions that follow specific rules before the next state is
reached. Depending on the logic of the state, those transitions switch their behavior time
dependently, at specific ratios, or via trigger. Those triggers can occur from the interactions
with other agents or the agent’s environment. These and other aspects as well as use cases
of AB modeling and simulation are well described in the literature [13,35,36].
The AB simulation paradigm was intended to simulate the dynamics of complex sys-
tems consisting of populations of autonomous, interacting agents or components. Nowa-
202
Pharmaceutics 2021, 13, 996
days, AB approaches are more and more common in production engineering [36]. AB
simulation in particular can be used to describe the interactions of machines and (interme-
diate) products represented by individual agents [37].
To the authors’ knowledge, AB simulation has not been used for pharmaceutical
processes generally or tableting specifically. AB simulation enables the analysis of the
development for specific product characteristics over the entire process chain, which is
an advantage in comparison to the DEM and FSS. Here, processes can be adapted in the
process of product development. Furthermore, the product quality can be controlled during
the production phase. In addition, discrete and continuous processes and sub-processes
can be considered together, which enhances the process understanding and is essential in
the transition to continuous pharmaceutical production.
203
Pharmaceutics 2021, 13, 996
The identified material agents are the blend in the four different sub-process steps
and especially the tablets at the end of the sub-process chain. Four different blend material
agents are considered so as to better model the process-material behavior, the intermediate
products, and the final tablet structures. The process parameters of the agents are listed
in Figure 5. For the ease of understanding and the distinction of original processes and
modeled agents, in the following all agents are written in italics.
As the CQAs of the final product are in the focus and the material flow rate is de-
termined by the number and weight of tablets that are produced per time, the process
chain and its parametrization is derived starting from the last material agent (tablet) and
ending at the first process agent (hopper) to reasonably define the sizes and interactions.
The agents in the agent group tablet are modeled as a passive data box, where the product
structures and properties for each tablet are collected. Even though several dies are filled
simultaneously on a rotary press, only one die is compressed at a time. Therefore, the
authors choose to model the blend in die as a single agent representing alternatingly all
dies of the rotary press. The twelve rotating stirrer blades on the paddle wheel in the
feed frame transport the powder inside their respective interspaces. These interspaces are
modeled as twelve compartments, each represented by one blend in feed frame agent. In
order to simulate the powder flow within the filling pipe, a group of agents is defined as
the blend in filling pipe. A flow profile inside the pipe with slower flow close to the wall
compared to the middle is expected. Therefore, the agent group consists of two agent
types–the midstream, representing the higher velocities in a central circular region, and the
outer ring (Figure 6), in which velocities are lower. Both types of agents possess the same
height in every vertical position, forming layers within the filling pipe, each containing a
midstream and an outer ring. The midstream agents travel faster than the outer ring agents
by a constant velocity difference.
Figure 6. Schematic visualization of the process and material agents with the relevant states of the
process agents.
In the state of equilibrium, the inflowing and outflowing powder streams in the feed
frame and the filling pipe are equal in volume. Thus, the volume of the powder agents
entering the feed frame from the filling pipe equals the powder volume filled to the die—
the dosing volume. Therefore, the summarized volumes of one outer ring and a defined
number of midstream blend in filling pipe agents, dependent on the velocity difference, are
equal to the dosing volume of the die. The hopper serves as a feed for the blend in this model
and the blend in hopper volume is reduced according to the remaining blend in the hopper.
204
Pharmaceutics 2021, 13, 996
3.2.1. Hopper
The hopper is used as a reservoir for the blend, which is initially introduced to the
tableting process. The blend in hopper agent’s structure is homogeneously distributed.
Within the ‘fill filling pipe’ state, the structure of the blend in hopper is transferred into the
newly generated blend in filling pipe agents. Thus, the mass fraction of the components
inside the filling pipe is determined by the feed exiting the hopper. Segregation phenomena
that may occur already in the hopper (e.g., during filling) will be the focus of future research
work and are not included in this paper.
1 xi
=∑ (1)
ρ f ill,mix i
ρbulk,i
Based on the powder weight filled into the die, the tablet weight m T can be calculated
considering the dosing height hdos :
2
Ddie
mT = π × × hdos × ρ f ill,mix (2)
2
205
Pharmaceutics 2021, 13, 996
the blend in die agent ρcomp,mix changes during compression, which can be calculated from
the minimal in-die height hmin and the tablet weight mT :
mT
ρcomp,mix = 2 (3)
Ddie
π× 2 × hmin
The compression curve (porosity over compression stress) can be applied to determine
the resulting stress (and calculate the respective force) that is applied to achieve the bespoke
compression density ρ P,max,mix . A wide range of models was developed, describing such
compressibility curves by mathematical equations [38,39]. The most prominent compres-
sion model is the model of Heckel [40]. Further compressibility models were developed
by Kawakita, Gurnham, Cooper, and Eaton, as well as Wünsch et al. [39,41–43]. However,
for describing the compressibility behavior of powder blends with known but varying
composition, an approach with an appropriate mixing rule is necessary. Busignies et al.
presented a volume-additive approach to predict the tablet density of a formulation using
the Kawakita model [44]. They transformed the classical equation into a model of the tablet
density where ρ0,i describes the density at low pressure (here 20 MPa) and ai and bi are
constants that need to be calibrated for the respective material:
1 x i 1 + ( 1 − a i ) × bi × P
=∑ (4)
ρ P,max,mix i
ρ0,i ( 1 + bi × P )
In this study, Equation (4) is converted to equal the compression stress P and solved
via the pq-formula to gain the compression stress of a blend with two components.
From a process agent’s point of view, the ejection state is used to create a new tablet
agent. After ejection, the tablet density changes due to the elastic recovery. Hirschberg
et al. developed an approach to predict the out-die density ρout-die based on the data of two
tablets compressed at high and low compression stresses [45]. Therefore, the measured
densities at maximum stress ρP,max , at zero axial stress ρP,0 , and out-of-die ρout-die are used.
The change in density due to the instantaneous elastic recovery (Δρin-die = ρP,max − ρP,0 ) is
expected to increase linearly with the compression stress while the change due to the slow
elastic recovery (Δρslow = ρP,0 − ρout-die ) is expected to be constant:
Δρslow ( P1 ) + Δρslow ( P2 )
Δρslow, o = (6)
2
ρout−die = ρ P, max − Δρin−die − Δρslow,o (7)
In order to predict the out-die density for mixtures of the excipients with varying
compositions, a mixing rule has to be introduced. For the mixtures, only the mass fraction
and the in-die density at the apparent stress ρ P, max,mix are known, so the elastic recovery
has to be described based on the data of the pure substances. Therefore, according to the
volumetric approach of Busignies et al., the densities at the different states (ρP,max , ρP,0 and
ρout-die ) for the high and low compression stress, investigated for the pure substances, are
calculated for the mixtures (according to Equation (1)). Following this, the change in in-die
density and out-die density (Equations (5) and (6)) can be calculated. Finally, the predicted
out-die density results from the apparent in-die density at maximum stress ρP,max minus
the corrective function Δρin-die at the respective compression stress and the corrective value
Δρslow ,ø (Equation (7)). Finally, the out-die density is corrected using the difference of the
predicted values and the measured out-die densities, as suggested by Hirschberg et al.
The final tensile strength of the tablets dominantly depends, besides other influences,
on the out-die porosity. In general, the lower the out-die porosity εout-die , the higher the
206
Pharmaceutics 2021, 13, 996
The coefficient σ0 describes the strength of a nonporous body of the same material,
while kb corresponds to the slope of the ln σ vs. εout-die curve.
207
Pharmaceutics 2021, 13, 996
Table 2. Overview of the states of rotary press process agents, their behaviors, and their interactions.
The whole AB model was implemented using the modeling and simulation software
AnyLogic® . In the following, the simulation results are compared with experimental results.
4.1. Materials
Anhydrous dicalcium phosphate (DCPA) in two grades (DI-CAFOS® A60, DI-CAFOS®
A150, Chemische Fabrik Budenheim KG, Budenheim, Germany) were used as model
materials. The characteristic particle sizes of DCPA A60 and DCPA A150, determined
by laser diffraction (Mastersizer 3000, Malvern Panalytical, Kassel, Germany), differ over
the whole distribution (Table 3, Figure 8). To enable their processability, DCPAs were
mixed with 1 wt% magnesium stearate (MgSt, Magnesia GmbH, Lüneburg, Germany) in a
cube blender (ERWEKA GmbH, Langen, Germany) for five minutes at 30 rpm. The bulk
ρb and tapped density ρt were determined according to the Ph. Eur. 9.3 2.9.34 using a
100 mL cylinder and a volumetric analyzer (Erich Tschacher Laboratoriumsbedarf, Bielefeld,
Germany). As the two grades consist of the same chemical material, the solid density
ρs , measured in triplicate with the helium pycnometer Ultrapyc 1200e (Quantachrome
Instruments, Boynton Beach, FL, USA), are practically identical (Table 3).
208
Pharmaceutics 2021, 13, 996
Table 3. Characteristic particle sizes, solid, bulk, and tapped density for DCPA A60 and A150.
Material x10 (μm) x50 (μm) x90 (μm) ρs (g/cm3 ) ρb (g/cm3 ) ρt (g/cm3 )
DCPA A60 34 64 116 2.849 1.33 1.51
DCPA A150 96 167 263 2.842 0.68 0.75
209
Pharmaceutics 2021, 13, 996
for the Kawakita model, the compressibility curves of the pure substances A60 and A150
were fitted in the range of 20 to 350 MPa. The starting pressure of 20 MPa was used, as the
Kawakita model is generally not suitable for very low pressure ranges [49]. The coefficients
can be found in Table 4.
Excipient ρ0 a b R2
DCPA A150 1.22954 0.4981 0.0072 0.9962
DCPA A60 1.62302 0.4891 0.0041 0.9959
In order to calibrate the model of Hirschberg, the densities in different states (ρP,max ,
ρP,0 and ρout-die ) of ten tablets of each component, DCPA A60 and A150, compressed at
30 and 400 MPa were used. The model of Ryshkewitch–Duckworth was calibrated using
the tensile strengths of tablets of all six compression stresses. In order to determine the
coefficients kb and σ0 and to develop a mixing rule for them, the compactability curves of
the pure substances and the 50:50 blend were fitted with the model-equation (Equation (8)).
The resulting progress of σ0 presents an exponential development with rising mass fraction
of DCPA A150 xA150 . Thus, for the mixtures σ0 is calculated as follows:
k b = m × x A150 + n (10)
expk −simk
n
simk
f = ∑ n
∀ k = experimental measuring points (11)
k =1
210
Pharmaceutics 2021, 13, 996
parameters are not directly assessable by experiments, so far. Firstly, the effect of the
theoretical diameter of the midstream, assuming a constant velocity ratio of 2:1 to the outer
ring of the filling pipe, is investigated within the simulation. Secondly, the filling pattern
from the feed frame fragments to the die, defining what fraction of the whole die volume is
filled by successive feed frame compartments passing over the die during one filling event
is also evaluated within the simulation.
The sensitivity of the simulation towards deviation of the repeated midstream di-
ameter increasing by always 6 mm shows that the relative error between the midstream
diameter of 36 mm and 30 mm has the lowest relative errors (Figure 9a). Therefore, the
mean of those two values (33 mm) is chosen as the midstream diameter. This proved well
suited, as its f value of 1.32% is the lowest in this data set (see Table 6).
Figure 9. Relative error between simulation and experimental results for (a) different midstream
diameter settings and (b) die filling ratio settings in relation to the tablet weight.
Table 6. Relative error of midstream model configuration regarding the tablet weight.
The sensitivity of the distribution of filling over three compartments was studied by
reducing the fraction filled by the first compartment in steps of 10%-points, distributing
the rest of the filling reasonably over the second and the third compartment with the
assumption that the ratio declines over the compartment number.
The lowest values are found for the ratio 40/35/25 and 34/33/33 (Table 7). This
finding shows that a distribution of the die filling is better described by filling from more
than one compartment of the feed frame. However, it must be born in mind that this
value most likely depends on the flowability of the formulation as well as on the process
parameters of turret and paddle speed. A descending pattern of 40/35/25 was used for the
case study.
Table 7. Relative error of die filling ratio configuration regarding the tablet weight.
1. Compartment 100 90 80 70 60 50 40 34
Filling ratio (%) 2. Compartment 0 5 15 25 35 40 35 33
3. Compartment 0 5 5 5 5 10 25 33
Mean Absolute Value of the Relative Error (%) 5.35 5.04 4.64 4.11 3.38 2.46 1.32 1.24
211
Pharmaceutics 2021, 13, 996
Figure 10. (a) Measured (solid lines), fitted (dashed lines), and calculated (dotted lines) compressibil-
ity curves of DI-CAFOS A60, A150, and mixtures, and (b) the relative error between the measured
and the predicted data.
212
Pharmaceutics 2021, 13, 996
Figure 11. (a) Measured (solid lines) and calculated (dotted lines) out-die porosities of DI-CAFOS
A60, A150, and mixtures, and (b) the relative error between the measured and the calculated data.
Finally, the tablet tensile strength was estimated in dependence on the out-die porosity
using the model of Ryshkewitch and Duckworth. As described in Section 4.3, a new mixing
rule was developed to predict the tensile strength of tablets consisting of two different
components. The coefficients of the model are calculated based on the mass fraction of
DCPA A150 (Equations (9) and (10)). The compactability curves of A150 and the mixture
25 wt.% A60 (Figure 12a) show a strong increase in tensile strength with decreasing out-die
porosity. With increasing mass fraction of A60, the curve flattens. The correct mathematical
description of the curves is challenging due to high changes in tensile strength for low
differences in the out-die porosity. This causes the high relative error between the measured
and the predicted tensile strength with about −30 up to 20% (Figure 12b). Although, the
calculated tensile strength curves fit the experimental values very well, as it can be seen in
a first approximation in Figure 12a. Nevertheless, it has to be taken into account that the
data for each composition is only relevant in a specific compression stress range and thus a
porosity range, as the stress changes with the composition and so does the out-die porosity.
The relevant relative errors of the tensile strength are marked for each blend composition
with hollow symbols (Figure 12b).
Figure 12. (a) Measured (solid lines) and calculated (dotted lines) tensile strength of DI-CAFOS A60,
A150, and mixtures, and (b) the relative error between the measured and the calculated data.
213
Pharmaceutics 2021, 13, 996
experiment, containing 100 wt.% DCPA A150. After about twelve minutes, the tablet
weight increases significantly as the DCPA A60 content starts to rise. This can be traced
back to the increasing apparent bulk density with higher DCPA A60 content as its small
particles fill pores and lead to a denser arrangement of the powder bed. Therefore, the
weight filled into the die increases, resulting in a higher tablet weight. With increasing mass
content of DCPA A60, the change in mass fraction as well as tablet weight flattens. Due
to the residence time distribution of DCPA A150, it takes a certain time to fill remaining
DCPA A150 particles into the dies, leading to change in content inside the feed frame.
Figure 13. Comparison of the simulative data with experimental results for (a) the mass fraction of
DCPA A150 and A60 and (b) the tablet weight.
The good agreement between the simulative data and the measured values for the
tablet weight and mass fraction (Figure 13) are particularly worth to mention. Only for high
DCPA A60 mass fractions, simulative data exhibit slightly lower tablet weights compared
to the measured values. This underestimation might be linked to die filling, where apparent
densities above the bulk density are possible due to particle rearrangements during forced
feeding and especially during dosing [47]. To meet the tablet weight, obtained by the
experiment, an apparent consolidated bulk density after filling of 0.73 g/cm3 had to be
used, which is considerably higher than the bulk density of 0.68 g/cm3 . As described in
the literature, good flowing powder as dicalcium phosphate consolidates at high paddle
speeds and low turret speeds inside the die. The process parameters selected here as well
as the material properties support this hypothesis. The same observations were made
for DCPA A60, so the apparent consolidated bulk density after filling was taken to be
1.41 g/cm3 compared to the bulk density of 1.33 g/cm3 .
Although no specific mixing model is introduced in the AB simulation, mixing and
the distribution of the newly entering DCPA A60 can be represented by the set-up of the
simulation itself. As described in Section 3, the midstream with a diameter of 33 mm
simulates a flow profile inside the filling pipe according to its geometry, by having double
the velocity as the surrounding powder flow (compare Section 3). Thus, a powder flow
similar to the actually expected is aspired. As soon as the powder, represented by the blend
in filling pipe, enters the feed frame, the mass fraction of the respective compartment is
recalculated. Therefore, the twelve compartments work as independent continuous stirred
tank reactors (CSTRs), as the mixing ratio is constant within the entire compartment in
each time step. Comparing this setting to the visual observations on the tablet press using
a transparent feed frame, a completely homogenous concentration of the excipients is not
realistic, especially not directly after new powder with a different composition entered
the feed frame. Mixing in the feed frame over time can be observed, while the powder
is not only mixed inside an interspace but is also able to change the compartment if the
particles are close to the bottom or flow over the paddles in real experiments. Due to the
high paddle speed of 60 rpm and the simultaneously low turret speed of 20 rpm, leading to
a high residence time, the powder can mix quickly inside the feed frame. Puckhaber et al.
214
Pharmaceutics 2021, 13, 996
investigated the residence time distribution on the XL 100 for a pure dicalcium phosphate
with similar powder properties as DCPA A150, showing strong intermixing for the process
parameters used in this study [50]. This improves the model quality in so far that the
simulated mass fraction, resulting from the compartments, modeled as CSTRs, is in good
agreement with the experimental data. In future work different combinations of paddle
and turret speed shall be looked at to investigate whether the concept of the CSTRs is still
valid or must be refined for combinations of free flowing with poorly flowing materials,
for example.
For the experimental data of the mass fraction, a very good correlation between
experimental and simulative values was found (Figure 13b). However, higher standard
deviations are observed for the experimentally determined mass fraction than for the tablet
weight, which shows standard deviations below 1.5%. The mass fraction is determined by
using dyed DCPA A60, which intrinsically shows higher standard deviations of the loading
as ten samples à 100 mg presented a relative standard deviation of 11.7%. Therefore, the
tablet weight (n = 6) is a more reliable and more facile to determine parameter to compare
the simulative and experimental data.
Besides the tablet weight, the out-die porosity and the tensile strength are of particular
interest regarding their possible correlation to disintegration time and the handling stability
of the tablets, respectively. The out-die porosity exhibited by the simulation fits very well
with the experimental data (Figure 14a). To achieve this degree of convergence, the models
used to describe the maximum compression stress and the total elastic recovery as function
of filling weight and blend density do present a very good applicability.
Figure 14. Comparison of the simulative data with experimental results for (a) the out-die porosity
and (b) the compression stress.
As shown for the validation of the process models, the in-die tablet density and thus
the in-die tablet porosity have a deviation between −2 and 4% above 50 MPa (Figure 10b).
As the tablets contain only DCPA A150 at a compression stress of about 60 Mpa in this
case, the apparent error due to the Kawakita model is close to 0%. As Figure 15a shows,
the resulting simulated compression stress in this study for 75 wt.% DCPA A60 is around
200 Mpa. In this stress range, the deviation for the respective content is small (Figure 11b,
hollow symbol). Although tablets with 50 wt.% and 25 wt.% DCPA A60 present high
relative errors around 400 Mpa (Figure 11b), they do not affect the simulation results as the
apparent compression stress is lower than 150 Mpa for these mass fractions. Therefore, the
total mean error for the out-die porosity is low with 2.13% as listed in Table 8.
215
Pharmaceutics 2021, 13, 996
Figure 15. (a) Calculated compression stress over the mass fraction of DCPA A60. (b) Comparison of
the simulative data with experimental results for the tensile strength.
For the comparison of the experimentally recorded compression stress on the rotary
press and the simulated values, a very good match can be observed (Figure 14b). In
the high stress range, the simulative data underestimates the experimental values. This
might be due to the slightly lower simulated tablet weight at high DCPA A60 content.
At such a high compression state, a low change in weight has a relatively strong effect
on the corresponding compression stress (compare Figure 10a). For further research,
the consolidation of the powder during and after filling has to be addressed in more
detail to better calculate the tablet weight and thus predict the compression stress even
more accurately.
Regarding the tensile strength, considerably higher deviations between the simulative
and the experimental data are observed (Figure 15b). While the measured values match
well for pure DCPA A150 and pure A60, the correct prediction of the tensile strength during
the change in content is challenging (Figure 16). The tensile strength obtains the highest
relative errors for the simulative data with over 100% with rising mass fraction of DCPA
A60 and about 70% with decreasing content (Figure 16a). The lower quality of the predicted
data of the tensile strength can be traced back to the model and especially the mixing rule
used. For the model validation, already relative errors of -35 up to 20% are obtained,
while no specific trend for the mass fraction is observable (Figure 12b). Nevertheless,
the model of Ryshkewitch uses the out-die porosity as input parameter to calculate the
corresponding tensile strength, which includes previous errors. Due to the strong increase
in strength for low changes in porosity, small differences in the porosity have a high impact.
A parabolic shape of the relative error over the mass content can be observed, which
indicates a systematic error for the determination of the tensile strength (Figure 16b).
216
Pharmaceutics 2021, 13, 996
Figure 16. Relative error of the simulative data compared to experimentally determined values for
the (a) tablet weight, out-die porosity, and tensile strength over time and (b) the mass fraction of
DCPA A60.
Interestingly, comparing the increasing and decreasing DCPA A60 fractions, the
relative errors of all three parameters present higher values with increasing DCPA A60
fraction than vice versa (Figure 16b). This might be due to the change in the apparent bulk
density as the actual volume of the powder is lower than the calculated one by adding the
respective volumes of each powder. Additionally, the higher deviation with rising DCPA
A60 content in comparison to the decreasing content can indicate different residence time
distributions and thus, different distribution profiles of the two excipients. Both will be
investigated in future research.
Considering the overall accuracy of the simulation results, the mean absolute relative
error f (compare Equation (11)) of the CQAs can be seen in Table 8.
Besides the tensile strength of the tablets, all other CQAs have a mean relative error of
less than or equal to 6.99%. However, the quality of the AB model can only be as good as
the quality of the physical models used to calculate product properties. For the calculation
of the tensile strength a better model and, moreover, a better bulk density determination of
the component mixture is necessary.
From a simulation point of view, the AB modeling approach allows the observation
of each agent over the complete simulation time. Furthermore, the input materials have
been digitally marked and allow a tracking and tracing over the whole process. This fact
allows to investigate the output in relation to the input on a specific matter. With this,
several processing questions can be answered, e.g., the characterization of tablets that have
fractions of a specific input material. This shows the potential of AB models to be capable
of model particle-based processes chains at low computational cost, prospectively making
their application in real time control of pharmaceutical production processes feasible.
217
Pharmaceutics 2021, 13, 996
Further research in this field should consider the effect of additional materials with dif-
ferent properties as well as predecessor and successor steps, e.g., blending, granulation and
coating. The main research questions consider the interconnection between these models
and the transfer of deep process knowledge into an AB simulation approach. This model
development is by now rather complex and further modeling strategies should be explored.
Moreover, the agent’s behavior should be enhanced to foster better and more accurate
simulation results. With this, several effects, as for example the segregation of components,
can be modeled within an AB model. Furthermore, environmental influences should be
integrated in AB models, e.g., the environmental relative humidity or temperature.
In order to contribute to the digitalization of manufacturing systems, the resulting AB
process chain models need to be empowered for process control, e.g., within cyber-physical
production systems. To this end, the development of an AB model can be based on or
assisted by QbD approaches and use inline data by coupling PAT with the AB model
to allow the forecast of product quality over complex process chains close to real time.
Additionally, the design space for new products might be more easily defined. New
formulations can already be evaluated in advance using the simulation before further
experiments are conducted, which saves time and costs. By now, the simulation approach
is able to perform descriptive as well as diagnostic analytics and in addition is much faster
than the experimental process. In future research, this model needs to enable predictive
or even prescriptive analytics. This will allow future manufacturing systems not to be
mandatorily centralized and bound to the necessary process knowledge but will be able to
work in decentralized locations with simulation models controlling the processes.
Author Contributions: Conceptualization, N.L.M., A.K.S., J.H.F., T.A., A.K. and C.H.; methodology,
N.L.M. and A.K.S.; software, N.L.M.; validation, A.K.S., N.L.M. and J.H.F.; formal analysis, A.K.S. and
N.L.M.; investigation, A.K.S. and N.L.M.; resources, N.L.M., A.K.S., J.H.F., T.A., A.K. and C.H.; data
curation, A.K.S. and N.L.M.; writing—original draft preparation, N.L.M. and A.K.S.; writing—review
and editing, J.H.F., T.A., A.K. and C.H.; visualization, N.L.M. and A.K.S.; supervision, A.K. and C.H.;
project administration, A.K. and C.H.; funding acquisition, N.L.M., J.H.F., A.K. and C.H. All authors
have read and agreed to the published version of the manuscript.
Funding: This paper evolved from the research project “Simulation of distributed product structures
in combined discrete and continuous production processes for solid, particulate products”, which is
funded by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation)–413141366.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: Not applicable.
Acknowledgments: We acknowledge support by the German Research Foundation and the Open
Access Publication Funds of Technische Universität Braunschweig. Furthermore, the authors ac-
knowledge and thank KORSCH AG for the support and the provision of the tablet rotary press XL
100 and Chemische Fabrik Budenheim KG for the provision of testing materials (DI-CAFOS® A150
and DI-CAFOS® A60).
Conflicts of Interest: The authors declare no conflict of interest.
References
1. Lee, S.L.; O’Connor, T.F.; Yang, X.; Cruz, C.N.; Chatterjee, S.; Madurawe, R.D.; Moore, C.M.V.; Yu, L.X.; Woodcock, J. Modernizing
Pharmaceutical Manufacturing: From Batch to Continuous Production. J. Pharm. Innov. 2015, 10, 191–199. [CrossRef]
2. Boukouvala, F.; Niotis, V.; Ramachandran, R.; Muzzio, F.J.; Ierapetritou, M.G. An integrated approach for dynamic flowsheet
modeling and sensitivity analysis of a continuous tablet manufacturing process. Comput. Chem. Eng. 2012, 42, 30–47. [CrossRef]
3. Thiede, S. Environmental Sustainability of Cyber Physical Production Systems. Procedia CIRP 2018, 69, 644–649. [CrossRef]
4. Borshchev, A.; Filippov, A. From System Dynamics and Discrete Event to Practical Agent Based Modeling: Reasons, Techniques,
Tools. In Proceedings of the 22nd International Conference of the System Dynamics Society, Oxford, UK, 25–29 July 2004; Kennedy,
M., Ed.; System Dynamics Society: Albany, NY, USA, 2004. ISBN 0974532908.
5. Forrester, J.W. System dynamics, systems thinking, and soft OR. Syst. Dyn. Rev. 1994, 10, 245–256. [CrossRef]
218
Pharmaceutics 2021, 13, 996
6. Hildebrandt, C. Improving Die Filling in Pharmaceutical Tableting by Experimental and Numerical Means. Ph.D. Thesis, Kiel
University, Kiel, Germany, 2018.
7. Hartge, E.-U.; Pogodda, M.; Reimers, C.; Schwier, D.; Gruhn, G.; Werther, J. Flowsheet Simulation of Solids Processes. KONA
Powder Part. J. 2006, 24, 146–158. [CrossRef]
8. Dosta, M.; Litster, J.D.; Heinrich, S. Flowsheet simulation of solids processes: Current status and future trends. Adv. Powder
Technol. 2020, 31, 947–953. [CrossRef]
9. Singh, R.; Ierapetritou, M.; Ramachandran, R. Hybrid Advanced Control of Flexible Multipurpose Continuous Tablet Manufac-
turing Process via Direct Compaction. In 23rd European Symposium on Computer Aided Process Engineering; Elsevier: Amsterdam,
The Netherlands, 2013; pp. 757–762. ISBN 9780444632340.
10. Ketterhagen, W.R.; Ende, M.T.A.; Hancock, B.C. Process Modeling in the Pharmaceutical Industry using the Discrete Element
Method. J. Pharm. Sci. 2009, 98, 442–470. [CrossRef]
11. Skorych, V.; Dosta, M.; Heinrich, S. Dyssol—An open-source flowsheet simulation framework for particulate materials. SoftwareX
2020, 12, 100572. [CrossRef]
12. Rogers, A.J.; Hashemi, A.; Ierapetritou, M.G. Modeling of Particulate Processes for the Continuous Manufacture of Solid-Based
Pharmaceutical Dosage Forms. Processes 2013, 1, 67–127. [CrossRef]
13. Macal, C.M.; North, M.J. Tutorial on agent-based modelling and simulation. J. Simul. 2010, 4, 151–162. [CrossRef]
14. Belič, A.; Škranjc, D.Z.-B.I.; Vrečer, F.; Karba, R. Artificial Neural Networks for Optimisation of Tablet Production. In Proceedings
of the 6th EUROSIM Congress on Modelling and Simulation, EUROSIM 2007, Ljubljana, Slovenia, 9–13 September 2007;
ISBN 978-3-901608-32-2.
15. Aksu, B.; Paradkar, A.; De Matas, M.; Özer, Ö.; Güneri, T.; York, P. Quality by Design Approach: Application of Artificial
Intelligence Techniques of Tablets Manufactured by Direct Compression. AAPS PharmSciTech 2012, 13, 1138–1146. [CrossRef]
16. Jia, R.; Mao, Z.; Wang, F.; He, D. Self-tuning final product quality control of batch processes using kernel latent variable model.
Chem. Eng. Res. Des. 2015, 94, 119–130. [CrossRef]
17. Westerhuis, J.A.; Coenegracht, P.M.J. Multivariate modelling of the pharmaceutical two-step process of wet granulation and
tableting with multiblock partial least squares. J. Chemom. 1997, 11, 379–392. [CrossRef]
18. Wang, Z.; Pan, Z.; He, D.; Shi, J.; Sun, S.; Hou, Y. Simulation Modeling of a Pharmaceutical Tablet Manufacturing Process via Wet
Granulation. Complexity 2019, 2019, 1–16. [CrossRef]
19. Kuentz, M.; Leuenberger, H. A new model for the hardness of a compacted particle system, applied to tablets of pharmaceutical
polymers. Powder Technol. 2000, 111, 145–153. [CrossRef]
20. Boukouvala, F.; Chaudhury, A.; Sen, M.; Zhou, R.; Mioduszewski, L.; Ierapetritou, M.G.; Ramachandran, R. Computer-Aided
Flowsheet Simulation of a Pharmaceutical Tablet Manufacturing Process Incorporating Wet Granulation. J. Pharm. Innov. 2013,
8, 11–27. [CrossRef]
21. Rogers, A.J.; Inamdar, C.; Ierapetritou, M.G. An Integrated Approach to Simulation of Pharmaceutical Processes for Solid Drug
Manufacture. Ind. Eng. Chem. Res. 2013, 53, 5128–5147. [CrossRef]
22. Baroutaji, A.; Bryan, K.; Sajjia, M.; Lenihan, S. Mechanics and Computational Modeling of Pharmaceutical Tabletting Process. In
Reference Module in Materials Science and Materials Engineering; Elsevier: Amsterdam, The Netherlands, 2017.
23. Lewis, R.W.; Gethin, D.T.; Yang, X.-S.; Rowe, R.C. A combined finite-discrete element method for simulating pharmaceutical
powder tableting. Int. J. Numer. Methods Eng. 2005, 62, 853–869. [CrossRef]
24. Cabiscol, R.; Finke, J.H.; Kwade, A. Calibration and interpretation of DEM parameters for simulations of cylindrical tablets with
multi-sphere approach. Powder Technol. 2018, 327, 232–245. [CrossRef]
25. Rantanen, J.; Khinast, J. The Future of Pharmaceutical Manufacturing Sciences. J. Pharm. Sci. 2015, 104, 3612–3638. [CrossRef]
26. Wu, C. DEM simulations of die filling during pharmaceutical tabletting. Particuology 2008, 6, 412–418. [CrossRef]
27. Guo, Y.; Kafui, K.D.; Wu, C.-Y.; Thornton, C.; Seville, J.P.K. A coupled DEM/CFD analysis of the effect of air on powder flow
during die filling. AIChE J. 2008, 55, 49–62. [CrossRef]
28. Bierwisch, C.; Kraft, T.; Riedel, H.; Moseler, M. Die filling optimization using three-dimensional discrete element modeling.
Powder Technol. 2009, 196, 169–179. [CrossRef]
29. Gopireddy, S.R.; Hildebrandt, C.; Urbanetz, N.A. Numerical simulation of powder flow in a pharmaceutical tablet press lab-scale
gravity feeder. Powder Technol. 2016, 302, 309–327. [CrossRef]
30. Ketterhagen, W.R. Simulation of powder flow in a lab-scale tablet press feed frame: Effects of design and operating parameters
on measures of tablet quality. Powder Technol. 2015, 275, 361–374. [CrossRef]
31. Siegmann, E.; Forgber, T.; Toson, P.; Martinetz, M.C.; Kureck, H.; Brinz, T.; Manz, S.; Grass, T.; Khinast, J. Powder flow and mixing
in different tablet press feed frames. Adv. Powder Technol. 2020, 31, 770–781. [CrossRef]
32. Mateo-Ortiz, D.; Méndez, R. Microdynamic analysis of particle flow in a confined space using DEM: The feed frame case. Adv.
Powder Technol. 2016, 27, 1597–1606. [CrossRef]
33. Cabiscol, R.; Finke, J.H.; Kwade, A. Assessment of particle rearrangement and anisotropy in high-load tableting with a DEM-based
elasto-plastic cohesive model. Granul. Matter 2019, 21, 98. [CrossRef]
34. Giannis, K.; Schilde, C.; Finke, J.H.; Kwade, A.; Celigueta, M.A.; Taghizadeh, K.; Luding, S. Stress based multi-contact model for
discrete-element simulations. Granul. Matter 2021, 23, 1–14. [CrossRef]
219
Pharmaceutics 2021, 13, 996
35. Macal, C.M.; North, M.J. Agent-Based Modeling and Simulation. In Proceedings of the 2009 Winter Simulation Conference (WSC),
Austin, TX, USA, 13–16 December 2009; pp. 86–98.
36. Macal, C.M. Everything you need to know about agent-based modelling and simulation. J. Simul. 2016, 10, 144–156. [CrossRef]
37. Schönemann, M.; Bockholt, H.; Thiede, S.; Kwade, A.; Herrmann, C. Multiscale simulation approach for production systems. Int.
J. Adv. Manuf. Technol. 2019, 102, 1373–1390. [CrossRef]
38. Celik, M. Overview of Compaction Data Analysis Techniques. Drug Dev. Ind. Pharm. 1992, 18, 767–810. [CrossRef]
39. Wünsch, I.; Finke, J.H.; John, E.; Juhnke, M.; Kwade, A. A Mathematical Approach to Consider Solid Compressibility in the
Compression of Pharmaceutical Powders. Pharmaceuticals 2019, 11, 121. [CrossRef]
40. Heckel, R.W. Density-Pressure Relationships in Powder Compaction. Trans. Metall. Soc. AIME 1961, 221, 671–675.
41. Kawakita, K.; Lüdde, K.-H. Some considerations on powder compression equations. Powder Technol. 1971, 4, 61–68. [CrossRef]
42. Gurnham, C.F.; Masson, H.J. Expression of Liquids from Fibrous Materials. Ind. Eng. Chem. 1946, 38, 1309–1315. [CrossRef]
43. Cooper, A.R.; Eaton, L.E. Compaction Behavior of Several Ceramic Powders. J. Am. Ceram. Soc. 1962, 45, 97–101. [CrossRef]
44. Busignies, V.; Mazel, V.; Diarra, H.; Tchoreloff, P. Prediction of the compressibility of complex mixtures of pharmaceutical powders.
Int. J. Pharm. 2012, 436, 862–868. [CrossRef]
45. Hirschberg, C.; Paul, S.; Rantanen, J.; Sun, C.C. A material-saving and robust approach for obtaining accurate out-of-die powder
compressibility. Powder Technol. 2020, 361, 903–909. [CrossRef]
46. Ryshkewitch, E. Compression Strength of Porous Sintered Alumina and Zirconia. J. Am. Ceram. Soc. 1953, 36, 65–68. [CrossRef]
47. Schomberg, A.K.; Kwade, A.; Finke, J.H. The Challenge of Die Filling in Rotary Presses—A Systematic Study of Material
Properties and Process Parameters. Pharmaceutics 2020, 12, 248. [CrossRef] [PubMed]
48. Fell, J.T.; Newton, J.M. Determination of Tablet Strength by the Diametral-Compression Test. J. Pharm. Sci. 1970, 59, 688–691.
[CrossRef]
49. Nicklasson, F.; Alderborn, G. Analysis of the Compression Mechanics of Pharmaceutical Agglomerates of Different Porosity and
Composition Using the Adams and Kawakita Equations. Pharm. Res. 2000, 17, 949–954. [CrossRef]
50. Puckhaber, D.; Eichler, S.; Kwade, A.; Finke, J.H. Impact of Particle and Equipment Properties on Residence Time Distribution of
Pharmaceutical Excipients in Rotary Tablet Presses. Pharmaceuticals 2020, 12, 283. [CrossRef] [PubMed]
220
pharmaceutics
Review
Reviewing the Impact of Powder Cohesion on Continuous
Direct Compression (CDC) Performance
Owen Jones-Salkey 1,2,† , Zoe Chu 1,2, *,† , Andrew Ingram 2 and Christopher R. K. Windows-Yule 2
Abstract: The pharmaceutical industry is undergoing a paradigm shift towards continuous pro-
cessing from batch, where continuous direct compression (CDC) is considered to offer the most
straightforward implementation amongst powder processes due to the relatively low number of
unit operations or handling steps. Due to the nature of continuous processing, the bulk properties
of the formulation will require sufficient flowability and tabletability in order to be processed and
transported effectively to and from each unit operation. Powder cohesion presents one of the greatest
obstacles to the CDC process as it inhibits powder flow. As a result, there have been many studies
investigating potential manners in which to overcome the effects of cohesion with, to date, little
consideration of how these controls may affect downstream unit operations. The aim of this literature
review is to explore and consolidate this literature, considering the impact of powder cohesion and
cohesion control measures on the three-unit operations of the CDC process (feeding, mixing, and
tabletting). This review will also cover the consequences of implementing such control measures
whilst highlighting subject matter which could be of value for future research to better understand
how to manage cohesive powders for CDC manufacture.
Keywords: continuous direct compression (CDC); powder cohesion; loss in weight (LIW) feeding;
Citation: Jones-Salkey, O.; Chu, Z.; blending; tabletting; continuous manufacture
Ingram, A.; Windows-Yule, C.R.K.
Reviewing the Impact of Powder
Cohesion on Continuous Direct
Compression (CDC) Performance. 1. Introduction
Pharmaceutics 2023, 15, 1587.
The pharmaceutical industry is currently undergoing a paradigm shift from batch to
https://doi.org/10.3390/
continuous processing in order to improve manufacturing efficiency [1]. Continuous direct
pharmaceutics15061587
compression (CDC) (Figure 1) is considered highly efficient because of the reduced number
Academic Editor: Colin Hare of unit operations involved [2,3]; however, there currently exists only a limited range of
formulations for which it is viable. Suitable formulations will generally have low drug
Received: 3 April 2023
Revised: 8 May 2023
load, a need for adequate flowability (as the material has to be transported through the
Accepted: 10 May 2023
system), and have sufficient tabletability, for the CDC process to work [4].
Published: 24 May 2023 There are several critical material attributes (CMAs) which impact the flow of pharma-
ceutical powders and hence the successful formulation of oral solid dosage forms, of which
cohesion is thought to be the most significant [5]. Powder cohesion refers to the affinity of
particles to adhere to each other. It manifests in the bulk as resistance to powder flow often
Copyright: © 2023 by the authors. accompanied by adhesive behaviour, which refers to the tendency of a material to bind to
Licensee MDPI, Basel, Switzerland. surfaces and other materials [6]. Factors which contribute to cohesion include:
This article is an open access article
• Liquid bridges (liquid bridges will not be discussed in this literature review as there
distributed under the terms and
conditions of the Creative Commons
already exist comprehensive reviews of this topic [7–9]). Our focus in the present
Attribution (CC BY) license (https://
work will also be on the effect of cohesion in CDC processes rather than the causes of
creativecommons.org/licenses/by/
cohesion) [10];
4.0/). • Van der Waals forces [11];
Figure 1. Schematic of the CDC process. Feeding: (where A, B, & X; are example labels of the
formulation’s constituents) powders are fed into the blending stage. Blending: Shows the constituents
entering and being blended by the agitator, before entering the hopper of the rotary tablet press.
Tabletting: (where F, PC, C, & E; corresponds to the Die Filling, Pre-Compression, Compression
and Ejection, respectively) the final formulation fills the die cavity, undergoes compression forces—
forming the tablet—before being ejected from the die.
222
Pharmaceutics 2023, 15, 1587
Active pharmaceutical ingredients (APIs) are typically the components of tablet for-
mulations that give rise to handling issues, due to their often very high cohesivity, low
bulk density, and highly aspherical geometries, in turn resulting in undesirable operating
performance for CDC [6,18]. Many APIs are needle-like and very fine (<10 μm) resulting
in a high contact area and cohesion (and thus poor flowability), whereas excipients are
typically coarser and have a more favourable (spherical) particle shape, leading to better
flowability. This difference in flow properties is perhaps unsurprising, as excipients are
inactive substances whose purpose is to aid the formulation process by supporting or
enhancing the stability and tabletability [19] of a pharmaceutical dosage form; as such,
cohesive excipients would be omitted during the experimental design stages [20] as they are
much easier to replace than APIs. Based on the above, it is clear that a high concentration
of API may heavily affect the processability of powders, rendering the CDC manufacturing
route non-viable [21].
In spite of the above, cohesion is not always a negative influence on CDC operations.
In some instances, moderate cohesion has been found to aid the proper functioning of
a given unit operation. For example, during powder feeding the addition of nano-sized
silica was shown to result in reduced adhesion to feeder surfaces and improved powder
flow as will be discussed later in this review [6,22]. However, in most cases of high
cohesivity, flow is inhibited [23,24], which is problematic for feeding [25], mixing [26] and
die filling [27,28]. Conversely, during the tabletting stage, powder cohesivity (specifically
interparticle adhesion) is generally desirable, inviting smaller particle sizes and higher
surface energy powders [29–36].
There are three main ways in which cohesion can be managed in tablet production:
1. Formulation modification, through the introduction of glidants or lubricants [37].
2. Operational process changes which are documented in many different studies for
each unit operation [38].
3. Granulation, which also manages cohesion as poor flowing powder fines are made
into coarser more uniform agglomerates. (This is beyond the scope of this literature
review as this unit operation is not a part of CDC processes [39]).
However, most literature does not consider how cohesion controls impact downstream
unit operations. As far as the authors are aware there are no literature reviews that encom-
pass the mitigation of the effects of cohesion and how those controls may affect certain unit
operations. Therefore, the primary aim of this literature review is to explore and consoli-
date current knowledge concerning the impact of powder cohesion and cohesion control
measures on all unit operations in the CDC process (feeding, mixing and compression),
discuss the consequences of cohesion control measures implemented, and establish what
further studies could be valuably conducted in the future.
2. Powder Feeding
2.1. Feeding Introduction
Powder feeding for CDC processes needs to be accurate and consistent for CDC
performance success. Feeding modes are commonly defined by the delivery and main-
tenance of material, in a given time frame. As such, there are two main categories for
feeders: volumetric and gravimetric, which supply material according to volume per unit
time, and mass per unit time, respectively [2,22,25,40,41]. Volumetric feeding is often
used to understand the characteristics of powder behaviour within the hopper, whereas
gravimetric feeding seeks to smooth out any volumetric perturbation through feedback
control, arguably resulting in gravimetric being the superior of the two modes for mass
delivery [2,22,25,40–42]. Powder feeding is the first unit of operation in the CDC process;
thus disturbance or variation from the set point can propagate downstream, fundamentally
impacting the tablet quality [25,43].
One of the tasks of the blending stage in ensuring that constituents are well-mixed is to
smooth out the perturbations from the feeder. The lower the intensity and frequency of these
perturbations, the lower the chance that content uniformity or manufacturability issues
223
Pharmaceutics 2023, 15, 1587
will arise downstream [22,44,45]. Moreover, it should be noted that not all disturbances
result in a detrimental impact. It is both dependant on the magnitude and duration of such
disturbances—Gyürkés et al. [46] aptly discuss this concept with the use of funnel plots.
Powder cohesion impacts feeding performance by introducing variation through
complex bulk behaviour, which results in the variation of flow behaviour [22,25]. This
contributes to both the restriction of power into the converting screws, and the resistance
to flow within the swept volume of the screws. In addition, these materials are prone to
electrostatic charging, causing attraction-repellent behaviour, which can lead to deposits of
material residing and adhering to surfaces post-feeding [15,47]. This buildup can migrate
into the bulk feeder, dosing it with a spike of material, therefore resulting in a compositional
imbalance [15,47]. The following section aims to discuss the work of several publications,
exploring the complexity surrounding powder feeding.
224
Pharmaceutics 2023, 15, 1587
Tran et al. studied the differences in silica loading for two different excipients using flow
characterisation, static image analysis, and particle size distribution. It was found that
materials with higher surface area (described as rugged) were able to accommodate higher
silica loads, and that the optimal amount of silica loading is dependent on both the surface
area and the size of the carrier particle but, most notably, their results discussed the possi-
bility of taking a ‘quality by design’ approach by determining silicate (glidant) loading by
carefully assessing the carrier powder’s physical properties.
Figure 2. (Left): Graphs highlighting the difference in characteristics (Flow Function (FFC), Cohe-
sion, and angle of internal friction) of the two excipients: Hypromellose and Dicalcium Phosphate
Anhydrous with and without colloidal silica. (Right): Scanning Electron Microscope (SEM) of the
two excipients mixed with 1% w/w colloidal silica. Adapted from Leung et al. [5].
Lopez et al. [55] published a DEM study on the effect of increasing cohesion on
feeding performance. The researchers began by calibrating the powder used in the system,
first identifying and assigning the physical properties and particle shape of paracetamol,
before performing a sensitivity analysis across different adhesive stiffness values (Kadh )
and conveying impeller rotation rates. When the conveying impeller was kept at a constant
rotation rate (10 rad/s), the particles saw an increase in translational velocity in conveying
barrel with increasing Kadh (ranging from 0 to 0.5). This effect was lesser but present on the
interface between the conveying barrel and hopper. On the other hand, keeping Kadh at a
225
Pharmaceutics 2023, 15, 1587
constant 0.2, across increasing conveying speeds (10, 20, and 50 rad/s) saw an increase in
the overall translational velocity of the particles. Perhaps unsurprisingly, for the range of
parameters explored, increasing conveying speed was observed to have a greater influence
on particle velocity in the hopper than altering cohesion.
The DEM simulation of Lopez et al. [55] also provides valuable insight into the
expected power draw; given the CDC system is continuous, it would mean the process
would be running for days, if not weeks, at a time, making this an important consideration.
The authors demonstrated the average mass discharged per Watt as a function of the Kadh
(between 0.1 to 0.5 Kadh ), which showed an exponential decay with increasing Kadh Figure 3.
Accordingly, Kadh values higher than 0.5 showed arching: a phenomenon which occurs
when the cohesive forces between particles are stronger than forces due to gravity. Arching
(also referred to as bridging) is a semi-stable instance of particles forming particle–particle
structures suspended in a hopper or opening, which are resistant to the displacement
of powder beneath them [25]. For additional information on the influence of particle
properties on arching, see the following papers: [25,49,56]. Lopez et al. described this
behaviour with absolute translational velocity where they revealed instances which lead
to inconsistent feeding of the conveying screw, which in turn lead to flow irregularities
Figure 3. The authors link this artefact of transient macro behaviour to the micro by stating
that inconsistencies in flow result in the improper filling of the screw pitch, giving rise to an
inconsistent mass flow rate. This is best demonstrated by the plot of mass per screw pitch
vs. Kadh shown in Figure 3). Finally, the researchers indicated what future work would be
of value, notably suggesting the mapping of the transient cohesive behaviour by altering
the particle properties, feeder’s geometry and rotational speed.
To build upon the above suggestions, it would be interesting to see the implementation
of a DEM study on a more commercially representative feeder, which possesses both a
hopper agitator and twin screw conveying elements. Allowing the exploration of concepts
such as agitator design and rotation speed to increase barrel/screw filling consistency or the
breakage/mitigation of arching. What is more, the research conducted by Leung et al. [5],
which discussed the minimal effect of friction in powder flow consistency, could utilise
simulation tools such as DEM to more directly support their claims.
Escotet-Espinoza et al. [6] present a publication showcasing the use of silication to
improve feeder performance. The researchers began by pre-blending the three different
APIs explored with 1% silica, before characterising the bulk behaviour of the pre-and post-
silicated API. The APIs (and their silicated counterparts) were then fed through the feeder,
whilst a catch scale continuously measured the mass exiting the feeder. The response was
then measured over time, with attention being paid to the consistency and accuracy of the
feeding process.
Like Leung et al. [5], Escotet-Espinoza et al. [6] also found that the introduction of
silica improved flowability by reducing the interparticle cohesion. Feeding improved with
the addition of silica, with each of the API case studies showing improved screw speed con-
sistency, a reduction of mass flow RSD (relative standard deviation), a reduction in powder
adhesion to the hopper surfaces, and a reduction of remaining mass. This suggests that the
addition of silica is hugely advantageous for feeding. The improvement can potentially be
attributed to an increase in flowability, allowing the powder to better fill the swept volume
of the screws, thus leading to an increase in mass per revolution with the addition of silica.
This observation is supported by the work of Lopez et al. [55] which shows that, with a
reduction of Kadh , there follows an increase in mass per screw pitch. Furthermore, with FFC
(as previously discussed) being a strong indicator for powder cohesion, it would suggest
that it would be useful to understand this screw-filling behaviour. This conveniently leads
to another suggestion to further Lopez et al’s work [55], whereby similar DEM studies could
be undertaken to understand the effect of additives on feeding and powder conveying.
Escotet-Espinoza et al. [6] also describe the material used generating electrostatic charge,
which enabled the API to stick to the agitator. However, in the silicated API case study,
226
Pharmaceutics 2023, 15, 1587
there was very low powder adhesion, suggesting that adding silica also suppressed the
effects of electrostatic adhesion.
Figure 3. (A): Mass per screw pitch as a function of cohesive stiffness (Kadh ). (B): Mass per Watt as
a function of cohesive stiffness (Kadh ). (D): Visualisation of the powder arching/bridging within
the hopper volume. The images show both the back and front of the hopper from two different
simulations using the same Kadh . (C): Irregular filling of the conveying volume due to cohesive
stiffness (Kadh ). The images compare Kadh = 0.3 and Kadh = 0.5. Adapted from Lopez et al. [55].
A study by Lumay et al. [57] explicitly investigated the influence and mechanisms
of mesoporous silica (MPS) on the electrostatic charge, again finding that the addition
of silica species improves powder flow. The experiments differed from those discussed
previously in several manners. Firstly, the study differed in the materials and methods
used—the authors tested three different grades of silica, varying in particle and pore size,
in three common excipients (microcrystalline cellulose, lactose, and maize starch) using a
rotating drum. Secondly, the authors applied different experimental analyses: the dynamic
effect of silication was evaluated in this case by measuring the cohesive index [58] of
materials (with and without MPS) across a range of rotation speeds. Finally, through the
use of a standardised charge density measurement technique [59], the charge density of the
material blends is shown in the presence of different types and amounts of MPS.
It was found that the MPS with the smallest particle size provided the greatest im-
provement in flowability. Furthermore, the addition of just 0.5% w/w made a considerable
impact on the cohesive index—see Figure 4A. The authors attribute this improvement to the
particle size of the MPS, explaining that there would be an increase in effective surface area,
which would suggest greater coverage, and thus intervention between, cohesive species.
An additional striking finding of the study was that the addition of silica was also
found to change the shear-dependent rheological properties of certain tested materials,
227
Pharmaceutics 2023, 15, 1587
with typically rheopectic materials becoming thixotropic when mixed with a small volume
of silica. Specifically, the study evaluated the cohesive index of the powder dynamically by
testing the excipients against their 2% w/w silica counterparts across a range of rotation
rates (see Figure 4B). Both materials, maize and lactose, were shown to be more cohesive
(thickening) with increasing rotating speed (ranging from 2 rpm to 10 rpm) and in the
presence of silica, less cohesive (thinning) with increasing rotating speed. The silicated
maize showed a greater decrease in cohesion than the silicated lactose with increasing
rotation speed.
Figure 4. (A) Comparing the differences in the dynamic cohesive index (σ) with the addition of
different types and weight percentages of Mesoporous Silicates (MPS). (B) Dynamic influence of
MPS on the cohesive index, demonstrated through plotting the cohesive index over different rotating
speeds. (C) Comparison of Charge Density (nC/g) in two excipients and resultant performance
with the addition of 2% weight silicates. Where: +S244 is Syloid® 244FP, +SAL1 is Syloid® AL–1FP,
and +SXDP is Syloid® XDP3050. Adapted from Lumay et al. [57].
Finally, the inclusion of MPS was shown to decrease the static charge density of
the powder—see Figure 4C. Across all materials and MPS grades, the addition of 1% (of
MPS) saw similar results to the addition of 2—suggesting 1% was sufficient for charge
mitigation, with any greater amount being surplus. MPS incorporated into maize saw a
decrease in charge after just 0.5% w/w, with the grade of MPS used making a minimal
difference. Conversely, MPS in lactose still produced a decrease in the magnitude of charge
density, though different grades responded differently. This suggests there may be some
complexity associated with the use of MPS with lactose. It is potentially important to note
that Lumay et al. [57] stated that the lactose varied in positive or negative charge on the
day, which was dependent on the air’s relative humidity. Given the magnitude is roughly
the same but the charge differs, if the S244 (Syloid® FP244) was positive it would suggest
that the grade of MPS used would have little effect on the charge density and vice versa for
SXDP (Syloid® XDP3050). Research by Ramires-Dorronsoro et al. [60] supports this: when
completing a similar study—with a different measurement technique—they also witnessed
spray-dried lactose exhibiting a positive charge density. Ultimately, this suggests that
the use of silica in poorly flowing hygroscopic powders greatly improves the flowability,
and decreases the sensitivity to triboelectric charging.
The previously discussed papers showcase the underpinning behaviour of cohesive
powders during feeding, and the discussion of silication demonstrates the advantage of
understanding the mechanism underlying a powder’s cohesive properties. Despite this,
when considering CDC in its entirety, the predictive capability is a large part of reducing
experimentation and understanding the process. Therefore it is desirable to be able to
predict the feeding performance from a few key powder characteristics [42,43,61], one of
which is cohesion.
228
Pharmaceutics 2023, 15, 1587
Van Snick et al. [43] and Garcia-Munoz et al. [45] provided some of the first in-depth
analysis of the end-to-end CDC process, thus by definition including aspects of feeding.
The authors discuss the correlation of bulk density, tapped bulk density and FFC to feeding
consistency (Figure 5). The experiments involved the comparison of different materials
running volumetrically (at a constant screw speed) from full to empty, whilst measuring
the feed factor response over time. The maximum ( f f max ) and minimum ( f f min ) feed
factors (g/revolution) were then taken and used for further calculation. Feed factors change
throughout the hopper fill as powder above the hopper bears down upon the powder below,
due to gravity. When the hopper is at low fill, there is not only a reduction of gravitational
bulk powder compaction but a potentially inconsistent feed into the conveying volume.
Figure 5. (Top): Two graphs detailing the feed factor throughout the volumetric empty, with abso-
lute values (left) and normalised values (right), the key indicates the material used. Where: FF316
is Lactose Fast Flo® 316, SD711 is Sodium Croscarmellose Ac-Di-Sol® SLS is Sodium Lauryl Sul-
phate, PH102 is MCC Avicel® PH-102, APAP is Acetaminophen, and MgSt is Magnesium Sterate.
(Bottom): Three graphs compare feeder performance to bulk characteristics. Adapted from Van
Snick et al. [43].
As a result, bulk density increased linearly with max feed factor, tapped bulk density
increased linearly with delta feed factor (where delta feed factor is: f f Δ = f f max − f f min ),
and finally flow rate variability exponentially decayed with increasing FFC. Since both
bulk density and FFC are common descriptors for powder cohesion, it then can be used
as a proxy to describe behaviour. The study highlighted FFC values greater than 3 would
give sub 1% flow rate variability, and that powders with low bulk density and low tapped
bulk density are prone to large f f Δ —which suggests instability when processing. Escotet-
Espinosa [22,62] also looked into both the influence of powder properties on feed factor
and modelling such behaviour using the following (simplified) equation:
where f f w is the feed factor as a function of w the weight, while f f max , f f min and β are
the maximum feed factor, minimum feed factor, and a fitting constant based on feeder
geometry, respectively.
The f f max and f f min were then plotted in a correlation heat map (Figure 6), relating
the responses to a plethora of recorded bulk characteristics. Interestingly, the highest
correlation was seen for cBD, conditioned bulk density (g/mL). cBD is akin to tapped bulk
density, obtained from the initial standardised conditioning step of a powder rheometer. It
involves a blade passing through a volume of powder with a set torque before measuring
229
Pharmaceutics 2023, 15, 1587
the weight of the known volume. Paying attention to the units, it is a given that there is a
correlation between the two density measurements, but the interest is with ‘compressibility
at 15 kPa’ which also scored high for f f min , which would be a method for attaining
higher bulk density. This suggests that powder that is capable of being compressed,
and has a high conditioned bulk density, will deliver a high f f min . Similar studies by
Shier et al. [63] and Bekaert et al. [42] similarly look to predict feeder performance using
bulk characteristic properties.
Figure 6. A correlation heat map acquired, with permission, from Escotet-Espinosa [62], detailing
the magnitude of linear correlation (R) between bulk powder characteristics and regressed feed
factor parameters.
Engisch & Muzzio [41] performed a study on varying hopper refilling conditions on
powders with varying cohesiveness and measured the resultant effect on feeding accuracy
and consistency. Typical Loss in Weight (LIW) feeder operation is gravimetric, until top-
ping up (or refilling), where the system operates volumetrically to ensure a consistent
feed is maintained whilst an influx of mass is registered on the feeder’s load cell. This
method results in a trade-off between the number of times the system has to enter the ‘less
accurate’ volumetric mode and ensuring the hopper is sufficiently full, such that the feed
factor is stable. For an example of the effect of decreasing fill on feed factor, see Figure 5.
The conclusion from Engisch & Muzzio is that refilling the feeders with less material,
and therefore more frequently, results in better overall performance. This should be kept
in mind when paying attention to the duration of time that the system enters volumetric
mode and reducing the overall disturbance from the setpoint over the whole duration of
running the feeder. Irrespective of the frequency and volume of fill, Engisch & Muzzio
state refilling should be gentle as this causes the least disruption to feeding. Furthermore,
it is shown that refilling sensitivity is much lower at higher refill levels, Figure 7. Refilling
at higher fill levels reduces the maximum set point deviation, time of deviation, and the
total amount of mass during deviation.
230
Pharmaceutics 2023, 15, 1587
Figure 7. (A–C) Shows the feeder performance for just semi-fine APAP & 0.25% silica over 3 refill
levels, each with 10 manual refills. (A) Maximum deviation from setpoint (kg/h), (B) Time of
deviation (s), and (C) Total deviation (kg). (D–F) Similarly, shows the comparison between zinc oxide
powder and semi-fine APAP & 0.25% silica blend. (D) Maximum deviation from setpoint (kg/h),
(E) Time of deviation (s), and (F) Total deviation (kg). The deviation is defined as the sum of the
surplus powder delivered during refill. Adapted from Engisch & Muzzio [41].
231
Pharmaceutics 2023, 15, 1587
Finally, Yadav et al. conducted a comprehensive study with a loss in weight feeder [64]
utilising a range of screw speeds, gear ratios, and screw types, across a range of materials.
The result was a PCA model which identified the influence of these processing parameters
and how they interacted with the materials’ bulk characteristics. The principal components
consisted of one comprising the powder’s bulk characteristics, a second representing the
feeder’s processing parameters, and a third representing an interaction component. Most
notably, the interaction component related screw free volume (cm3 ) to density, particle
size (d10 , d50 , d90 ), wall friction angle, and Feed Factor. The authors note that a different
screw type will have a different screw-free volume, meaning that the comparison between
a smaller or larger volume to dispense per revolution must be taken into consideration
when evaluating the feed factor. Similar findings are seen with Engisch et al. [2], where
their suggestion is the use of both self-cleaning concave screws and the removal of any
outlet screens for very cohesive materials.
232
Pharmaceutics 2023, 15, 1587
develop a control proposition for shear thickening powders involving a sudden reduction
in RPM followed by a slow build-up of screw speed back to setpoint.
3. Blending
3.1. Blending Introduction
Despite diverse designs available commercially and in the academic literature, con-
tinuous blenders can, in general, be reduced to four main components: an inlet, an outlet,
a mixing volume, and a mechanism for promoting blending whilst in the volume of the
blender. Typically, this last component will be some form of rotating agitator, the geometry
of which may differ from mixer to mixer, or even formulation to formulation. Recognising
the different types and geometries of mixers that exist within this space is fundamental
when evaluating the role of powder mixing in CDC performance, as the features that
differentiate one mixer from another can have significant impacts on the behaviour of the
powders being blended, and thus may require different measures to handle the blending of
cohesive powders.
Achieving good mixing is integral to the performance of any given pharmaceutical
tablet; the therapeutic efficacy and mechanical properties of the tablet will suffer without
the effective distribution of its constituent materials, ultimately compromising the safety of
the product [43,71–75]. Thus, mixing inherently defines the final quality of the formulation
(albeit that the final product cannot be delivered without the consistency and accuracy of
the upstream and downstream processes [43,45]).
Practically defining whether a formulation is well-mixed is a complex endeavour with
issues such as frequency of sampling [71,76,77], location of sampling [71,76,77], and scrutiny
(size) of the sample [71,76,77]. However, describing the mechanisms of mixing may be
ever-so-slightly easier, and is of more direct relevance to the present work.
The two commonly-defined types of mixing—macro- and micro-mixing—are both es-
sential when looking to create ideal mixtures [71,72,75,78,79]. Macro-mixing is responsible
for large, global movements, of material—or bulk transport—allowing the material to flow
and circulate around the vessel [71,75,79–82]. Micro-mixing, on the other hand, involves
the movement and displacement of material at the particle scale, i.e., the local interactions
between particle species [71,79]. Accordingly, micro-mixing achieves the detail of the mix-
ing action due to the mixing happening at the smallest possible scale: the particle–particle
level [71,79], whereas macro-mixing serves to smooth out the upstream mass flow rate
perturbations, due to the flow and bulk motion of the powder [71,79].
Fan et al. [83], succinctly discuss the evaluation of mixing in terms of cohesive mixtures.
Figure 8 and Equation (2) encapsulate the majority of this discussion. It is shown that by
increasing the surface force—which is also a function of particle size, and therefore surface
area—it is possible to reduce the influence factor defined in Equation (2). A low influence
factor means surface forces dominate, and the particle species adhere to the surface of
another particle species, creating a perfectly-ordered blend. Small, less dense particles
would be the ones to adhere to the more dense ‘carrier’ particles- due to gravitational forces.
Then, if the surface forces are minimal, gravity dominates, and thus the mixture would
involve non-cohesive species, following more typical solids mixing criteria.
Gravitational Force
In f luence Factor = (2)
Sur f ace Force
233
Pharmaceutics 2023, 15, 1587
be worked into powder, through the triboelectric effect, creating attractive or repulsive
forces between different species [26,77]. These interactions underpin much of the following
discussion of how cohesion affects the process of mixing and the competition between
mixing and de-mixing.
Figure 8. Cohesive mixtures described by homogeneity vs influence factor (Equation (2)). Adapted
from Fan et al. [83].
234
Pharmaceutics 2023, 15, 1587
at (+30◦ ) incline there was an increase in mean residence time which led to a decrease in the
relative standard deviation (RSD) of Acetaminophen (APAP) concentration in two grades
of lactose. This demonstrates the effect of longer mean residence time on macro mixing and
a reduction in APAP RSD. The grades of lactose differed in terms of average particle size
and size range: Lactose 100 (70–250 μm, average 130 μm) and Lactose 125 (55 μm), whereas
APAP (36 μm) remained the same in both tests, and made up 3% w/w of the mass flow rate.
They also discussed that, despite typical behaviour, the smaller (and more cohesive) grade
of lactose did not affect mixing performance at either the high or the low rotation speeds. It
was observed that while agglomerates were readily formed by the cohesive material, these
were relatively weak and agitation from the impeller was sufficient in breaking the clusters
up, dispersing the powder.
Figure 9. Example linear blender used in Portillo et al. [91]. Profile view of the blender (A) view of
the agitator shaft and blade angle orientation (B).
Later work by the same group [78] discussed the influence of cohesivity on powder
behaviour in a different continuous mixer (albeit similar to the blender shown in Figure 9),
inclined at +17º. The study used two different grades of lactose, Fast Flo and Edible, which
displayed flow indices (measured using a Gravitational Displacement Rheometer) of 24.9
and 34.8 respectively, suggesting them both to be highly flowable powders. In addition,
the bulk density tapped density of the Fast Flo Lactose were 0.626 (g/mL) and 0.704 (g/mL),
respectively, while the Edible Lactose measured 0.629 (g/mL) and 0.981 (g/mL), respec-
tively. Portillo et al. [78] found that cohesivity negatively impacted axial mixing, becoming
a statistically significant factor at higher levels of agitation; thus demonstrating cohesion’s
effect in resisting the separation or relative movement of particles. In addition, the authors
observed that particles with high cohesion experience longer residence times, with similar
path lengths, suggesting the particles move at a slower velocity through the blender. This
postulates that methods which increase the shear rate, and therefore the relative particle
velocity, may combat the effects of cohesion by increasing dispersion [92], and consequently
improve micro-mixing. This could also be described as increasing the dynamic granular
temperature as a function of the increased shear rate [93,94]. This finding may, at face
value, seem to contradict the findings discussed in the previously mentioned study by
Lopez et al. [55], in which the translational velocity of the powder going through the convey-
ing portion of the system was observed to be higher with increasing cohesivity. However,
despite the net velocity being higher, the relative velocity (which drives micro-mixing) may
nonetheless be lower.
It should be noted that any residence time distribution (RTD) data will be a func-
tion of both the volume % of the tracer and the interaction of the tracer within the bulk
that it aims to measure. For more detail on RTD behaviour in continuous blenders see
235
Pharmaceutics 2023, 15, 1587
Escotet-Espinoza et al. [61,80], where the authors provide a comprehensive two-part study,
analysing the behaviour of tracers with different properties.
Vanarase et al. [89] investigated the blending performance of a horizontal linear
blender, both with and without the addition of a co-mill. The authors stated that under op-
timal macro-mixing conditions, micro-mixing was identified as the limiting factor, leading
to poor mixing performance. The inverse was also found for optimum micro-mixing condi-
tions. This highlights the independence of these two types of mixing, and the consequent
requirement, to balance a blending system’s capability to deliver both sufficient micro and
macro mixing.
Interestingly, research by Portillo et al. [91] contains contradictory findings, when
compared to Vanarase et al.’s study [89]. They found that using an incline blender increases
the overall residence time when compared to a horizontal system—see Figure 10. Thus,
for the same residence time as was demonstrated by Vanarase et al. and therefore sim-
ilar macro behaviour, Portillo et al. showed that the agitator speed could be increased,
which in turn provides greater shear, achieving improved micro-mixing conditions with
cohesive powders.
Figure 10. Acetaminophen residence time distribution plots (gathered with permission) from Por-
tillo et al. [91], demonstrating the effect of blender volume inclination and RPM on RTD; (a) 78 RPM,
(b) 16 RPM.
Van Snick et al. [43] investigated the performance of different mixing blade configura-
tions (Figure 11, graphically demonstrates the orientation of blades along the agitator’s axis)
and found that despite the ‘P16’ blade configuration not delivering the highest dispersion,
it offered the best consistency (see Figure 12). The ‘P16’ blade config’ utilised a series of
two sets of transport blades (45º), followed by a set of radial mixing blades (90º). Whereas
236
Pharmaceutics 2023, 15, 1587
the other example—‘D8’—delivered the highest powder dispersion, with the downside of
higher variability. This is supported by Vanarase et al.’s [89] findings, with both surmising
that a balance is required between micro and macro mixing. Based on ‘P16’s performance,
in Van Snick et al. [43], it seems as if the ‘D8’ configuration provides a surplus of dispersion-
lending to poor macro conditions. The authors also attribute the good performance of ‘P16’
to the increased mass hold-up within the blender’s volume, as the mixing blades are shown
to increase fill level, and therefore total mass hold-up [43]. Ultimately, the study suggests
that varying position, number, and angle of the blades could aid in controlling API within
the blending phase, and that different blade configurations may be selected according to
the formulations’ properties, such that the desired throughput is attained, whilst output
variation is minimised.
Figure 11. Blade configurations refer to a series of different blade orientations along the agitator’s axis.
(Top left): side-by-side comparison of transport blade and radial mixing blade (RMB). (Top right): ex-
ample of a blade combination that would take a position along the agitator’s axis, denoted by a
number in the configurations D8 and P16 below. D8 and P16 depict blade configurations; all positions
that are not labelled as RMBs are transport blades, and each collar position sees the blade combination
rotate 60◦ clockwise to the direction of the impeller. Adapted from Van Snick et al. [43].
237
Pharmaceutics 2023, 15, 1587
Figure 12. Tablet uniformity corresponding to the blade configurations seen in Figure 11 (D8 and
P16); Upper graphs: content uniformity as a function of impeller speed and mass rate (green
triangle = 24 kg/h, blue square = 30 kg/h, red triangle = 36 kg/h). Lower graphs: relative density
distribution based on the target API dosage at the centre point of the experimental design (250 rpm
and 30 kg/h). Gathered with permission from Van Snick et al. [43].
Figure 13. Periodic section of a continuous blender including two blades (a) Isometric View, (b) Side
profile view with dimensions. Gathered with permission from Gao et al. [95].
It is worth noting that, for all non-cohesive powders used in Gao et al.’s work, when
going from high to moderate fill levels (i.e., from 75% to 40%) at 100 rpm, normalised axial
velocity increases. Cohesion is still shown to have an effect at higher rpms, demonstrated
by the increased distance between the contours. This finding corroborates with the rpm-
cohesion-significance result from Portillo et al. [78], clearly showing that the relative effect
of cohesion will depend on operating conditions (speed and fill/mass rate), and not
necessarily in a predictable way, reinforcing the need for more work in this area. This
finding, surrounding the three non-cohesive powders, was further supported by the work
238
Pharmaceutics 2023, 15, 1587
of Sarkar and Wassgren [96], where the normalised axial velocity presented the same trend,
despite different geometry, particle properties, and operating conditions.
Gao et al. [95] also mapped the continuous blending rate (k c ), which is the ratio of the
normalised mean axial velocity to the time-dependent RSD decay rate, where k c essentially
details the ratio between the axial (macro) and radial (micro) behaviour of the system.
The results obtained were a similar contour plot to Figure 14, but in terms of k c . Across
the different particle properties, it was shown that high RPM and low fill produced the
greatest k c . Moreover, there was little difference seen in k c at <125 rpm irrespective of
fill level. For rpm > 125: going from high-to-low fill showed a sluggish increase in k c ,
as the space between contours and colour changes progressed slowly. This highlights that
cohesion does not only affect the axial aspect of the mixing process, but also contributes
to the hampering of radial mixing. Finally, Gao et al. [95] suggest further investigation on
segregating materials, which would prompt several questions on a constituent’s experience
within the bi-or-polydisperse system. Particularly, it is worth considering what might
be the best excipients to pair with different levels of cohesive materials, in order to best
aid their macro- and micro-mixing. This has the potential to result in a CDC formulation,
of the same drug, differing from their original batch formulation. Different excipient(s)
may offer a significantly improved CDC performance, without impacting the tablet’s
therapeutic efficacy.
Figure 14. Contour plots of normalized mean axial velocity: (a) control, (b) lower density, (c) larger
particle size, and (d) cohesive particles. Gathered with permission from Gao et al. [95].
Tomita et al. [97] utilised an atypical blending set-up, operating with both an impeller
and a scraper in a horizontal mixer (see Figure 15). The inclusion of the scraper allowed for
the control of the mean residence time in a manner that—unlike the previously-discussed
systems—was quasi-independent of the magnitude of the impeller’s agitation. In this
system, the rpm of the axial impeller was driven separately to the ribbon-shaped scraper,
which circumscribed the internal wall of the blender. Figure 15B shows the RTD curves of
spiked acetaminophen using different scraper rpm speeds, whilst running at a feed rate of
10 kg/s and constant axial agitator speed of 3000 rpm. There is a considerable difference in
the RTD performance between the slowest (5 rpm), and highest (50 rpm) scraper speed,
which delivered a mean residence time of 84.3 s and 12.1 s, respectively. The mixer provides
239
Pharmaceutics 2023, 15, 1587
Figure 15. (A) Impeller and scraper continuous mixer with arrows detailing rotation direction,
(B) APAP RTD spike response. Adapted from Tomita et al. [97].
Table 1. Material properties of the APAP grades used in Palmer et al. [75].
240
Pharmaceutics 2023, 15, 1587
Figure 16. Tablet assay RSD plotted against Strain (see Equation (3)), an exponential decay model is
fitted to the data. Gathered with permission from Palmer et al. [75].
Figure 17. Peclet Number (Pe) and Strain relationship plotted on a log–log graph, grouped by
throughput. Gathered with permission from Palmer et al. [75].
Figure 18. Contour plots mapping the impeller speed, number of mixing blades, and fill level at
a throughput of 26.25 kg/h (a) Micronized APAP, (b) powder APAP, (c) special granular APAP.
Gathered with permission from Palmer et al. [75].
Comparing the bulk characteristics in Table 1, Micronized APAP (mAPAP) has closer
bulk characteristics to Powdered APAP (pAPAP), but shows a unique shape contour
in Figure 18. Despite pAPAP and Special Granular APAP (sgAPAP) being opposed in
flowability indices, they present similar contour shapes. This indicates a similar fill-level,
and therefore bulk powder response, to the number of radial mixing blades (see Figure 11)
241
Pharmaceutics 2023, 15, 1587
across a range of RPMs. For the mAPAP, however, increasing the number of radial mixing
blades has more impact on the fill level at lower RPMs, and this impact reduces with
increasing RPM. Van Snick et al. [43], discussed the necessity of sufficient mass hold-up (fill
level) to attain good levels of macro mixing. The combination of these two studies suggests
(i) that it is critical to determine the fill level landscape for the desired formulation to achieve
and maintain sufficient mixing, and (ii) that different highly cohesive powders would be
expected to exhibit different complex behaviours. What is more, using the mAPAP contour
from Palmer et al. [75], considerations have to be made regarding what variables can be
controlled in real-time. Since the radial mixing blades cannot be changed in real-time, it
could be inferred from Figure 18, that using the highest number of mixing blades would be
more optimum, as it would allow the system to have the largest range of fill levels.
Furthermore, Figure 16, from Palmer et al. [75], models the exponential decay of
content uniformity RSD with increasing Strain (Strain, Equation (3)). Strain is a confounded
variable for this system (an incline (+15º) linear blender), as increasing the RPM decreases
the mean residence time, meaning that the number of blade passes also affects the mean
residence time [75,91]. Thus, strain is dependent on the throughput, RPM, and the pow-
der’s material properties, making it essential to understand how the powder responds
to processing parameters. Understanding in this area can be gleaned from the work of
Portillo et al. [78,91], Gao et al. [90], and Lopez et al. [55] regarding powder velocity, fill
level, and resultant changes in residence time due to cohesion. More cohesive powders
would be subject to longer residence times, but lower relative velocities. Therefore, more
cohesive powders should be more resistant to increasing RPM in turn reducing cohesion’s
effectiveness; Figure 18 demonstrates the added complexity. Ultimately, strain is a great,
yet elusive, measure of effective work done to the powder. The model shows there is an
optimum amount of strain required (around 1500) to attain a high-quality blend.
Lastly, Palmer et al. [75] shows a linear log–log relationship between Peclet Number
(Pe) and strain, see Figure 17. Pe is a dimensionless number describing the ratio between
advective and diffusive mixing in the system (see Equation (4)). To give an example,
a high Peclet number would be expected to produce a narrow RTD. The model, at higher
throughputs (and therefore, higher residence masses), showed reduced scattering and more
linear behaviour, implying that the model may be favoured in predicting systems with high
residence mass, whilst also implicitly suggesting that systems with low residence mass may
present unstable micro-mixing capability and therefore may be unfavourable. Despite this,
across all throughputs, it was seen that with reducing strain, the Peclet number increased.
where MRT is the mean residence time and ω is the agitator rotation rate.
advective transport uL
Peclet Number = Pe = = (4)
di f f usive transport D
where u is the flow velocity, L is the characteristic length and D is the mass diffusion coefficient.
centripetal f orce ω2 R
Froude Number = Fr = = (5)
gravitational f orce g
where ω is the blade rotation rate, and R is the geometrically relevant length-scale—i.e.,
the radial distance from the centre axis to the tip of a blade.
While, as outlined above, strain is considered to be beneficial to powder blending, it
has also been shown that powders which experience high levels of strain (number of blade
passes) are also more likely to be subject to triboelectric charging [13]. Naturally, the effect
will present some interest in how electrostatics influence/interact with powder cohesion.
Karner and Urbanetz [13] investigated the influence of pharmaceutical mixing in a batch
mixer and found that particle size, the fraction of fine particles, and mixing volume were
significantly impacting the charging behaviour of the powder, stating that small particles
242
Pharmaceutics 2023, 15, 1587
were more prone to generate higher charge density. How these findings would translate to
a continuous blender represents a potentially interesting and valuable line of inquiry.
Beretta et al. [15,47] performed a two-part study on the triboelectric effect of charging
powder via. powder transport in twin-screw feeding. This was followed by a study using
relative humidity (RH) to subdue the electrostatic effect. Given that it is typical to maintain
a RH between 30–60% in ISO-graded GMP environments [16] and that additional air
moisture can increase the presence of liquid bridging, promoting cohesion [57], it becomes
difficult to assess purely on the grounds of cohesion. The downside of controlling relative
humidity is that it becomes difficult and expensive to maintain HVAC systems, especially
as temperature and humidity fluctuate across the year. There are other options to consider,
such as silication, outlined by Lumay et al’s study [57], which was shown to be capable
of electrostatic dissipation and localised relative humidity control. Therefore, the use of
silicates becomes a much cheaper and simpler option, despite what relative humidity
may offer. However, this assumption is limited by current knowledge and awareness
of the author(s) at the time of composing this review—novel literature may render this
assumption moot.
Beretta et al. [15,47] stated that the charging forces are primarily recruited through
particle–particle friction, while particle–wall friction did not directly contribute to charging.
In addition, the authors made considerations for impacts on powder mixing; interest-
ingly, the discussion touched on utilising these static forces to aid the mixing process [47].
Huang et al. [70] provide evidence of this, as dry powder coating is seen in a high-shear
co-mill. Huang et al. [70] investigated the effect of dry powder coating of different APIs
and excipients and found that the coating of fine colloidal silica improved both bulk density
and FFC performance. SEM images displaying the silica coating are shown in Figure 19.
Therefore, charge- or cohesion-based coating could provide an avenue to attain what
Fan et al. [83] (see Figure 8) would describe as perfect ordered mixtures.
Figure 19. SEM images showing the surface coatings of (top): excipient (lactose) and (bottom): API
(Ascorbic Acid). (Left): Non-coated and (right): 1% w/w Aerosil® R972 colloidal silica coating.
Adapted from Huang et al. [70].
243
Pharmaceutics 2023, 15, 1587
244
Pharmaceutics 2023, 15, 1587
4. Tabletting
4.1. Tabletting Introduction
The tabletting unit operation consists of three main steps:
• Die filling;
• Compaction;
• Tablet ejection.
The tabletting unit operation starts with the die-filling stage which fills the die with a
consistent amount of powder mixture to achieve an adequate fill depth ready for the com-
paction step while also keeping good drug content uniformity [101,102]. The upper punch
then lowers into the die increasing the compressional pressure applied to loose powder in
the tabletting die; the powder consolidates as particles rearrange, filling void spaces and
finally, as the availability of void spaces reduces, particles deform elastically, and ultimately
plastically, before beginning to fragment [21] once the target pressure has been reached
the upper die lifts releasing the pressure off the compact. Finally, the lower punch lifts
the tablet to the top of the die ejecting the tablet. However, die filling and compaction are
strongly influenced by cohesion whereas tablet ejection is not [103]. Qu et al. [103] study
showed that additives such as silica which lowered cohesion did not lower ejection stresses
effectively. Therefore, as this literature review focuses on cohesion, this section will only
discuss die filling and compaction.
245
Pharmaceutics 2023, 15, 1587
die, promoting the movement of the powder into the die [106] which removes the
effect of air entrapment [38,107] (Figure 20).
Similar to Sections 2 and 3, to have a good die-filling performance, the powder will
need to have lower cohesive properties [3] These include larger particle size, spherical
particle shape powder with lower surface energy. (Lower surface energy will mean that
powder will find it more difficult to form strong interparticle bonds [103,108]), which all
contribute to improved flowability [27,38,107]. To have content uniformity, the segregation
must be kept to a minimum during the die-filling stage so as not to undo the work done
during the mixing stage [107,109]. This section will focus on the die fill performance where
the mass, segregation and fill depth are important to ensure the compression process will
be successful.
i)
ii)
Figure 20. (A) Schematic diagram of air flow mechanism in (i) gravitational die-filling mechanism
(ii) suction fill mechanism in a shoe and die system obtained from Mills and Sinka [38]. (B) Schematic
of a rotary die-filling system, a type of force feeder, typically found in a turret tabletting machine
obtained from Tang et al. [105].
246
Pharmaceutics 2023, 15, 1587
Figure 21. Graphs obtained from Zakhvatayeva et al. [101] showing the change of critical velocity
as a function of: average particle size (top left), specific energy (top right) and cohesion at 3 kPa
consolidation stress measured using the FT4 (bottom) of a variety of different materials.
The effect of paddle speed in forced feeders was investigated by Goh et al. [107] which
saw with a higher paddle speed there was an increase in die fill weight and reduced die
variation. However, it was seen in other papers that the paddle speed only had an effect on
the weight variability, not the fill weight itself [84] and only affected fair flowing materials
such as MCC [113]. Therefore, a higher paddle speed can allow for cohesive powders
247
Pharmaceutics 2023, 15, 1587
to ‘flow better’ as they can essentially move as a solid block. On the other hand, in the
study of Peeters et al. [114], results showed that the paddle speed affected the tensile
strength. However, for the formulation with just microcrystalline cellulose, the tensile
strength outcome was not affected by paddle speed. The only formulations that were
affected were with magnesium stearate that saw a decrease in tensile strength with a higher
paddle speed. Lubricants can have a negative effect on the tabletability if excessive shear
force and higher levels of mixing occur. This is often referred to as ‘over-lubrication’ in
the literature [115]. Therefore, the decrease in tensile strength is most likely due to these
lubrication effects which is promoted with a higher paddle speed [116]. To understand
the effects of lubrication in a forced feeder with higher paddle speeds, work should be
conducted comparing the tensile strength outcomes of formulations with and without
magnesium stearate. However, it should be noted that brittle materials are reported to not
be as susceptible to the effects of lubricants as much as plastic materials [117]. Therefore,
formulations with majority plastic powders will most likely be affected more heavily than
formulations with a majority of brittle powders.
Another method to increase the flowability is by introducing air into the system,
as flowing air gives a lubrication effect allowing particles to flow more easily with less
resistance, increasing critical velocity [27]. The air prevents the percolation of fine particles,
which reduces vertical segregation [118]. However, the presence of air can also create an
adverse pressure gradient which resists the motion of particles, and pressure built up can
further oppose the flow of the powder [27]. Suction filling has also seen positive effects with
overcoming poorly flowable powders [38,106,109,119] and helps with the air entrapment
which hinders powder flow as it creates a negative pressure gradient [38,107]. Furthermore,
suction fill is generally known to lower the risk of segregation and improve the packing
density [38]. This is seen in Figure 22 obtained from the study of Zakhvatayeva et al. [109].
The degree of segregation was measured using a sampling unit which separates the die
into five sections; the process was repeated twice. The degree of segregation was calculated
using Equation (6) where the segregation index indicates the level of uniformity in the
powder blend in the die. Although the conclusion in Zakhvatayeva et al.’s [109] study is
that suction filling improves die-filling efficiency and reduces segregation, in Figure 22
it is seen that gravitation filling with a fast die-filling velocity had a lower segregation
index compared to suction filling. However, the slow die-filling velocity may not be
representative of a real-life scenario in industry. Another interesting observation from
Figure 22 is that with acetylsalicylic acid concentration of 10 percent, this had the worst
overall segregation index across all cases which was explained as generally lower API
concentrations are known to have a higher risk of segregation [120].
n
ci − c t
Segregation Index = ∑ ct
(6)
i
where ci is the starting concentration of the blend and ct is the concentration of the anal-
ysed sample.
All these results show that there are promising ways to overcome the effect of particle
cohesion on die-filling processes; however, it is often concluded that these results are
highly dependent on the particles’ material properties [102,107]. Active pharmaceutical
ingredients (APIs) are new and complex molecules that will often have unknown material
properties [18]. Further studies should measure bulk properties and be able to understand
what process parameters need to be utilised during the die-filling stage. APIs are known to
be very cohesive and poorly flowing which gives the limitation that formulations for direct
compression must have a maximum of 30% drug content [21].
248
Pharmaceutics 2023, 15, 1587
Figure 22. Segregation index of three different blends of Acetylsalicylic acid (ASA) with mannitol
using gravity and suction filling where suction filling mostly has a lower segregation index compared
to gravity from Zakhvatayeva et al. [109] the fast scenario is where the die-filling velocity is 260 mm/s
and slow has a die-filling velocity of 70 mm/s.
A possible method to allow formulations with more than 30% drug content to undergo
direct compression is spherical agglomeration/crystallisation, which will improve both
processability and tabletability of the API. This is completed by changing the way the API
is crystallised by either modifying the solvent addition rate and/or cooling rate [3]. Spher-
ical agglomeration allows the API to crystallise spherically (Figure 23), which increases
flowability and therefore lowers cohesion [3]. There are many promising studies using
spherical agglomeration/crystallisation such as a study by Chen et al. [18] which saw an
increase in flowability and therefore die-filling performance. Furthermore, the tabletability
also improved (Figure 23) for the spherical agglomerated/crystallised produced using the
quasi-emulsion solvent diffusion (QESD) method. The tensile strength was considerably
higher compared to the ‘as received’ ferulic acid exhibits a needle-like elongated shape
(Figure 23f). While spherical agglomeration/crystallisation is beneficial this should not
be confused with the negative kind of agglomeration that occurs during secondary man-
ufacture that can have a poor effect on blending, feeding and mixing. It is important to
distinguish the two.
On the other hand, there are some disadvantages to spherical agglomeration/
crystallisation; for example, solvent selection is a difficult process and may be limiting due
to the growing concerns of using more green solvents in industry [121]. This will need to be
further researched and optimised in order to be used as a commercial technique [121,122].
Furthermore, the optimisation of process parameters required for the spherical agglom-
eration/crystallisation method is difficult and will need considerable work to scale up
the processes [123]. Therefore, although this technique will help produce many more
formulations suitable for direct compression for most drugs to take place, the preparation
times at this current state will probably outweigh the benefits of just granulating. However,
this technique is very promising and something future research should focus on.
249
Pharmaceutics 2023, 15, 1587
e f
Figure 23. (a–d): Scanning electron microscope images of ferulic acid (FA) particles (a) Quasi-
emulsion solvent diffusion (QESD) at 50× magnification, (b) QESD powder at 2000× magnification,
(c) as-received FA powder at 50× magnification, and (d) as-received FA powder at 2000× mag-
nification. (e,f): Figures to show improved tabletability and flowability with spherical agglomer-
ated/crystallised (QESD) ferulic acid obtained from Chen et al. [18].
250
Pharmaceutics 2023, 15, 1587
under strict humidity controls [100], this section will just be focussing on surface energy
and particle size and shape. These three particle properties contribute heavily to the
magnitude of cohesion [128]. Although cohesion effects for other process operations
in the tabletting manufacturing process are known to have a mostly detrimental effect
(see Sections 2 and 3), in contrast, cohesion properties are important to the compression as
cohesivity helps produce a strong solid compact. Therefore, it is important to understand
how the cohesion properties contribute to the tabletting stage so potential compromises
can be taken to help the formulation perform well throughout the manufacturing process.
Figure 24. Johansson and Alderborn [30] G = granules (irregular), P = pellets (nearly spherical) the
number is the compressional pressure it was compressed to e.g., G200 = granules compressed to
200 MPa.
251
Pharmaceutics 2023, 15, 1587
Figure 25. (Top): SEM micrographs of the paracetamol samples made (prepared and raw). The pre-
pared samples were dissolved in ethanol and mixed at a set RPM to achieve the modified shapes.
(Bottom): Heckel plot of different sizes and shapes of paracetamol. Both figures obtained from
Šimek et al. [133].
252
Pharmaceutics 2023, 15, 1587
Although the general trends of particle size and shape’ effects on tabletting is men-
tioned in the previous section, there are some conflicting results regarding how they affect
the compressibility and therefore tablet strength. This can be differentiated by material
properties or the dominating deformation mechanism of the powder. Brittle materials such
as lactose and paracetamol deform by fragmentation [135]. In studies, the strength of tablet
consisting of brittle materials have been seen to be affected by particle size [29,128,136,137]
and shape [138]. Skelbæk-Pedersen et al. [137] found that larger brittle powder particles
such as lactose and calcium hydrogen phosphate dehydrate (DCP) fragmented more. In
Figure 26, with increasing particle size in DCP and lactose, the peaks are more flatlines for
the lines corresponding to the higher compressional pressure which means the particles
fragmented into smaller particle sizes more readily. This agrees with De Boer et al. [29]
Sun and Grant, [130] and Skelbæk-Pedersen et al. [137]. The more extensive fragmentation
allows the void spaces to be filled with the smaller particles which further densifies the
powder bed to reduce its negative influence of tensile strength. Nonetheless, smaller parti-
cles have higher tensile strength outcomes after compression [130]. Surprisingly, in contrast,
Almaya and Aburub [117] found that dibasic calcium phosphate dihydrate did not have a
difference with different particle sizes. Suggestions into why this may be occurring will be
discussed below.
0.1 1 10 100 1000 10,000 0.1 1 10 100 1000 10,000 0.1 1 10 100 1000 10,000
0.1 1 10 100 1000 10,000 0.1 1 10 100 1000 10,000 0.1 1 10 100 1000 10,000
Figure 26. Skelbaek-Pedersen [137] PSD of calcium hydrogen phosphate dihydrate (top) and lac-
tose (bottom) with different initial particle size distribution in the graph titles. The plots show the
change in PSD with increasing levels of compressional pressure.
Plastically deforming powders are known not to be affected by particle size [117,128,136]
due to the way they deform. McKenna and McCafferty [128] found that different sizes
of microcrystalline cellulose (MCC) did not affect the tensile strength. However, Herting
and Kleinebudde [116] found contradicting results that decreasing particle size of MCC and
theophylline resulted in higher tensile strength. Wunsch et al. [132] found that there was
an increase in tensile strength with decreasing particle size for all materials investigated
(MCC and vildagliptin) (Figure 27) However, with vildagliptin, internal defects such as cracks
during elastic recovery may have occurred resulting in large error bars. No fundamental
253
Pharmaceutics 2023, 15, 1587
understanding of these differing results is given in the papers, particularly regarding the
contradictions with past literature.
Figure 27. Compactability graphs of coarse, medium and fine grades of MCC, Ibuprofen and
vildagliptin showing the change of compactability with particle size, obtained from Wunsch [132].
254
Pharmaceutics 2023, 15, 1587
Figure 28. (A) Tabletability curves showing the difference of tabletability of three polymorphs of
sulfamerazine I, II(A) and II(B). (B) Schematic of crystal structure of polymorphs: I and II. Hydrogen
bonding are the broken lines. These crystal structure diagrams show that, even though the hydrogen
bonding connectivity is the same for both polymorphs, the secondary structures are different. All
figures obtained from Sun and Grant [130].
Future studies should pay more attention to crystal structure and understanding the
corresponding mechanical properties of the materials being investigated before and after
they are processed to a certain particle size which could affect the mechanical properties of
the powder. For example, Simek et al. [133] modified their paracetamol samples in ethanol
and changed the stirring speed to achieve the desired particle shape (Figure 25). Under-
standing these discrepancies could allow potential compromises to be made. For example,
if a certain material is very cohesive but the particle size is independent of the tabletability
properties, a larger particle size could be used to reduce cohesivity while not affecting the
tabletability of the tablet formulation.
Surface energy is the excess energy on the surface compared to the bulk [141] which is
affected by a number of intermolecular forces such as van der Waals [142]. A more cohesive
powder will have a higher surface energy with interparticle bonding that will hinder the
powder flow [143]. However, it is well documented that surface energy has a profound ef-
fect on producing successful tablets after powder compaction. Many results have suggested
higher surface energy leads to stronger bonds forming between powder particles, which
form higher tensile strength tablets [33–36]. For example, Wünsch et al. [132] suggested
that formation of hydrogen bonds between MCC particles is why there was a considerable
255
Pharmaceutics 2023, 15, 1587
increase of tensile strength compared to ibuprofen and vildagliptin, as seen in Figure 29.
However, the way that MCC, ibuprofen and vildagliptin deforms was not considered in
this discussion point.
Figure 29. Graphs to show the specific plastic (solid lines) and elastic energies (dotted lines) of fine,
medium and coarse grades of MCC, ibuprofen (Ibu) and vildagliptin (Vil) and paracetamol (para)
obtained from Wunsch et al. [132] These graphs show that coarse grades have lower specific surface
energy across all cases.
To intrinsically measure the effect of surface energy of the powder on the tensile
strength is immensely difficult. Fichtner et al. [144] and Chen et al. [145] have attempted to
change the surface energy of the powder by dry coating with polysorbate and Aerosil 200 re-
spectively. This showed a decrease in tensile strength with the decrease of surface energy in
both studies. However, the effect of changing particle size was not well documented and as
previously stated in Section 4.2, depending on the specific material, particle size is another
factor that changes the tablet outcome after compression. Ho et al. [146] counteracted this
issue by changing the surface energy without changing any other powder particle property
by modifying the functional groups on the surface of the powder without disrupting the
surface morphology and particle size by silanisation. This treatment profoundly decreased
the tensile strength of the tablet as the surface energy was reduced [147].
Another method, which was also mentioned in Section 2.1 to decrease the surface
energy of a powder and improve flowability is to add a glidant, which is seen to increase
the flowability by reducing cohesion [6]. How these additives affect the tabletability of the
powder formulation was looked at in many different studies. Magnesium stearate (MgSt)
is known to improve flowability [148]. However, it has a detrimental effect on the tablet
strength especially when magnesium stearate is mixed for too long and vigorously. This is
an effect which is known as ‘over-lubrication’ [116,149]. The MgSt is theorised to create a
thin hydrophobic layer around powder particles preventing bonding between the powder
particles to take place and therefore decreasing the tensile strength [149]. Furthermore,
as MgSt is hydrophobic, the powder being ‘over-lubcricated’ will effect the dissolution [150].
256
Pharmaceutics 2023, 15, 1587
On the other hand, MgSt has excellent lubrication properties and is commonly used in
tablet formulation as a lubricant [144] rather than a glidant.
A glidant that is discussed frequently in literature is silicon dioxide. Apeji and
Olowosulu [67] used talc and colloidal silica which appears to have had a negative influ-
ence on the tensile strength of the tablet. This paper did not have any comparison to a
control data without any glidant to compare with. What the previous papers mentioned
do not include is whether they are using hydrophobic or hydrophilic silicon dioxide. Kun-
nath et al. [151] looked at both hydrophilic and hydrophobic nano-silica and their impact
on tablet tensile strength and dissolution (Figure 30). The nano-silica was dry coated
onto the API with similar methods to those implemented by Chen et al. [145], mentioned
previously. This saw an increase in tensile strength for both hydrophilic and hydrophobic
nano-silica and did not see an effect on the dissolution profile. The possible explanation
of this is that the dry coating of API led to deagglomeration of the powder and increased
the total contact area where interparticle bonding can take place. On the other hand,
in Mužíková et al.’s [66] study, when colloidal silica (Aerosil 200 and 255) was added, this
saw a decrease in tensile strength, therefore the disintegration time also decreased. The max
compressional pressure used to produce the tablets was very low (26.37 MPa) compared to
what is normally used (200–250 MPa). However, Wunsch [132] found that it is the number
of bonds which determine the strength of the tablet, not the bonding strength. Similar to
the particle size conclusion, the fundamentals of the mechanical properties and therefore
deformation properties of the powder are vital to understand as enhancing the bonding
area may be more important than increasing the surface energy. Furthermore, this could
mean a compromise can be taken on surface energy which will lower the cohesivity of the
powder but would not affect the tablet outcomes if the contact area between particles is suf-
ficient enough to gain the number of bonds needed. This further points to the importance
of understanding how the powder deforms and how that effects and governs the contact
area between particles [19].
There have been promising results for silicon dioxide alternatives such as magnesium
aluminosilicate (MAS) and tricalcium phosphate (TCP). Both of these novel glidants have a
positive effect on flowability [98]. However, there is just a single study investigating the
tabletability of MAS, and none with TCP; in which, Hentzchel et al. [152] compared the
tabletability of MAS with other silicates, when mixed with MCC of different concentrations.
MAS was the only silicate that did not decrease in tablet strength, with increasing concen-
tration, which is a positive result. Although the tablet strength is not affected by MAS,
a study on how these novel glidants affect the dissolution rate at different concentrations
will need to be conducted, as this is a critical quality attribute of the tablet. A study by
Khunawattanakul et al. [153] saw the MAS in a tablet film coat increased the dissolution
time, suggesting a balance will be required to ensure the right concentration to aid flowa-
bility is achieved, without negatively affecting dissolution. Although Tran et al. [98] shows
that MAS aids flowability, this will need to be further investigated into whether MAS
does have an impact on the feeding and blending parts of the CDC process. As far as the
authors are aware there are no extensive studies that analyse the effect of these alternative
glidants on these unit operations, essentially showing that additional studies are required
to understand the complex interactions between not only CDC processing but the resultant
in vivo performance of the tablet. The aforementioned publication Tran et al. [54] takes a
perspective of using a range of measurement techniques to inform quantities or types of
flow modifiers, practically taking a quality by design approach. A similar methodology
could be applied in terms of glidants (namely, MAS) and their subsequent impact on tablet
quality CQAs such as tensile strength of binary mixtures with excipients other than MCC.
257
Pharmaceutics 2023, 15, 1587
Figure 30. Tensile strength of the tablets made from (a) coarse acetaminophen (cAPAP), (b) micronised
acetaminophen (mAPAP), and (c) micronised ibuprofen (mIBU) where Aerosil R972P nano-silica
(hydrophobic) and CAB-O-SIL M5P nano silica (hydrophilic) obtained from Kunnath et al. [151].
258
Pharmaceutics 2023, 15, 1587
mixtures with different commonly used excipients with different deformation properties
than MCC.
5. Summary Table
Table 2 summarises the effect cohesion has on the unit operation and what control
measures are outlined in the discussion. How the control measures from one-unit operation
can affect another downstream has yet to be discussed. This section will outline how
some cohesion controls that have positive effects on unit operations can promote issues in
downstream unit operations in the continuous direct compression line.
The consistency and accuracy of which powders are dispensed from a feeder will
directly influence the downstream unit operations [43,45]. Therefore, it is of utmost im-
portance that understanding is gained on how to maintain a consistent feed rate for very
cohesive materials [2,41,55]. The use of silica shows great promise, as it seemingly both
improves flowability and reduces electrostatic charging, leading to better screw filling and
thus resulting in improved feeding consistency [5,47,57]. Lumay et al. [57] also showed
silica’s ability to influence the thixotropic behaviour; in future work, it would be interesting
to see a development on this in the context of LIW feeding. However, the addition of silica
has potentially negative effects on tabletting [66,67], while its effect on blending has not
been explicitly described. Bridging/arching within the hopper volume has been shown
to suffocate the flow of powder into the conveying volume. As a result, both additives
and refilling strategy show ideas for reducing the processing variability [2,41,55,65]. Fur-
thermore, using DEM for complex agitator design would further help the development of
arching minimisation.
Despite the benefits of silication, attention should be paid to the end-to-end impact
of formulation additives. For instance, nano-silicas were seen to have a lesser impact,
compared to MPS, on the tablet strength, but greatly improved the flowability [57]. Alterna-
tively, novel glidants, such as magnesium aluminosilicates, show promise in minimising the
respective trade-off of improving flowability and affecting the tablet’s tensile strength [152].
Blending seeks to intimately mix all of a given formulation’s constituents and deliver
them to the final, tabletting stage of the process with minimal variation [45,154]. Cohesion,
in most cases, serves as a direct barrier to blending, as it provides resistance to intimate
mixing [75,78,90]. Thus, higher levels of agitation are required to break apart these cohesive
clusters, but systems must devise their own method to balance both the micro and the
macro aspects of mixing [43,88]. In addition, complex bulk behaviour occurs when there
is a moderate amount of cohesive particle–particle contacts [90]. Thus, it is of interest
to develop a greater understanding of the mechanisms which govern this complex bulk
behaviour. Moreover, despite the viewpoint to increase RPM to deal with cohesion, it is
not understood what these new shear environments look like, and whether they have an
impact on tabletability.
Spherical Agglomeration, which was discussed in Section 4.2, is a very promising
technique to control the amount of cohesivity APIs add to tablet formulations without com-
promising tabletability to the same extent as the addition of glidants and lubricants which
is beneficial to the whole CDC process. Further work will, however, need to be carried
out to improve the efficiency of selecting the right solvent and process parameters [123].
Process parameters especially will need to be focussed on as some parameters are high-
lighted in studies that may be difficult to scale up. However, there have been good strides
looking into the scale-up process and making spherical agglomeration/crystallisation
continuous [155,156]. Ensuring the API crystals are not too large will be important to
ensure content uniformity within tablets [123].
259
Table 2. Summary table with all the unit operations discussed in this literature review and their respective cohesion effects and controls. Note:
Cohesion compromise is denoted with a , future work with ◦, considerations .
Unit Operation Cohesion Effects Results in . . . Cohesion Control or Compromise Future Work ◦ /Considerations
Modifying bulk density and cohesive particle-particle contact
Inconsistent screw filling through additives, such as Silica [5,47,57] Understand and be able to modify thixotropic powder ◦ behaviour
Feeding Increase/Decrease conveying screw speeds [57]
Refill Strategy [41,65] DEM simulations exploring optimum hopper agitator design ◦
Bridging/Arching
Pharmaceutics 2023, 15, 1587
Agitator Design
Increase in rpm (Fr) balanced by changes to system
Reduced micro-mixing
geometry to maintain macro [75,78,90]
260
Tabletting then resist the motion of particles. Find the ranges when air
lubricant [27]
introduction is beneficial by Quality by Design experiments
Understand the impacts of different types of nano-silica.
Nano-silica and other novel glidants (magnesium aluminosilicate)
Magnesium aluminosilicate will need to be tabletted
Better tabletability; generally lower can help promote better flowability and has little to no effect
with other common excipients and dissolution testing will
particle size and higher surface energy on tablet strength [151]
to be conducted to assess the limitations◦
Compaction results in stronger tablets but
can have bad effects on dissolution if Number of bonds is more important than the strength of
Understanding crystal structure fundamentally and how that changes
the tablet strength is too high [157] the bond could be able to compromise lower surface
material properties. This may be what is causing discrepancies
energy/cohesivity if there is sufficient contact area gained
regarding particle size and the tablet strength outcome ◦ [130]
during compaction
Pharmaceutics 2023, 15, 1587
6. Conclusions
This literature review consolidates the impact of powder cohesion and mitigation
for each of the unit processes in CDC, whilst considering the consequences of upstream/
downstream processing. For a tablet formulation to be viable for the CDC process, the for-
mulation must possess both suitable flowability and tabletability [19,25]. Cohesion is known
to cause powders to resist flow [23,24] which has a negative impact on the performance
of the feeding [25], mixing [85,87], and die filling [3] unit operations involved in the CDC
process whereas, for the compression stage, cohesivity is desired [29,30,33,35].
Understanding the cohesive nature of APIs and excipients and then manipulating
them, where possible, is useful for improving the overall performance of the CDC process.
The use of additives such as glidants and lubricants may aid the flow of powders, thus
improving the feeding [6], mixing [37] and die filling [110] performance; however, these
additives can have a negative impact on compaction [29]. Flow variability should be
managed (specifically during the feeding stage) as any perturbations will be propagated
into successive unit operations [2,158].
Despite the volume of literature in this area, there remain many important open
questions and thus valuable future research avenues:
• Silication has benefits similar to glidants and lubricants as it can lower the cohesive
behaviour of the powder and therefore increase flowability [6]. However, there are
contradictions on whether silication negatively affects tabletability; this will need to
be further investigate [67,145]. Alternative glidants, e.g., magnesium aluminosilicates,
are a promising option however, more work needs to be carried out to understand the
impact on tablet performance.
• Discrete Element Method (DEM) modelling, allowing researchers to create a digital
twin(s) of an existing experiment and conduct statistical analysis on the simulated pow-
der behaviour would allow researchers to gather metrics to quantify mixing/feeding
performance (as discussed in Escotet-Espinoza et al. [80]) that were not accessible
experimentally.
• Utilising triboelectric charging for the blending stage: static surface charge is de-
veloped on the surface of particles due to the strain given to the powder during
mixing/transport [47], potentially allowing a formulation to be altered to promote
attraction/repulsion between constituent species in order to gain a more-ordered
well-mixed system.
• Spherical agglomeration/crystallisation can improve the flowability and tabletability of
APIs, which is normally the most cohesive component of the tablet formulation [3,18].
However, further work needs to be performed to improve the solvent selection and
process parameters to allow for scale-up [123].
• The many discrepancies in the literature regarding the manner in which particle
size affects the tabletability of the powder should be comprehensively addressed,
as this represents an important gap in the fundamental understanding of powder
compression [39,128]. Understanding which types of powders are affected by particle
size and which are not could lead to compromises to have larger particle sizes to lower
cohesivity and therefore improve flow while not affecting the tabletability.
Author Contributions: Conceptualization, O.J.-S. and Z.C.; formal analysis, O.J.-S. and Z.C.; investi-
gation, O.J.-S. and Z.C.; writing—original draft preparation, O.J.-S. and Z.C.; writing—review and
editing, C.R.K.W.-Y. and A.I.; visualization, O.J.-S. and Z.C.; supervision, C.R.K.W.-Y. and A.I.; project
administration, C.R.K.W.-Y. and A.I.; funding acquisition, C.R.K.W.-Y. and A.I. All authors have read
and agreed to the published version of the manuscript.
Funding: The authors would like to thank the Engineering Doctorate programme funded by EPSRC
through the Centre for Doctoral Training in Formulation Engineering (grant no. EP/L015153/1),
and from AstraZeneca plc.
261
Pharmaceutics 2023, 15, 1587
Acknowledgments: The authors would like to also thank Adedoyin Yahyi for the initial insight and
discussion surrounding powder feeding.
Conflicts of Interest: The authors declare no conflict of interest. The company had no role in
the design of the study; in the collection, analyses, or interpretation of data; in the writing of the
manuscript, and in the decision to publish the results.
Abbreviations
The following abbreviations are used in this manuscript:
References
1. Vercruysse, J.; Delaet, U.; Van Assche, I.; Cappuyns, P.; Arata, F.; Caporicci, G.; De Beer, T.; Remon, J.P.; Vervaet, C. Stability and
repeatability of a continuous twin screw granulation and drying system. Eur. J. Pharm. Biopharm. 2013, 85, 1031–1038. [CrossRef]
[PubMed]
2. Engisch, W.E.; Muzzio, F.J. Loss-in-Weight Feeding Trials Case Study: Pharmaceutical Formulation. J. Pharm. Innov. 2015,
10, 56–75. [CrossRef]
3. Chattoraj, S.; Sun, C.C. Crystal and Particle Engineering Strategies for Improving Powder Compression and Flow Properties to
Enable Continuous Tablet Manufacturing by Direct Compression. J. Pharm. Sci. 2018, 107, 968–974. [CrossRef] [PubMed]
4. Liu, L.X.; Marziano, I.; Bentham, A.C.; Litster, J.D.; White, E.T.; Howes, T. Influence of particle size on the direct compression of
ibuprofen and its binary mixtures. Powder Technol. 2013, 240, 66–73. [CrossRef]
5. Leung, L.Y.; Mao, C.; Srivastava, I.; Du, P.; Yang, C.Y. Flow Function of Pharmaceutical Powders Is Predominantly Governed by
Cohesion, Not by Friction Coefficients. J. Pharm. Sci. 2017, 106, 1865–1873. [CrossRef] [PubMed]
6. Escotet-Espinoza, M.S.; Scicolone, J.V.; Moghtadernejad, S.; Sanchez, E.; Cappuyns, P.; Van Assche, I.; Di Pretoro, G.; Ierapetritou,
M.; Muzzio, F.J. Improving Feedability of Highly Adhesive Active Pharmaceutical Ingredients by Silication. J. Pharm. Innov. 2021,
16, 279–292. [CrossRef]
7. Visser, J. Van der Waals and other cohesive forces affecting powder fluidization. Powder Technol. 1989, 58, 1–10. [CrossRef]
8. Juarez-Enriquez, E.; Olivas, G.I.; Zamudio-Flores, P.B.; Perez-Vega, S.; Salmeron, I.; Ortega-Rivas, E.; Sepulveda, D.R. A review on
the influence of water on food powder flowability. J. Food Process Eng. 2022, 45, e14031. [CrossRef]
9. Zafar, U.; Vivacqua, V.; Calvert, G.; Ghadiri, M.; Cleaver, J.S. A review of bulk powder caking. Powder Technol. 2017, 313, 389–401.
[CrossRef]
262
Pharmaceutics 2023, 15, 1587
10. Wei, Z.; Zhao, Y.P. Growth of liquid bridge in AFM. J. Phys. D Appl. Phys. 2007, 40, 4368. [CrossRef]
11. Hou, Q.; Dong, K.; Yu, A. DEM study of the flow of cohesive particles in a screw feeder. Powder Technol. 2014, 256, 529–539.
[CrossRef]
12. Allenspach, C.; Timmins, P.; Lumay, G.; Holman, J.; Minko, T. Loss-in-weight feeding, powder flow and electrostatic evaluation
for direct compression hydroxypropyl methylcellulose (HPMC) to support continuous manufacturing. Int. J. Pharm. 2021,
596, 120259. [CrossRef] [PubMed]
13. Karner, S.; Urbanetz, N.A. Arising of electrostatic charge in the mixing process and its influencing factors. Powder Technol. 2012,
226, 261–268. [CrossRef]
14. Abouzeid, A.Z.M.; Fuerstenau, D.W. Effect of Humidity on Mixing of Particulate Solids. Ind. Eng. Chem. Process Des. Dev. 1972,
11, 296–301.
15. Beretta, M.; Hörmann, T.R.; Hainz, P.; Hsiao, W.K.; Paudel, A. Investigation into powder tribo–charging of pharmaceuticals. Part
II: Sensitivity to relative humidity. Int. J. Pharm. 2020, 591, 120015. [CrossRef] [PubMed]
16. Haycocks, N.; Goldschmidt, N.A.; Thomsen, U. Temperature & Humidity Requirements in Pharmaceutical Facilities. 2023.
Available online: https://ispe.org/pharmaceutical-engineering/september-october-2021/temperature-humidity-requirements-
pharmaceutical (accessed on 31 March 2023).
17. Sun, C.C. Mechanism of moisture induced variations in true density and compaction properties of microcrystalline cellulose. Int.
J. Pharm. 2008, 346, 93–101. [CrossRef]
18. Chen, H.; Aburub, A.; Sun, C.C. Direct Compression Tablet Containing 99% Active Ingredient—A Tale of Spherical Crystallization.
J. Pharm. Sci. 2019, 108, 1396–1400. [CrossRef]
19. Sun, C.C. Decoding powder tabletability: Roles of particle adhesion and plasticity. J. Adhes. Sci. Technol. 2011, 25, 483–499.
[CrossRef]
20. Haywood, A.; Glass, B.D. Pharmaceutical excipients—Where do we begin? Aust. Prescr. 2011, 34, 112–114. [CrossRef]
21. Jivraj, M.; Martini, L.G.; Thomson, C.M. An overview of the different excipients useful for the direct compression of tablets.
Pharm. Sci. Technol. Today 2000, 3, 59–63. [CrossRef]
22. Escotet-Espinoza, M.S. Phenomenological and Residence Time Distribution Models for Unit Operations in a Continuous
Pharmaceutical Manufacturing Process. Ph.D. Thesis, Rutgers University-School of Graduate Studies, New Brunswick, NJ, USA,
2018; p. 312.
23. Shah, U.V.; Olusanmi, D.; Narang, A.S.; Hussain, M.A.; Tobyn, M.J.; Hinder, S.J.; Heng, J.Y. Decoupling the contribution of surface
energy and surface area on the cohesion of pharmaceutical powders. Pharm. Res. 2015, 32, 248–259. [CrossRef]
24. Shah, R.B.; Tawakkul, M.A.; Khan, M.A. Comparative evaluation of flow for pharmaceutical powders and granules. AAPS
PharmSciTech 2008, 9, 250–258. [CrossRef] [PubMed]
25. Engisch, W.E.; Muzzio, F.J. Method for characterization of loss-in-weight feeder equipment. Powder Technol. 2012, 228, 395–403.
[CrossRef]
26. Mazumder, M.K.; Sims, R.A.; Biris, A.S.; Srirama, P.K.; Saini, D.; Yurteri, C.U.; Trigwell, S.; De, S.; Sharma, R. Twenty-first century
research needs in electrostatic processes applied to industry and medicine. Chem. Eng. Sci. 2006, 61, 2192–2211. [CrossRef]
27. Wu, C.Y.; Dihoru, L.; Cocks, A.C. The flow of powder into simple and stepped dies. Powder Technol. 2003, 134, 24–39. [CrossRef]
28. Freeman, R.; Fu, X. Characterisation of power bulk, dynamic flow and shear properties in relation to die filling. Powder Metall.
2008, 51, 196–201. [CrossRef]
29. De Boer, A.H.; Vromans, H.; Leur, C.F.; Bolhuis, G.K.; Kussendrager, K.D.; Bosch, H. Studies on tableting properties of lactose—
Part III. The consolidation behaviour of sieve fractions of crystalline α-lactose monohydrate. Pharm. Weekbl. Sci. Ed. 1986,
8, 145–150. [CrossRef]
30. Johansson, B.; Alderborn, G. The effect of shape and porosity on the compression behaviour and tablet forming ability of granular
materials formed from microcrystalline cellulose. Eur. J. Pharm. Biopharm. 2001, 52, 347–357. [CrossRef] [PubMed]
31. Chang, S.Y.; Sun, C.C. Insights into the effect of compaction pressure and material properties on interfacial bonding strength of
bilayer tablets. Powder Technol. 2019, 354, 867–876. [CrossRef]
32. Cabiscol, R.; Shi, H.; Wünsch, I.; Magnanimo, V.; Finke, J.H.; Luding, S.; Kwade, A. Effect of particle size on powder compaction
and tablet strength using limestone. Adv. Powder Technol. 2020, 31, 1280–1289. [CrossRef]
33. El Gindy, N.A.; Samaha, M.W. Tensile strength of some pharmaceutical compacts and their relation to surface free energy. Int. J.
Pharm. 1982, 13, 35–46. [CrossRef]
34. Fichtner, F.; Mahlin, D.; Welch, K.; Gaisford, S.; Alderborn, G. Effect of surface energy on powder compactibility. Pharm. Res.
2008, 25, 2750–2759. [CrossRef] [PubMed]
35. Luangtana-Anan, M.; Fell, J.T. Bonding mechanisms in tabletting. Int. J. Pharm. 1990, 60, 197–202. [CrossRef]
36. Etzler, F.M.; Pisano, S. Tablet tensile strength: Role of surface free energy. Adv. Contact Angle Wettability Adhes. 2015, 2, 397–418.
37. Pingali, K.; Mendez, R.; Lewis, D.; Michniak-Kohn, B.; Cuitino, A.; Muzzio, F. Mixing order of glidant and lubricant—Influence
on powder and tablet properties. Int. J. Pharm. 2011, 409, 269–277. [CrossRef]
38. Mills, L.A.; Sinka, I.C. Effect of particle size and density on the die fill of powders. Eur. J. Pharm. Biopharm. 2013, 84, 642–652.
[CrossRef]
39. Herting, M.G.; Kleinebudde, P. Roll compaction/dry granulation: Effect of raw material particle size on granule and tablet
properties. Int. J. Pharm. 2007, 338, 110–118. [CrossRef]
263
Pharmaceutics 2023, 15, 1587
40. Kehlenbeck, V.; Sommer, K. Possibilities to improve the short-term dosing constancy of volumetric feeders. Powder Technol. 2003,
138, 51–56. [CrossRef]
41. Engisch, W.E.; Muzzio, F.J. Feedrate deviations caused by hopper refill of loss-in-weight feeders. Powder Technol. 2015,
283, 389–400. [CrossRef]
42. Bekaert, B.; Van Snick, B.; Pandelaere, K.; Dhondt, J.; Di Pretoro, G.; De Beer, T.; Vervaet, C.; Vanhoorne, V. In-depth analysis of
the long-term processability of materials during continuous feeding. Int. J. Pharm. 2022, 614, 121454. [CrossRef]
43. Van Snick, B.; Holman, J.; Vanhoorne, V.; Kumar, A.; De Beer, T.; Remon, J.P.; Vervaet, C. Development of a continuous direct
compression platform for low-dose drug products. Int. J. Pharm. 2017, 529, 329–346. [CrossRef]
44. Simonaho, S.P.; Ketolainen, J.; Ervasti, T.; Toiviainen, M.; Korhonen, O. Continuous manufacturing of tablets with PROMIS-
line—Introduction and case studies from continuous feeding, blending and tableting. Eur. J. Pharm. Sci. 2016, 90, 38–46.
[CrossRef]
45. García-Muñoz, S.; Butterbaugh, A.; Leavesley, I.; Manley, L.F.; Slade, D.; Bermingham, S. A flowsheet model for the development
of a continuous process for pharmaceutical tablets: An industrial perspective. AIChE J. 2018, 64, 511–525. [CrossRef]
46. Gyürkés, M.; Madarász, L.; Köte, Á.; Domokos, A.; Mészáros, D.; Beke, K.; Nagy, B.; Marosi, G.; Pataki, H.; Nagy, Z.K.; et al.
Process design of continuous powder blending using residence time distribution and feeding models. Pharmaceutics 2020, 12, 1119.
[CrossRef] [PubMed]
47. Beretta, M.; Hörmann, T.R.; Hainz, P.; Hsiao, W.K.; Paudel, A. Investigation into powder tribo-charging of pharmaceuticals. Part
I: Process-induced charge via twin-screw feeding. Int. J. Pharm. 2020, 591, 120014. [CrossRef]
48. Windows-Yule, C.R.; Neveu, A. Calibration of DEM simulations for dynamic particulate systems. Pap. Phys. 2022, 14, 140010.
[CrossRef]
49. Rhodes, M.J. Introduction to Particle Technology; John Wiley & Sons: Hoboken, NJ, USA, 2008.
50. Windows-Yule, K.; Nicuşan, L.; Herald, M.T.; Manger, S.; Parker, D. Comparison with other techniques. In Positron Emission
Particle Tracking: A Comprehensive Guide; IOP Publishing: Bristol, UK, 2022.
51. Thornton, C. Granular Dynamics, Contact Mechanics and Particle System Simulations: A DEM Study; Springer’s International
Publishing: Berlin/Heidelberg, Germany, 2015; Volume 24, pp. 1–195. [CrossRef]
52. Kojima, T.; Elliott, J.A. Effect of silica nanoparticles on the bulk flow properties of fine cohesive powders. Chem. Eng. Sci. 2013,
101, 315–328. [CrossRef]
53. Meyer, K.; Zimmermann, I. Effect of glidants in binary powder mixtures. Powder Technol. 2004, 139, 40–54. [CrossRef]
54. Tran, D.T.; Majerová, D.; Veselý, M.; Kulaviak, L.; Ruzicka, M.C.; Zámostný, P. On the mechanism of colloidal silica action to
improve flow properties of pharmaceutical excipients. Int. J. Pharm. 2019, 556, 383–394. . [CrossRef]
55. López, A.; Vivacqua, V.; Hammond, R.; Ghadiri, M. Analysis of screw feeding of faceted particles by discrete element method.
Powder Technol. 2020, 367, 474–486. [CrossRef]
56. Krantz, M.; Zhang, H.; Zhu, J. Characterization of powder flow: Static and dynamic testing. Powder Technol. 2009, 194, 239–245.
[CrossRef]
57. Lumay, G.; Pillitteri, S.; Marck, M.; Monsuur, F.; Pauly, T.; Ribeyre, Q.; Francqui, F.; Vandewalle, N. Influence of mesoporous silica
on powder flow and electrostatic properties on short and long term. J. Drug Deliv. Sci. Technol. 2019, 53, 101192. [CrossRef]
58. Neveu, A.; Francqui, F.; Lumay, G. Measuring powder flow properties in a rotating drum. Measurement 2022, 200, 111548.
[CrossRef]
59. Patel, D.; Francqui, F.; Defrêne, R.J.L.; Neveu, A. Influence of additives on the electrostatic charge build-up of excipients. Powder
Technol. 2003, 132, 192–198. [CrossRef]
60. Ramirez-Dorronsoro, J.C.; Jacko, R.B.; Kildsig, D.O. Chargeability measurements of selected pharmaceutical dry powders to
assess their electrostatic charge control capabilities. AAPS PharmSciTech 2006, 7, 103. [CrossRef] [PubMed]
61. Escotet-Espinoza, M.S.; Moghtadernejad, S.; Oka, S.; Wang, Z.; Wang, Y.; Roman-Ospino, A.; Schäfer, E.; Cappuyns, P.; Van Assche,
I.; Futran, M.; et al. Effect of material properties on the residence time distribution (RTD) characterization of powder blending
unit operations. Part II of II: Application of models. Powder Technol. 2019, 344, 525–544. [CrossRef]
62. Escotet-Espinoza, M.S.; Moghtadernejad, S.; Scicolone, J.; Wang, Y.; Pereira, G.; Schäfer, E.; Vigh, T.; Klingeleers, D.; Ierapetritou,
M.; Muzzio, F.J. Using a material property library to find surrogate materials for pharmaceutical process development. Powder
Technol. 2018, 339, 659–676. [CrossRef]
63. Shier, A.P.; Kumar, A.; Mercer, A.; Majeed, N.; Doshi, P.; Blackwood, D.O.; Verrier, H.M. Development of a predictive model for
gravimetric powder feeding from an API-rich materials properties library. Int. J. Pharm. 2022, 625, 122071. [CrossRef]
64. Yadav, I.K.; Holman, J.; Meehan, E.; Tahir, F.; Khoo, J.; Taylor, J.; Benedetti, A.; Aderinto, O.; Bajwa, G. Influence of material
properties and equipment configuration on loss-in- weight feeder performance for drug product continuous manufacture. Powder
Technol. 2019, 348, 126–137. [CrossRef]
65. De Souter, L.; Waeytens, R.; Van Hauwermeiren, D.; Grymonpré, W.; Bekaert, B.; Nopens, I.; De Beer, T. Elucidation of the powder
flow pattern in a twin-screw LIW-feeder for various refill regimes. Int. J. Pharm. 2023, 631, 122534. [CrossRef]
66. Mužíková, J.; Louženská, M.; Pekárek, T. A study of compression process and properties of tablets with microcrystalline cellulose
and colloidal silicon dioxide. Acta Pol. Pharm.—Drug Res. 2016, 73, 1259–1265.
67. Apeji, Y.E.; Olowosulu, A.K. Quantifying the effect of glidant on the compaction and tableting properties of paracetamol granules.
J. Res. Pharm. 2020, 24, 44–55. [CrossRef]
264
Pharmaceutics 2023, 15, 1587
68. Uzunović, A.; Vranić, E. Effect of magnesium stearate concentration on dissolution properties of ranitidine hydrochloride coated
tablets. Biomol. Biomed. 2007, 7, 279–283. [CrossRef] [PubMed]
69. Bolhuis, G.K.; Anthony Armstrong, N. Excipients for Direct Compaction—An Update. Pharm. Dev. Technol. 2006, 11, 111–124.
[CrossRef]
70. Huang, Z.; Scicolone, J.V.; Gurumuthy, L.; Davé, R.N. Flow and bulk density enhancements of pharmaceutical powders using a
conical screen mill: A continuous dry coating device. Chem. Eng. Sci. 2015, 125, 209–224. [CrossRef]
71. Bridgwater, J. Fundamental powder mixing mechanisms. Powder Technol. 1976, 15, 215–236. [CrossRef]
72. Cooke, M.H.; Stephens, D.J.; Bridgwater, J. Powder mixing—A literature survey. Powder Technol. 1976, 15, 1–20. [CrossRef]
73. Laurent, B.; Bridgwater, J. Influence of agitator design on powder flow. Chem. Eng. Sci. 2002, 57, 3781–3793. [CrossRef]
74. Van Snick, B.; Grymonpré, W.; Dhondt, J.; Pandelaere, K.; Di Pretoro, G.; Remon, J.P.; De Beer, T.; Vervaet, C.; Vanhoorne, V.
Impact of blend properties on die filling during tableting. Int. J. Pharm. 2018, 549, 476–488. [CrossRef]
75. Palmer, J.; Reynolds, G.K.; Tahir, F.; Yadav, I.K.; Meehan, E.; Holman, J.; Bajwa, G. Mapping key process parameters to the
performance of a continuous dry powder blender in a continuous direct compression system. Powder Technol. 2020, 362, 659–670.
[CrossRef]
76. Venables, H.J.; Wells, J.I. Powder mixing. Drug Dev. Ind. Pharm. 2001, 27, 599–612. [CrossRef]
77. Asachi, M.; Nourafkan, E.; Hassanpour, A. A review of current techniques for the evaluation of powder mixing. Adv. Powder
Technol. 2018, 29, 1525–1549. [CrossRef]
78. Portillo, P.M.; Vanarase, A.U.; Ingram, A.; Seville, J.K.; Ierapetritou, M.G.; Muzzio, F.J. Investigation of the effect of impeller
rotation rate, powder flow rate, and cohesion on powder flow behavior in a continuous blender using PEPT. Chem. Eng. Sci. 2010,
65, 5658–5668. [CrossRef]
79. Poux, M.; Fayolle, P.; Bertrand, J.; Bridoux, D.; Bousquet, J. Powder mixing: Some practical rules applied to agitated systems.
Powder Technol. 1991, 68, 213–234. [CrossRef]
80. Sebastian Escotet-Espinoza, M.; Moghtadernejad, S.; Oka, S.; Wang, Y.; Roman-Ospino, A.; Schäfer, E.; Cappuyns, P.; Van Assche,
I.; Futran, M.; Ierapetritou, M.; et al. Effect of tracer material properties on the residence time distribution (RTD) of continuous
powder blending operations. Part I of II: Experimental evaluation. Powder Technol. 2019, 342, 744–763. [CrossRef]
81. Gao, Y.; Vanarase, A.; Muzzio, F.; Ierapetritou, M. Characterizing continuous powder mixing using residence time distribution.
Chem. Eng. Sci. 2011, 66, 417–425. [CrossRef]
82. Vanarase, A.U.; Muzzio, F.J. Effect of operating conditions and design parameters in a continuous powder mixer. Powder Technol.
2011, 208, 26–36. [CrossRef]
83. Fan, L.; Chen, Y.m.; Lai, F. Recent developments in solids mixing. Powder Technol. 1990, 61, 255–287. [CrossRef]
84. Rosato, A.D.; Windows-Yule, C. Segregation in Vibrated Granular Systems; Academic Press: Cambridge, MA, USA, 2020;
pp. 197–218.
85. Rennie, P.R.; Chen, X.D.; Hargreaves, C.; MacKereth, A.R. Study of the cohesion of dairy powders. J. Food Eng. 1999, 39, 277–284.
[CrossRef]
86. De Villiers, M.M. Description of the kinetics of the deagglomeration of drug particle agglomerates during powder mixing. Int. J.
Pharm. 1997, 151, 1–6. [CrossRef]
87. Mehrotra, A.; Muzzio, F.J.; Shinbrot, T. Spontaneous separation of charged grains. Phys. Rev. Lett. 2007, 99, 058001. [CrossRef]
[PubMed]
88. Bekaert, B.; Grymonpré, W.; Novikova, A.; Vervaet, C.; Vanhoorne, V. Impact of blend properties and process variables on the
blending performance. Int. J. Pharm. 2022, 613, 121421. [CrossRef] [PubMed]
89. Vanarase, A.U.; Osorio, J.G.; Muzzio, F.J. Effects of powder flow properties and shear environment on the performance of
continuous mixing of pharmaceutical powders. Powder Technol. 2013, 246, 63–72. [CrossRef]
90. Gao, Y.; Ierapetritou, M.; Muzzio, F. Periodic section modeling of convective continuous powder mixing processes. AIChE J. 2012,
58, 69–78.
91. Portillo, P.M.; Ierapetritou, M.G.; Muzzio, F.J. Characterization of continuous convective powder mixing processes. Powder
Technol. 2007, 182, 368–378. [CrossRef]
92. Werner, D.; Davison, H.; Robinson, E.; Sykes, J.; Seville, J.; Wellings, A.; Bhattacharya, S.; Monsalve, D.S.; Wheldon, T.K.;
Windows-Yule, C. Effect of system composition on mixing in binary fluidised beds. Chem. Eng. Sci. 2023, 271, 118562. [CrossRef]
93. Goldhirsch, I. Introduction to granular temperature. Powder Technol. 2008, 182, 130–136. . [CrossRef]
94. Ogawa, S. Multitemperature theory of granular materials. In Proceedings of the US-Japan Seminar on Continuum Mechanical
and Statistical Approaches in the Mechanics of Granular Materials, Sendai, Japan, 5–9 June 1978; pp. 208–217.
95. Gao, Y.; Muzzio, F.J.; Ierapetritou, M.G. Optimizing continuous powder mixing processes using periodic section modeling. Chem.
Eng. Sci. 2012, 80, 70–80. [CrossRef]
96. Sarkar, A.; Wassgren, C.R. Simulation of a continuous granular mixer: Effect of operating conditions on flow and mixing. Chem.
Eng. Sci. 2009, 64, 2672–2682. [CrossRef]
97. Tomita, Y.; Nagato, T.; Takeuchi, Y.; Takeuchi, H. Control of residence time of pharmaceutical powder in a continuous mixer with
impeller and scraper. Int. J. Pharm. 2020, 586, 119520. [CrossRef]
98. Tran, D.; Komínová, P.; Kulaviak, L.; Zámostnỳ, P. Evaluation of multifunctional magnesium aluminosilicate materials as novel
family of glidants in solid dosage products. Int. J. Pharm. 2021, 592, 120054. [CrossRef] [PubMed]
265
Pharmaceutics 2023, 15, 1587
99. Toson, P.; Siegmann, E.; Trogrlic, M.; Kureck, H.; Khinast, J.; Jajcevic, D.; Doshi, P.; Blackwood, D.; Bonnassieux, A.; Daugherity,
P.D.; et al. Detailed modeling and process design of an advanced continuous powder mixer. Int. J. Pharm. 2018, 552, 288–300.
[CrossRef] [PubMed]
100. Kale, D.P.; Zode, S.S.; Bansal, A.K. Challenges in translational development of pharmaceutical cocrystals. J. Pharm. Sci. 2017,
106, 457–470. [CrossRef] [PubMed]
101. Zakhvatayeva, A.; Zhong, W.; Makroo, H.A.; Hare, C.; Wu, C.Y. An experimental study of die filling of pharmaceutical powders
using a rotary die filling system. Int. J. Pharm. 2018, 553, 84–96. [CrossRef] [PubMed]
102. Schomberg, A.K.; Kwade, A.; Finke, J.H. The challenge of die filling in rotary presses—A systematic study of material properties
and process parameters. Pharmaceutics 2020, 12, 248. [CrossRef] [PubMed]
103. Qu, L.; Stewart, P.J.; Hapgood, K.P.; Lakio, S.; Morton, D.A.; Zhou, Q.T. Single-step coprocessing of cohesive powder via
mechanical dry coating for direct tablet compression. J. Pharm. Sci. 2017, 106, 159–167. [CrossRef] [PubMed]
104. Baserinia, R.; Sinka, I.C. Powder die filling under gravity and suction fill mechanisms. Int. J. Pharm. 2019, 563, 135–155. [CrossRef]
[PubMed]
105. Tang, X.; Zakhvatayeva, A.; Zhang, L.; Wu, Z.F.; Sun, P.; Wu, C.Y. Flow behaviour of pharmaceutical powders during rotary die
filling with a paddle feeder. Int. J. Pharm. 2020, 585, 119547. [CrossRef]
106. Jackson, S.; Sinka, I.C.; Cocks, A.C. The effect of suction during die fill on a rotary tablet press. Eur. J. Pharm. Biopharm. 2007,
65, 253–256. [CrossRef]
107. Goh, H.P.; Heng, P.W.S.; Liew, C.V. Understanding effects of process parameters and forced feeding on die filling. Eur. J. Pharm.
Sci. 2018, 122, 105–115. [CrossRef]
108. Han, X.; Jallo, L.; To, D.; Ghoroi, C.; Davé, R. Passivation of high-surface-energy sites of milled ibuprofen crystals via dry coating
for reduced cohesion and improved flowability. J. Pharm. Sci. 2013, 102, 2282–2296. [CrossRef]
109. Zakhvatayeva, A.; Hare, C.; Wu, C.Y. Size-induced segregation during die filling. Int. J. Pharm. X 2019, 1, 100032. [CrossRef]
[PubMed]
110. Wu, C.Y. DEM simulations of die filling during pharmaceutical tabletting. Particuology 2008, 6, 412–418. [CrossRef]
111. Mittal, B. How to Develop Robust Solid Oral Dosage Forms: From Conception to Post-Approval; Academic Press: Cambridge, MA, USA,
2016.
112. Bansal, A.K.; Balwani, G.; Sheokand, S. Critical Material Attributes in Wet Granulation; Academic Press: Cambridge, MA, USA,
2019; pp. 421–453.
113. Eidevåg, T.; Abrahamsson, P.; Eng, M.; Rasmuson, A. Modeling of dry snow adhesion during normal impact with surfaces.
Powder Technol. 2020, 361, 1081–1092. [CrossRef]
114. Peeters, E.; Beer, T.D.; Vervaet, C.; Remon, J.p.; Peeters, E.; Beer, T.D.; Vervaet, C.; Remon, J.p. Reduction of tablet weight
variability by optimizing paddle speed in the forced feeder of a high-speed rotary tablet press forced feeder of a high-speed
rotary tablet press. Drug Dev. Ind. Pharm. 2015, 41, 530–539. [CrossRef] [PubMed]
115. Kushner, J.; Moore, F. Scale-up model describing the impact of lubrication on tablet tensile strength. Int. J. Pharm. 2010, 399, 19–30.
[CrossRef]
116. Mosig, J.; Kleinebudde, P. Critical evaluation of root causes of the reduced compactability after roll compaction/dry granulation.
J. Pharm. Sci. 2015, 104, 1108–1118. [CrossRef] [PubMed]
117. Almaya, A.; Aburub, A. Effect of Particle Size on Compaction of Materials with Different Deformation Mechanisms with and
without Lubricants. Aaps Pharmscitech 2008, 9, 414–418. [CrossRef] [PubMed]
118. Guo, Y.; Wu, C.Y.; Thornton, C. The effects of air and particle density difference on segregation of powder mixtures during die
filling. Chem. Eng. Sci. 2011, 66, 661–673. [CrossRef]
119. Schneider, L.C.; Sinka, I.C.; Cocks, A.C. Characterisation of the flow behaviour of pharmaceutical powders using a model
die-shoe filling system. Powder Technol. 2007, 173, 59–71. [CrossRef]
120. Kukkar, V.; Anand, V.; Kataria, M.; Gera, M.; Choudhury, P.K. Mixing and formulation of low dose drugs: Underlying problems
and solutions. Thai J. Pharm. Sci 2008, 32, 43–58.
121. Byrne, F.P.; Jin, S.; Paggiola, G.; Petchey, T.H.M.; Clark, J.H.; Farmer, T.J.; Hunt, A.J.; Robert McElroy, C.; Sherwood, J. Tools and
techniques for solvent selection: Green solvent selection guides. Sustain. Chem. Process. 2016, 4, 7. [CrossRef]
122. Javadzadeh, Y.; Vazifehasl, Z.; Maleki Dizahj, S.; Mokhtarpor, M. Spherical crystallization of drugs. Acta Pharm. 2012, 62, 1–14.
[CrossRef]
123. Pitt, K.; Peña, R.; Tew, J.D.; Pal, K.; Smith, R.; Nagy, Z.K.; Litster, J.D. Particle design via spherical agglomeration: A critical review
of controlling parameters, rate processes and modelling. Powder Technol. 2018, 326, 327–343. [CrossRef]
124. Quodbach, J.; Kleinebudde, P. A critical review on tablet disintegration. Pharm. Dev. Technol. 2016, 21, 763–774. [CrossRef]
125. Nicolas, V.; Chambin, O.; Andrès, C.; Rochat-Gonthier, M.H.; Pourcelot, Y. Preformulation: Effect of moisture content on
microcrystalline cellulose (Avicel PH-302) and its consequences on packing performances. Drug Dev. Ind. Pharm. 1999,
25, 1137–1142. [CrossRef]
126. Tye, C.K.; Sun, C.; Amidon, G.E. Evaluation of the effects of tableting speed on the relationships between compaction pressure,
tablet tensile strength, and tablet solid fraction. J. Pharm. Sci. 2005, 94, 465–472. [CrossRef] [PubMed]
127. Osei-Yeboah, F.; Zhang, M.; Feng, Y.; Sun, C.C. A formulation strategy for solving the overgranulation problem in high shear wet
granulation. J. Pharm. Sci. 2014, 103, 2434–2440. [CrossRef]
266
Pharmaceutics 2023, 15, 1587
128. McKenna, A.; McCafferty, D.F. Effect of particle size on the compaction mechanism and tensile strength of tablets. J. Pharm.
Pharmacol. 1982, 34, 347–351. [CrossRef] [PubMed]
129. Hallam, C.N.; Gabbott, I.P. Increasing tensile strength by reducing particle size for extrudate-based tablet formulations. J. Drug
Deliv. Sci. Technol. 2019, 52, 825–830. [CrossRef]
130. Sun, C.; Grant, D.J. Influence of crystal structure on the tableting properties of sulfamerazine polymorphs. Pharm. Res. 2001,
18, 274–280. [CrossRef] [PubMed]
131. Eriksson, M.; Alderborn, G. The effect of particle fragmentation and deformation on the interparticulate bond formation process
during powder compaction. Pharm. Res. 1995, 12, 1031–1039. [CrossRef] [PubMed]
132. Wünsch, I.; Finke, J.H.; John, E.; Juhnke, M.; Kwade, A. The influence of particle size on the application of compression and
compaction models for tableting. Int. J. Pharm. 2021, 599, 120424. [CrossRef]
133. Šimek, M.; Grünwaldová, V.; Kratochvíl, B. Comparison of compression and material properties of differently shaped and sized
paracetamols. KONA Powder Part J. 2017, 2017, 197–206. [CrossRef]
134. Choi, D.H.; Kim, N.A.; Chu, K.R.; Jung, Y.J.; Yoon, J.h.; Jeong, S.H. Material Properties and Compressibility Using Heckel and
Kawakita Equation with Commonly Used Pharmaceutical Excipients. J. Pharm. Investig. 2010, 40, 237–244. [CrossRef]
135. Prasad, K.V.; Sheen, D.B.; Sherwood, J.N. Fracture property studies of paracetamol single crystals using microindentation
techniques. Pharm. Res. 2001, 18, 867–872. [CrossRef]
136. Roberts, R.; Rowe, R.C. The compaction of pharmaceutical and other model materials—A pragmatic approach. Chem. Eng. Sci.
1987, 42, 903–911. [CrossRef]
137. Skelbæk-pedersen, A.L.; Vilhelmsen, T.K.; Wallaert, V. Investigation of the e ff ects of particle size on fragmentation during
tableting. Int. J. Pharm. 2020, 576, 118985. [CrossRef]
138. Wong, L.W.; Pilpel, N. The effect of particle shape on the mechanical properties of powders. Int. J. Pharm. 1990, 59, 145–154.
[CrossRef]
139. Upadhyay, P.; Khomane, K.S.; Kumar, L.; Bansal, A.K. Relationship between crystal structure and mechanical properties of
ranitidine hydrochloride polymorphs. CrystEngComm 2013, 15, 3959–3964. [CrossRef]
140. Khomane, K.S.; Bansal, A.K. Effect of particle size on in-die and out-of-die compaction behavior of ranitidine hydrochloride
polymorphs. AAPS PharmSciTech 2013, 14, 1169–1177. [CrossRef] [PubMed]
141. Packham, D.E. Surface energy , surface topography and adhesion. Int. J. Adhes. Adhes. 2003, 23, 437–448. [CrossRef]
142. Annamalai, M.; Gopinadhan, K.; Han, S.A.; Saha, S.; Park, H.J.; Cho, E.B.; Kumar, B.; Patra, A.; Kim, S.W.; Venkatesan, T. Surface
energy and wettability of van der Waals structures. Nanoscale 2016, 8, 5764–5770. [CrossRef]
143. Li, Q.; Rudolph, V.; Weigl, B.; Earl, A. Interparticle van der Waals force in powder flowability and compactibility. Int. J. Pharm.
2004, 280, 77–93. [CrossRef]
144. Fichtner, F.; Rasmuson, A.; Alderborn, G. Particle size distribution and evolution in tablet structure during and after compaction.
Int. J. Pharm. 2005, 292, 211–225. [CrossRef] [PubMed]
145. Chen, L.; Ding, X.; He, Z.; Huang, Z.; Kunnath, K.T.; Zheng, K.; Davé, R.N. Surface engineered excipients: I. improved functional
properties of fine grade microcrystalline cellulose. Int. J. Pharm. 2018, 536, 127–137. [CrossRef] [PubMed]
146. Ho, R.; Hinder, S.J.; Watts, J.F.; Dilworth, S.E.; Williams, D.R.; Heng, J.Y. Determination of surface heterogeneity of d-mannitol
by sessile drop contact angle and finite concentration inverse gas chromatography. Int. J. Pharm. 2010, 387, 79–86. [CrossRef]
[PubMed]
147. Shinebaum, R. Investigating the Mechanical Properties of Pharmaceutical Excipients in Granule and Tablet Production. Ph.D.
Thesis, Univerisity of Birmingham, Edgbaston, UK, 2021.
148. Morin, G.; Briens, L. The effect of lubricants on powder flowability for pharmaceutical application. AAPS PharmSciTech 2013,
14, 1158–1168. [CrossRef] [PubMed]
149. Lakio, S.; Vajna, B.; Farkas, I.; Salokangas, H.; Marosi, G.; Yliruusi, J. Challenges in detecting magnesium stearate distribution in
tablets. AAPS PharmSciTech 2013, 14, 435–444. [CrossRef] [PubMed]
150. Abe, H.; Otsuka, M. Effects of lubricant-mixing time on prolongation of dissolution time and its prediction by measuring near
infrared spectra from tablets. Drug Dev. Ind. Pharm. 2012, 38, 412–419. [CrossRef]
151. Kunnath, K.; Huang, Z.; Chen, L.; Zheng, K.; Davé, R. Improved properties of fine active pharmaceutical ingredient powder
blends and tablets at high drug loading via dry particle coating. Int. J. Pharm. 2018, 543, 288–299. [CrossRef]
152. Hentzschel, C.M.; Alnaief, M.; Smirnova, I.; Sakmann, A.; Leopold, C.S. Tableting properties of silica aerogel and other silicates.
Drug Dev. Ind. Pharm. 2012, 38, 462–467. [CrossRef]
153. Khunawattanakul, W.; Puttipipatkhachorn, S.; Rades, T.; Pongjanyakul, T. Novel chitosan-magnesium aluminum silicate
nanocomposite film coatings for modified-release tablets. Int. J. Pharm. 2011, 407, 132–141. [CrossRef] [PubMed]
154. Holman, J.; Tantuccio, A.; Palmer, J.; van Doninck, T.; Meyer, R. A very boring 120 h:15 million tablets under a continuous state of
control. Powder Technol. 2021, 382, 208–231. [CrossRef]
155. Tahara, K.; Kono, Y.; Myerson, A.S.; Takeuchi, H. Development of Continuous Spherical Crystallization to Prepare Fenofibrate
Agglomerates with Impurity Complexation Using Mixed-Suspension, Mixed-Product Removal Crystallizer. Cryst. Growth Des.
2018, 18, 6448–6454. [CrossRef]
156. Chen, Y.; Yao, Z.; Tang, S.; Tong, H.; Yanagishima, T.; Tanaka, H.; Tan, P. Morphology selection kinetics of crystallization in a
sphere. Nat. Phys. 2021, 17, 121–127. [CrossRef]
267
Pharmaceutics 2023, 15, 1587
157. Van Snick, B.; Holman, J.; Cunningham, C.; Kumar, A.; Vercruysse, J.; De Beer, T.; Remon, J.P.; Vervaet, C. Continuous direct
compression as manufacturing platform for sustained release tablets. Int. J. Pharm. 2017, 519, 390–407. [CrossRef]
158. Ierapetritou, M.; Muzzio, F.; Reklaitis, G. Perspectives on the continuous manufacturing of powder-based pharmaceutical
processes. AIChE J. 2016, 62, 1846–1862. [CrossRef]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual
author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to
people or property resulting from any ideas, methods, instructions or products referred to in the content.
268
MDPI
St. Alban-Anlage 66
4052 Basel
Switzerland
www.mdpi.com
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are
solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s).
MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from
any ideas, methods, instructions or products referred to in the content.
Academic Open
Access Publishing