Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

International Journal of Pharmaceutics 608 (2021) 121110

Contents lists available at ScienceDirect

International Journal of Pharmaceutics


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

Systematic study of paracetamol powder mixtures and granules


tabletability: Key role of rheological properties and dynamic image analysis
Oliver Macho a, Ľudmila Gabrišová a, Jana Brokešová b, Petra Svačinová b, Jitka Mužíková b,
Paulína Galbavá c, d, Jaroslav Blaško c, Zdenka Šklubalová b, *
a
Institute of Process Engineering, Faculty of Mechanical Engineering STU in Bratislava, Námestie slobody 17, 812 31 Bratislava 1, Slovak Republic
b
Charles University, Faculty of Pharmacy in Hradec Králové, Department of Pharmaceutical Technology, Ak. Heyrovského 1203, 500 05 Hradec Králové, Czech
Republic
c
Department of Analytical Chemistry, Faculty of Natural Sciences, Comenius University in Bratislava, Ilkovičova 6, 842 15, Bratislava 4, Slovak Republic
d
Hermes lab systems, Puchovska 12, 83106 Bratislava 3, Slovak Republic

A R T I C L E I N F O A B S T R A C T

Keywords: The aim of this systematic study was to analyze the granulometric and rheological behavior of tableting mixtures
Dynamic image analysis in relation to tabletability by single tablet and lab-scale batch compression with an eccentric tablet machine.
Powder rheology Three mixtures containing 33, 50, and 66% of the cohesive drug paracetamol were prepared. The high
Heckel analysis
compressibility of the powder mixtures caused problems with overcompaction or lamination in the single tablet
High shear granulation
Tabletability
compression method; due to jamming of the material during the filling of the die, the lab-scale batch compression
Paracetamol was impossible. Using high shear granulation, the flow properties and tabletability were adjusted. A linear
relationship between the span of granules and the specific energy measured by FT4 powder rheometer was
detected. In parallel, a linear relationship between conditioned bulk density and the tensile strength of the tablets
at lab-scale batch tableting was noted. The combination of dynamic image analysis and powder rheometry was
useful for predicting the tabletability of pharmaceutical mixtures during the single tablet (design) compression
and the lab-scale batch compression.

1. Introduction particle shape, surface area and structure, density, and interparticle
interactions, but also environmental factors such as humidity or elec­
Tablets currently represent the largest portion of all pharmaceutical trostatic charge as well as the equipment used also have to be considered
preparations. Due to the variable physicochemical properties, the (Hirschberg et al., 2019).
development of solid dosage forms is generally an empirical and time- Powder flow is not an inherent material property which means that
consuming process (Capece et al., 2015) as the resulting properties are no single descriptor enables a complete decription of flow behavior (Van
qualitatively and quantitatively influenced by all components (Patel and Snick et al., 2018). There are several static and dynamic methods for
Bansal, 2011). To produce pharmaceutical products of the highest determining the flow properties of powder materials and granules, such
quality, it is necessary to properly understand their composition and as determination of the angle of repose, flow through the orifice,
manufacturing process at the stage of dosage form development (Kalaria Hausner ratio, Carr index, avalanche behavior, and shear cell mea­
et al., 2020). surements (Blanco et al., 2020), (Ghadiri et al., 2020), (Tan et al., 2015).
Flowability and compressibility represent the key properties, Shear testers provide the most detailed description of the behavior of
particularly in the direct compaction process (Worku et al., 2017). The powder material on a particle–particle basis. Based on the relationship
flow and compaction properties of pharmaceutical materials are influ­ of normal and shear stress, many detailed characteristics can be esti­
enced by many factors: such as particle size and size distribution, mated from the yield curve by Mohr circle analysis (Carson and Wilms,

Abbreviations: AIF, Angle of internal friction; AIFE, Effective angle of internal friction; PEG, Polethylethylene glycol; API, Active pharmaceutical ingredient; PVP,
Polyvinylpyrrolidone; BFE, Basic Flowability Energy; SE, Specific Energy; CBD, Conditioned bulk density; SI, Stability index; DIA, Dynamic image analysis; WFA,
Wall friction angle; FRI, Flow Rate Index.
* Corresponding author.
E-mail address: sklubalova@faf.cuni.cz (Z. Šklubalová).

https://doi.org/10.1016/j.ijpharm.2021.121110
Received 23 June 2021; Received in revised form 7 September 2021; Accepted 14 September 2021
Available online 20 September 2021
0378-5173/© 2021 Elsevier B.V. All rights reserved.
O. Macho et al. International Journal of Pharmaceutics 608 (2021) 121110

2006). In addition to traditional shear testers, powder rheometers are the potential technological problems in the production process.
now increasingly used, the main advantage of which is to perform
various static and dynamic tests of the materials under investigation. 2. Materials and methods
Several studies have focused on comparing different types of rheometers
(Koynov et al., 2015), (Berry et al., 2015), (Slettengren et al., 2016) and 2.1. Materials
(Salehi et al., 2017). Rheometer FT4 (Freeman Technology) has become
often used in pharmaceutical preformulation studies (Wang et al., Paracetamol (Acetamidophenol, API, Glentham Life Science, UK)
2016a), (Wang et al., 2016b), (Chen et al., 2018a), (Majerová et al., was used as an Active Pharmaceutical Ingredient (API). Microcrystalline
2016), (Nan et al., 2017), (Tran et al., 2019). cellulose (Avicel® PH 101, FMC Biopolymer, Ireland) was used as a dry
Inadequate properties of tableting mixtures can result in problems in binder / filler, maltitol (Maltilite® P200 Pharma, Tereos, France) with
the manufacturing process such as feeding of material into a die during fine particles and sorbitol (Merisorb® 200, Tereos, France) with coarse
the compaction process, but further this also affects the properties of the particles were used as the sweetening fillers, and magnesium stearate
final product such as tensile strength, disintegration time or dissolution (Glentham Life Science, UK) as a lubricant. An aqueous solution of
performance (Megarry et al., 2019). Particularly whenever the active polyvinylpyrrolidone (Povidone K30, PVP, AppliChem, Germany) or
pharmaceutical ingredient is cohesive, the adjustment of its flowability polethylethylene glycol (PEG 8000, AppliChem, Germany) were used as
requires considerable effort in the development of solid dosage forms. granulation liquids. Acetonitrile (Merck, Germany), HMDS (1,1,1,3,3,3-
The analgetic and antipyretic drug paracetamol represents a widely used hexamethyldisilazane, Merck, Germany), trifluoroacetic acid (Fluo­
drug known for poor flowability, bad tabletability, and reduced plastic rochem, UK) and BSTFA (N, O-bis (trimethylsilyl) trifluoroacetamide,
deformation during compaction (Taipale-Kovalainen et al., 2018), (Rose Fluorochem, UK) were used in analytical trials for uniformity of content.
and Kaialy, 2019). In order to improve material flowability and tablet­
ability, various types of excipients are added, such as fillers, binders, 2.2. Methods
disintegrants, glidants or lubricants (Chen et al., 2018b).
If flow properties of tableting mixtures remain poor even after the All measurements and handling of powders, granules or tablets took
addition of excipients, dry or wet granulation processing is necessary. In place at an ambient temperature of 22 ± 2 ◦ C and a relative humidity of
the manufacturing of pharmaceuticals, high shear granulation (HSG) is 22 ± 3%. (608 - H1 Hygrometer Testo, China).
the most common method. The reduction of fine particle content
improving the flow properties, increase in density and homogeneity of 2.2.1. Preparation of powder mixtures
tablet components, and better tabletability are the primary aims. Even Individual substances were consecutively mixed in a cube mixer
though HSG of powder mixtures containing paracetamol has already KB15S (Erweka, Germany) at 30 rpm, for 10 min in total. Finally,
been investigated by several authors (Arp et al., 2011), (Albertini et al., magnesium stearate was added and mixed for a further 2 min. Three
2004), (Fayed et al., 2017), (Han et al., 2019), understanding material formulations of powder tableting mixtures F1 to F3 were prepared,
variability and variance in this unit operation is key to ensuring differing according to the concentration of the active pharmaceutical
adequate consistency of the final product (Stauffer et al., 2019) and ingredient Paracetamol: F1 (33% API), F2 (50% API), F3 (66% API). The
granulometric parameters of the tableting mixtures should be correctly composition of powder tableting mixtures is shown in Table 1.
characterized to avoid mechanical arching in the feeding hopper orifice.
The dynamic image analysis (DIA) method is attractive for its relative 2.2.2. Particle size analysis
low cost, measurement speed, and ability to provide a wide range of The particle size and size distribution of the individual powder
morphological parameters (Czajkowska et al., 2015), (De Simone et al., components of the mixtures were analyzed using a Malvern Mastersizer
2018), (Whiting et al., 2019). However, to date, no study has been 3000 Laser Diffraction Analyzer (Malvern Instruments, UK) by the dry
published analyzing the parameters obtained from the DIA in relation to cell method. Malvern software was used to determine d10, d50 and d90
the data obtained from the powder rheometer. values, which are the particle sizes corresponding to 10, 50 and 90% of
During compaction, the material generally passes through three the cumulative volume distribution curve, and Sauter diameter d32.
phases: an initial stage of particle rearrangement at low compaction Experiments were repeated 3 times and the average characteristic di­
pressure, plastic deformation and fragmentation at medium pressure, mensions are shown in Table 1.
and finally elastic deformation at high pressures (Heckel, 1961). In
addition, the tabletability of the proposed tableting mixtures should be 2.2.3. Preparation of granules by high shear granulation
necessarily identified during the development of new dosage forms and In the second step, the powder mixtures were used to prepare
particularly before the continuous tablet compaction process to avoid granules. An aqueous solution of polyvinylpyrrolidone, or polethyl­
economic loss. The use of batch tablet compaction currently represents a ethylene glycol respectively, both at concentrations of 1, 3 and 5% (w/
useful method. In the simplest experimental arrangement of compress­ w), were used as binders. The granules GF1 to GF3 were coded in
ibility measurement, the confined sample is exposed to uniaxial load and agreement with the powder mixture composition and the binder used as
the relation between the applied force and the piston position is recor­ illustrated: GF1 PVP 1% meaning the granules prepared from formula­
ded (Penkavova et al., 2019). Generally, the preliminary processes and tion F1 using 1% solution of povidone (Table 2).
data are enlarged empirically, experimentally validated and further High shear granulation of the mixtures was carried out in a batch in
optimized in a scaled-up process to achieve successful continuous pro­ an experimental granulator constructed by the author (Macho et al.,
duction and essential quality parameters of the final products (Bou­ 2020). The device is vertically oriented with a three-blade bottom
kouvala et al., 2012). driven impeller and a five-blade horizontal chopper. A constant chopper
In this regard, the aim of this systematic study was to investigate the speed of 1000 rpm was used throughout the granulation process. The
dependence between the granulometric and rheological parameters of batch of experimental material was 300 g.
paracetamol-containing powder mixtures and granules. Three types of Dry mixing of the powder mixture described above was performed
mixtures with percentage content of paracetamol 33% to 66% were for 2 minfollowed by wetting the mixture, both at an impeller speed of
prepared and evaluated by dynamic image analysis. The powder mix­ 300 rpm. The amount of binder solution was determined experimentally
tures and granules prepared by high shear granulation using two types of to produce granules of about 1–1.5 mm average size: 120 g of a binder
granulation liquids with different concentrations were compacted in the solution was used in the granulation of F1 mixture which represents a
single tablet design machine and the lab-scale eccentric tablet machine liquid to solid ratio L/S = 40%, 100 g or 40 g of a binder solution was
(batch compression) and analyzed by Heckel analysis in order to predict used for F2 (L/S = 33.3%) or F3 (L/S = 13.3%), respectively. Wet

2
O. Macho et al. International Journal of Pharmaceutics 608 (2021) 121110

Table 1
Composition of the powder mixtures and characteristic dimensions of the individual components for formulations F1 to F3.
Component Formulation Diameter

F1 F2 F3 d10(μm) d50(μm) d90(μm) d32(μm)

Paracetamol (g) 100 150 200 5.66 35.7 185 13.5


Microcrystalline cellulose (g) 137 87 37 19.8 57.4 132 39.5
Sorbitol (g) 30 30 30 118 232 399 186
Maltitol (g) 30 30 30 5.05 29.6 156 11.1
Magnesium stearate (g) 3 3 3 2.76 10.3 28.6 5.32

PartAn 3D (Macho et al., 2019).


Table 2
Summary of experimental conditions for prepared granules.
2.2.5. Rheological properties
Formulation Coding of Polyethylene glycol Polyvinylpyrrolidone A FT4 powder rheometer (Freeman Technology, UK) was used to
granules solution (%) solution (%)
determine the flow properties of powder mixtures and granules in
F1 GF1 PEG 1 agreement with ASTM-D7891-15,2015 standard (Shi et al., 2018). The
1%*
device consists of different measurement vessels and accessories such as
GF1 PEG 3% 3
GF1 PEG 5% 5
a blade, a vented piston and a shear head.The powder rheometer was
GF1 PVP 1 calibrated for force, torque and height. The powder mixtures F1 to F3
1%* and the prepared granules GF1 to GF3 were analyzed in the following
GF1 PVP 3% 3 tests i) compressibility, ii) shear cell test, iii) stability and variable flow
GF1 PVP 5% 5
rate, and iv) wall friction (Freeman, 2007). The tests were carried out in
F2 GF2 PEG 1
1%* a 50 mm diameter measuring vessel.
GF2 PEG 3% 3
GF2 PEG 5% 5 FT4 compressibility test. The batch of 60 g of powder material or 100 g of
GF2 PVP 1
granules, respectively, was filled into an 85-mL measuring vessel. The
1%*
GF2 PVP 3% 3 powders and granules were filled manually into the vessel. Prior to
GF2 PVP 5% 5 measuring the compressibility, 3 conditioning cycles were performed
F3 GF3 PEG 1% 1 using a 48 mm blade to achieve a homogeneously packed powder bed.
GF3 PEG 3% 3
The pretreated sample was split into a defined volume and the excess
GF3 PEG 5% 5
GF3 PVP 1% 1
material was carefully removed. Then, the mass of sample was recorded.
GF3 PVP 3% 3 Subsequently, the blade was replaced with a vented piston (diameter =
GF3 PVP 5% 5 48 mm) and the normal stress in a range of 0.5 to 15 kPa (9 tests) was
*
“blank” samples used for comparison. Explanation in text. applied to the sample. Compressibility was calculated from the change
in volume of tested sample after compression, expressed as a percentage
(Dudhat et al., 2017). Compressibility measurements were performed 3
massing was run for 2 min at 700 rpm, apart from GF3 PVP samples
times for each sample and the results were interpreted as the mean ±
where 800 rpm was used.
standard deviation.
Using a sieve with a diameter of 2 mm, excess locally formed coarse
granules (agglomerates) were separated immediately after the granu­
lation process. Coarse granules were only found in GF2 granules pre­ Shear test. The batch of 60 g of powder material or 100 g granules was
pared with 5% solution of PEG. The granules were then oven-dried at filled into a measuring vessel and the material was preconditioned by a
60 ◦ C for 24 h. conditioning cycle as decribed above. The shear measurement appa­
In total, 18 types of granules were prepared, i.e., 6 samples for each ratus, consisting of two 85 mL vessels, was the same as for compress­
powder formulation F1 to F3. Two batches of granules prepared from ibility measurement, but it was built with a serrated base to prevent the
powder mixtures F1 and F2 with the lowest concentration of PVP and entire sample from rotating as the blade passed through. Then, the
PEG, both having the mean size of about 1 mm, were used as the “blank” vented piston was used for preconsolidation of the sample bed and the
samples for the comparison reason (Table 2). excess material was removed.
Samples were examined under preconsolidation stress at preshear of
2.2.4. Dynamic image analysis 3, 6 and 9 kPa. The test consists of three steps: 1. shear head pre­
The size distribution of the dried granules was determined using a consolidation, 2. preshear until steady state achieved, and 3. shear test
PartAn 3D particle size and shape analyzer (Microtrac, Germany) by measurement (Freeman et al., 2009). Preshear was applied after each
dynamic image analysis (DIA). The granules were characterized volu­ shear measurement to achieve steady state shear stress and shearing
metrically by digitizing a 3D image of each falling granule particle photo processes. Shear test measurements were repeated 3 times for all
in agreement with ISO standards (13322–2, 9276–6). consolidation stresses. The measurement results were interpreted as
The Area Equivalent Diameter Da (μm) characterising the dimension mean ± standard deviation.
of a granule was calculated by (Eq. (1)): From the normal and shear stress points, a yield locus line was
generated using Mohr circle analysis and parameters such as cohesion,
Da = ((4 × A)/π )0.5 (1) major principal stress MPS (kPa), unconfined yield strength UYS (kPa),
angle of internal friction AIF (◦ ) and effective angle of internal friction
where A is the area of projected image (average from a sequence of
AIFE (◦ ) were derived. The flow function coefficient ffc (-) was calculated
3D images). The width of distribution (span) was estimated by Eq. (2)
according to Eq. (3). (Jenike, 1964).
where d10, d50 and d90 are the particle sizes as explained above.
ff c = MPS/UYS (3)
span = (d90 − d10 )/d50 (2)
Finally, the mean granular sphericity ϕmean was obtained using the Stability and variable flow rate test. Stability and variable flow rate tests

3
O. Macho et al. International Journal of Pharmaceutics 608 (2021) 121110

were performed using a 160 mL measurement vessel with a serrated base and the wall friction angle WFA (◦ ) was determined. The experiments
and an 85 mL split vessel. Prior to the actual test, the material bed was were repeated 3 times for each sample. The measurement results were
preconditioned by a blade. This was followed by splitting off the excess interpreted as mean ± standard deviation.
material ensuring an equal volume of all samples examined. The
conditioned bulk density CBD (g/mL) were calculated in regard to the 2.2.6. Single tablet compression
weight of the sample before and after the splitting off, and the known Single tablet compression experiments were performed on a Kistler
vessel volume of 85 mL. NCFN 60 (Kistler Eastern Europe s.r.o.) electromechanical press. The
The test consists of 11 cycles. The first seven cycles (stability test) are tablets were compacted in a steel jig formed by a die and a punch with a
accomplished by rotating the blade at 100 mm/s counter-clockwise diameter of 6 mm loaded to 100 ± 5 mg (Nimbus NBL 254i analytical
when the blade moves from top to bottom, and clockwise when the balance, readability 0.1 mg, Adam Equipment, UK) of each sample. The
blade moves upward. The following four tests were performed at blade tablets were compacted using six compaction pressures in a range of 50
tip speeds of 100, 70, 40 and 10 mm/s, respectively, to evaluate sensi­ to 300 MPa, with a punch speed of 5 mm/s. Ten flat faced tablets were
tivity to different flow rates. prepared from each sample, examined and analyzed further.
Changes in torque and force on the moving blade were recorded, and To characterize compressibility of material in the single tablet
together with the blade position information, were used to calculate the compression, in-die Heckel analysis Eq. (6) (Heckel, 1961) was applied.
flow energy E (mJ) to provide powder movement according to Eq. (4).
ln1/(1 − D) = kP + A (6)
∫ H( )
T
E= + F dH (4) where D is the relative density of the compact at pressure P and k is
R × tan α
0 the slope of the straight-line portion of the Heckel plot. A is a constant
where T is torque (mN/m), F is axial force (N), R is blade radius characterizing the rearrangement of particles. A yield pressure Py = 1/k
(mm), α is helix angle (◦ ) and H is penetration (mm) (Nan et al., 2017). (Heckel parameter of plasticity) was estimated.
Torque and force signals are represented as total flow energy, the energy The relative density D is defined as the ratio between the density of
required to move the blade through the sample from the top to the the compact at pressure P and the true density of the solid particles. The
bottom of the powder column. Because the values of torque and force are true density of powder samples was measured in duplicate using
constantly changing, it is necessary to calculate the energy for each Ultrapyc 1200e helium pycnometer (Quantachrome Ltd, UK). The
small distance travelled. The energy gradient is energy measured for pressure P and movement of punches were measured using TraceControl
each millimetre of blade travel. The area under the energy gradient V1.1.1.4.0 software.
curve yields a total energy (mJ), which value represents the resistance of
the powder to being made to flow in a dynamic state. 2.2.7. Lab-scale batch compression
Stability and variable flow rate measurements were performed 3 Lab-scale batch compression was accomplished using the eccentric
times with 85 g of the powder sample and 6 times with 140 g of the tablet press EP-1 (Erweka, Germany). By the position of the lower
granular material. The exception was sample GF3 (the smallest gran­ punch, the press was adjusted so that the resulting tablet mass was
ules) where 120 g was used. approximately 200 mg. The compaction pressure was set to 70 MPa.
Some other parameters characterizing the material behavior were Using a compaction rate of 13 tablets per minute, seventy flat faced
derived from the Stability and Variable Flow Rate tests such as the basic tablets of 7 mm diameter were prepared from each powder mixture and
flowability energy BFE (mJ) which is the energy calculated from the granule samples for further testing.
work that the blade has to perform when passing through the layer of
material to be examined from top to bottom during test No. 7, at a blade 2.2.8. Tensile strength of tablets
tip speed of 100 mm/s. In order to eliminate differences in densities of In accordance with the European Pharmacopoeia 10.0, tablet
granules and powder mixtures and to compare samples, the BFE value breaking force, the diameter and the height were determined using an 8
divided by the weight of the material examined was used (Lu et al., M hardness tester (Dr. Schleuniger Pharmatron). Using Eq. (7), the
2017). tensile strength TS (MPa) of 10 tablets produced by single tablet and lab-
The stability index SI (-), calculated as the ratio between the blade scale batch compression, respectively, was calculated (Fell and Newton,
energy in test No.7 and test No.1, and the specific energy SE (mJ/g) were 1970).
estimated. Unlike BFE, this measurement is made by moving the blade of
TS = 2 × FB /(π × d × h) (7)
material upwards in a clockwise direction (Tank et al., 2018). It char­
acterizes the flow of powder material at low load when its value is where FB (N) is the radial force to break the tablet, d (mm) is the
mostly affected by cohesion. The specific energy value SE per gram of diameter of the tablet and h (mm) is the height of the tablet.
material after the splitting process is calculated according to Eq. (5). The tensile strength measurement was carried out immediately after
production and repeated again 24 h after manufacture.
(UpEnergyTest 6 + UpEnergyTest 7)/2
SE = (5)
Split Mass 2.2.9. Uniformity of mass
The Flow Rate Index FRI (-) is calculated as the energy ratio of tests Ten tablets from each lab-scale batch compression were individually
No. 11 and test No. 8; it represents the powder sensitivity to changes in weighed using a Nimbus NBL 254i analytical scale. Results were
the tip speed of the blade from 100 to 10 mm/s as it passes down through expressed as mean ± standard deviation.
the material.
2.2.10. Uniformity of content
Wall friction. The batch of 60 g of powder material or 100 g granules, A validated analytical procedure was used to estimate the drug
respectively, was filled into an 85-mL measuring vessel. The material content. The tablet was carefully crushed and 10 mg of this powder were
batches were pre-consolidated to 3 kPa and the measurements were dissolved in 300 µL of acetonitrile (2 h at 50 ◦ C). Then, 300 µL of HMDS
carried out at the normal stress in a range of 3, 2, 1.75, 1.5, 1.25, and 1.0 and 2 µL of trifluoroacetic acid were added to the solution. The solution
kPa. Similarly, to the shear cell test, the measurement vessel with a was stirred and heated at 50 ◦ C for 30 min and, subsequently, 400 µL of
serrated base was used but a friction disc having a roughness of Ra = 1.2 BSTFA was added.
μm was applied instead of the shear head. By plotting the shear stresses The sample was centrifuged and the concentration of the drug in
and the associated normal stresses, a wall friction locus was obtained a sample supernatant was measured by flame ionization detector (GC-
FID) gas chromatography under the following conditions: Analysis was

4
O. Macho et al. International Journal of Pharmaceutics 608 (2021) 121110

performed on an Agilent Technologies 6850 Network GC System gas With the increase in the active pharmaceutical ingredient content,
chromatograph on a SPB-1 column (100% methyl siloxane) 24 m × 250 however, a reduction in Avicel® PH 101 content was necessary
µm × 0.25 µm. The injected sample volume was 1 μL into the injector (Table 1). While API concentrations of 33 % and 50 % (GF1, GF2)
operating in split mode (split ratio 10:1). Hydrogen with constant flow resulted in similar granule size trends, significant differences were
was used as a carrier gas. Inlet temperature was set to 280 ◦ C. The oven observed for GF3 with the highest (66%) API content. The lowest Avi­
temperature was set at 60 ◦ C and gradually increased to 310 ◦ C at a rate cel® content resulted in the lowest consumption of granulation liquid to
of 20 ◦ C/min and held for 4 min. form granules of the required size of aproximately 1 mm and it was
necessary to use the higher impeller speed (800 rpm). This produced
2.2.11. Data processing greater shear stresses which led to the formation of small granules in
OriginPro 9.0 software (OriginLab Corporation, USA) was used to GF3 PVP having a low value of the median particle size d50 (0.5–0.59
process the experimental data. Using this software, statistical regression mm). Even though the sphericity of the granules increased proportion­
analysis of the data and calculation of 95% confidence and prediction ally to binder concentration, the differences were small.
bands were performed. To allow the comparison of samples, the mean size diameter dmean
(mm) was expressed using the following formula (Eq. (8)). Percent sum
3. Results and discussion in class is related to number of particles.
∑ /
% in class × mid.class size
3.1. Dynamic image analysis of prepared granules dmean = 100 (8)
number of classes

The dried granules were analyzed using a PartAn 3D device for Apart from GF3 PVP 5% sample, the span for all granules decreased
particle size, the size distribution and the sphericity. DIA is a technique
that characterizes granules or particles in motion by digitizing photos of
each falling particle (Wei et al., 2020). The characteristic dimensions of
granules GF1 to GF3 in relation to binder concentration are shown in
Fig. 1.
Mean granule size increased with the increase in binder concentra­
tion regardless of its type in all samples. The higher viscosity of higher
concentration binder solution resulted in stronger liquid bridges and
larger granules were obtained. Granules formed by using PEG were
larger in most experiments than granules prepared by using PVP solu­
tions due to its higher viscosity, particularly in the case of the highest
concentration (5 % w/w). Similar results have been reported in the
granulation of micronized paracetamol with povidone solutions (El-
gindy et al., 1988). In the case of F1 granules comprising of the lowest
API content, the same mean granule size of 0.94 mm was detected with a
concentration of 1 % for both binders. These samples were used for
comparison reason with powder mixtures in further investigation as
discussed below.
Except for GF3 PVP, the width of the particle size distribution (span)
of the formed granules decreased with increasing binder concentration
Fig. 2. Dependence between size distribution width (span) and dmean for all 18
as illustrated in Fig. 1. In contrast, granules made from F3 by using PVP
types of granules.
solutions showed the largest span (1.72–2.14).

Fig. 1. Characteristic dimensions, sphericity and span of granules prepared from powder mixtures, depending on the concentration of the granulation liquid. The red
line represents the granule size of 1 mm.

5
O. Macho et al. International Journal of Pharmaceutics 608 (2021) 121110

in linear proportion with the increasing mean size as illustrated in Fig. 2. indicates better flowability of the granules.
The relationship was pronounced by regression: span = 2.25–1.033 *
dmean with the coefficient of determination R2 = 0.904. The results for 3.2.2. Shear test of powder mixtures and granules
GF3 PVP 5% sample were not included for the wider distribution due to During manipulation, the material is generally exposed to various
the larger particles that affected the d90 value. consolidation stresses that can cause changes in mechanical inter­
particulate forces and its density. To initiate the material flow, it is
necessary to overcome the yield point of the powder bed which is mainly
3.2. Rheology of the samples affected by the material interparticle forces resulting from the particle
size, shape, and surface area as well as the moisture content (Kunnath
To maintain the desired quality of the pharmaceutical product, it is et al., 2018). The external forces to which the material is exposed during
essential to pre-determine the flowability of particulate matter (Garg handling or processing can be examined with a FT4 rheometer using the
et al., 2018). Especially in direct tableting production, the good flow­ shear cell test. The measurements present data that permit a better un­
ability of tableting mixtures provides trouble-free manufacture of derstanding of the material flow characteristics derived by Mohr circle
products and results in good properties of tablets such as mass unifor­ analysis of the relationship between the normal and shear stresses
mity and dose accuracy (Takeuchi et al., 2018). As already mentioned, (Carson and Wilms, 2006).
rheometer FT4 has found many applications in the analysis of powder Shear properties of the powder mixtures F1 to F3 were evaluated.
and granular materials in the pharmaceutical industry particularly due The results obtained at consolidation stresses 3, 6, and 9 kPa are shown
to use of static and dynamic tests. in Table 3.
The values of major principal stress (MPS), representing the stress
3.2.1. Compressibility of powder mixtures and granules state during consolidation, had higher values with increasing API con­
A compressibility test measures how the density of a bulk powder tent. MPS values also increased with increasing the consolidation stress.
material changes as a function of applied stress. It is defined as the The angle of internal friction (AIF) characterizes the internal friction
percentage change in the volume of the sample. Although this test is not in the shear plane and it is a measure of the ability to withstand shear
a direct measure of flowability, it can indicate whether the material is stress. It represents the angle between the normal and shear force (the
free-flowing or cohesive (Freeman and Fu, 2008). In common, the linearized yield locus). The obtained values of AIF decreased with
cohesive powders show a larger change in the volume (and the density) increasing API content in the mixtures (Table 3). A higher value means
than the non-cohesive ones (Tran et al., 2021). The granular material more pronounced friction connected with the presence of interparticle
has generally lower compressibility, higher density and better flow interaction (Komínová et al., 2021). As Avicel® PH101 has a higher AIF
properties than bulk powder. value (38.1◦ ) compared to paracetamol (32.6◦ ), this phenomenon was
The compressibility test with a FT4 rheometer was used for all simply related to its content. Only a mild effect of the used consolidation
powder samples studied in this work. An increase in compressibility stress on the AIF values of powder samples was detected.
with the increasing paracetamol content was generally detected. For The effective angle of internal friction (AIFE) estimated from the line
example, the compressibility of 17.92 %, 22.58 %, and 23.99 % were running through the τ-σ graph origin tangentially to the larger Mohr
observed for powder mixtures F1, F2, and F3, respectively, at the applied stress circle showed a decreasing trend with increasing consolidation
load of 15 kPa. In comparison, the compressibility of pure paracetamol (preshear) stress (Rohilla et al., 2018), (Barletta and Poletto, 2019). The
and Avicel® PH 101 was 36.3% and 15.6%, respectively, at the same highest value of AIFE = 46.3◦ was observed at a load of 3 kPa in the case
load. of formulation F3 (Table 3).
A significant reduction in the compressibility achieved by high shear Table 3 summarizes also the values of cohesion estimated from the
granulation is illustrated in Fig. 3. The FT4 compressibility profile of intersection of the yield locus with the ordinate (Schulze, 2006). By
powder mixtures F1, F2, F3 is compared with the profile of “blank” GF1 increasing the API content in the powder mixture as well as by
granules (Table 2) having the same mean size dmean = 0.94 mm. Five increasing consolidation stress, the value of cohesion increased; the
times lower value in the case of GF1 PEG 1% (3.47 %) and even almost highest cohesion value 1.43 kPa was found for F3 at consolidation stress
eight times lower value for GF1 PVP 1% granules (2.32 %) were noted in of 9 kPa. By the use of a quality by design approach to study flow
contrast to the powder mixtures. This reduction in the compressibility behavior of pharmaceutical blends, (Wang et al., 2016b) detected a
linear relationship between cohesion and Unconfined Yield Strength
(UYS). Based on these findings, they proposed classification of the flow
properties of materials according to the value of cohesion. In agreement,
F1 mixture having the lowest content of paracetamol represents easy
flowing material (0.42 kPa < cohesion < 1.05 kPa) while powder
samples F2 at 9 kPa and F3 at 6 and 9 kPa are classified as cohesive
materials (1.05 < cohesion < 2.10).
In common, the material flow properties can also be classified using
the parameters of the flow function ffc generated from shear tests
(Schulze, 2006). At the lowest consolidation stress, the value of the ffc
for all powder mixtures is within the range for cohesive materials (2 <
ffc < 4). At higher consolidation stresses (6 and 9 kPa), F1 formulation
became easy flowing (4 < ffc < 10). F2 formulation remained cohesive at
6 kPa changing to easy flowing (ffc = 4.13) at the highest stress. F3
formulation with the highest paracetamol content was classified as a
fully cohesive material. Similarly, the decreased ffc values in relation to
increasing API content were observed in binary mixtures of paracetamol
and various excipients by (Chen et al., 2019) and (Capece et al., 2016).
However, as the ffc parameter does not include the density of the ma­
Fig. 3. The course of the compressibility test depending on the applied normal terial, some denser powders with lower ffc may flow better because they
stress for powder mixtures (F1, F2, F3) and granules GF1 prepared by the use of are more effectively influenced by gravity.
1% solutions of PEG or PVP (n = 3). In addition, the theoretical value of the size of hopper opening able

6
O. Macho et al. International Journal of Pharmaceutics 608 (2021) 121110

Table 3
Properties of experimental powder mixtures from Shear Cell Test and Stability and Variable Flow Rate. The table shows mean values ± standard deviation (n = 3).
Sample F1 F2 F3

Consolidation stress (kPa) 3 6 9 3 6 9 3 6 9

MPS (kPa) 6.47 ± 0.14 12.33 ± 0.23 18.47 ± 0.21 6.48 ± 0.14 12.8 ± 0.17 18.73 ± 0.15 6.82 ± 0.31 12.73 ± 0.25 18.90 ± 0.43
AIF (◦ ) 36.6 ± 1.5 37.4 ± 0.7 37.4 ± 0.4 36.7 ± 1.3 37.3 ± 0.5 36.4 ± 0.2 35.14 ± 1.00 35.3 ± 0.3 34.8 ± 0.3
AIFE (◦ ) 43.6 ± 0.7 42.2 ± 0.1 42.0 ± 0.2 45.2 ± 0.6 43.9 ± 0.2 42.6 ± 0.2 46.3 ± 0.9 43.8 ± 0.6 42.5 ± 0.5
Cohesion (kPa) 0.44 ± 0.04 0.59 ± 0.09 0.85 ± 0.04 0.53 ± 0.05 0.83 ± 0.07 1.14 ± 0.03 0.71 ± 0.02 1.06 ± 0.06 1.43 ± 0.09
ffc (-) 3.69 ± 0.29 5.26 ± 0.74 5.34 ± 0.18 3.10 ± 0.20 3.82 ± 0.22 4.13 ± 0.08 2.49 ± 0.01 3.12 ± 0.12 3.45 ± 0.15
BFE (mJ/g) 16.28 ± 0.44 13.99 ± 0.19 13.79 ± 0.82
SI (-) 0.94 ± 0.07 0.91 ± 0.06 0.94 ± 0.06
SE (mJ/g) 9.75 ± 0.21 9.96 ± 0.21 10.23 ± 0.50
FRI (-) 1.27 ± 0.10 1.41 ± 0.03 1.43 ± 0.02

F1, F2, F3 - powder mixtures, AIF - angle of internal friction, AIFE - effective angle of internal friction, ffc - flow function parameter, BFE - basic flowability energy, SI -
stability index, SE - specific energy, FRI - flow rate index

to produce arch free flow was calculated using Freeman software. For 3.2.3. Evaluation of stability and variable flow rate test
axi-symmetric hopper with 20◦ half angle, the detected values of hopper In addition, Table 3 summarizes the results of Stability and variable
outlet were 0.47 m (F1), 0.54 m (F2), and 0.74 m for F3 mixture. This flow rate test for powder mixtures F1 to F3. This test is used to describe
indicates the worsening of flow in context to the increase in API content. the material behaviour at different speeds of the blade and to predict
In Fig. 4, the graphical dependence of shear parameters (AIF, AIFE, possible problems due to attrition or segregation. Material is firstly
cohesion, and UYS) of powder mixtures on MPS is shown. With preconditioned by the passing of a blade through a sample bed to ach­
increasing the value of MPS, the parameter UYS and cohesion increased, ieve stability of flow energy; then, 11 measurement cycles (tests) follow
the values of AIF did not change significantly, on the contrary, the value and the energy changes are recorded (Freeman, 2007).
of AIFE decreased. As the AIFE is associated with adhesive forces The basic flowability energy (BFE) is mainly influenced by the
contribution to the total particle–particle surface contact friction and compressibility of the material. When moving downwards under a helix
interclocking and as this effect decreases with increase in major angle of 5◦ , the blade inclination affects the value of the BFE as the blade
consolidation stress, AIFE decreases in agreement (Garg et al., 2018). pushes the examined material against the bottom of the vessel. If the
Highly compressible and cohesive materials tend to cause problems material is more compressible, the BFE value is generally higher. In
when filling the dies as well as when compacting the powder mixtures contrast to cohesion, the BFE parameter decreased with increasing API
into tablets. In order to compare the results of shear test for powder content. This is pronounced with the value of BFE = 16.28 mJ/g for F1
mixtures and the granules, the “blank” samples GF1 (Table 2) were mixture as opposed to 13.79 mJ/ g for the more compressible and more
selected again. Although the AIF values for the granules GF1 PEG 1% cohesive F3 sample. The API particles form a packing state as the par­
and GF1 PVP 1% were the same (40.5◦ and 40.7◦ ) at a consolidation ticles settle on top of each other, resulting in the formation of air
stress of 9 kPa, at lower consolidation stresses they differed similarly to pockets. Due to this fact, the blade passes easily between the particles
powder mixtures. The AIF values 35.7◦ and 40.4◦ at 3 kPa and 6 kPa, which leads to a reduction in the BFE value (Trivedi and Dave, 2014).
respectively, were detected for GF1 PEG 1%, while 41.8◦ and 37.0◦ were Stability index SI (unitless) assesses the instabilities of the flow of the
registered for GF1 PVP 1% under the same consolidation stresses. De­ examined material, whether due to attrition, segregation, moisture ef­
viations resulted from different values of span due to the presence of fect or overblending with an additive, e.g., magnesium stearate. For the
polydisperse particles. Easy flowing behaviour was observed for GF1 powder mixtures, SI was calculated as the ratio between the blade en­
PEG 1% and GF1 PVP 1% with the cohesion values between 0.03 and ergy in Test No.7 and Test No.1. The differences between the values
0.25 kPa in the range of used consolidation loads (cohesion < 0.42 - were small. Based on the SI values within the range 0.9 < SI < 1.1
according to (Wang et al., 2016b) classification). This was also (Table 3), the investigated powder mixtures are characterized as stable,
confirmed by ffc values above 10 (35.31–50.11) for both granule not prone to particle segregation or attrition. However, it is not possible
samples. to exclude fragmentation of coarse particles during the measurement
that might not be detected by the test. The decrease in SI and BFE values
of powder mixtures compared to pure API was attributed to the presence
of magnesium stearate. (Dudhat et al., 2017) found that the presence of
0.5% magnesium stearate already reduced the BFE values and the spe­
cific energy SE (discussed below) over the entire range of paracetamol
concentrations.
Moving the blade through the material bed upwards in a clockwise
direction, specific energy SE (mJ/g) can be estimated (Tank et al.,
2018). The values characterize the flow of powder at low load and are
closely connected to material cohesion. While the SE values above 10
mJ/g are typical for highly cohesive material, material having 5 < SE <
10 mJ/g is moderately cohesive. As could be seen in Table 3, the values
of SE increased with the increase in content of paracetamol. The results
for SE measurement corresponded well to the aforementioned charac­
terization of powder mixtures based on the ffc and cohesion values. The
value of SE = 10.23 mJ/g detected for F3 shows possible problems in
manufacturing processes, such as die filling.
In order to characterize powder sensitivity to changes in the tip speed
of the blade, the flow rate index FRI (unitless) is estimated. At lower
rotation speeds, the energy that the blade has to expend to pass through
Fig. 4. Dependence of shear properties of shear parameters AIF, AIFE, cohesion
and UYS on MPS for powder mixtures F1, F2 and F3 (n = 3).
the material increases. Generally, the values FRI > 3 identify higher

7
O. Macho et al. International Journal of Pharmaceutics 608 (2021) 121110

material cohesion while values FRI < 1 are typical for powders con­
taining flow enhancers (Leturia et al., 2014). According to the FRI
values, all examined powder samples had average flow rate sensitivity
(Table 3). Similarly, to SE and cohesion results, FRI increased with
increasing API content. The cohesive powders are more compressible
due to the presence of a higher air content. Due to the air being easily
released around the blade, the powder bed becomes stiffer and more
resistant to flow (Freeman, 2007).
The results of Stability and variable flow rate test for granules GF1 to
GF3 are shown in Table 4 except for GF3 PEG granules showing similar
results to GF1 PVP due to the similar size distribution, and GF2 PEG 5%
with the largest granule size dmean = 1.6 mm.
The values of the BFE parameter of granules were greater than those
registered for the powder mixtures (Table 3). This phenomenon was
caused by an increase in conditioned bulk density (CBD, Table 4). When
compared with the F1 to F3 mixtures (CBD in the range of 0.412–0.432
g/mL), the blade movement was hindered by a particle bed having a
higher bulk density. The lowest CBD values were observed for GF3
granules which were also characterised with a low value of BFE. As the Fig. 5. Comparison of total energy during stability (1–7) and variable flow rate
BFE generally decreases with the decrease in the mean size of granules (8–11) tests for powder mixtures (F1, F2, F3), and granules GF3 formed with 1,
(Mohylyuk et al., 2019), the lowest value 16.84 mJ/g was found at GF3 3 and 5% solution of PVP. (n = 3 for powder mixtures, n = 6 for granules). Total
PVP 3% with the smallest granule size. Simultaneously, the particle size energy - area under the energy gradient curve yields.
influenced the BFE standard deviations (SD) at repeated measurements
as illustrated with the highest observed SD value of 10.71 mJ/g for the per mm of blade travel) and resistance of the powder material can be
largest granules (dmean = 1.42 mm) GF2 PVP 5%. The BFE results are estimated (Freeman et al., 2016). In contrast to powder mixtures F1 to
similar to the study of (Narang et al., 2016) who investigated granules F3, the increase in energy is visible for the GF3 that shows unstable flow
prepared with the use of 1, 3 and 5% hydroxypropyl cellulose solutions. of granular material, which would result in segregation during transport
The comparison of the total flow energy (mJ) that the blade had to and handling.
expend when passing through the bed of material during the first seven Stability index SI values were >1 for all granule samples; the highest
stability and variable flow rate tests is shown in Fig. 5 for powder average SI value 2.88 was observed for GF2 PVP 5% granules. This
mixtures and granules GF3 PVP (the smallest particle size). Measuring phenomenon was caused by particle segregation; the larger granules
the values of torque and force required, the energy gradient (the energy move upwards when the blade forces its way through the bed.
Conversely, the “blank” granules GF1 PVP 1% and GF1 PEG 1% showed
the same SI value. This resulted in the partial conclusion that the SI
Table 4 parameter increases with the mean size of the granules. The relationship
Stability and variable flow rate test parameters for selected granules. The table between specific energy SE and particle size distribution (span) is
shows mean values ± standard deviation (n = 6).
illustrated in Fig. 6. The negative linear proportion is pronounced by
Form. Granulation BFE CBD (g/ SI (-) FRI (-) SE (mJ/ coefficient of determination R2 = 0.85. As shown in a small inserted
liquid (mJ/g) mL) g) graph in contrast, SE increased with the mean particle size (dmean). The
GF1 PEG1% 21.27 ± 0.819 ± 1.47 ± 0.72 ± 7.77 ± smallest values of the SE parameter were found in the GF3 with the
0.16 0.006 0.06 0.05 1.01 smallest granules.
PEG3% 17.11 ± 0.777 1.18 ± 0.66 ± 7.30 ±
±
As summarized in Table 4, the values of SE > 10 mJ/g were mainly
1.09 0.007 0.24 0.19 0.24
PEG5% 26.59 ± 0.775 ± 1.57 ± 1.19 ± 11.09 ± registered for larger granules prepared with the highest concentration of
2.40 0.006 0.40 0.17 0.97 binder. In upward stability tests, the blade overcomes larger particles
PVP1% 25.82 ± 0.773 ± 1.47 ± 1.26 ± 11.85 ±
1.09 0.000 0.09 0.37 1.74
PVP3% 23.56 ± 0.764 ± 2.27 ± 1.02 ± 9.33 ±
2.23 0.011 0.34 0.08 1.79
PVP5% 21.50 ± 0.770 ± 1.93 ± 1.20 ± 11.93 ±
2.91 0.011 0.35 0.13 0.66
GF2 PEG1% 25.27 ± 0.699 ± 1.76 ± 1.02 ± 8.72 ±
2.64 0.006 0.18 0.08 1.14
PEG3% 31.07 ± 0.678 ± 2.20 ± 1.13 ± 13.11 ±
5.08 0.011 0.09 0.02 3.27
PVP1% 35.44 ± 0.721 ± 2.18 ± 0.94 ± 11.34 ±
8.50 0.010 0.22 0.02 1.55
PVP3% 28.89 ± 0.711 ± 2.29 ± 1.17 ± 11.79 ±
5.64 0.010 0.52 0.29 3.13
PVP5% 32.75 ± 0.707 ± 2.88 ± 1.14 ± 14.27 ±
10.71 0.010 1.21 0.17 1.17
GF3 PVP1% 18.47 ± 0.630 ± 1.57 ± 1.03 ± 6.49 ±
1.40 0.021 0.16 0.00 0.19
PVP3% 16.84 ± 0.623 ± 1.68 ± 1.04 ± 6.31 ±
1.54 0.016 0.11 0.04 0.47
PVP5% 17.10 ± 0.626 ± 1.39 ± 1.05 ± 6.29 ±
1.54 0.025 0.03 0.05 0.30

GF1, GF2, GF3 – granules prepared from powder mixtures F1 to F3, PEG -
polyethylene glycol, PVP - polyvinylpyrrolidone, BFE - basic flowability energy,
CBD - conditioned bulk density, FRI - flow rate index, SI - stability index, SE - Fig. 6. Dependence of the SE parameter on the span value and the dmean size of
specific energy the granules for all analyzed granule samples (n = 6).

8
O. Macho et al. International Journal of Pharmaceutics 608 (2021) 121110

which is more energy consuming. Sample GF3 PVP 5% was the excep­
tion as these granules were the smallest. Although the SE parameter is a
representation of flowability for non-constrained powder such as during
low-speed mixing or low-pressure filling, larger granules with larger
span can be affected by the segregation process as mentioned above.
The FRI values showing sensitivity to blade speed change were lower
for all granules when compared to those observed for powder mixtures.
This reflected the lower cohesion of the granules discussed above. As
generally cohesive powders are more sensitive to changes in flow rate
(Leturia et al., 2014), the granular material shows lower sensitivity. FRI
values < 1.5 reflect materials containing, for example, larger particles or
particles surface-treated in order to improve the flow properties.

3.2.4. Evaluation of wall friction


The measurement of wall friction characterizes the relationship of
frictional forces between the particulate material and the surface of the
container in which the material is stored or processed. The wall friction
head of FT4 rheometer induces the required normal stress in the powder
bed and then slowly shifts to induce shear stresses. Fig. 7. Heckel plot for F1 powder mixture and granules GF1 PVP 1% and GF1
The wall friction angle (WFA) of the investigated materials was PEG 1% (only one curve from each compacted material was selected
measured using a stainless steel friction disc with a roughness of Ra = for clarity).
1.2 μm. For powder mixtures F1 to F3, WFA decreased with the increase
in the API content (26.37◦ , 24.03◦ , and 23.23◦ , respectively). This during the relaxation phase of the compaction process, higher tensile
phenomenon indicates lower resistance and adhesion between the ma­ strengths of tablets can be measured (Davis et al., 2018).
terial and the friction disc due to the higher paracetamol content. Ac­ However, it is difficult to completely separate the contributions of
cording to the WFA values, all mixtures can be classified as moderately plastic and elastic deformation mechanisms to the strength of a compact;
flowable. In agreement with (Søgaard et al., 2014) study, the WFA value the strength is mainly influenced by the formation of physical bonds
of granules decreased with increasing particle size of materials. For GF1 between particles due to plastic deformation (Hiestand, 1997). Using
PEG 1% granules with dmean = 0.94 mm, the WFA value was = 23.27◦ . Eq. (6), yield pressures Py = 1/k (Heckel parameter of plasticity)
reflecting the particle deformability during the compression were
3.3. Single tablet compression calculated. In Fig. 8, the results for above mentioned samples F1, F2,
GF1 PVP 1%, and GF1 PEG 1% are shown. Samples F3 were not analyzed
Preparation of tablets includes two primary steps: filling a die with due to significant lamination over the entire range of compaction pres­
material, and the compaction. During the second step, a compact having sures (Fig. 9a). The Py values increased with compaction pressure. As the
suitable properties such as mechanical resistance and tensile strength electromechanical press in single tablet compression allows the regis­
should be obtained. As the quality of the final product is highly sensitive tration of all results even though no tablets of sufficient hardness could
to manufacturing scale (Ghazi et al., 2019), tablets were prepared by be prepared, the results of the F2 mixture compaction at pressures of 50
two different methods: i) a single tablet compression method illustrating and 100 MPa are shown.
production of one tablet and ii) a lab-scale compression method illus­ A lower yield pressure generally indicates a higher plasticity of the
trating small scale, partially continuous production. material (Sun and Grant, 2001). In agreement with generally accepted
Direct compaction of powder tableting mixtures presents a relatively plasticity of MCC (Desai et al., 2018), the lower Py value was observed
fast and efficient manufacturing process. However, poorly compactible for powder mixture F1 with a higher MCC content (i.e. lower amount of
materials, such as e.g. those containing paracetamol (Taipale-Kovalai­ paracetamol) in comparison to F2 mixture. The exception is only at the
nen et al., 2018), can produce serious problems. Compressibility of the highest compaction pressures when the F1 mixture became
powder mixtures F1 to F3 and the selected granules, GF1 and GF2, was
first investigated in the single tablet compression experiment using
pressures in a range between 50 and 300 MPa.

3.3.1. Compressibility study by Heckel analysis


Three phases of the compaction process are generally referred to: i)
an initial stage of particle rearrangement, ii) plastic deformation and
fragmentation, and iii) elastic deformation. The consolidation mecha­
nisms of powder and granule samples were studied in this work by in-die
Heckel analysis (Heckel, 1961).
The differences between the Heckel curve observed for F1 and the
GF1 PVP 1% and GF1 PEG 1% “blank” granules are illustrated in Fig. 7.
In general, the significantly easier rearrangement of particles (detail in
inserted graph) and transition to the linear region of plastic deformation
are visible for the powder mixtures. Contrarily, the granular material
underwent plastic deformation easily and rapidly. During the
compression, the granules were fractured into small ones which facili­
tated further rearrangement and, moreover, this resulted in a better
formation of physical bonds between the particles leading to an increase
in the tablet strength as will be discussed later. In addition, the differ­
ences in area of elastic deformations for the powder mixtures and Fig. 8. Comparison of yield pressure for powder mixtures (F1, F2) and granules
granules were observed in Fig. 7. If elastic deformations are not released (GF1, GF2) prepared with 1% solutions of PEG and PVP, respectively (n = 10).

9
O. Macho et al. International Journal of Pharmaceutics 608 (2021) 121110

In the powder mixtures F1 - F3, the compaction process was influ­


enced by the API content. During compaction of the powder tableting
mixture F1 (the lowest amount of paracetamol), the tensile strength of
the tablets increased with increasing compaction pressure up to 250
MPa, when the tablet strength value was 1.75 MPa (Fig. 9a). By further
increasing the compaction pressure, overcompaction occurred reducing
the tensile strength. Similar overcompaction at higher compaction
pressures has been reported by (Li et al., 2019), and (Cheng et al., 2020).
As mentioned above, it was not possible to produce tablets up to a
compaction pressure of 150 MPa from the F2 formulation (50 % of
paracetamol) and the low tensile strength value of 0.55 MPa was
detected even at a compaction pressure of 300 MPa. Compaction of F3
mixture (highest content of paracetamol) was accompanied by a sig­
nificant lamination of tablets over the entire range of compaction
pressures (Fig. 9a) and such tablets were too brittle for hardness testing.
Paracetamol is widely known to be elastic in compression behaviour
thus cause capping in tabletting (Anuar and Briscoe, 2010) and our re­
sults were consistent with this. The tablets showing elastic expansion
can be easily damaged during ejection.
In Fig. 9b and 9c, the results for granules GF1 and GF2 prepared
using 1% solutions of PVP and PEG are shown. The processing of the
material by wet granulation did not show a significant influence on the
tabletability of formulation F1. However, the overcompaction of the
sample at compaction pressures of 300 MPa was avoided. The tensile
strength of tablets made with GF1 PVP 1% was higher (2.46 MPa) than
for GF1 PEG 1 % (1.61 MPa) at 300 MPa which is the result of larger
particle size and lower value of Py for GF1 PVP granules.
The positive effect of the granulation on the tabletability and tensile
strength of the tablets was observable over the entire range of
compaction pressures for GF2 when compared with the F2 mixture.
Material was already compressible at a compaction pressure of 50 MPa.
The tensile strength of GF2 PVP 1% tablets was higher than of GF2 PEG
Fig. 9. Tabletability of powder mixtures and granules, a-powder mixtures (F1 1% and similar results were detected for other used concentrations of
to F3), b-comparison of F1 and granules GF1 prepared by 1% solutions of PVP binders.
and PEG, c- comparison of F2 and granules GF2 prepared by 1 % PVP and PEG The results are consistent with the Heckel analysis decribed above.
(n = 10). The lower tensile strength of tablets corresponded to the higher Py
values for powder materials. This phenomenon was caused by the fact
overcompressed as will be discussed in Chapter – 3.4.2. that powder materials were able to achieve greater packing in the die
The lower Py values were generally registered for granules showing and hence, exhibited lower movement as compaction pressure was
higher plasticity (Sun and Grant, 2001). At low compaction pressure (50 applied (Patel et al., 2006). The increase in the tensile strength is
MPa), the differences in Py were not observable but they reached generally due to the plastic deformation and formation of interparticle
considerable levels at compression pressures above 100 MPa. The values bonds. This was confirmed for PVP granules. The Heckel plot showed
of Py decrease with the increasing size of the compressed particles significant differences in the compressibility of powder mixtures and
(Cabiscol et al., 2020). The Py values were lower for the PVP granules granular materials due to fragmentation, attrition and particle rear­
than those produced by PEG solutions in most experiments. The fluc­ rangement. A more detailed description of the relationship between
tuation in Py values resulted from the actual differences in the sphericity tensile strength and the Py parameter can be found in the work (Ghori,
and width of the distribution of the granules prepared (Veronica et al., 2016).
2018).
3.4. Lab-scale batch compression
3.3.2. Tabletability of powder mixtures and granules
Tablets should have a suitable mechanical resistance to withstand In order to illustrate batch tablet production, the behaviour of the
subsequent processing, transport and handling by the patients (Reynolds tabletting material during the lab-scale compression was monitored
et al., 2017). Two tests are recommended to determine the mechanical using an eccentric Erweka EP1 automatic press. For the powder mixtures
strength of tablets, namely the friability of uncoated tablets and the F1 - F3, some problems were anticipated due to their cohesive behav­
resistance to crushing based on the diametral compression test (Juban iour, the high compressibility and poor flow properties (Table 3). As
et al., 2015). The latter was used in this study; the breaking force was expected, fast jamming occured when filling the die with the F1
measured immediately after production and the tensile strength of formulation. Even though this powder mixture showed the lowest
tablets was calculated using Eq. (7). The tensile strength of tablets cohesion value and a moderate value of WFA, the formation of an arch in
prepared from granules was estimated immediately after production and the feeding hopper (Fig. 10a) and the shoe (Fig. 10b) was observed.
repeated again after 24 h. As the changes in tensile strengths after 24 h Similar behaviour was observed in F2 and F3. As only a few tablets could
were below 0.5%, the immediatelly estimated values are used instead. be formed from each powder sample they were not analyzed further.
The tensile strengths (MPa) obtained are illustrated in Fig. 9 for The better flow properties of the GF1 to GF3 granules positively
powder mixtures F1 - F3; the results for F1 and F2 are also compared to affected the batch compression and confirmed the general importance of
granules GF1 or GF2 prepared with 1% solutions of PEG and PVP, the high shear granulation process. In addition to the HSG, however, it
respectively. should be mentioned that direct compression of paracetamol can be
done if the formulation consists of suitable directly compressible grades

10
O. Macho et al. International Journal of Pharmaceutics 608 (2021) 121110

(0.54 MPa) having the same mean particle size dmean = 0.94 mm. In
contrast, the same tablet tensile strengths 1.06 MPa and 1.09 MPa were
noted for GF2 PEG 1% and GF2 PVP 1%, respectively. Higher tensile
strength in tablets of GF2 PVP sample was also recorded in single tablet
compression (Fig. 9).
As can be seen in Fig. 11, the highest tensile strength was detected at
GF3 PVP 5% granules (2.029 MPa). Eichie and Kudehinbu, 2009
observed that tensile strength increases with the decrease in granule
size. Our results were in good agreement as the particle size dmean =
0.63 mm was the smallest out of all granule samples.
The tensile strength of the tablets in lab-scale batch compression was
mainly influenced by the CBD values. In Fig. 12, the linear relationship
between CBD of granules GF1 to GF3 and tablet tensile strength char­
acterized with the coefficient of determination R2 = 0.86 is illustrated.
As CBD is related to particle size, it confirmed again that GF3 samples
with the smallest particles (Fig. 1) and lowest CBD values (Table 4)
showed higher tensile strength values. In contrast, GF1 samples with
higher CBD values achieved lower tensile strength values.
Apart from dmean, it has to be noted, that the particle size distribution
influenced the tensile strength values as well. Tablets compacted from
Fig. 10. Arching of F1 mixture in lab-scale batch compression with an eccentric
the small granules (GF3) with the wider size distribution showed the
tablet machine: a - top view of the feeding hopper, b - bottom view of the wiper
shoe of the tablet press. higher values of tensile strength at the same compaction pressures as
these small granules not only filled the die space quickly, but also easily
penetrated into interparticle voids. However, this observation could not
of excipients that are both fragmenting and plastic.
be applied to all experiments.
The experiments with powder material showed significant differ­
ences between single tablet and lab-scale batch compression. Since the
3.4.1. Mass and content uniformity of lab-scale batch produced tablets
material was directly poured into the die during single tablet (design)
Ten tablets of a targeted mass of 200 mg, were weighed and the mean
compression, the inapropriate flow properties of samples influenced the
and standard deviations were expressed. A smallest difference in mass
compaction process only occasionally such as e.g., at higher concen­
uniformity (0.39 %) was detected for tablets produced from GF1 PEG 3%
trations of API (lamination). Contrarily, good flow properties are a key
with lowest CBD. With the increasing span of granules, the mass uni­
parameter for the proper operation of the continuous tableting process.
formity of tablets decreased; the largest deviations (2.07 %) were
(Wu et al., 2003) found that the level of die filling can also be affected by
detected with tablets prepared from GF3 PVP 5% granules.
the speed of filling itself. In order to evaluate the properties of samples,
Lab-scale batch prepared tablets from GF1 granules with a target
the same filling speed was used in all compaction experiments. The re­
value 33 % of paracetamol were also analyzed for content uniformity by
sults obtained at a compaction pressure of 70 MPa are shown in Fig. 11.
gas chromatography. (Muthancheri and Ramachandran, 2020) found
The values increased with the content of binder used in the HSG
that the size and size distribution of the granules, significantly affected
similarly to the work (Autamashih et al., 2011). In the compaction of
content uniformity of paracetamol tablets. Our results were consistent
GF1, a higher tensile strength was observed for tablets obtained for GF1
with this as uniformity in API content decreased with increasing dmean of
PEG compared to GF1 PVP granules. This is illustrated by comparing
granules. The smallest differences in API content (0.2 %) were detected
“blank” GF1 PEG 1% granules (0.74 MPa) and GF1 PVP 1% granules
for both granules GF1 PVP 1% and GF1 PVP 3% contrary to GF1 PEG 1%
and GF1 PEG 3% granules showing 2.7% deviation in API content. In
contrast, the largest deviations 5.99 % and 5.77% were registerred for
GF1 PEG 5% granules and GF1 PVP 5%, respectively.

Fig. 11. Tensile strength of lab-scale batch compacted tablets produced from
selected granules GF1, GF2, and GF3 at compaction pressure of 70 MPa (n Fig. 12. Dependence of tensile strength on CBD in lab-scale batch compression
= 10). of granules GF1, GF2 and GF3.

11
O. Macho et al. International Journal of Pharmaceutics 608 (2021) 121110

4. Conclusions the work reported in this paper.

Good flow properties of particulate matter are important to avoid


problems in routine tablet production. Based on the results of dynamic Acknowledgements
image analysis as well as the static and dynamic tests available with
powder rheometer, this article describes the flow properties, The authors would like to thank IMCD Czech Republic, s.r.o. Prague,
compressibility and tabletability of three powder tableting mixtures for their cooperation in the research and Grant Scheme to support
containing various concentrations of paracetamol (33–50 − 66 %) and excellent teams of young researchers under the conditions of the Slovak
the equivalent granules prepared by high shear granulation. Despite the University of Technology in Bratislava.
addition of magnesium stearate, poor flow properties associated with
high compressibility and cohesion were observed for powder mixtures.
The higher the content of API the worse properties were detected. This Funding
negatively influenced both methods of tablet production leading to
overcompaction or lamination in the single tablet compression or to This work was supported by the Funding Agency of Charles Uni­
arching and blockage of the feeding hopper in the lab-scale batch versity [Grant No. 268120/2020; Grant No. SVV 260 547]; European
compression with an eccentric tablet machine. Regional Development Fund [ITMS 26240220084], [313021X329];
Using high shear granulation, granulometric and flow properties of Slovak Research and Development Agency [APVV-18-0348], [APVV-18-
mixtures were positively affected. Even though rheology parameters 0282], [KEGA 016STU-4/2019, KEGA 036STU-4/2020].
worsened generally with the increasing in API content, the reduction in
compressibility and cohesivity resulted in trouble-free single tablet as References
well as lab-scale batch tablet production for all formulations using
compaction pressures in a range of 50–300 MPa. The differences be­ Albertini, B., Cavallari, C., Passerini, N., Voinovich, D., González-Rodrı́guez, M.L.,
Magarotto, L., Rodriguez, L., 2004. Characterization and taste-masking evaluation of
tween powder mixtures and granules were demonstrated by in die
acetaminophen granules: Comparison between different preparation methods in a
Heckel analysis. The lower Py values registered for granules showed high-shear mixer. Eur. J. Pharm. Sci. 21 (2-3), 295–303. https://doi.org/10.1016/j.
their higher plasticity, particularly at compaction pressures above 100 ejps.2003.10.017.
Anuar, M.S., Briscoe, B.J., 2010. Detrimental consequences of the Paracetamol tablet
MPa.
elastic relaxation during ejection. Drug Dev. Ind. Pharm. 36 (8), 972–979. https://
The results for granules were interpreted with respect to the particle doi.org/10.3109/03639041003610807.
morphology obtained from dynamic image analysis. The granule size Arp, Z., Smith, B., Dycus, E., O’grady, D., 2011. Optimization of a high shear wet
increased proportionally with the increase in binder (either PEG or PVP) granulation process using focused beam reflectance measurement and particle vision
microscope technologies. J. Pharm. Sci. 100 (8), 3431–3440. https://doi.org/
concentration in a range of 1–5 % while the width of sample particle size 10.1002/jps.22556.
distribution (span) decreased. Dependencies were found between the Autamashih, M., Isah, A.B., Allagh, T.S., 2011. Compressional characteristics and drug
parameters obtained from the DIA and the data from the powder release profile of tablets of the crude leaves extract of Vernonia galamensis. Int. J.
Green Pharm. https://doi.org/10.4103/0973-8258.82095.
rheometer, as well as the data from the rheometer on the compression Barletta, D., Poletto, M., 2019. A new split cell for the measurement of the horizontal to
parameters. A linear relationship was detected between the span of vertical stress ratio of powders. Powder Technol 351, 273–281. https://doi.org/
granules and the specific energy (R2 = 0.85). This parameter charac­ 10.1016/j.powtec.2019.04.041.
Berry, R.J., Bradley, M.S.A., McGregor, R.G., 2015. Brookfield powder flow tester -
terizes the flow of material at low load such as e.g., die filling; the Results of round robin tests with CRM-116 limestone powder. Proc. Inst. Mech. Eng.
smaller the granules the smaller the values of the SE were observed. In Part E J. Process Mech. Eng. https://doi.org/10.1177/0954408914525387.
agreement with the increasing particle size, the Py values obtained by Blanco, D., Antikainen, O., Räikkönen, H., Mah, P.T., Healy, A.M., Juppo, A.M.,
Yliruusi, J., 2020. Image-based characterization of powder flow to predict the
Heckel analysis were mostly lower for the PVP granules than for those
success of pharmaceutical minitablet manufacturing. Int. J. Pharm. 581, 119280.
produced by PEG solutions. A linear relationship with the coefficient of https://doi.org/10.1016/j.ijpharm.2020.119280.
determination R2 of 0.86 was displayed between the tensile strength of Boukouvala, F., Niotis, V., Ramachandran, R., Muzzio, F.J., Ierapetritou, M.G., 2012. An
integrated approach for dynamic flowsheet modeling and sensitivity analysis of a
the tablets at lab-scale batch compression and CBD values of GF1 to GF3
continuous tablet manufacturing process. Comput. Chem. Eng. 42, 30–47. https://
granules. This also clearly demonstrated the influence of particle size; doi.org/10.1016/j.compchemeng.2012.02.015.
GF3 samples with the lowest CBD values and smallest particle size Cabiscol, R., Shi, H., Wünsch, I., Magnanimo, V., Finke, J.H., Luding, S., Kwade, A.,
showed the higher tensile strength values as opposed to GF1 granules 2020. Effect of particle size on powder compaction and tablet strength using
limestone. Adv. Powder Technol. 31 (3), 1280–1289. https://doi.org/10.1016/j.
with higher CBD and lower tensile strength values. The importance of apt.2019.12.033.
particle size was finally demonstrated during the evaluation of the mass Capece, M., Ho, R., Strong, J., Gao, P., 2015. Prediction of powder flow performance
and content uniformity. With an increase in the span of granules, the using a multi-component granular Bond number. Powder Technol. 286, 561–571.
https://doi.org/10.1016/j.powtec.2015.08.031.
mass uniformity of tablets decreased, and simultaneously with an in­ Capece, M., Silva, K.R., Sunkara, D., Strong, J., Gao, P., 2016. On the relationship of
crease in the size of granules, uniformity in content decreased. inter-particle cohesiveness and bulk powder behavior: Flowability of pharmaceutical
powders. Int. J. Pharm. 511 (1), 178–189. https://doi.org/10.1016/j.
ijpharm.2016.06.059.
CRediT authorship contribution statement Carson, J.W., Wilms, H., 2006. Development of an international standard for shear
testing. Powder Technol. 167 (1), 1–9. https://doi.org/10.1016/j.
Oliver Macho: Conceptualization, Writing – original draft, Investi­ powtec.2006.04.005.
Chen, L., Ding, X., He, Z., Fan, S., Kunnath, K.T., Zheng, K., Davé, R.N., 2018a. Surface
gation, Methodology. Ľudmila Gabrišová: Visualization, Investigation,
engineered excipients: II. Simultaneous milling and dry coating for preparation of
Conceptualization. Jana Brokešová: Data curation, Investigation. fine-grade microcrystalline cellulose with enhanced properties. Int. J. Pharm. 546
Petra Svačinová: Writing – review & editing, Methodology. Jitka (1-2), 125–136. https://doi.org/10.1016/j.ijpharm.2018.05.019.
Chen, L., Ding, X., He, Z., Huang, Z., Kunnath, K.T., Zheng, K., Davé, R.N., 2018b.
Mužíková: Writing – review & editing, Methodology. Paulína Galbavá:
Surface engineered excipients: I. improved functional properties of fine grade
Data curation, Investigation. Jaroslav Blaško: Data curation, Formal microcrystalline cellulose. Int. J. Pharm. 536 (1), 127–137. https://doi.org/
analysis. Zdenka Šklubalová: Supervision, Conceptualization, Writing 10.1016/j.ijpharm.2017.11.060.
– review & editing. Chen, L., He, Z., Kunnath, K.T., Fan, S., Wei, Y., Ding, X., Zheng, K., Davé, R.N., 2019.
Surface engineered excipients: III. Facilitating direct compaction tableting of binary
blends containing fine cohesive poorly-compactable APIs. Int. J. Pharm. 557,
Declaration of Competing Interest 354–365. https://doi.org/10.1016/j.ijpharm.2018.12.055.
Cheng, H., Wei, Y., Wang, S., Qiao, Q., Heng, W., Zhang, L., Zhang, J., Gao, Y., Qian, S.,
2020. Improving Tabletability of Excipients by Metal-Organic Framework-Based
The authors declare that they have no known competing financial Cocrystallization: a Study of Mannitol and CaCl2. Pharm. Res. 37 (7) https://doi.
interests or personal relationships that could have appeared to influence org/10.1007/s11095-020-02850-8.

12
O. Macho et al. International Journal of Pharmaceutics 608 (2021) 121110

Czajkowska, M., Sznitowska, M., Kleinebudde, P., 2015. Determination of coating Kunnath, K., Huang, Z., Chen, L., Zheng, K., Davé, R., 2018. Improved properties of fine
thickness of minitablets and pellets by dynamic image analysis. Int. J. Pharm. 495 active pharmaceutical ingredient powder blends and tablets at high drug loading via
(1), 347–353. https://doi.org/10.1016/j.ijpharm.2015.08.102. dry particle coating. Int. J. Pharm. 543 (1-2), 288–299. https://doi.org/10.1016/j.
Davis, M.T., Potter, C.B., Walker, G.M., 2018. Downstream processing of a ternary ijpharm.2018.04.002.
amorphous solid dispersion: The impacts of spray drying and hot melt extrusion on Leturia, M., Benali, M., Lagarde, S., Ronga, I., Saleh, K., 2014. Characterization of flow
powder flow, compression and dissolution. in: Particle Technology Forum 2018 - properties of cohesive powders: A comparative study of traditional and new testing
Core Programming Area at the 2018 AIChE Annual Meeting. methods. Powder Technol. 253, 406–423. https://doi.org/10.1016/j.
De Simone, V., Caccavo, D., Lamberti, G., d’Amore, M., Barba, A.A., 2018. Wet- powtec.2013.11.045.
granulation process: phenomenological analysis and process parameters Li, Y., Yu, J., Hu, S., Chen, Z., Sacchetti, M., Sun, C.C., Yu, L., 2019. Polymer Nanocoating
optimization. Powder Technol. 340, 411–419. https://doi.org/10.1016/j. of Amorphous Drugs for Improving Stability, Dissolution, Powder Flow, and
powtec.2018.09.053. Tabletability: The Case of Chitosan-Coated Indomethacin. Mol. Pharm. 16 (3),
Desai, P.M., Anbalagan, P., Koh, C.J.N., Heng, P.W.S., Liew, C.V., 2018. Evaluation of 1305–1311. https://doi.org/10.1021/acs.molpharmaceut.8b0123710.1021/acs.
tablet punch configuration on mitigating capping by a quality by design approach. molpharmaceut.8b01237.s001.
Drug Deliv. Transl. Res. 8 (6), 1635–1643. https://doi.org/10.1007/s13346-017- Lu, X.Y., Chen, L., Wu, C.Y., Chan, H.K., Freeman, T., 2017. The effects of relative
0425-0. humidity on the flowability and dispersion performance of lactose mixtures.
Dudhat, S.M., Kettler, C.N., Dave, R.H., 2017. To Study Capping or Lamination Tendency Materials (Basel). 10, 1–9. https://doi.org/10.3390/ma10060592.
of Tablets Through Evaluation of Powder Rheological Properties and Tablet Macho, O., Čierny, M., Gabrišová, L., Juriga, M., Ružinský, R., Peciar, P., 2019. Dynamic
Mechanical Properties of Directly Compressible Blends. AAPS PharmSciTech 18 (4), Image Analysis to Determine Granule Size and Shape, for Selected High Shear
1177–1189. https://doi.org/10.1208/s12249-016-0576-1. Granulation Process Parameters. Stroj. Cas. https://doi.org/10.2478/scjme-2019-
Eichie, F.E., Kudehinbu, A.O., 2009. Effect of particle size of granules on some 0043.
mechanical properties of paracetamol tablets. African J. Biotechnol. https://doi.org/ Macho, O., Kabát, J., Gabrišová, Ľ., Peciar, P., Juriga, M., Fekete, R., Galbavá, P.,
10.5897/AJB09.859. Blaško, J., Peciar, M., 2020. Dimensionless criteria as a tool for creation of a model
El-gindy, N.A., Samaha, M.H., El-maradny, H.A., 1988. Evaluation of binder activities on for predicting the size of granules in high-shear granulation. Part. Sci. Technol. 38
the physical properties and compression characteristics of granules prepared by two (3), 381–390. https://doi.org/10.1080/02726351.2018.1548531.
different modes. Drug Dev. Ind. Pharm. 14 (7), 977–1005. https://doi.org/10.3109/ Majerová, D., Kulaviak, L., Růžička, M., Štěpánek, F., Zámostný, P., 2016. Effect of
03639048809151915. colloidal silica on rheological properties of common pharmaceutical excipients. Eur.
Fayed, M.H., Abdel-Rahman, S.I., Alanazi, F.K., Ahmed, M.O., Tawfeek, H.M., Al- J. Pharm. Biopharm. 106, 2–8. https://doi.org/10.1016/j.ejpb.2016.04.025.
Shedfat, R.I., 2017. High-shear granulation process: Influence of processing Megarry, A.J., Swainson, S.M.E., Roberts, R.J., Reynolds, G.K., 2019. A big data
parameters on critical quality attributes of acetaminophen granules and tablets using approach to pharmaceutical flow properties. Int. J. Pharm. 555, 337–345. https://
design of experiment approach. Acta Pol. Pharm. - Drug Res. 74, 235–248. doi.org/10.1016/j.ijpharm.2018.11.059.
Fell, J.T., Newton, J.M., 1970. Determination of tablet strength by the diametral- Mohylyuk, V., Styliari, I.D., Novykov, D., Pikett, R., Dattani, R., 2019. Assessment of the
compression test. J. Pharm. Sci. 59 (5), 688–691. https://doi.org/10.1002/ effect of Cellets’ particle size on the flow in a Wurster fluid-bed coater via powder
jps.2600590523. rheology. J. Drug Deliv. Sci. Technol. 54, 101320. https://doi.org/10.1016/j.
Freeman, R., 2007. Measuring the flow properties of consolidated, conditioned and jddst.2019.101320.
aerated powders - A comparative study using a powder rheometer and a rotational Muthancheri, I., Ramachandran, R., 2020. Mechanistic understanding of granule growth
shear cell. Powder Technol 174 (1-2), 25–33. https://doi.org/10.1016/j. behavior in bi-component wet granulation processes with wettability differentials.
powtec.2006.10.016. Powder Technol 367, 841–859. https://doi.org/10.1016/j.powtec.2020.04.016.
Freeman, R., Fu, X., 2008. Characterisation of powder bulk, dynamic flow and shear Nan, W., Ghadiri, M., Wang, Y., 2017. Analysis of powder rheometry of FT4: Effect of
properties in relation to die filling. Powder Metall. 51 (3), 196–201. https://doi.org/ particle shape. Chem. Eng. Sci. 173, 374–383. https://doi.org/10.1016/j.
10.1179/174329008X324115. ces.2017.08.004.
Freeman, R.E., Cooke, J.R., Schneider, L.C.R., 2009. Measuring shear properties and Narang, A.S., Sheverev, V., Freeman, T., Both, D., Stepaniuk, V., Delancy, M., Millington-
normal stresses generated within a rotational shear cell for consolidated and non- Smith, D., Macias, K., Subramanian, G., 2016. Process Analytical Technology for
consolidated powders. Powder Technol 190 (1-2), 65–69. https://doi.org/10.1016/j. High Shear Wet Granulation: Wet Mass Consistency Reported by In-Line Drag Flow
powtec.2008.04.084. Force Sensor Is Consistent with Powder Rheology Measured by At-Line FT4 Powder
Freeman, T., Vom Bey, H., Hanish, M., Brockbank, K., Armstrong, B., 2016. The influence Rheometer®. J. Pharm. Sci. 105 (1), 182–187. https://doi.org/10.1016/j.
of roller compaction processing variables on the rheological properties of granules. xphs.2015.11.030.
Asian J. Pharm. Sci. https://doi.org/10.1016/j.ajps.2016.03.002. Patel, S., Bansal, A.K., 2011. Prediction of mechanical properties of compacted binary
Garg, V., Mallick, S.S., Garcia-Trinanes, P., Berry, R.J., 2018. An investigation into the mixtures containing high-dose poorly compressible drug. Int. J. Pharm. 403 (1-2),
flowability of fine powders used in pharmaceutical industries. Powder Technol 336, 109–114. https://doi.org/10.1016/j.ijpharm.2010.10.039.
375–382. https://doi.org/10.1016/j.powtec.2018.06.014. Patel, S., Kaushal, A.M., Bansal, A.K., 2006. Compression physics in the formulation
Ghadiri, M., Pasha, M., Nan, W., Hare, C., Vivacqua, V., Zafar, U., Nezamabadi, S., development of tablets. Crit. Rev. Ther. Drug Carrier Syst. 23 (1), 1–66. https://doi.
Lopez, A., Pasha, M., Nadimi, S., 2020. Cohesive powder flow: Trends and challenges org/10.1615/CritRevTherDrugCarrierSyst.v23.i110.1615/
in characterisation and analysis. KONA Powder Part. J. 37 (0), 3–18. https://doi. CritRevTherDrugCarrierSyst.v23.i1.10.
org/10.14356/kona.2020018. Penkavova, V., Kulaviak, L., Ruzicka, M.C., Puncochar, M., Grof, Z., Stepanek, F.,
Ghazi, N., Liu, Z., Bhatt, C., Kiang, S., Cuitino, A., 2019. Investigating the Effect of APAP Schongut, M., Zamostny, P., 2019. Compression of anisometric granular materials.
Crystals on Tablet Behavior Manufactured by Direct Compression. AAPS Powder Technol. 342, 887–898. https://doi.org/10.1016/j.powtec.2018.10.031.
PharmSciTech 20 (5). https://doi.org/10.1208/s12249-019-1369-0. Reynolds, G.K., Campbell, J.I., Roberts, R.J., 2017. A compressibility based model for
Ghori, M.U., 2016. Powder Compaction: Compression Properties of Cellulose Ethers. Br. predicting the tensile strength of directly compressed pharmaceutical powder
J. Pharm. https://doi.org/10.5920/bjpharm.2016.09. mixtures. Int. J. Pharm. 531 (1), 215–224. https://doi.org/10.1016/j.
Han, J.K., Shin, B.S., Choi, D.H., 2019. Comprehensive Study of Intermediate and Critical ijpharm.2017.08.075.
Quality Attributes for Process Control of High-Shear Wet Granulation Using Rohilla, L., Garg, V., Mallick, S.S., Setia, G., 2018. An experimental investigation on the
Multivariate Analysis and the Quality by Design Approach. Pharmaceutics 11, 252. effect of particle size into the flowability of fly ash. Powder Technol. 330, 164–173.
https://doi.org/10.3390/pharmaceutics11060252. https://doi.org/10.1016/j.powtec.2018.02.013.
Heckel, R.W., 1961. Density-Pressure Relationships in Powder Compaction. Trans. Rose, A.A., Kaialy, W., 2019. Improved tableting behavior of paracetamol in the presence
Metall. Soc, AIME. of polyvinylpyrrolidone additive: Effect of mixing conditions. Particuology 43, 9–18.
Hiestand, E., 1997. Principles, tenets and notions of tablet bonding and measurements of https://doi.org/10.1016/j.partic.2018.01.010.
strength. Eur. J. Pharm. Biopharm. 44 (3), 229–242. https://doi.org/10.1016/ Salehi, H., Barletta, D., Poletto, M., 2017. A comparison between powder flow property
S0939-6411(97)00127-6. testers. Particuology 32, 10–20. https://doi.org/10.1016/j.partic.2016.08.003.
Hirschberg, C., Sun, C.C., Risbo, J., Rantanen, J., 2019. Effects of Water on Powder Schulze, D., 2006. Flow properties of powders and bulk solids. Braunschweig/Wolfenbu
Flowability of Diverse Powders Assessed by Complimentary Techniques. J. Pharm. ttel, Ger. Univ. 1–21.
Sci. 108 (8), 2613–2620. https://doi.org/10.1016/j.xphs.2019.03.012. Shi, H., Mohanty, R., Chakravarty, S., Cabiscol, R., Morgeneyer, M., Zetzener, H., Ooi, J.
Jenike, A., 1964. Storage and Flow of Solids, Bulletin No. 123. Utah Eng. Exp. Stn. Y., Kwade, A., Luding, S., Magnanimo, V., 2018. Effect of particle size and cohesion
Juban, A., Nouguier-Lehon, C., Briancon, S., Hoc, T., Puel, F., 2015. Predictive model for on powder yielding and flow. KONA Powder Part. J. 2018, 226–250. https://doi.
tensile strength of pharmaceutical tablets based on local hardness measurements. org/10.14356/kona.2018014.
Int. J. Pharm. 490 (1-2), 438–445. https://doi.org/10.1016/j.ijpharm.2015.05.078. Slettengren, K., Xanthakis, E., Ahrné, L., Windhab, E.J., 2016. Flow Properties of Spices
Kalaria, D.R., Parker, K., Reynolds, G.K., Laru, J., 2020. An industrial approach towards Measured with Powder Flow Tester and Ring Shear Tester-XS. Int. J. Food Prop. 19
solid dosage development for first-in-human studies: Application of predictive (7), 1475–1482. https://doi.org/10.1080/10942912.2015.1083576.
science and lean principles. Drug Discov. Today. 25 (3), 505–518. https://doi.org/ Søgaard, S.V., Pedersen, T., Allesø, M., Garnaes, J., Rantanen, J., 2014. Evaluation of ring
10.1016/j.drudis.2019.12.012. shear testing as a characterization method for powder flow in small-scale powder
Komínová, P., Kulaviak, L., Zámostný, P., 2021. Stress-Dependent Particle Interactions of processing equipment. Int. J. Pharm. 475 (1-2), 315–323. https://doi.org/10.1016/j.
Magnesium Aluminometasilicates as Their Performance Factor in Powder Flow and ijpharm.2014.08.060.
Compaction Applications. Materials (Basel). 14, 900. https://doi.org/10.3390/ Stauffer, F., Vanhoorne, V., Pilcer, G., Chavez, P.-F., Schubert, M.A., Vervaet, C., De
ma14040900. Beer, T., 2019. Managing active pharmaceutical ingredient raw material variability
Koynov, S., Glasser, B., Muzzio, F., 2015. Comparison of three rotational shear cell during twin-screw blend feeding. Eur. J. Pharm. Biopharm. 135, 49–60. https://doi.
testers: Powder flowability and bulk density. Powder Technol. 283, 103–112. org/10.1016/j.ejpb.2018.12.012.
https://doi.org/10.1016/j.powtec.2015.04.027.

13
O. Macho et al. International Journal of Pharmaceutics 608 (2021) 121110

Sun, C., Grant, D.J.W., 2001. Influence of crystal structure on the tableting properties of Van Snick, B., Dhondt, J., Pandelaere, K., Bertels, J., Mertens, R., Klingeleers, D., Di
sulfamerazine polymorphs. Pharm. Res. https://doi.org/10.1023/A: Pretoro, G., Remon, J.P., Vervaet, C., De Beer, T., Vanhoorne, V., 2018.
1011038526805. A multivariate raw material property database to facilitate drug product
Taipale-Kovalainen, K., Karttunen, A.P., Ketolainen, J., Korhonen, O., 2018. Lubricant development and enable in-silico design of pharmaceutical dry powder processes.
based determination of design space for continuously manufactured high dose Int. J. Pharm. 549 (1-2), 415–435. https://doi.org/10.1016/j.ijpharm.2018.08.014.
paracetamol tablets. Eur. J. Pharm. Sci. 115, 1–10. https://doi.org/10.1016/j. Veronica, N., Goh, H.P., Kang, C.Y.X., Liew, C.V., Heng, P.W.S., 2018. Influence of spray
ejps.2017.12.021. nozzle aperture during high shear wet granulation on granule properties and its
Takeuchi, Y., Tomita, T., Kuroda, J., Kageyu, A., Yonekura, C., Hiramura, Y., Tahara, K., compression attributes. Int. J. Pharm. 553 (1-2), 474–482. https://doi.org/10.1016/
Takeuchi, H., 2018. Characterization of mannitol granules and powder: A j.ijpharm.2018.10.067.
comparative study using two flowability testers. Int. J. Pharm. 547 (1-2), 106–113. Wang, Y., Koynov, S., Glasser, B.J., Muzzio, F.J., 2016a. A method to analyze shear cell
https://doi.org/10.1016/j.ijpharm.2018.05.061. data of powders measured under different initial consolidation stresses. Powder
Tan, G., Morton, D., Larson, I., 2015. On the Methods to Measure Powder Flow. Curr. Technol. 294, 105–112. https://doi.org/10.1016/j.powtec.2016.02.027.
Pharm. Des. https://doi.org/10.2174/1381612821666151008125852. Wang, Y., Snee, R.D., Meng, W., Muzzio, F.J., 2016b. Predicting flow behavior of
Tank, D., Karan, K., Gajera, B.Y., Dave, R.H., 2018. Investigate the effect of solvents on pharmaceutical blends using shear cell methodology: A quality by design approach.
wet granulation of microcrystalline cellulose using hydroxypropyl methylcellulose as Powder Technol. 294, 22–29. https://doi.org/10.1016/j.powtec.2016.01.019.
a binder and evaluation of rheological and thermal characteristics of granules. Saudi Wei, H., Zhao, T., Meng, Q., Wang, X., Zhang, B., 2020. Quantifying the Morphology of
Pharm. J. 26 (4), 593–602. https://doi.org/10.1016/j.jsps.2018.02.007. Calcareous Sands by Dynamic Image Analysis. Int. J. Geomech. 20 (4), 04020020.
Tran, D.T., Komínová, P., Kulaviak, L., Zámostný, P., 2021. Evaluation of multifunctional https://doi.org/10.1061/(ASCE)GM.1943-5622.0001640.
magnesium aluminosilicate materials as novel family of glidants in solid dosage Whiting, J.G., Tondare, V.N., Scott, J.H.J., Phan, T.Q., Donmez, M.A., 2019. Uncertainty
products. Int. J. Pharm. 592, 120054. https://doi.org/10.1016/j. of particle size measurements using dynamic image analysis. CIRP Ann 68 (1),
ijpharm.2020.120054. 531–534. https://doi.org/10.1016/j.cirp.2019.04.075.
Tran, D.T., Majerová, D., Veselý, M., Kulaviak, L., Ruzicka, M.C., Zámostný, P., 2019. On Worku, Z.A., Kumar, D., Gomes, J.V., He, Y., Glennon, B., Ramisetty, K.A., Rasmuson, Å.
the mechanism of colloidal silica action to improve flow properties of C., O’Connell, P., Gallagher, K.H., Woods, T., Shastri, N.R., Healy, A.-M., 2017.
pharmaceutical excipients. Int. J. Pharm. 556, 383–394. https://doi.org/10.1016/j. Modelling and understanding powder flow properties and compactability of selected
ijpharm.2018.11.066. active pharmaceutical ingredients, excipients and physical mixtures from critical
Trivedi, M.R., Dave, R.H., 2014. To study physical compatibility between dibasic calcium material properties. Int. J. Pharm. 531 (1), 191–204. https://doi.org/10.1016/j.
phosphate and cohesive actives using powder rheometer and thermal methods. Drug ijpharm.2017.08.063.
Dev. Ind. Pharm. 40 (12), 1585–1596. https://doi.org/10.3109/ Wu, C.-Y., Dihoru, L., Cocks, A.C.F., 2003. The flow of powder into simple and stepped
03639045.2013.838576. dies. Powder Technol 134 (1-2), 24–39. https://doi.org/10.1016/S0032-5910(03)
00130-X.

14

You might also like