You are on page 1of 10

European Journal of Pharmaceutical Sciences 115 (2018) 223–232

Contents lists available at ScienceDirect

European Journal of Pharmaceutical Sciences


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

Breakage and drying behaviour of granules in a continuous fluid bed dryer: T


Influence of process parameters and wet granule transfer
F. De Leersnydera,1, V. Vanhoorneb,1, H. Bekaertb, J. Vercruysseb, M. Ghijsc, N. Bostijna,
M. Verstraetena, P. Cappuynsd, I. Van Assched, Y. Vander Heydene, E. Ziemonsf, J.P. Remonb,

I. Nopensc, C. Vervaetb, T. De Beera,
a
Laboratory of Pharmaceutical Process Analytical Technology, Ghent University, Belgium
b
Laboratory of Pharmaceutical Technology, Ghent University, Belgium
c
BIOMATH, Department of Mathematical Modelling, Statistics and Bioinformatics, Ghent University, Belgium
d
Department of Pharmaceutical Development, Johnson & Johnson Pharmaceutical Research and Development, Janssen Pharmaceutica, Belgium
e
Department of Analytical Chemistry and Pharmaceutical Technology, VUB, Belgium
f
CIRM, Laboratory of Analytical Chemistry, University of Liege, Belgium

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

Keywords: Although twin screw granulation has already been widely studied in recent years, only few studies addressed the
Continuous manufacturing subsequent continuous drying which is required after wet granulation and still suffers from a lack of detailed
Fluid bed drying understanding. The latter is important for optimisation and control and, hence, a cost-effective practical im-
Critical quality attributes plementation. Therefore, the aim of the current study is to increase understanding of the drying kinetics and the
Twin screw granulation
breakage and attrition phenomena during fluid bed drying after continuous twin screw granulation. Experiments
Process understanding
Particle size distribution
were performed on a continuous manufacturing line consisting of a twin-screw granulator, a six-segmented fluid
Moisture content bed dryer, a mill, a lubricant blender and a tablet press. Granulation parameters were fixed in order to only
examine the effect of drying parameters (filling time, drying time, air flow, drying air temperature) on the size
distribution and moisture content of granules (both of the entire granulate and of size fractions). The wet
granules were transferred either gravimetrically or pneumatically from the granulator exit to the fluid bed dryer.
After a certain drying time, the moisture content reached an equilibrium. This drying time was found to depend
on the applied airflow, drying air temperature and filling time. The moisture content of the granules decreased
with an increasing drying time, airflow and drying temperature. Although smaller granules dried faster, the
multimodal particle size distribution of the granules did not compromise uniform drying of the granules when
the target moisture content was achieved. Extensive breakage of granules was observed during drying. Especially
wet granules were prone to breakage and attrition during pneumatic transport, either in the wet transfer line or
in the dry transfer line. Breakage and attrition of granules during transport and drying should be anticipated
early on during process and formulation development by performing integrated experiments on the granulator,
dryer and mill.

1. Introduction However, it is essential to gain fundamental knowledge about the me-


chanisms occurring during continuous processes to allow process opti-
Traditionally, batch processes are used to produce pharmaceutical misation and control for reliable real-time release of produced products.
products. However, the pharma industry is willing to invest in con- Tablets are the most commonly used solid dosage forms in the
tinuous processes since many benefits are associated with continuous pharmaceutical industry (Gohel and Jogani, 2005). Direct compression
manufacturing (e.g. flexibility, smaller equipment, cost, time-to-market is the preferred continuous manufacturing method for production of
reduction and improved product quality) (FDA, 2004; Plumb, 2005; tablets, but often an intermediate agglomeration step is required to
Teżyk et al., 2015; Vercruysse et al., 2013; Vervaet and Remon, 2005). improve the flowability, homogeneity and compressibility (Armstrong,

Abbreviations: L/S, liquid-to-solid; LOD, loss-on-drying; PBM, Population Balance Model; V0, bulk volume; V500, tapped volume after 500 taps; ρb, bulk density; ρt, tapped density after
500 taps

Corresponding author at: Ghent University, Laboratory of Pharmaceutical Process Analytical Technology, Ottergemsesteenweg 460, 9000 Ghent, Belgium.
E-mail address: Thomas.debeer@ugent.be (T. De Beer).
1
Shared first authorship.

https://doi.org/10.1016/j.ejps.2018.01.037
Received 28 September 2017; Received in revised form 12 December 2017; Accepted 20 January 2018
0928-0987/ © 2018 Elsevier B.V. All rights reserved.
F. De Leersnyder et al. European Journal of Pharmaceutical Sciences 115 (2018) 223–232

Fig. 1. Horizontal ConsiGma™-25 system (left) and vertical ConsiGma™-25 system (right) with 1. loss-in-weight feeder, 2. twin screw granulator, 3. wet transfer line (only in horizontal
set-up), 4. six-segmented fluid bed dryer, 5. dry transfer line and 6. product control unit with option for milling.

Fig. 2. Visual presentation of the filling cycle in the six-segmented fluid bed dryer of a ConsiGma™-25 line.

Table 1 While twin screw granulation was widely studied in recent years,
Overview of the granulation parameters. only few studies addressed subsequent drying on a segmented fluid bed
dryer which is required after wet granulation. Fonteyne et al. and
Screw speed (rpm) 900
Barrel temperature (°C) 25 Chablani et al. reported on the implementation of Raman and NIR
Throughput (kg/h) 20 probes in the six-segmented fluid bed dryer unit of the ConsiGma™-25
L/S ratio 0.15 for monitoring of moisture content and polymorphism while Vercruysse
et al. evaluated the stability and repeatability of the granulation and
drying units during long runs with constant process parameters
2007; Gohel and Jogani, 2005; Vervaet and Remon, 2005). (Chablani et al., 2011; Fonteyne et al., 2014a; Vercruysse et al., 2013).
Several equipment manufacturers currently offer integrated from- However, these studies did not focus on the influence of dryer para-
powder-to-tablet lines for continuous production of tablets via con- meters on critical quality attributes of granules and did not aim to ex-
tinuous twin screw granulation (e.g. QbCon® by L.B. Böhle, plore and understand breakage and attrition phenomena taking place
Granuformer® by Freund Vector, ConsiGma™ by GEA Pharma Systems, during drying.
MODCOS system by Glatt). Whereas in the Granuformer® line (Freund Although batch-wise fluid bed drying has been intensively studied
Vector) a spiral dryer is implemented, the other continuous lines and the influence of most parameters (air flow, drying air temperature
pneumatically transfer the wet granules from the granulator to a seg- and drying time) on the granule quality is expected to be similar during
mented fluid bed dryer (Kleinebudde, 2017; Byrn et al., 2014). How- continuous fluid bed drying, the set-up of a batch fluid bed dryer (with
ever, on the ConsiGma™ system (GEA Pharma systems) gravimetric a single drying cell) is fundamentally different in comparison to a
instead of pneumatic granule transfer is also possible. Therefore two segmented fluid bed dryer (with multiple drying cells). Whereas an
different set-ups can be distinguished: (1) a horizontal set-up (granu- entire batch of wet granules is introduced at the same moment into a
lator is positioned next to the dryer) with pneumatic granule transfer batch fluid bed dryer, granules are continuously transferred to a specific
via a wet transfer line, (2) a vertical set-up (granulator is positioned cell of the segmented fluid bed dryer for a set time (filling time) after
above the dryer) with gravimetric transfer of wet granules to the dryer which the next cell of the segmented fluid bed dryer is filled. Therefore,
(Fig. 1). the filling time of a dryer cell is an additional variable inherent to
Segmented fluid bed dryers are currently most often implemented in drying in a segmented fluid bed dryer after continuous granulation.
continuous manufacturing lines. Although other designs of continuous However, the effect of the filling time has not yet been investigated.
fluid bed dryers are available such as horizontal fluid bed dryers (e.g. Therefore, the effect of filling time, in addition to other drying para-
GF series by Glatt, Niro Contipharm granulator, Heinen drying tech- meters (air flow, drying air temperature, drying time), was evaluated in
nologies) or the AGT fluid bed dryer by Glatt, these systems are typi- the current study.
cally not applicable in the pharmaceutical industry due to the long The impact of the transfer of wet granules after continuous granu-
material residence time and lack of plug-flow (Gotthardt et al., 1999; lation was also neglected up to now. Using an in-line particle imaging
Vervaet and Remon, 2009; Teżyk et al., 2015). Short and controllable tool, Kumar et al. observed that tray dried granules showed wider
residence times are of utmost importance during granulation of phar- granule size distributions in comparison to fluid bed dried granules
maceutical products for traceability and to avoid product degradation. which was attributed to breakage and attrition in the segmented fluid

224
F. De Leersnyder et al. European Journal of Pharmaceutical Sciences 115 (2018) 223–232

Table 2 suboptimal dryer settings or due to pneumatic granule transfer, in the


Overview of performed experiments (*: no granules could be collected on vertical set-up). current study the effect of drying parameters and granule transfer
(gravimetric or pneumatic) on the granule quality was evaluated.
Experimental Drying air Airflow Filling Drying
conditions temperature (°C) (m3/h) time (s) time (s) Efforts were made by Mortier et al. to gain fundamental under-
standing of the physics determining the drying behaviour of granules in
A 40 360 60 80 the six-segmented fluid bed dryer via application of mass-energy bal-
100
ances and mechanistic models (Mortier et al., 2012, 2013, 2014). Cal-
200
300 culating mass and energy balances, Mortier et al. could predict the
324 moisture content of granules, produced with different granulator and
B 60 360 60 80 dryer settings, based on the in-line data logged by the ConsiGma™-25
100 line (Mortier et al., 2014). Although mass and energy balances proved
200
to be powerful to monitor the overall granule moisture content in-line,
300
324 obtaining detailed insight of the drying process and kinetics at granule
C 40 440 60 80 level proved to be a tedious endeavor (Mortier et al., 2014). Therefore,
100 a mechanistic model was developed by Mortier et al. describing the
200
drying of a single granule in a dryer cell of the ConsiGma™-25 system
300
324 (Mortier et al., 2012). This model was calibrated and validated using
D 60 440 60 80 experimental data for a pharmaceutical formulation and allowed to
100 gain insight in the drying behaviour of a single granule. In a next step,
200 this model was rewritten as a Population Balance Model (PBM) to si-
300
mulate the moisture content of a population, or batch, of granules
324
E* and J* 50 400 120 140 (Mortier et al., 2013). In this PBM, the evolution of the moisture con-
200 tent of granules with different initial moisture contents during drying
300 can be described at the same time. This generates a distribution of
400
granules during the drying process in function of their current moisture
679
F 40 360 180 250*
content, which is the output of the model. Another feature of the PBM is
300 that the effect of the continuous filling of the dryer cell can be in-
400 corporated, which implies that different granules are not at the same
700 stage in the drying process (Mortier et al., 2012).The aforementioned
1034
PBM has finally been extended by Mortier et al. to describe the evo-
G 60 360 180 200*
250 lutions of particle size distribution and moisture content distribution of
300 the granules simultaneously during drying (Mortier et al., 2013). These
400 PBM models still need to be calibrated and validated to experimental
700
moisture content distribution data. However, moisture content dis-
1034
H 40 440 180 250*
tribution data is exhaustive to gather. Because of this, the model for
300 single granule drying by Mortier et al. is being focused on for calibra-
400 tion and validation to different formulations (Mortier et al., 2012).
700 Therefore, in the current study, experiments were setup to better
1034
understand the effect of the drying parameters (filling time, drying
I 60 440 180 200*
250 time, drying air temperature and air flow) and wet granule transfer on
300 the critical quality attributes of granules (particle size and moisture
400 content). Moisture content and granule size distribution are critical
700 quality attributes of granules as they determine the flowability, segre-
1034
gation tendency and stability of the granules (Crouter and Briens, 2014;
El Hagrasy et al., 2013; Emery et al., 2009). Additionally, the relation
bed dryer or during pneumatic granule transfer from the twin screw between particle size and moisture content was investigated in the
granulator to the dryer (Kumar et al., 2015). As the study of Kumar current study by determination of the moisture content of different
et al. could not differentiate between breakage and attrition due to sieve fractions in function of drying time at different drying conditions.

Fig. 3. Total moisture content of the granules produced on the horizontal ConsiGma™-25 set-up in function of drying for filling times of 60 s (left) and 180 s (right).

225
F. De Leersnyder et al. European Journal of Pharmaceutical Sciences 115 (2018) 223–232

2. Materials and methods

2.1. Materials

The formulation consisted of two active ingredients, maize starch


(Cargill, Vilvoorde, Belgium) and powdered cellulose (Solka-Floc,
International Fiber Corporation, Ohio, USA) were used as fillers and
pregelatinised starch (National Starch & Chemical company,
Westchester, USA) and sodium starch glycolate (Roquette, Lestrem,
France) were used as binder and desintegrant, respectively.

2.2. Methods

2.2.1. Description of the ConsiGma™-25 system


Granulation and drying experiments were performed on a
ConsiGma™-25 system (GEA Pharma Systems, Wommelgem, Belgium).
Fig. 4. Total moisture content of the granules produced on the horizontal set-up in This continuous line consists of a twin screw granulator, a six-seg-
function of drying time for the duplicate experiments at intermediate drying conditions. mented fluid bed dryer, a granule conditioning unit with a mill, a
continuous blender for addition of for instance a lubricant such as
The results of these experiments can in the future be used to apply the magnesium stearate and a tablet press (latter not shown in Fig. 1). As
mechanistic model for single granule drying of Mortier et al. for si- fluid bed drying is not an inherently continuous process, the fluid bed
mulation of the moisture content evolution for different granule sizes in dryer unit of the ConsiGma™-25 system consists of six cells, which are
function of drying time and drying conditions (Mortier et al., 2012). consecutively fed with wet granules produced by twin screw granula-
tion. After filling a dryer cell for a set filling time, granules are fed to the
next cell, while the previous cells complete their drying cycle. De-
pending on the applied filling and drying times, two to six cells can
contain granules simultaneously. This process is schematically pre-
sented in Fig. 2. It is clear that once filling and drying of the granules in
cell 4 starts, cell 1, 2 and 3 still contain granules in their drying phase.

Fig. 5. Moisture content of the granules produced on horizontal set-up in function of drying time, drying air temperature and airflow with blue and red circles representing short (60 s)
and long (180 s) filling times for different granule size fractions respectively (smaller and larger size fractions are represented by smaller and larger circles, respectively). (For inter-
pretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

226
F. De Leersnyder et al. European Journal of Pharmaceutical Sciences 115 (2018) 223–232

Fig. 6. Fraction of fines, yield and oversized for experiments A–D (filling time 60 s) performed on the horizontal set-up, in function of drying time. Drying time 0: granules directly after
wet granulation.

The design of the dryer unit implies that after the drying operation in consisting of 4 kneading elements (L = D/4 for each kneading element)
the ConsiGma™-25 system, mini-batches are created, which are suc- at an angle of 60 degrees. Both kneading zones were separated by a
cessively transported to the granule conditioning unit, lubricant blender conveying element (L = 1.5D). An extra conveying element (L = 1.5D)
and the tablet press. was implemented after the second kneading block together with 2
Depending on the position of the six-segmented fluid bed dryer and narrow kneading elements (L = D/6 for each kneading element) in
granulator, two different set-ups of the ConsiGma™-25 can be dis- order to reduce the amount of oversized agglomerates, as reported by
tinguished: (1) a horizontal set-up (the granulator is positioned next to Van Melkebeke et al. (2008).
the dryer) including pneumatic transport of wet granules from the exit After granulation, the wet granules were transferred pneumatically
of the granulator to the dryer via a wet transfer line and (2) a vertical by a wet transfer line to the six-segmented fluid bed dryer unit of the
set-up (the granulator is positioned above the dryer) which comprises horizontal ConsiGma™-25 set-up (wet transfer line indicated in red in
gravimetric addition of wet granules to the dryer. Fig. 1) or transferred gravimetrically to the six-segmented fluid bed
dryer of the vertical ConsiGma™-25 set-up (Fig. 1).
After drying the granules were pneumatically transferred to the
2.2.2. Preparations of granules product control unit via the dry transfer line (indicated in green in
A preblend of the formulation was prepared in a tumbling blender Fig. 1) and were collected at the exit of the product control unit. In the
(Inversina-Bioengineering, Wald, Switzerland) during 15 min at 25 rpm current study no mill was implemented in the product control unit.
and transferred to the loss-in-weight feeder (KT20, K-Tron Soder, Additionally, wet granules were collected at the exit of the granulator
Niederlenz, Switzerland) of the ConsiGma™-25 system (GEA Pharma and tray dried in an oven during 24 h at 40 °C prior to granule char-
Systems, Wommelgem, Belgium). acterization.
The barrel of the twin screw granulator (length-to-diameter ratio: The filling time, drying time, air flow and drying air temperature
20:1) can be divided into a feed segment consisting of conveying ele- were varied during the experiments on the horizontal and vertical
ments and a work segment where the powder is intensively mixed with ConsiGma™-25 set-up. An overview of the experiments is presented in
granulation liquid by kneading elements. The granulation liquid was Table 2. Filling times of 60 (A–D) and 180 s (F–I) per cell were used. At
pumped into the barrel just before the first kneading element via a least four different drying times were applied, varying between the
double liquid addition port, injecting granulation liquid on top of each shortest and longest drying time applicable at a set filling time. The
screw. The equipment has a built-in torque gauge which monitors the minimal drying time needed to discharge the wet granules from the
torque with a frequency of 1 Hz. The applied granulation parameters for dryer unit, depended on the drying parameters and was set as shortest
current formulation were optimised by Vercruysse et al. and are listed drying time. For example in Table 2, it is indicated that experiments F
in Table 1 and were kept constant during all runs (Vercruysse et al., and H could not be performed with a drying time of 200 s, while
2013). The screw configuration was composed of 2 kneading zones each

227
F. De Leersnyder et al. European Journal of Pharmaceutical Sciences 115 (2018) 223–232

Fig. 7. Fraction of fines, yield and oversized for experiments F–I (filling time 180 s) performed on the horizontal set-up, in function of drying time. Drying time 0: granules directly after
wet granulation (no granules could be collected after 200 s of drying time for experiments F and H).

Fig. 8. Fraction of fines, yield and oversized for experiment E and J (replicates) on the horizontal set-up, in function of drying time. Drying time 0: granules directly after wet granulation.

experiment G and I could be performed. This could be attributed to process repeatability. No granules could be collected on the vertical
application of a higher drying air temperature resulting in faster drying ConsiGma™-25 set-up when applying drying times of 200 s in experi-
and therefore less sticky granules. The longest drying time corre- ments G and I and 250 s in experiments F and H because a blockage in
sponded to six times the applied filling time minus the time needed for the dry transfer line (i.e. the transfer line transporting the granules from
cell discharge. All experiments were performed at low (360 m3/h) or the dryer to the product control unit) occurred. The granules were too
high (440 m3/h) airflow in combination with a low (40 °C) or high wet and sticky and larger than when they were discharged from the
(60 °C) drying air temperature (Table 2). Additionally, experiments on dryer unit in the horizontal set-up (discussed in Sections 3.2.1 and
the horizontal ConsiGma™-25 set-up were performed in duplicate (E 3.2.2).
and J) at intermediate conditions (120 s filling time, 50 °C drying air
temperature and 400 m3/h intermediate airflow) to estimate the

228
F. De Leersnyder et al. European Journal of Pharmaceutical Sciences 115 (2018) 223–232

Fig. 9. Fraction of fines, yield and oversized for experiments A–D (filling time 60 s) on the vertical set-up, in function of drying time. Drying time 0: granules directly after wet
granulation.

2.2.3. Characterization granules granules. The bulk volume (V0) of 50 g granules was measured in a
2.2.3.1. Determination of the overall moisture content and moisture content 250 ml graduated cylinder as well as the tapped volume after 500 taps
of different size fractions. The residual moisture content of granules (V500) using a tapping machine (J. Englesman, Ludwigshafen,
collected after the product control unit was determined via loss-on- Germany). Experiments were performed in duplicate. Bulk and
drying (LOD) using a moisture analyser (Mettler LP16, Mettler-Toledo, tapped densities were calculated as 50 g/V0 and 50 g/V500,
Zaventem, Belgium) including an infrared dryer and a balance. A respectively. The Hausner ratio was calculated from the bulk (ρb) and
sample of 1 g was dried at 105 °C until the weight was constant for 30 s. tapped (ρt) densities using the following equation:
Additionally, the granules produced on both the horizontal and the ρt
vertical set-up were sieved on a series of sieves (150, 300, 500, 850, HR =
ρb
1000, 2000 μm) and the moisture content of each sieve fraction was
determined. The target moisture content for the formulation was 1.5%
measured by LOD.
3. Results and discussion

2.2.3.2. Particle size analysis. The granule size of the tray dried 3.1. Moisture content
granules and of the granules collected after the product control unit
(after fluid bed drying) was analysed via dynamic image analysis using The total moisture content of the granules in function of drying time
the QICPIC™ system (Sympatec, Clausthal-Zellerfeld, Germany) was evaluated to gain better understanding of the drying kinetics. The
equipped with a vibrating feeder system (Vibri/L™) for gravimetrical results were similar on the horizontal and vertical set-ups (data of
feeding of the granules. Samples of 20 g were measured in duplicate and vertical set-up not shown). After a certain drying time, the moisture
the averages of volume granule size distributions were reported, content did not further decrease. Depending on the air flow and drying
WINDOX 5 software (Sympatec, Clausthal-Zellerfeld, Germany) was air temperature, a constant moisture was reached after a specific drying
used to calculate the granule size distributions. The amounts of fines time. This is illustrated in Fig. 3 for the experiments with a filling time
and oversized granules were defined as the fractions < 150 μm of 60 (A–D) and 180 s (F–I). A constant moisture content was reached
and > 1000 μm, respectively. The yield of the process was defined as after 200–300 and 400–700 s, respectively.
the percentage of granules between 150 and 1000 μm. Intense drying conditions (D, I: high airflow and high drying air
temperature) resulted in faster drying of granules whereas milder
2.2.3.3. Flowability testing. The flowability of the granules collected drying conditions (A, F: low airflow and low drying air temperature)
after the product control unit was evaluated by the Hausner ratio. The resulted in slower drying of the granules (Fig. 3). Consequently, the
Hausner ratio was calculated from the bulk and tapped densities of the moisture content after the longest possible drying time (324 and 1034 s

229
F. De Leersnyder et al. European Journal of Pharmaceutical Sciences 115 (2018) 223–232

Fig. 10. Fraction of fines, yield and oversized for experiments F–I (filling time 180 s) on the vertical set-up, in function of drying time. Drying time 0: granules directly after wet
granulation.

after filling times of 60 and 180 s, respectively) was slightly lower if the same cell (at a filling time of 180 s, the first granule entering a dryer
intense drying conditions were applied in comparison to milder drying cell is dried 180 s longer in comparison to the last granule entering the
conditions. The moisture content results of replicate experiments E and dryer cell) do not compromise uniform drying of the granules if suffi-
J (at intermediate airflow, drying air temperature and filling time on ciently large drying times are applied.
horizontal set-up) are shown in Fig. 4 and demonstrate that the results
were highly repeatable. 3.2. Particle size
Next, the evolution of the moisture content of the granules was
evaluated in function of granule size and drying time (Fig. 5). The re- In this section, the granule size distributions were evaluated in
sults of the size fraction > 2000 μm were excluded as the moisture function of the applied filling time, drying time, drying air temperature
content determination by LOD of 1 g of granules was not reproducible and air flow to study breakage and attrition phenomena during fluid
since 1 g of granules often consisted of only 2–4 granules. Fig. 5 shows bed drying. The granule size distributions of tray dried granules col-
that smaller size fractions of the granules, and especially the fines lected at the outlet of the granulator were also determined and are
fraction, dried faster than the larger size fractions. included in the graphs (Figs. 6–10) as the granule size distributions
During the initial drying phase, water from the surface of the after 0 s of drying time. It was decided to represent the granule size
granules is evaporated and the surface/volume ratio of the granules in distributions by the fines (< 150 μm) fraction, yield (150–1000 μm)
the smaller size fractions is larger than of those in the larger size and oversized fraction (> 1000 μm) to make the interpretation less
fractions, resulting in faster drying of the former. Similar observations complex and more relevant from a pharmaceutical point of view.
were reported by Fonteyne et al. studying the end-point determination However, conclusions made were also valid when more size classes
of the fluid bed drying process (Fonteyne et al., 2014b). However, the were taken into account.
differences in moisture content between the different size fractions
disappeared at longer drying times. Additionally, the differences in
3.2.1. Horizontal set-up
moisture content of the different size fractions disappeared faster when
Drying of the granules in the horizontal set-up resulted in extensive
applying intense drying conditions (D, I: high airflow and drying air
breakage and attrition of the granules (i.e. loss in oversized and gain in
temperature) in comparison to milder drying conditions (A, F: low
yield and fines, respectively). However, breakage and attrition of
airflow and drying air temperature). However, for all drying conditions,
granules exclusively occurred during the first drying period of 0–80 s
there was no relevant difference in the moisture content of the different
for filling times of 60 s (Fig. 6) and of 0–200 s for filling times of 180 s
size fractions once the target moisture content of 1.5% was approached.
(Fig. 7). When comparing the granule size distributions of the granules
This signifies that the multimodal particle size distribution of the
obtained after longer drying times, no relevant differences in granule
granules or the difference in drying time between granules filled into
size distributions could be detected.

230
F. De Leersnyder et al. European Journal of Pharmaceutical Sciences 115 (2018) 223–232

It could be concluded that granules are most prone to breakage times. Although this seems contradictory at first instance, this can be
during the first seconds of drying. However, between the exit of the explained considering the moisture content of granules. Shortly dried
granulator (0 s) and discharge from the dryer after the shortest drying granules are still wet and therefore less resistant to pneumatic transport
time (80 s drying time after a filling time of 60 s and 200 s drying time in the dry transfer line (i.e. transfer line between the fluid bed dryer and
after a filling time of 180 s) not only drying of the granules occurred but the product control unit) as less solid bonds can be formed by crystal-
also pneumatic transport of granules from the exit of the granulator to lisation of solubilised material. Therefore, this phenomenon occurring
the inlet of the dryer via a wet transfer line took place. During this in the dry transfer line of the vertical set-up is comparable to the
transport, wet granules could break up by air friction, interparticular breakage and attrition of wet granules in the wet transfer line of the
collisions and collisions with the wet transfer line. As the granules were horizontal set-up. It can be concluded that a critical drying time on the
wet, solid bonds between the particles were not yet formed. vertical set-up is necessary for the granules to acquire sufficient
Additionally, exiting the wet transfer line, the granules entered the strength by formation of strong solid bonds. Further drying after the
dryer at high velocity and could consequently collide with the wall or critical drying time, did not affect the granule size distributions (Figs. 9
bottom of the dryer unit with high impact. and 10). The effect of drying time on breakage and attrition was less
Comparison of the granule size distributions of experiments A and B pronounced for the experiments with long filling times (F–I). This could
to C and D and comparison of experiments F and G to H and I for filling be attributed to the dryer concept of the segmented dryer as the filling
times of 60 and 180 s, respectively, showed that application of higher time is included in the drying time. Application of 180 s filling time and
airflow resulted in slightly less oversized granules and a slightly higher 200 s drying time signifies that the first granule entering the cell was
yield (Figs. 6 and 7). This could be attributed to more interparticle dried during 200 s whereas the last granule added to the cell was dried
collisions and more collisions between the particles and the stainless during only 20 s.
steel dryer wall when a higher airflow was applied. In contrast, drying Similar to the horizontal set-up, the drying temperature did not
air temperature did not affect the granule size distributions. influence the granule size distributions. Additionally, the airflow did
The granule size distribution of the granules produced in duplicate not influence the granule size distributions although on the horizontal
(experiments E and J) were compared to ensure the repeatability of the set-up a minor, non-relevant, effect of the airflow was observed.
experimental set-up. The experiments were highly repeatable (Fig. 8). In conclusion, breakage and attrition mainly occurred when wet
granules were exposed to higher shear force caused by pneumatic
3.2.2. Vertical set-up transport in the wet and/or dry transfer line. Integrated optimisation
The granulator settings applied on the horizontal and vertical set-up experiments varying granulation, drying and milling parameters in the
of the continuous line were identical to allow direct comparison of both same experimental design are recommended to obtain the highest yield
set-ups and indeed tray dried granules produced on both set-ups prior to tableting.
showed identical granule size distributions.
Similarly as on the horizontal set-up, breakage and attrition of the 3.3. Flowability
granules was observed comparing the tray dried granules collected at
the granulator outlet and the fluid bed dried granules (Figs. 9 and 10). The flowability of the granules was assessed by their Hausner ratio.
However, the extent of breakage and attrition was much smaller for the Granules produced on both horizontal and vertical set-ups of the con-
granules dried on the vertical set-up in comparison to the horizontal tinuous lines were classified as fairly to good flowing according to the
set-up. classification of the European Pharmacopeia (Council of Europe, 2011),
Whereas fluid bed drying of the granules on the horizontal set-up as Hausner ratios ranging from 1.15 to 1.24 and from 1.12 to 1.25 were
resulted in a 40–65% decrease in oversized granules, a 50–60% increase obtained after processing on the horizontal and vertical set-up, re-
in yield and a 4–8% increase in fines, fluid bed drying of the granules on spectively. No influence of the drying parameters on the flowability of
the vertical set-up resulted in a 20–50% decrease in oversized granules, the granules could be detected even though coarser granules were ob-
a 20–40% increase in yield and 1–6% increase in fines. As clearly less tained at longer drying times on the vertical set-up.
breakage and attrition of granules occurred during processing on the
vertical set-up, the more extensive breakage and attrition observed 3.4. General discussion
during processing on the horizontal set-up was mainly attributed to the
pneumatic transfer of the wet granules in the wet transfer line. Possibly, breakage and attrition of the granules during drying in the
However, as granule breakage and attrition also occurred on the ver- continuous line could be reduced if a binder was added via the gran-
tical set-up (although to a lesser extent), breakage and attrition was also ulation liquid as then no time for binder activation is required during
caused by the fluid bed drying process itself (i.e. collisions of the granulation. Alternatively, a higher concentration of binder or a more
granules with the dryer wall and interparticular collisions). potent binder could be included in the formulation to increase the
As a result of less breakage occurring during fluid bed drying on the granule strength and to reduce breakage and attrition during dynamic
vertical set-up, the yield percentages obtained on the vertical set-up drying of the granules. The formulation used in the current study was
were lower in comparison to the yield obtained on the horizontal set-up established for batch granulation. However, in contrast to traditional
as more oversized granules remained after drying. This signifies that the batch granulation processes, the residence time and therefore also the
granulation parameters on the vertical set-up should be optimised to time available for bond formation during twin screw granulation is very
decrease the oversized fraction of granules formed during granulation short, typically 5–20 s in twin-screw granulation versus minutes for
and consequently to increase the yield. Initially, the granulation para- batch granulation (El Hagrasy et al., 2013; Kumar et al., 2014).
meters were set based on preliminary experiments on the horizontal set- Therefore, more potent binders could be necessary for continuous
up anticipating extensive breakage and attrition during fluid bed granulation in comparison to batch granulation.
drying. However, as less breakage and attrition occurred on the vertical The results of this experimental study will also be used to calibrate
set-up, the granulation parameters could be optimised or a mill could be the mechanistic model for drying of single pharmaceutical granules.
implemented after the drying step to increase the yield on the vertical This will allow gaining more fundamental knowledge on the drying
set-up. behaviour and breakage and attrition phenomena occurring during
Interestingly, drying time affected the obtained granule size dis- drying of granules on a continuous line. For the calibration of the PBM
tributions after fluid bed drying on the vertical set-up. Applying short on granule moisture content, as well as the two-dimensional PBM in-
drying times (80–100 s for A–D and 300–400 s for F–I) resulted in more volving granule size and moisture content distribution, both developed
breakage and attrition of the granules in comparison to longer drying by Mortier et al., moisture content distribution data is required,

231
F. De Leersnyder et al. European Journal of Pharmaceutical Sciences 115 (2018) 223–232

involving categorisation of granules of a specific sieve fraction ac- Thien, M.P., Trout, L.T., 2014. Achieving continuous manufacturing for final dosage
cording to moisture content (Mortier et al., 2012, 2013). formation: challenges and how to meet them. J. Pharm. Sci. 104, 792–802.
Chablani, L., Taylor, M.K., Methotra, A., Rameas, P., Stagner, W.C., 2011. Inline real-time
near-infrared granule moisture measurements of a continuous granulation–drying–-
4. Conclusions milling process. AAPS PharmSciTech 12, 1050–1055.
Council of Europe, 2011. European Pharmacopoeia 7.0. pp. 308–311.
Crouter, A., Briens, L., 2014. The effect of moisture on the flowability of pharmaceutical
The moisture content of the granules decreased with an increasing excipients. AAPS PharmSciTech 15, 65–74.
drying time, airflow and drying temperature. Smaller granules dried El Hagrasy, A.S., Hennenkamp, J.R., Burke, M.D., Cartwright, J.J., Litster, J.D., 2013.
faster, resulting in a lower moisture content due to their relative larger Twin screw wet granulation: influence of formulation parameters on granule prop-
erties and growth behavior. Powder Technol. 238, 108–115.
surface area. When drying time increased, differences between moisture Emery, E., Oliver, J., Pugsley, T., Sharma, J., Zhou, J., 2009. Flowability of moist phar-
content of larger and smaller granules disappeared. Therefore the maceutical powders. Powder Technol. 189, 409–415.
multimodal particle size distribution of the granules or the difference in FDA, U.S.D. of H. and H.S, 2004. Guidance for Industry PAT — A Framework for
Innovative Pharmaceutical Development, Manufacuring, and Quality Assurance.
drying time between granules filled into the same cell (which is in-
vol. 16.
herent to the principle of the segmented fluid bed dryer) did not Fonteyne, M., Arruabarrena, J., de Beer, J., Hellings, M., Van Den Kerkhof, T.,
compromise uniform drying of the granules if a sufficiently long drying Burggraeve, A., Vervaet, C., Remon, J.P., De Beer, T., 2014a. NIR spectroscopic
time is applied. method for the in-line moisture assessment during drying in a six-segmented fluid bed
dryer of a continuous tablet production line: validation of quantifying abilities and
Extensive breakage and attrition of granules observed on the hor- uncertainty assessment. J. Pharm. Biomed. Anal. 100, 21–27.
izontal set-up was mainly due to the pneumatic transfer of wet granules Fonteyne, M., Gildemyn, D., Peeters, E., Mortier, S.T.F.C., Vercruysse, J., Gernaey, K.V.,
between the granulator and the dryer. In the vertical set-up, less Vervaet, C., Remon, J.P., Nopens, I., De Beer, T., 2014b. Moisture and drug solid-state
monitoring during a continuous drying process using empirical and mass balance
breakage and attrition of granules was observed, since the wet granules models. Eur. J. Pharm. Biopharm. 87, 616–628.
were not subjected to pneumatic transport during transport to the six- Gohel, M.C., Jogani, P.D., 2005. A review of co-processed directly compressible ex-
segmented dryer. A higher degree of breakage and attrition of the cipients. J. Pharm. Pharm. Sci. 8, 76–93.
Gotthardt, S., Knoch, A., Lee, G., 1999. Continuous wet granulation using fluidized-bed
granules was observed when shorter drying times were applied on the techniques I. Examination of powder mixing kinetics and preliminary granulation
granules. When applying short drying times, granules were still wet experiments. Eur. J. Pharm. Biopharm. 48, 189–197.
when entering the dry transfer line, causing breakage and attrition in Kleinebudde, P., 2017. Continuous manufacturing of pharmaceutical products. In: 8th
International Granulation Workshop; Granulation Conference, Sheffield, UK, 28–30th
the dry transfer line due to pneumatic transport. Breakage and attrition June 2017.
of granules during fluid bed drying should be taken into account early Kumar, A., Vercruysse, J., Toiviainen, M., Panouillot, P.E., Juuti, M., Vanhoorne, V.,
on during process and formulation development on a continuous pro- Vervaet, C., Remon, J.P., Gernaey, K.V., De Beer, T., Nopens, I., 2014. Mixing and
transport during pharmaceutical twin-screw wet granulation: experimental analysis
duction line for tablets by performing integrated experiments on all unit
via chemical imaging. Eur. J. Pharm. Biopharm. 87, 279–289.
operations of the line. Additionally, attention should be given to Kumar, A., Dhondt, J., De Leersnyder, F., Vercruysse, J., Vanhoorne, V., Vervaet, C.,
granule transfer as it was demonstrated that less breakage and attrition Remon, J.P., Gernaey, K.V., De Beer, T., Nopens, I., 2015. Evaluation of an in-line
occurred after gravimetric granule transfer between the granulator and particle imaging tool for monitoring twin-screw granulation performance. Powder
Technol. 285, 80–87.
the dryer. Mortier, S.T.F.C., De Beer, T., Gernaey, K.V., Vercruysse, J., Fonteyne, M., Remon, J.P.,
Overall, the segmented dryer proved to be a robust system towards Vervaet, C., Nopens, I., 2012. Mechanistic modelling of the drying behaviour of single
granule size distribution and flowability as air flow and drying air pharmaceutical granules. Eur. J. Pharm. Biopharm. 80, 682–689.
Mortier, S.T.F.C., Gernaey, K.V., De Beer, T., Nopens, I., 2013. Development of a popu-
temperature only showed a negligible influence on these granule lation balance model of a pharmaceutical drying process and testing of solution
characteristics. Therefore, small deviations in these settings during long methods. Comput. Chem. Eng. 50, 39–53.
term manufacturing are not expected to influence the granule size Mortier, S.T.F.C., Gernaey, K.V., De Beer, T., Nopens, I., 2014. Analysing drying unit
performance in a continuous pharmaceutical manufacturing line by means of mass -
distribution or flowability of the granules, which are important critical energy balances. Eur. J. Pharm. Biopharm. 86, 532–543.
quality attributes for further downstream processing. Plumb, K., 2005. Continuous processing in the pharmaceutical industry: changing the
mind set. Chem. Eng. Res. Des. 730–738.
Teżyk, M., Milanowski, B., Ernst, A., Lulek, J., 2015. Recent progress in continuous and
Acknowledgement
semi-continuous processing of solid oral dosage forms: a review. Drug Dev. Ind.
Pharm. 9045, 1–20.
The authors would like to acknowledge GEA Pharma Systems for Van Melkebeke, B., Vervaet, C., Remon, J.P., 2008. Validation of a continuous granula-
tion process using a twin-screw extruder. Int. J. Pharm. 356, 224–230.
offering the possibility to use the vertical ConsiGma™-25 set-up at their
Vercruysse, J., Delaet, U., Van Assche, I., Cappuyns, P., Arata, F., Caporicci, G., De Beer,
facilities in Wommelgem. T., Remon, J.P., Vervaet, C., 2013. Stability and repeatability of a continuous twin
screw granulation and drying system. Eur. J. Pharm. Biopharm. 85, 1031–1038.
References Vervaet, C., Remon, J.P., 2005. Continuous granulation in the pharmaceutical industry.
Chem. Eng. Sci. 60, 3949–3957.
Vervaet, C., Remon, J.P., 2009. Continuous granulation. In: Swarbrick, J. (Ed.),
Armstrong, N.A., 2007. Tablet manufacture by direct compression. In: Swarbrick, J. (Ed.), Handbook of Pharmaceutical Granulation Technology. 2009. Informa Healthcare,
Encyclopedia of Pharmaceutical Technology, 3rd ed. Informa Healthcare, New York, New York, pp. 308–322.
pp. 3673–3683.
Byrn, S., Futran, M., Thomas, H., Jayjock, E., Maron, N., Meyer, R.F., Myerson, A.S.,

232

You might also like