Anda di halaman 1dari 10

PHASCI-03397; No of Pages 10

European Journal of Pharmaceutical Sciences xxx (2015) xxx–xxx

Contents lists available at ScienceDirect

European Journal of Pharmaceutical Sciences

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

Roller compaction scale-up using roll width as scale factor and laser-based determined
ribbon porosity as critical material attribute
Morten Allesø a,b,⁎, René Holm a, Per Holm a
a
H. Lundbeck A/S, Biologics and Pharmaceutical Science, Ottiliavej 9, DK-2500, Valby, Denmark
b
NNE Pharmaplan A/S, Process Technology Consulting, Nybrovej 80, DK-2820 Gentofte, Denmark

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

Article history: Due to the complexity and difficulties associated with the mechanistic modeling of roller compaction process for
Received 31 August 2015 scale-up, an innovative equipment approach is to keep roll diameter fixed between scales and instead vary the
Accepted 1 November 2015 roll width. Assuming a fixed gap and roll force, this approach should create similar conditions for the nip regions
Available online xxxx
of the two compactor scales, and thus result in a scale-reproducible ribbon porosity. In the present work a non-
destructive laser-based technique was used to measure the ribbon porosity at-line with high precision and high
Keywords:
Roller compaction
accuracy as confirmed by an initial comparison to a well-established volume displacement oil intrusion method.
Ribbon porosity The ribbon porosity was found to be scale-independent when comparing the average porosity of a group of rib-
Solid fraction bon samples (n = 12) from small-scale (Mini-Pactor®) to large-scale (Macro-Pactor®). A higher standard devi-
Volume displacement ation of ribbons fragment porosities from the large-scale roller compactor was attributed to minor variations in
Direct laser measurement powder densification across the roll width.
Microcrystalline cellulose With the intention to reproduce ribbon porosity from one scale to the other, process settings of roll force and gap
size applied to the Mini-Pactor® (and identified during formulation development) were therefore directly trans-
ferrable to subsequent commercial scale production on the Macro-Pactor®. This creates a better link between for-
mulation development and tech transfer and decreases the number of batches needed to establish the parameter
settings of the commercial process.
© 2015 Elsevier B.V. All rights reserved.

1. Introduction process parameters as well as equipment design alterations; the latter


typically realistic only with new equipment as a part of its installation
Roller compaction for dry granulation has gained increasing interest and qualification. Proper understanding of scale-up parameters stems
in the pharmaceutical industry over the past decades. Improvements in from mechanistic insight into the given process technology and its var-
the technology such as better control over the gap, leading to more uni- ious scales. The aim of a successful scale up is therefore to create similar
form ribbon and granule properties, and improved roller side sealing process conditions for the powder material across scales through the
and thus containment of fines, are key reasons for its increased popular- adjustment of process parameters. For roller compaction, assuming a
ity (Kleinebudde, 2004; Miller, 2005). fixed granulator sieve size between pilot and commercial scale, ribbon
Roller compaction is a continuous unit operation, meaning that a pri- porosity is the most important quality descriptor of the granules' com-
mary batch may be split into several sub-batches and processed at var- pression properties and size distribution, as discussed by Allesø et al.
ious combinations of process parameters. This approach carries the (2013). The proper scale-factor for compaction pressure may be found
potential of efficient formulation development as well as flexibility to through empirical testing at the larger scale or through the application
shift between various scales by either adjusting press roller speed of powder mechanics theory (Johanson, 1965) combined with experi-
and/or gap size; however, only when their impact on granule quality mental data (Reynolds et al., 2010; Rowe et al., 2013). Mechanics theory
is fully understood. can be applied to determine the nip angle/in the compaction zone when
Scale-up from lab/pilot to commercial scale can pose challenges that switching to larger roll diameters. In roller compaction, the nip angle in-
are difficult to foresee as well as mitigate; not least because these issues dicates the compaction zone from start to end of densification, which
occur late in development where the formulation composition is often therefore dictates the total pressure applied to the powder (at a given
fixed. This restricts scale-up control strategies to the adjustment of roll force) and ultimately impacts ribbon porosity. There are studies
demonstrating the use of instrumented roller compactors to acquire
⁎ Corresponding author at: NNE Pharmaplan A/S, Nybrovej 80, DK-2820 Gentofte,
the pressure profile of the powder material in the nip region
Denmark. (Bindhumadhavan et al., 2005; Miguélez-Moran et al., 2008; Nesarikar
E-mail address: MnAq@nnepharmaplan.com (M. Allesø). et al., 2012) as well as a recent example of roller compaction modeling

http://dx.doi.org/10.1016/j.ejps.2015.11.001
0928-0987/© 2015 Elsevier B.V. All rights reserved.

Please cite this article as: Allesø, M., et al., Roller compaction scale-up using roll width as scale factor and laser-based determined ribbon porosity
as critical material attrib..., European Journal of Pharmaceutical Sciences (2015), http://dx.doi.org/10.1016/j.ejps.2015.11.001
2 M. Allesø et al. / European Journal of Pharmaceutical Sciences xxx (2015) xxx–xxx

by computational FEM combined with experimental data (Cunningham higher speed than the previously described, well-established volume
et al., 2010). As an alternative to first-principle understanding, Liu et al. displacement approaches. The laser technique provides a result in ap-
(2011) modeled historic data from a small-scale roller compactor by proximately 2 min per ribbon, avoids the use of any consumable sup-
Joint-Y PLS regression and applied it to predict the scale factor for roll plies and is simple to utilize. Although the technique is commercially
pressure on a larger scale compactor, having a larger roll diameter. available, literature demonstrating its performance is still sparse. There-
Parts of the scale-up challenges in roller compaction may also be fore, one aim of this study was to compare the laser-based porosity
averted through innovative designing of the different equipment scales. method against a volume displacement method using low viscosity oil
The flagship models of Gerteis® compactors include the Mini-Pactor® (medium chain triglyceride, MCT). Another aim of the present study
and the Macro-Pactor®. A scale-comparable front-view of the two com- was to investigate the hypothesized scale independence of ribbon po-
pactors is shown in Fig. 1. The Mini-Pactor® is the R&D and pilot-scale rosity, when roller compaction scale-up was done by means of a roll
machine although the continuous nature of this unit operation makes width increase from 25 mm to 100 mm on Gerteis® compactors.
it suitable for production scale also depending on the processing time.
The Macro-Pactor® is Gerteis®' production scale compactor. It is very
similar to the Mini-Pactor®, the two most notable differences being 2. Materials and methods
four times wider press rollers (25 mm vs. 100 mm) and a feeding
system with two tamp augers instead of one in addition to the single 2.1. Materials
feed auger that both machines have. The fact that rollers have identical
diameters between the two scales ensures an identical nip area and thus Microcrystalline cellulose (Avicel PH102) and croscarmellose sodium
a similar impact of roll pressure and gap size on the conveyed powder (Ac-Di-Sol) were sourced from FMC Health and Nutrition (Philadelphia,
material. Consequently, one may claim ribbon porosity to be scale inde- PA, USA) and used as filler/binder and superdisintegrant, respectively.
pendent for these particular roller compactor types. To the best of our Magnesium stearate (Peter Greven GmbH & Co.KG, Bad Münstereifel,
knowledge there are no published studies demonstrating this. Germany) was added as lubricant to reduce friction between the powder
Determination of ribbon porosity is not as trivial to determine as for and metal surfaces of the roller compactor. Medium chain triglyceride
a tablet. A tablet has well defined dimensions and, using punch draw- was obtained from Delios V (Illertissen, Germany).
ings, for example, the volume of the tablet can be calculated at a specific
tablet thickness. When the tablet mass and the density of the particles 2.2. Experimental design
are known, the porosity can easily be acquired. Ribbons on the other
hand are irregular in shape and a frequently used method to determine All samples consisted of microcrystalline cellulose (diluent) and
the volume of the inter-particular voids is a volume displacement meth- 0.75% magnesium stearate. A two-factor, two level design was carried
od where a medium of known density is absorbed into the pores of the out at each roller compactor equipment scale, with the factor gap size
sample. A long-established technique in this category is mercury at levels 2.0 and 4.0 mm and the factor, roll force at levels 3.0 and 11.0
porosimetry, its major drawback being the toxicity of the mercury. kN/cm. The study included a center point with a gap of 3.0 mm and
Environmentally friendly alternatives are intrusion with low viscosity roll force of 7.0 kN/cm. For the comparison of the two porosity measure-
oil (Allesø et al., 2013; Khorasani et al., 2015) or a fine-particulate ment techniques (laser and oil; see below), a total of six ribbons were
solid material such as the Dry Flo® medium used for the GeoPyc® analyzed per experiment (five in total) per equipment scale (two), i.e.
(www.micromeretics.com). 6 × 5 × 2 = 60 ribbons in total. Every ribbon was first analyzed by the
Nkansah et al. (2008) have described a simple throughput approach non-destructive laser method and finally subjected to the destructive
to solid fraction1 determination. Ribbons are collected in real-time with- oil intrusion method.
in a certain process time frame and weighed. The volume of ribbon For comparison of roller compactor equipment scales, the sample
material produced in that time frame is calculated from the process population was increased from six to n = 12, the additional six ribbons
parameters gap and roll speed as well as a correction for viscoelastic only being analyzed by the laser method.
expansion. The mass and volume (per time unit) divided provide the An additional four experiments were conducted separately with the
at-gap ribbon density. An obvious drawback of this technique is the intention to sample an entire 100 mm ribbon from the Macro-Pactor®
large sample size required for accurate porosity determination. The di- (see below) and assess the within-ribbon porosity variability. Four com-
mensional method is perhaps the most simple of all, where the volume binations of gap and roll force where investigated, but this time at rela-
of the ribbon is determined by cutting the ribbon to a rectangular shape tively high roll pressures, since whole ribbons are difficult to sample
with a trimming knife. This method, however, has a very limited use when they get too fragile. The investigated settings were 2.0 and
since fragile and small ribbons, which often are the case when API is 4.0 mm gap and 7.0 and 11.0 kN/cm.
present in the formulation, cannot be cut properly. Furthermore the di-
mensional approach does not take into account thickness variations of
ribbons manufactured using knurled or serrated rollers. Spectroscopic 2.3. Roller compaction
techniques have also been described in the literature, though they do
not directly measure the density of the compact but rather a property The microcrystalline cellulose was mixed with magnesium
correlated with the density. These methods include near-infrared stearate in a Bohle LM40 bin blender of suitable size (L.B. Bohle,
(NIR) spectroscopy (Lim et al., 2011; Acevedo et al., 2012; Khorasani Ennigerloh, Germany). Dry granulation was carried out on the small-
et al., 2015; Souihi et al., 2015), where the impact of light scattering er scale Mini-Pactor® (Gerteis® Maschinen + Processengineering
on the baseline is considered. Recently terahertz (THz) spectroscopy AG, Jona, Switzerland) equipped with 25 mm wide rolls as well as
has been described, which measures the effective refractive index of on its larger scale equivalent, the Macro-Pactor®, equipped with
the ribbon that can be correlated to the density (Sullivan et al., 2015). 100 mm wide rolls. Gap size and roll force were varied according to
The use of a laser-based direct volume measurement device is a re- the design described in the previous section. Roll speed was
cently introduced technique for determination of ribbon porosity (Iyer 6.0 rpm and the feed/tamp auger ratio was 250%. Auger speed was
et al., 2014a,b). The large number of ribbons needed for scale-up inves- automatically adjusted by the gap control feature in order to ensure
tigations, call for a method that efficiently analyzes ribbon porosity at a a stable gap during processing. A stable gap was confirmed for all in-
dividual experiments of this study. Sample collection (see below)
was performed once the roller compactor had reached its target set-
1
Porosity = (1 − solid fraction) ∗ 100% ting for gap and roll force.

Please cite this article as: Allesø, M., et al., Roller compaction scale-up using roll width as scale factor and laser-based determined ribbon porosity
as critical material attrib..., European Journal of Pharmaceutical Sciences (2015), http://dx.doi.org/10.1016/j.ejps.2015.11.001
M. Allesø et al. / European Journal of Pharmaceutical Sciences xxx (2015) xxx–xxx 3

Fig. 1. Scale-comparable front view of Gerteis®’ A) Mini-Pactor® and B) Macro-Pactor® roller compactors. The rolls, and thus produced ribbons, are four times wider for the Macro-
Pactor® as shown in the bottom of the picture.

2.4. Sample collection during roller compaction 2.5. Ribbon porosity determination

2.4.1. Mini-Pactor® (small scale) 2.5.1. Laser-based technique


The process was run without granulator to avoid clogging of the The porosity of the ribbons was measured using a laser-based direct
space between the rolls and the granulator region and to minimize volume measurement device (Solid Fraction Measurement Systems,
wear on the screen mesh. This was particularly relevant in the experi- Centerbrook, CT, USA). The approach uses two laser displacement sen-
ments where hard ribbons (i.e. low porosity) were produced. Samples sors capable of measuring the distances to the top and bottom of the rib-
of ribbons from the Mini-Pactor® were collected in real-time from a bon. Since the distance between the two lasers is known, it is possible to
sample collection port, located underneath the rollers and the left-side compute the thickness of the object being measured at the point of mea-
scraper. Of the collected ribbons from all five batches, most were al- surement. The various distances relevant to the thickness measurement
ready of a size suitable for porosity measurement by the laser and oil in- are illustrated in Fig. 2.
trusion method (see below). Given this relationship, the thickness of the sample (T) at the laser
beam is given by:

2.4.2. Macro-Pactor® (large scale) T ¼ D−U−L


Even after scraping off the left side roller, the ribbons were still too
large to sample while the equipment was running. Since representative
sampling is best ensured through sampling in real-time, it was decided where U and L are the distances measured by the Upper and Lower la-
to configure the granulator with an open star rotor, however, operating sers to the top and the bottom of the sample, respectively, and D is
it without any screen support. That way the granulator broke the large the known distance between the lasers.
100 mm ribbons into smaller ribbon fragments that could be sampled By making measurements of the thickness of the sample at a large
in real-time from the granulator outlet. Most of these fragments were number of points on the surface of the sample, and then integrating
of suitable size for the following porosity measurements by the laser this thickness data, one can determine the volume of the sample being
and oil intrusion method (see below). tested. In practice, the ribbon is held by a set of jaws attached to a com-
The four experiments in the intra-ribbon variability study required puter controlled precision XYZ stage, the XY direction controlling the
sampling of an entire ribbon. To achieve this, the process was allowed movement during scanning of a ribbon.
to run to steady-state (i.e. target settings of gap and roll force) and After the volume measurements are complete, the ribbon is trans-
thereafter brought to immediate stop. A 100 mm wide ribbon attached ferred (automatically) to a precision weighing module within the
to the left-side roller was manually removed and kept for later fragmen- laser-based system, where its mass (mribbon) is determined. Given the
tation with a trimming knife and porosity determination by the laser mass of the ribbon and the volume of the sample, the density (ρribbon)
method. of the sample can be computed as ρribbon = mribbon/Vribbon.

Please cite this article as: Allesø, M., et al., Roller compaction scale-up using roll width as scale factor and laser-based determined ribbon porosity
as critical material attrib..., European Journal of Pharmaceutical Sciences (2015), http://dx.doi.org/10.1016/j.ejps.2015.11.001
4 M. Allesø et al. / European Journal of Pharmaceutical Sciences xxx (2015) xxx–xxx

oil. The vacuum was then released and the sample was allowed to ab-
sorb oil into the internal ribbon pores for 2 h. Excess oil was removed
by wiping the ribbons carefully with a tissue. The amount of oil
absorbed (moil) was calculated by measuring the increase in mass of
the sample. Given that the density of the oil (ρoil) and the density of
the particles (ρp) were known, the volume of the oil (Voil) and the vol-
ume of the particles (Vp) could be calculated as basis for the calculation
of ribbon porosity (εribbon):
 
V oil
εribbon ¼  100%
V oil þ V p

where Voil = moil/ρoil and Vp = mribbon/ρp.


Medium chain triglyceride oil was used in this study and its density
at room temperature is 0.942 g/cm3

2.6. Statistical analysis

Statistical comparison of the medians from two groups of samples


(Mini-Pactor® vs Macro-Pactor®) was performed by the Mann–Whitney
U test using SigmaPlot (ver. 11.2.0.5, Systat Software, Inc., San Jose CA,
USA). The results were considered significant if P b 0.05

3. Results and discussion

3.1. Performance of the laser method for ribbon porosity determination

Ribbon porosities measured by the laser and oil intrusion method


are listed in Tables 1 and 2, reflecting samples processed on the small-
scale Mini-Pactor® and large-scale Macro-Pactor®, respectively. Due
to varying combinations of roll force and gap size, the five batches
span a relatively broad range of ribbon porosities, batch #1 including
the most fragile ribbons with the highest porosity (approximately
50%), while the hardest ribbons (approximately 20% porosity) were
found in batch #3. Overall the laser and oil intrusion method provided
very similar results and both corroborated the effect of gap size and
roll force on ribbon porosity that one would expect. Increasing the roll
force from 3.0 to 11.0 kN/cm, at a fixed gap of 4.0 mm, increased the
powder densification and consequently reduced ribbon porosity from
Fig. 2. Illustration of the basic principle of the laser-based ribbon volume measurement. 51.1% (laser) to 25.2% (laser) on the Mini-Pactor® (Table 1), with com-
parable numbers on the Macro-Pactor® being 50.6% and 26.0%, respec-
tively. At a fixed gap of 2.0 mm, a similar increase in roll force decreased
Given the operator's input of the density of the particles (ρp), which the laser-determined ribbon porosity from 47.1% to 21.2% (Mini-
were compressed to form the ribbon, the ribbon porosity (εribbon) can be Pactor®), and 44.5% to 21.6% (Macro-Pactor®), i.e. approximately the
computed as: same absolute loss in porosity irrespective of gap size suggesting that
gap size and force do not interact for this placebo formulation. The effect
!
ρribbon of gap size was positively correlated to the ribbon porosity, although
ε ribbon ¼ 1−  100%: much less pronounced than the effect of force. Here an increased gap
ρp
size from 2.0 to 4.0 mm increased the porosity by approximately 6% at
both a 3.0 kN/cm roll force (batches #2 and #1) and at 11.0 kN/cm
The reader is referred to the work by Iyer et al. (2014a,b) for a more roll force (batches #3 and #4). This negative effect of roll force on ribbon
detailed description of the system. porosity is logical and follows the same principles of powder densifica-
In the current study, a scan step size of 0.7 mm in the Y-direction tion as for any compaction process, e.g. tablet compression. While stud-
(resolution) was applied to almost all ribbon measurements, providing ies investigating the effect of gap can be found, they are less common
a total analysis time (from ribbon loading to display of result) of less than those related to roll force since many studies operate the roller
than 2 min. In the Macro-Pactor® intra-ribbon variability study (see compactor at a fixed gap. Rambali et al. (2001) conducted dry granula-
the Experimental Design section), a few ribbon fragments, regarded as tion trials on a compactor with movable gap and inclined roll configura-
exceptionally small in size, were scanned at a 0.4 mm step size. The tion, i.e. similar to the type used in this study, and found that an increase
scan step size in the X-direction was fixed at 0.01 mm. The value for in gap from 2.0 to 3.0 mm slightly, yet significantly, increases the tensile
ρp (true density) of the MCC and magnesium stearate mixture used in strength of the resulting tablets compressed from the roller compactor
the current study was 1.565 g/cm3 and was determined using an (RC) granules. Although they did not assess the ribbon density, the
AccuPyc 1330 (Micromeritics, Norcross, GA, USA). higher tablet strength suggests better recompactability of the MCC-
based granules, which ultimately is linked to higher ribbon porosity,
2.5.2. Oil intrusion method and thus a higher number of inter-particulate bonding sites available
The sample ribbon (with known weight) was kept under vacuum for subsequent tablet compression. Bindhumadhavan et al. (2005) in-
(approximately 1 mBar) in a desiccator and immersed in a low viscosity vestigated the effect of gap size using an instrumented roller compactor

Please cite this article as: Allesø, M., et al., Roller compaction scale-up using roll width as scale factor and laser-based determined ribbon porosity
as critical material attrib..., European Journal of Pharmaceutical Sciences (2015), http://dx.doi.org/10.1016/j.ejps.2015.11.001
M. Allesø et al. / European Journal of Pharmaceutical Sciences xxx (2015) xxx–xxx 5

Table 1
Mini-Pactor® ribbon porosity results from the oil-based method and laser method. Each ribbon was first measured by the non-destructive laser-based system, then followed by an oil
intrusion measurement, the latter being destructive.

Mini-Pactor®

Batch # Technique Ribbon porosity, % (six ribbons) Avg. St.dev. ΔPorosity


(%) (%) (oil-laser)
1 2 3 4 5 6

#1 Oil 51.1 49.9 50.8 51.8 51.1 51.4 51.0 0.6 −1.0%
Gap: 4.0 mm Laser 50.5 51.8 52.2 52.4 52.5 52.4 52.0 0.8
Force: 3.0 kN/cm
#2 Oil 44.5 44.8 44.4 44.5 45.8 45.7 45.0 0.6 −2.1%
Gap: 2.0 mm Laser 46.2 46.7 47.7 46.3 48.0 47.5 47.1 0.8
Force: 3.0 kN/cm
#3 Oil 18.9 18.5 18.1 18.0 18.5 19.0 18.5 0.4 −2.7%
Gap: 2.0 mm Laser 21.3 21.0 21.4 21.0 21.1 21.4 21.2 0.2
Force: 11.0 kN/cm
#4 Oil 25.3 25.1 24.9 25.2 25.4 25.2 25.2 0.2 −2.0%
Gap: 4.0 mm Laser 27.2 27.7 26.8 26.8 27.1 27.5 27.2 0.4
Force: 11.0 kN/cm
#5 (CP) Oil 33.1 31.8 32.3 32.9 31.8 32.7 32.4 0.6 −2.0%
Gap: 3.0 mm Laser 35.9 33.9 34.1 34.4 33.6 34.9 34.5 0.8
Force: 7.0 kN/cm

and found that the average maximum pressure between the rolls de- porosities subtracted from the oil intrusion porosities are depicted in
creases at higher roll gap, eventually leading to more porous ribbons. Fig. 4 as quartiles (n = 6) in a box plot against the average ribbon poros-
The ribbon porosities shown in Tables 1 and 2 are depicted graphically ity determined by the laser method. For the 25 mm ribbons (Mini-
in Fig. 3. The linear correlation between the laser- and the oil-based poros- Pactor®), the biggest underestimation of the oil-based method was
ity data was excellent (R2 = 0.99) for both the Mini- and Macro-Pactor® found for the strongest ribbons of low average porosity (21.2%, batch
datasets. For the six samples from each batch manufactured on the Mini- #3). These samples were underestimated by as much as 3%, whereas
Pactor®, both porosity measurement techniques provided very low the samples in-between approximately 27 and 47% porosity (average
standard deviations (SD's), where the laser SD was slightly less than values) were generally underestimated by 2% (batches #2, #4 and
0.8% and the oil-based SD was slightly less than 0.6%. The porosity distri- #5). In contrast, for the most fragile ribbons (51.5% porosity, batch
bution of the six samples collected from the large scale Macro-Pactor® #1) the differences in porosities were close to 0% confirming better
was broader, with SDs from 1.9% to 3.0% and 1.6% to 3.1% for the laser agreement between the two techniques than with the other ribbon
and oil-based methods, respectively. The excellent correlation between batches. A similar pattern of underestimations was identified in the
the two methods for the Macro-Pactor® (same ribbons analyzed), irre- Macro-Pactor® dataset, although the lowest porosity ribbons (21.6%)
spective of the porosity variability within a batch, confirmed that the were underestimated by approximately 2%, which was in the same re-
higher SDs were in fact attributed to the roller compaction process and gion as for the ribbon batches of 28 and 45% porosity (average values).
not a lack of accuracy of the porosity determinations. This impact of It can be argued as to whether the discrepancy between the two
scale on the porosity distribution is further investigated in the following methods represents an underestimation of the oil intrusion method or
section. an overestimation by the laser method. However, the nature of the vol-
Although the two porosity determinations performed equally well in ume displacement method supports the hypothesis that this method is
their ability to detect systematic and random porosity variations en- underestimating the results slightly. Both the lasers and the balance of
countered during processing, the positive offset of the regression line the laser system were calibrated against certified standards, and the vol-
of 3.6% and 3.2% for the Mini- and Macro-Pactor® suggests a tendency ume and mass provided can therefore be assumed to be highly accurate.
of the oil intrusion method to underestimate the porosity relative to Volume displacement methods, such as the one used in the present
the laser method. The average value (n = 6) of the laser-based study, rely on the complete absorption of the liquid medium into the

Table 2
Macro-Pactor® ribbon porosity results from the oil-based method and laser method. Each ribbon was first measured by the non-destructive laser-based system, then followed by an oil
intrusion measurement, the latter being destructive.

Macro-Pactor®

Batch # Technique Ribbon porosity, % (six ribbons) Avg. St. dev. ΔPorosity
(%) (%) (oil-laser)
1 2 3 4 5 6

#1 Oil 50.3 49.2 49.9 49.8 50.0 53.7 50.5 1.6 −0.1%
Gap: 4.0 mm Laser 50.8 49.9 48.2 49.4 51.5 53.9 50.6 2.0
Force: 3.0 kN/cm
#2 Oil 44.9 45.5 39 40.7 43.8 43.6 42.9 2.5 −1.6%
Gap: 2.0 mm Laser 46.0 46.8 41.3 42.3 45.6 44.9 44.5 2.2
Force: 3.0 kN/cm
#3 Oil 19.0 16.2 21.1 22.4 17.7 21.1 19.6 2.4 −2.0%
Gap: 2.0 mm Laser 22.2 18.3 22.9 22.8 20.4 23.1 21.6 1.9
Force: 11.0 kN/cm
#4 Oil 22.5 25.5 28.1 23.5 27.5 28.9 26.0 2.6 −1.5%
Gap: 4.0 mm Laser 24.0 27.7 28.8 24.7 29.3 30.5 27.5 2.6
Force: 11.0 kN/cm
#5 (CP) Oil 36.7 35.2 33.5 29.7 29.6 29.9 32.4 3.1 −2.1%
Gap: 3.0 mm Laser 38.5 38 34.4 32.0 32.2 32.0 34.5 3.0
Force: 7.0 kN/cm

Please cite this article as: Allesø, M., et al., Roller compaction scale-up using roll width as scale factor and laser-based determined ribbon porosity
as critical material attrib..., European Journal of Pharmaceutical Sciences (2015), http://dx.doi.org/10.1016/j.ejps.2015.11.001
6 M. Allesø et al. / European Journal of Pharmaceutical Sciences xxx (2015) xxx–xxx

Fig. 3. Plot of ribbon porosities determined by oil intrusion against the laser-based results Fig. 4. Box plot showing the distribution of differences between ribbon porosities deter-
for (A) Mini-Pactor® ribbons and (B) Macro-Pactor® ribbons. Each sample coordinate in mined by oil intrusion and the laser method at given mean porosity (measured by the
the plot represents the same ribbon analyzed first by the laser system and then by oil laser). Data are shown for A) Mini-Pactor® and B) Macro-Pactor®. A negative value indi-
intrusion. cates an underestimation by the oil-based method. The bottom and top of each box are the
first (25%) and third (75%) quartiles, and the band inside is the second quartile or median
(50%). The red dashed line is the mean value (n = 6).

pores of the compact, and the liquid will not be able to enter closed
pores of the compact. Incomplete intrusion would result in an underes- Collectively, the data suggested that the laser system was very suit-
timation of the pore volume and thus porosity, a probability that is log- able for further investigations of the Mini- and Macro-Pactor® ribbons,
ically expected to be higher for hard compacts and lower for fragile hence it was the only porosity measurement technique used in the fol-
compacts. The experimental results plotted in Fig. 4 support this lowing scale-up investigations.
explanation.
Finally, the non-destructive nature of the laser method enabled the 3.2. Impact of roller compaction scale on ribbon porosity
assessment of repeatability, an analytical parameter that cannot be ac-
quired for the oil intrusion method. Three single ribbons, from batches As evident from Tables 1 and 2, ribbon fragments from the Macro-
which represented the edges of the experimental space (batches #1 Pactor® runs yielded a higher standard deviation compared to ribbons
and #3) and the center point (batch #5), were selected for repeatability collected from the Mini-Pactor®. This was not unexpected since every
analysis. Each of these ribbons was measured six consecutive times with collected ribbon sample from the Mini-Pactor® represented the entire
the laser system. The porosity results, including the raw data volume 25 mm roll width (see picture of ribbon in Fig. 1). In contrast, a sample
and mass, are presented in Table 3. With a standard deviation of merely population of n = 6 ribbons from the Macro-Pactor®, although sampled
0.2% across the entire porosity range from 21 to 53%, the laser system in real-time, consisted of random fragments from several 100 mm rib-
clearly performs consistently with a high precision. Interestingly, the bons, which resulted in a broader porosity distribution. The variability
raw data reveal a very minor wear of the ribbons during the repeated of 25 mm ribbons collected from the Micro-Pactor® may be considered
performance of the six measurements. This was seen as a reduction in to be a true measure of between-ribbon variability, whilst for the
the initial mass from start to end by approximately 6 mg for the fragile Macro-Pactor® the measurement variability represents a mixture of
ribbons and 2 mg for the strong ribbons. This wear was mainly a conse- within-ribbon (i.e. fragments taken from the same 100 mm primary rib-
quence of the jaws grasping the ribbon, as this was the only stress ap- bon) and between-ribbon variability. Hence, to improve sampling in an
plied to the ribbons. This, however, does not impact the validity of the attempt to obtain a better measurement of the average porosity of the
results since the ribbon volume correspondingly was measured to be 100 mm primary ribbon, the sample population was increased from
lower by the lasers. six to twelve. This increase in population size was done for both the

Please cite this article as: Allesø, M., et al., Roller compaction scale-up using roll width as scale factor and laser-based determined ribbon porosity
as critical material attrib..., European Journal of Pharmaceutical Sciences (2015), http://dx.doi.org/10.1016/j.ejps.2015.11.001
M. Allesø et al. / European Journal of Pharmaceutical Sciences xxx (2015) xxx–xxx 7

Table 3
Results demonstrating the repeatability of the laser-based technique. Six repeated measurements were performed on a single ribbon selected from each of three separate.

Measurement no. Batch #1 Batch #3 Batch #5


(fragile ribbon) (intermediate ribbon) (hard ribbon)

Volume Mass Porosity Volume Mass Porosity Volume Mass Porosity


(cm3) (mg) (%) (cm3) (mg) (%) (cm3) (mg) (%)

1 528.4 391.5 52.7 1020.0 1033.2 35.3 533.4 652.9 21.8


2 523.2 390.0 52.4 1021.6 1031.8 35.5 529.5 652.2 21.3
3 527.9 389.5 52.9 1019.4 1030.7 35.4 529.0 651.9 21.3
4 526.8 388.3 52.9 1019.4 1030.0 35.4 528.1 651.7 21.2
5 522.5 385.7 52.8 1018.7 1029.4 35.4 527.9 651.2 21.2
6 522.1 385.2 52.9 1011.2 1028.9 35.0 528.8 651.0 21.3
Average 52.8 Average 35.3 Average 21.4
St. dev. 0.2 St. dev. 0.2 St. dev. 0.2

Mini-Pactor® and Macro-Pactor® samples, and again the former twelve level. The data of Table 4 are depicted graphically in Fig. 5. The excellent
samples were complete 25 mm ribbons, and the latter twelve samples linear correlation (R2 = 0.996) between ribbon porosities obtained on
were twelve fragments chosen randomly from the 100 mm primary rib- the two compactor scales confirm the expected 1:1 scalability of the
bon. Table 4 summarizes the average porosities from the 12 measure- Mini- and Macro-Pactor®.
ments per batch and also shows the raw data, ribbon volume and To gain more insight into the variability between ribbons
mass, constituting the basis of the reported porosity results. The in- manufactured on the Macro-Pactor®, four additional batches were
crease in sample population did not improve the ribbon porosity distri- manufactured, with settings of force and gap providing average ribbon
bution of the Macro-Pactor®, but rather slightly increased the standard porosities which ranged from approximately 28% to 37%. While in gen-
deviation (now up to 3.3%). Standard deviations were generally very eral 100 mm ribbons break up into fragments as they leave the roller,
low in the ribbons from the Mini-Pactor®, 1.0% or lower, with the ex- the relatively low porosity of these special batches allowed collection
ception of batch #1. Although not shown in Table 4, a single measure- of complete 100 mm wide ribbons. Three whole ribbons per batch
ment from Mini-Pactor® batch #1 was primarily responsible for the were sectioned into eight segments with a trimming knife (Fig. 6) and
larger SD. (11 out of 12 measurements fell within the 50.5% to 52.3% po- the porosity of each individual segment was determined using the
rosity range, with a single outlier measurement at 45.1%. Omitting the laser-based method giving a total of 3 × 8 measurements per batch.
outlier from the dataset provides an average Mini-Pactor® ribbon po- The porosity data are listed in Table 5 also showing the total measured
rosity for batch #1 (n = 11) of 51.9% and brings the SD down to 0.6%, volume per 100 mm ribbon and its corresponding mass. The average
which was comparable to the SDs of the all other ribbon batches from porosity of each of the three ribbons within a batch (n = 8) were very
the Mini-Pactor®. It was decided, however, not to discard the outlier similar, which confirms that the roller compaction process, in spite of
measurement, since triplicate measurements on that same ribbon some slight heterogeneity within a ribbon, performed well and that
gave similar porosities (45.1, 45.3, and 45.5%). the process was robust. The SDs (n = 8), for the 8 segments of the rib-
There was an excellent agreement between the average ribbon po- bon, was in the range of 1 to 3%, which was comparable to the SDs ob-
rosity obtained on the Mini- and Macro-Pactor® when operating the tained in the experiments previously described, based on 12 randomly
compactor at identical settings of roll force and gap size. A statistical collected ribbon fragments. There was no apparent relationship be-
comparison of medians did not suggest any difference between the tween the chosen gap size and the within-ribbon standard deviation.
two compactor scales for batches #1, #2, #4 and #5 at 95% confidence. To investigate whether any pattern exists in the porosity distribution
From a practical point of view the average ribbon porosities of batch #3 of a 100 mm ribbon, the porosity data for each ribbon fragment in
were very similar for the two scales, 20.8% for the Mini-Pactor® and Table 5 are presented graphically in Fig. 7. As the sectioning was done
22.0% for the Macro-Pactor®, but due to low SDs on both scales (1.0% manually by hand, the dimensions of the individual fragments varied
and 1.7%) they became statistically different at the 95% confidence and comparisons should be made with some caution. However, the

Table 4
Ribbon measurement results (n = 12) from the Mini- and Macro-Pactor® trials, respectively. All measurements were done by the laser system and the table shows raw data (volume and
mass) and the calculated ribbon porosity. The P-value is derived from the Mann–Whitney U test for comparison of two medians.

Batch # Equipment scale Ribbon raw data Ribbon porosity, %

Total volume, mm3 Total mass, mg Average St. dev. P-value


(n = 12) (n = 12) (n = 12) (n = 12)

#1 Mini-Pactor® 10,623 8,050 51.3 2.0 0.795


Gap: 4.0 mm Macro-Pactor® 12,675 9,631 51.5 1.8
Force: 3.0 kN/cm
#2 Mini-Pactor® 6,940 5,738 47.2 0.7 0.214
Gap: 2.0 mm Macro-Pactor® 7,206 6,152 45.6 3.3
Force: 3.0 kN/cm
#3 Mini-Pactor® 8,088 10,038 20.8 1.0 0.017
Gap: 2.0 mm Macro-Pactor® 6,722 8,195 22.0 1.7
Force: 11.0 kN/cm
#4 Mini-Pactor® 7,491 8,554 27.0 0.4 0.525
Gap: 4.0 mm Macro-Pactor® 6,166 7,005 27.5 3.0
Force: 11.0 kN/cm
#5 (CP) Mini-Pactor® 10,646 10,977 34.1 0.7 0.272
Gap: 3.0 mm Macro-Pactor® 10,824 11,013 35.0 2.7
Force: 7.0 kN/cm

Please cite this article as: Allesø, M., et al., Roller compaction scale-up using roll width as scale factor and laser-based determined ribbon porosity
as critical material attrib..., European Journal of Pharmaceutical Sciences (2015), http://dx.doi.org/10.1016/j.ejps.2015.11.001
8 M. Allesø et al. / European Journal of Pharmaceutical Sciences xxx (2015) xxx–xxx

well as process conditions (gap size, roll speed, roll width, and compac-
tor design) can influence the density distribution across the ribbon
width. Miguélez-Moran et al. (2009) applied micro-indentation,
among others, to measure the change in relative density across a
200 mm ribbon width manufactured on a lab scale instrumented roller
compactor using gravity feeding. The authors demonstrated a tendency
of all ribbons to be most dense in the middle (least porous) relative to
both edges. This difference was most pronounced for ribbons
manufactured at the largest gap (here 1.0 mm compared against
0.9 mm) and high roll speeds (6 rpm). The authors attributed this to
the friction between the powder and fixed side plate, inhibiting powder
pull-in at the edges and consequently less powder compaction at the
edges since the gap was fixed. The importance of the friction between
powder and side sealing material was confirmed by the addition of lu-
bricant, which improved the density distribution across the ribbon
width (Miguélez-Moran et al., 2008). Using NIR chemical imaging (at-
line) Khorasani et al. (2015) investigated the difference in relative den-
sity of an MCC-based formulation processed on a pilot scale roller com-
pactor with rolls mounted in the vertical position and powder feeding
with a single screw. They found ribbons to be more heterogeneous
when compacted at higher forces, with the edges being the least
Fig. 5. Plot showing the relationship between ribbon porosities from the Mini-Pactor® and dense. It is, however, difficult to interpret the density distribution
Macro-Pactor® experiments. Ribbons from each of the five levels have been manufactured maps since these were provided as relative values based on the changes
at identical process parameter settings on the Mini- and Macro-Pactor® (roll force and
in the slope and shift of the NIR spectrum baseline (i.e. scatter effects).
gap). The two dashed lines outline the 95% confidence interval of the regression line.
Guigon and Simon (2003) investigated the heterogeneity in density of
sodium chloride ribbons processed on a lab-scale roller compactor
graphs are useful for revealing potential systematic differences in the equipped with 130 mm smooth rolls. They found the density distribu-
porosity between the middle (e.g. fragments 3 to 6) and the edges tion to be periodical and linked to the period of the screw feeder, and ul-
(e.g. fragments 1, 2 and 7, 8) of a ribbon. Although there was a tendency timately the uneven distribution of feed pressure. In a very recent study
for some ribbons to be more porous in the middle and less porous at the Sullivan et al. (2015) demonstrated the application of terahertz spec-
edges, this pattern was not consistently observed within a batch. For ex- troscopy for density mapping (via refractive index, RI) of ribbons
ample, two of the ribbons of batch #4 show this pattern while a third manufactured on a roller compactor with inclined roll configuration,
ribbon shows the exact opposite pattern. Therefore, for this experimen- i.e. similar to the one used in the present work. Information on process
tal setup the heterogeneity in porosity seemed to be randomly defined parameter settings in this study is very sparse, but translation of the RI
during the roller compaction process. This observation was in contrast profiles into porosity generally revealed lower porosity in the middle
to other studies, which revealed some repeatable patterns in ribbon relative to the edges, which is in accordance with the research of others
density distribution across the roll width. However, direct comparisons described above, but not the present. The pattern, however, did not
with studies from different research groups are difficult, since formula- seem to be constant as a function of process time, which is probably
tion factors like filler compression properties and powder lubrication as due to the period of the feed screw described by Guigon and Simon

Fig. 6. Photograph of a Macro-Pactor® ribbon before and after sectioning with a trimming knife. Each ribbon segment has been numbered so that ribbons 1 and 8 represent the edges of the
whole ribbon with 4–5 representing the center (approximately).

Please cite this article as: Allesø, M., et al., Roller compaction scale-up using roll width as scale factor and laser-based determined ribbon porosity
as critical material attrib..., European Journal of Pharmaceutical Sciences (2015), http://dx.doi.org/10.1016/j.ejps.2015.11.001
M. Allesø et al. / European Journal of Pharmaceutical Sciences xxx (2015) xxx–xxx 9

Table 5
Porosities of the eight ribbon fragments, covering the entire 100 mm width of a Macro-Pactor ribbon. Results from 4 combinations of roll force and gap (both at 2 levels) are shown. (Three
ribbons were fragmented and analyzed per experiment, and all measurements were done by the laser system.) The table shows raw data (volume and mass), and the calculated ribbon
porosity.

Batch # Ribbon # Ribbon porosity, % Ribbon raw data

Ribbon segment # Average St. dev. Total volume, mm3 Total mass, mg
(n = 8) (n =8) (n = 8) (n = 8)
1 2 3 4 5 6 7 8
(edge) (edge)

#1 A 30.5 31.6 32.2 32.5 32.2 32.3 32.1 29.9 31.7 1.0 6,080 6,502
Gap: 2.0 mm B 30.6 31.2 31.4 31.5 32.0 32.3 32.9 30.4 31.5 0.8 6,483 6,951
Force: 7.0 kN/cm C 29.2 31.5 33.8 34.1 31.4 32.9 30.4 28.2 31.4 2.1 6,318 6,779
#2 A 33.9 36.8 37.4 37.2 38.9 37.8 35.8 34.0 36.5 1.8 7,910 7,873
Gap: 4.0 mm B 37.3 38.2 30.4 34.7 35.0 37.5 39.0 38.4 36.3 2.8 5,849 5,783
Force: 7.0 kN/cm C 37.5 37.1 35.9 35.5 37.9 37.4 36.7 36.9 36.9 0.8 4,912 4,854
#3 A 24.1 22.2 23.1 21.4 21.0 20.6 20.5 20.8 21.7 1.3 6,082 7,437
Gap: 2.0 mm B 22.2 22.7 24.9 23.1 23.0 21.0 19.0 18.1 21.8 2.3 6,665 8,143
Force: 11.0 kN/cm C 16.6 18.9 23.0 24.6 24.0 25.2 22.5 19.9 21.8 3.1 6,668 8,168
#4 A 26.8 27.0 30.0 30.6 30.0 30.8 26.9 20.3 27.8 3.5 5,989 6,741
Gap: 4.0 mm B 27.3 28.4 29.1 28.8 28.7 28.4 27.5 21.9 27.5 2.4 4,964 5,616
Force: 11.0 kN/cm C 32.7 29.5 29.1 26.3 26.5 26.8 27.2 29.6 28.5 2.2 5,578 6,267

(2003). In the present investigation, the period of the feed screw may 4. Conclusion
also explain the lack of a repeatable pattern in the porosity distribution
of the 100 mm ribbons plotted in Fig. 7. Considering the inherent high The major aim of this study was to investigate the hypothesis of rib-
sensitivity of a plastic material to deformation upon pressure, like the bon porosity being scale-independent when using roll width as scale up
MCC used here, these within-ribbon variations were deemed minor parameter, more specifically, that ribbon porosity is not affected by a
from a practical point of view and the average porosity value of a change of roll width from 25 mm to 100 mm at constant RC parameter
100 mm ribbon was very repeatable. settings. A direct and rapid laser-based method was used to determine

Fig. 7. Porosities of the eight ribbon segments, covering the entire 100 mm width of a Macro-Pactor ribbon. Results from 4 combinations of roll force and gap (both at 2 levels) are shown.
(Three ribbons were sectioned and analyzed per experiment).

Please cite this article as: Allesø, M., et al., Roller compaction scale-up using roll width as scale factor and laser-based determined ribbon porosity
as critical material attrib..., European Journal of Pharmaceutical Sciences (2015), http://dx.doi.org/10.1016/j.ejps.2015.11.001
10 M. Allesø et al. / European Journal of Pharmaceutical Sciences xxx (2015) xxx–xxx

the porosity of the ribbons at-line. With a repeatability of 0.2%, and Guigon, P., Simon, O., 2003. Roll press design—influence of force feed systems on compac-
tion. Powder Technol. 130, 41–48.
excellent agreement with a well-established volume displacement Iyer, R.M., Hegde, S., DiNunzio, J., Singhal, D., Malick, W., 2014a. The impact of roller com-
method, the laser-based method was indeed a powerful and elegant ap- paction and tablet compression on physicomechanical properties of pharmaceutical
proach to acquire ribbon porosity rapidly (less than 2 min per ribbon). excipients. Pharm. Dev. Technol. 19, 583–592.
Iyer, R.M., Hegde, S., Singhal, D., Malick, W., 2014b. A novel approach to determine solid
The ribbon porosity was found to be scale-independent when com- fraction using a laser-based direct volume measurement device. Pharm. Dev. Technol.
paring the average porosity of five groups of ribbon samples (each with 19, 577–582.
n = 12 samples) from the small-scale (Mini-Pactor®) to the large-scale Johanson, J.R., 1965. A rolling theory for granular solids. ASME J. Appl. Mech. Ser. E 32 (4),
842–848.
(Macro-Pactor®). The higher standard deviation of the measured po- Khorasani, M., Amigo, J.M., Sonnergaard, J., Olsen, P., Bertelsen, P., Rantanen, J., 2015. Visu-
rosities of ribbons segments from the large-scale roller compactor was alization and prediction of porosity in roller compacted ribbons with near-infrared
attributed to minor variations in powder densification across the roll chemical imaging (NIR-CI). J. Pharm. Biomed. Anal. 109, 11–17.
Kleinebudde, P., 2004. Roll compaction/dry granulation: pharmaceutical applications. Eur.
width (100 mm), and is in agreement with reports from the literature.
J. Pharm. Biopharm. 58, 317–326.
This within-ribbon variability was not assessed on the smaller scale Lim, H., Dave, V.S., Kidder, L., Lewis, E.N., Fahmy, R., Hoag, S.W., 2011. Assessment of the
Mini-Pactor® ribbons since each ribbon measured naturally represent- critical factors affecting the porosity of roller compacted ribbons and the feasibility
ed the entire roll width (25 mm). of using NIR chemical imaging to evaluate the porosity distribution. Int. J. Pharm.
Sci. 410, 1–8.
With the intention to reproduce ribbon porosity from one scale to Liu, Z., Bruwer, M.-J., MacGregor, J.F., Rathore, S.S.S., Reed, D.E., Champagne, M.J., 2011.
the other, process settings of roll force and gap size applied to the Scale-up of a pharmaceutical roller compaction process using a joint-Y partial least
Mini-Pactor® (and identified during formulation development) are squares model. Ind. Eng. Chem. Res. 50, 10696–10706.
Miguélez-Moran, A.M., Wu, C.-Y., Seville, J.P.K., 2008. The effect of lubrication on density
therefore directly transferrable to subsequent commercial scale produc- distributions of roller compacted ribbons. Int. J. Pharm. Sci. 362, 52–59.
tion on the Macro-Pactor® with same roll diameter. This creates a better Miguélez-Moran, A.M., Wu, C.-Y., Dong, H., Seville, J.P.K., 2009. Characterisation of density
link between formulation development and technology transfer, and distributions in roller-compacted ribbons using micro-indentation and X-ray micro-
computed tomography. Eur. J. Pharm. Biopharm. 72, 173–182.
decreases the number of batches needed to establish the parameters Miller, R.W., 2005. Roller compaction technology. In: Parikh, D.M. (Ed.), Handbook of
of the commercial process. pharmaceutical granulation technology. Taylor & Francis Group, New York,
pp. 159–190.
Nesarikar, V.V., Patel, C., Early, W., Vatsaraj, V., Sprockel, O., Jerzweski, R., 2012. Roller
Acknowledgments compaction process development and scale up using Johanson model calibrated
with instrumented roll data. Int. J. Pharm. Sci. 436, 486–507.
The authors are thankful to Uffe Teis Svare and Jacob West for their Nkansah, P., Wu, S.-J., Sobotka, S., Yamamoto, K., Shao, Z.J., 2008. A novel method for
estimating solid fraction of roller-compacted ribbons. Drug Dev. Ind. Pharm. 34,
skilled contributions to the experimental work. Terry Chriss and Mi-
142–148.
chael Hanisch are thanked for the technical discussions. Rambali, B., Baert, L., Jans, E., Massart, D.L., 2001. Influence of the roll compactor parame-
ter settings and the compression pressure on the buccal bio-adhesive tablet proper-
References ties. Int. J. Pharm. Sci. 220, 129–140.
Reynolds, G., Ingale, R., Roberts, R., Kothan, S., Gururajan, B., 2010. Practical application of
Acevedo, D., Muliadi, A., Giridhar, A., Litster, J.D., Romanach, R.J., 2012. Evaluation of three roller compaction process modeling. Comput. Chem. Eng. 34, 1049–1057.
approaches for real-time monitoring of roller compaction with near-infrared spec- Rowe, J.M., Crison, J.R., Carragher, T.J., Vatsaraj, N., McCann, R.J., Nikfar, F., 2013. Mechanis-
troscopy. AAPS PharmSciTech 13, 1005–1012. tic insights into the scale-up of the roller compaction process: A practical and dimen-
Allesø, M., Torstenson, A.S., Bryder, M., Holm, P., 2013. Presenting a rational approach to sionless approach. J. Pharm. Sci. 102, 3586–3595.
QbD-based pharmaceutical development: a roller compaction case study. Eur. Souihi, N., Nilsson, D., Josefson, M., Trygg, J., 2015. Near-infrared chemical imaging (NIR-
Pharm. Rev. 6, 17–24. CI) on roll compacted ribbons and tablets — multivariate mapping of physical and
Bindhumadhavan, G., Seville, J.P.K., Adams, M.J., Greenwood, R.W., Fitzpatrick, S., 2005. chemical properties. Int. J. Pharm. Sci. 483, 200–211.
Roll compaction of a pharmaceutical ingredient: experimental validation of rolling Sullivan, M., Heaps, D., McKay, R., Kato, E., Zhou, X.H., Wu, C.-Y., Pei, C.-L., Zhang, J.-Y.,
theory for granular solids. Chem. Eng. Sci. 60, 3891–3897. Schiano, S., 2015. Density mapping of roller-compacted ribbons using terahertz spec-
Cunningham, J.C., Winstead, D., Zavaliangos, A., 2010. Understanding variation in roller troscopy. Tablets & Capsules: granulation analysis, pp. 26–32 (January).
compaction through finite element-based process modeling. Comput. Chem. Eng.
34, 1058–1071.

Please cite this article as: Allesø, M., et al., Roller compaction scale-up using roll width as scale factor and laser-based determined ribbon porosity
as critical material attrib..., European Journal of Pharmaceutical Sciences (2015), http://dx.doi.org/10.1016/j.ejps.2015.11.001

Anda mungkin juga menyukai