Sei sulla pagina 1di 15

Pharmaceutical Development and Technology, 10:423–437, 2005

Copyright D 2005 Taylor & Francis Inc.


ISSN: 1083-7450 print / 1097-9867 online
DOI: 10.1081/PDT-200054462

Impact of Solid-State Properties on Lubrication Efficacy


of Magnesium Stearate
Pharmaceutical Development and Technology Downloaded from informahealthcare.com by University of Guelph on 05/04/12

K. Phanidhara Rao, Garima Chawla, Aditya M. Kaushal, and Arvind K. Bansal


Department of Pharmaceutical Technology (Formulations), National Institute of Pharmaceutical
Education and Research, Punjab, India

INTRODUCTION
The advent of high-speed tableting and slug capsule-filling
machines has ushered in an increasingly important role for the Most tablet formulations contain an additive designed
lubricants to enact during manufacturing of dosage forms. to reduce the intergranular and die wall friction, called
Although lubricants help in processing, they can also adversely lubricants. Magnesium stearate is the most common
affect the flow properties and dissolution profile of the drug. It is choice among lubricants.[1] These additives function by
thus critical to maintain a balance between these two behaviors,
interposing a film of low shear strength at interface be-
by understanding the underlying mechanisms and using their
tween the tableting mass and the die wall. Lubricants with
For personal use only.

optimum concentration in the formulation. The source and


manufacturing process inculcate different solid-state properties low shear strength but cohesive tendencies perpendicular
to magnesium stearate, the most commonly used lubricant, to shear plane serve the purpose legitimately.[2] Lubri-
leading to variations in its lubrication efficacy. However, there cants, besides plummeting frictional forces during manu-
has been no complete study relating the lubrication efficacy of facturing of compressed solid dosage forms, can affect
magnesium stearate to various levels of solid state. Hence, this many other quality parameters such as weight variation,
study was aimed at comprehensively scrutinizing the role of hardness, disintegration, time and dissolution rate.[3] The
molecular, particle, and bulk level properties of solid state on the role of lubricants has assumed greater significance with
lubrication efficacy of magnesium stearate. A method based on the introduction of high-speed tableting machines, as
net work done during compression using texture analyzer, was shorter dwell times and faster compression rate pose
developed and validated to analyze its performance. Particle and
greater demands on their performance.[4] Similarly, their
bulk-level properties were studied using microscopy, particle
size analysis, and particle surface area determination, and
lubrication efficiency critically affects weight variation in
molecular level was characterized using thermal, spectroscopic, dosator- and tamping-based capsule-filling machines.[5]
and crystallographic methods. Interplay of solid-state character- Magnesium stearate improves the performance of adhe-
istics such as particle size, degree of agglomeration, and crystal sive materials (lactose, calcium carbonate, etc.), prevents
habit were found to markedly influence the lubrication potential material buildup on tampering faces and shafts, and
of magnesium stearate. decreases the ejection force leading to reduced wear of
these automatic slug-forming equipments. However,
Keywords magnesium stearate, lubrication efficacy, net work magnesium stearate is known to adversely affect the flow
of compression, solid-state characterization properties and dissolution rate, both in tablets and
capsules, due to its hydrophobic nature. Higher levels of
magnesium stearate in capsules may lead to the formation
of hydrophobic powder beds that do not disperse after the
dissolution of capsule shell. Hence, there is a need to
control the levels of magnesium stearate in the formulation
Received 6 October 2004, Accepted 13 December 2004.
Address correspondence to Arvind K. Bansal, Department of
to satisfy both the content uniformity and dissolution
Pharmaceutical Technology (Formulations), National Institute of acceptance criteria, by understanding the parameters
Pharmaceutical Education and Research, Sector—67, S.A.S. controlling the performance of magnesium stearate.
Nagar, Punjab 160 062, India; Fax: 91-0172-2214692; E-mail: Magnesium stearate most closely resembles an ideal
akbansal@niper.ac.in lubricant with low friction coefficient, large ‘‘covering

423
Order reprints of this article at www.copyright.rightslink.com
424 K. P. Rao et al.

potential,’’ and a high melting point.[6] However, it is Wallingstown, Ireland), dicalcium phosphate (Engranule,
reported to suffer from source, manufacturer-to-manufac- Enar Chemie Pvt. Ltd., Gujarat, India), and six commer-
turer, and batch-to-batch variation in their physicochem- cial samples from different manufacturers (designated as I
ical properties that are mainly attributable to their solid- to VI) of magnesium stearate were used for the study. Sam-
state properties.[7 – 9] Solid-state properties of a substance ple I was of British Pharmacopoeia (BP) grade; samples II,
can be classified into three levels—molecular (crystalline/ IV, and V were of National Formulary (NF) grade; and
amorphous forms), particle (habit, size distribution), and samples III and VI were of Indian Pharmacopoeia (IP) grade.
bulk level (density, porosity). The relationship between
the three levels is such that an intentional or unintentional
Pharmaceutical Development and Technology Downloaded from informahealthcare.com by University of Guelph on 05/04/12

alteration at one level may significantly influence the METHODS


other two, thereby affecting the physicochemical and
physicotechnical performance. This could also translate Development of Method for Testing
into severe processing problems, keeping in view the Lubrication Efficacy
multiple functions performed by the lubricant. Attempts
have been made to understand the critical contributors to Settings of Texture Analyser
the performance of lubricant, and there are numerous
contradictory reports interrelating surface area, particle All experiments were done with a 50-kg load cell
size,[10,11] and lamellar structure[12] to the lubrication using a 10-mm, flat-faced, die punch set, and texture
efficacy. However, the previous reports do not provide a analyser (TA-XT 2i, Stable Microsystems, Godalming,
collective picture of all solid-state levels and how change Surrey, UK) was calibrated for both the distance and the
at one level may be interrelated to the effect seen at the load cell before starting an experiment. The following two
other level. Therefore, there is a need to assess the types of methods were used.
lubrication efficacy by a suitable method and to be aware
For personal use only.

of the critical parameters contributing to variation in Method 1. The force required to travel a particular
lubrication efficacy, thus allowing selection of optimum distance (4 mm) was measured. Net work done was
grade and quantity of the lubricant. calculated from area under the curve. Initial position of the
In this study, a simple method based on the net work upper punch was kept constant in all the experiments using
done using texture analyzer has been developed to the calibration scale. The other parameters were fixed as-
evaluate the lubrication efficacy, and an attempt has been punch travel speed of 2 mm/sec before and after the test,
made to characterize the ‘‘best’’ and the ‘‘worst’’ 0.1 mm/sec during the test, and a trigger/stimulus was set
performers, and relate their performance to solid-state to ‘‘activate after traveling a distance of 1 mm.’’
properties. Compaction involves two distinct stages—
compression (reduction in bulk volume due to displace- Method 2. Distance traveled by the punch to attain
ment of gaseous phase) and consolidation (increase in a particular force (475 N) was determined. Net work done
mechanical strength due to cold welding and fusion was calculated as in method 1. Stimulus was set to 0.05 N
bonding between the particles).[2] Any lubricant needs to (‘‘just after touching the bed’’), which triggered the
perform during both the stages. This method based on use recording. The remaining texture analyser settings were
of texture analyzer mimics the initial compression stage similar to that of method 1.
of tableting and during slug formation in dosator- and
tamping-type, slug-forming, capsule-filling machines. Six Selection of Suitable Excipient
commercially available samples of magnesium stearate
were studied for their performance as a lubricant. All Three diluents—dicalcium phosphate, lactose, and
three levels of solid-state properties viz. molecular, microcrystalline cellulose—were tested for their suitabil-
particular, and bulk were investigated systematically to ity for use in further experiments. Accurately weighed
identify the parameters critical for the lubricant efficacy quantities of dicalcium phosphate, lactose, and microcrys-
of magnesium stearate. talline cellulose were compressed either neat or with 1%
(w/w) magnesium stearate (grade II) using texture
analyser, according to method 1. The diluents were mixed
EXPERIMENTAL in geometric progression with magnesium stearate and
then blended for 5 min in a double cone blender (VDM
Materials 4SP, Kalweka, Ahmedabad, India) at 150 rpm prior to
compression. Excipient showing minimum variation
Lactose (Meggle, Germany), microcrystalline cellu- (%RSD) based on 10 experiments was selected for fur-
lose (PH 102, FMC International, Little Island, Cork, ther experimentation.
Solid-State Properties and Magnesium Stearate 425

Validation of Texture Analyser-Based Method plotting bulk density and tapped density curves against
the number of taps. Tapped density was plotted against
Intraday and interday variability were determined, for number of taps by performing the tap density experiment
both methods 1 and 2, using the work of compression on blends filled in a graduated measuring cylinder. The
keeping the maximum distance traveled by upper punch cylinder was subjected to 2000 taps at a rate of 300 taps/
constant (4 mm) and measuring the distance traveled to min on Electrolab1 tap density apparatus (ETD-1020,
attain a fixed force (475 N), respectively. Reproducibility Mumbai, India). The volume was recorded at 2, 4, 8, 16,
and repeatability were checked by performing the 50, 100, 300, and 750 taps. The profiles were compared
experiment on 3 successive days and thrice on a day. with that of neat dicalcium phosphate.
Pharmaceutical Development and Technology Downloaded from informahealthcare.com by University of Guelph on 05/04/12

Comparative Performance of Various Samples of


Determination of Flow Rate
Magnesium Stearate
Flow rate of the blend was determined using Erweka
The effect of different magnesium stearate samples
Granulate tester (Type: GT; Erweka, Heusenstamm,
on intergranular friction, and on both the intergranular and
Germany). Time taken for 100 g quantity of blend to
die wall friction, was determined by different methods.
flow through a 6-mm orifice was recorded in triplicates.
Method A for Intergranular Friction. Method 1
was followed where the maximum distance traveled was Solid-State Characterization
4 mm during the compaction to study the intergranular
friction. A weight of 580± 5 mg (weight enough to fill Solid-state characterization of commercially sourced
a 10-mm die) of each blend having different samples of samples was studied at each of the three levels of solid state.
magnesium stearate (1% w/w) was filled into the die
Particle and Bulk Level
For personal use only.

cavity and compressed. First, five compressions were done


to coat the die wall with magnesium stearate. Using the Microscopy. Microscopic studies were done using
same blend, an additional six compressions were given to optical microscope (Leica DMLP, Leica, Wetzlar,
determine the average of net work done required to travel a Germany) attached to an imaging system (DC-300,
distance of 4 mm for that blend. The amount of net work Leica). In optical microscope, images were acquired at
done needed to compress only dicalcium phosphate was a magnification of 630  (63  10) using the Image
used as the control while switching over to the other blend Manager (IM50, V1.20 Twain module, Leica, Germany)
with different magnesium stearate sample. software. Crystal habit, state of agglomeration, and
relative size were observed.
Method B for Both Intergranular and Die Wall Particle Size Analysis. Particle size analysis was
Friction. This experiment was done to study the effect carried out using laser particle sizer (Analysete 22, Fritsch
of different magnesium stearate samples on the combined GmbH, Idar Oberstein, Germany) having a measuring
effect of intergranular and die wall frictions. The ex- range of 0.18 to 102.89 mm, with a helium-neon laser
periment was done using method 2, where the maximum source with less than 5 mW power and Version-C software
force to be achieved was set at 475 N. One percent (w/w) to compute the particle size distribution. Around 20 mg of
blends of different magnesium stearate samples with the sample was dispersed in water, ultrasonicated with
dicalcium phosphate were prepared in the similar fashion stirring at 75 rpm to disrupt the agglomerates and pumped.
as in method A. The method of experimentation was The equipment works on the principle of diffraction
identical to method A, but without any preliminary patterns (Fraunhofer rings) produced when a monochro-
compression cycle to coat the die wall. The die wall was matic laser light strikes the boundary edge of the two
cleaned with cotton swabs dipped in isopropyl alcohol to media with a different refractive index. The diameter of
remove a layer of magnesium stearate deposited from the particle is inversely proportional to the radius of the
previous experiment. first circle of the light produced due to diffraction and is
calculated using the following equation:
Particle diameter ¼ 1:81  f  l=Ro
Flowability Test
where f, l, and Ro are the focal length of the imaging lens,
Density Measurements wavelength of laser light, and radius of the Fraunhofer
ring, respectively.
The effect of different magnesium stearate samples Particle Surface Area Determination. Specific
on the improvement of flow properties was studied by surface area was determined for the selected samples
426 K. P. Rao et al.

using Coulter surface area determination apparatus RESULTS AND DISCUSSION


(Coulter SA 3100, version 2.11; Fullerton, CA, USA),
which works on the BET (Brunauer, Emmett, and Teller) Method Development for Testing
principle. The experiment was done using both single- Lubrication Efficacy
point and multipoint methods. Weighed quantity (about
200 mg) of sample was degassed at 40C for 1 hr and the Selection of Suitable Excipient
weight of sample after degassing was taken as the initial
weight. Amount of nitrogen adsorbed was calculated by The %RSD of maximum force achieved on the punch
the difference in weight before and after adsorption. The to travel a certain distance showed that dicalcium
Pharmaceutical Development and Technology Downloaded from informahealthcare.com by University of Guelph on 05/04/12

specific surface area was calculated using the phosphate showed minimum degree of variation and
BET principle. was therefore selected for further experimentation. The
results are shown in Figure 1.
Molecular Level
Thermogravimetric Analysis. Thermogravimetric Validation of Texture Analyser-Based Method
analysis (Mettler Toledo, TGA/SDTA 851, Greifensee,
Switzerland, Mettler STAR, version 5.1) was done for The validation exercise confirms that the method is
each sample placed in alumina crucible, from a suitable for use and produces results that are accurate and
temperature range of 30C to 210C at the rate of 5C/ precise (i.e., the system works as it ought to work). The
min under nitrogen purge (20 mL/min). texture analyser-based methods were validated for
Differential Scanning Calorimetry. Differential performance by checking interday and intraday variations.
scanning calorimetry (DSC; 821 Mettler Toledo, Grei- Both methods 1 and 2 showed a %RSD of < 3%; interday
fensee, Switzerland, operating with STAR, version 5.1) %RSD 2.52 and 1.01, respectively, and intraday %RSD
was done between temperatures  10C to 210C at a 2.15 and 0.75, respectively.
For personal use only.

heating rate of 5C/min and nitrogen purge of 100 mL/


min. Samples (2 – 5 mg) were tested in sealed aluminum Comparative Performance of Various Samples of
pans with a pinhole, and an empty pan was used as a Magnesium Stearate
reference. The temperature axis and cell constant of
DSC were previously calibrated with a high-purity Lubrication efficacy was studied using method A (for
standard of Indium. intergranular friction) and method B (for both intergran-
Hot Stage Microscopy. Optical microscope was ular and die wall friction) to select the magnesium
attached with a hot stage (Leica LMW 50, Leica, Ger- stearate samples (‘‘most effective performer’’ and ‘‘least
many) to observe the thermal events. The samples were
observed with and without silicone oil, covered with a
glass cover slip and heated. Images were acquired using the
Image Manager (IM50, V1.20 Twain module) software.
FTIR Studies. Fourier-transform infrared (FTIR)
spectra were recorded using the KBr pellet method in the
region 400 to 4000 cm 1 on PerkinElmer spectropho-
tometer (FTIR multiscope/synthetic monitoring system,
PerkinElmer, Buckinghamshire, UK), equipped with
Spectrum v3.02 software. Background spectrum was
collected under identical conditions. The spectrum for
each sample (an average of 16 coadded scans) was
recorded at a resolution of 4 cm 1, Fourier transformed
and ratioed against background interferogram.
Powder XRD Studies. Powder x-ray diffraction
Figure 1. Relative standard deviation (n = 10) of maximum
(XRD) patterns of samples were recorded at room
force achieved on the punch to travel a certain distance by
temperature on x-ray powder diffractometer (Seifert, different excipients and excipient – magnesium stearate blend.
Iso-Debyeflex 2002, Ahrensburg, Germany) using Cu LAC, lactose; MCC, microcrystalline cellulose; DCP, dicalcium
Ka radiation (l =1.54184’) at 30 kV, 20 mA passing phosphate; LAC + MS, lactose – magnesium stearate blend;
through nickel filter. Data were collected in a continuous MCC + MS, microcrystalline cellulose – magnesium stearate
scan mode with a sweep of 1.2/min over an angular blend; DCP + MS, dicalcium phosphate – magnesium stearate
range of 5 to 70 2q. blend.
Solid-State Properties and Magnesium Stearate 427

effective performer’’) for further characterization and to The results of net work done were subjected to one-
evaluate the role of various solid-state levels on the way analysis of variance (ANOVA) to check whether
performance of magnesium stearate. Work of compres- there was any significant difference in the net work done,
sion can be used as one of the parameters for estimating and hence, the lubrication efficacy of different blends.
lubrication efficacy. Net work done was determined either The data were found to show normal distribution even at
by measuring the maximum achievable force or maxi- 99% confidence limits. Both Tukey (more conservative—
mum distance traveled by upper punch. One parameter is less likely to show statistically significant difference) and
determined to keep the other constant. Method A provides Student-Newman-Keuls (less conservative—more likely
an assessment of intergranular friction, as the initial stage to show statistically significant difference) tests illustrated
Pharmaceutical Development and Technology Downloaded from informahealthcare.com by University of Guelph on 05/04/12

of compression involves rearrangement and repacking of the difference in the mean net work done among the
particles, with the contribution of die wall friction being ‘‘effective performers’’ (samples I and II) and the ‘‘less
negligible. Method B studies the later stages of compres- effective performers’’ (samples III and IV) was more than
sion, wherein contribution of both intergranular and die that expected by chance (P< 0.05), hence showing a
wall friction need to be accounted for. Figure 2a and 2b statistically significant difference between them.
show the net work done to travel a distance of 4 mm and The results obtained from method B, by measuring
the distance traveled by upper punch at constant force, the distance traveled by the upper punch to achieve a
respectively, and the order of efficacy of different samples specified force, were in close agreement to that of method
of magnesium stearate. The less work needed for a blend, A (Figure 2b). Because the upper punch moves with a
the better is its lubrication efficacy, as seen in sample I, constant speed, time to achieve the maximum force can be
which was found to be superior to all other samples. used to judge the lubrication efficacy. The longer the
time, the more the distance traveled and the better the
lubrication efficacy. To circumvent the problem of
differences in the bulk densities of different samples, all
For personal use only.

blends including dicalcium phosphate were precom-


pressed to attain a constant bed volume. This also helps
normalize the effects caused due to differences in
particulate arrangement after filling the die cavity. Two
best (samples I and II) and two worst (samples III and IV)
performers were selected for further characterization to
recognize the solid-state properties imperative for perfor-
mance as a lubricant.

Flowability Test

Flow characteristics of blends were characterized in


their static and dynamic stages by performing density and
flow rate experiments, respectively.

Density Measurements

The plot of tapped density against number of taps is


shown in Figure 3. Magnesium stearate enhanced the rate
of packing of blends on tapping. Therefore, it can be
concluded that magnesium stearate facilitates in reaching
an ultimate value of density in a shorter period. However,
no appreciable difference was found between different
samples of magnesium stearate.
Figure 2. Measurement of lubrication efficacy of magnesium
stearate samples. (a) Method A: net work done by the upper Flow Rate Measurements
punch to travel a distance of 4 mm. (b) Method B: distance
traveled by upper punch at force of 475 N. Note: The error bars Results of the flow rate were found to be interesting
represent the %RSD of n = 6. because ‘‘poor performers’’ in a lubrication efficacy test
428 K. P. Rao et al.
Pharmaceutical Development and Technology Downloaded from informahealthcare.com by University of Guelph on 05/04/12

Figure 3. Density measurements – comparative behavior of different blends on tapping.

improved the flow rate to a greater extent as compared with reverse correlation between the lubrication efficacy and
the ‘‘better performers.’’ Samples I (33.5 s) and II (33.9 s) tendency to improve flow of magnesium stearate samples
took longer time to flow, in contrast to samples III (32.0 s) was seen (Figure 4). During flow of a powder, unlike as in
and IV (31.8 s). The data were treated with one-way compression, much space is available around the particles,
ANOVA and were found to show normal distribution. facilitating rolling and sliding of particles over each other
For personal use only.

Tukey and Student-Newman-Keuls tests exemplified that with less resistance and friction. In contrast, during
there was a statistically significant difference in the flow compression, movement of particles is restricted, and the
rate of samples I and II versus samples III and IV blends, at relative movement of static and dynamic planes in opposite
99% confidence intervals with P<0.05. These differences direction generate resistance, restricting the rolling move-
can be attributed to the intrinsic cohesiveness reflected as ment. Therefore, the particle characteristics required to
agglomeration of the lubricants. Optical microscopy improve flowability and to diminish the frictional forces
(discussed later) revealed that sample III contained more during compression are bound to be different.
agglomerates, which might cause a ‘‘ball bearing’’-like Flow of powders involves the rolling or sliding of
effect, facilitating rolling of particles over each other. A particles over one another, with continuous change in

Figure 4. Lubrication and flow performance of good performer and bad performer blends of magnesium stearate. The bars show
the average net work done by the upper punch to travel a distance of 4 mm, and the points depict the results of flowability test (the
values are the time in seconds taken for 100 g of blend to flow through the sample holder [orifice diameter = 6 mm]). Note: The error
bars represent the %RSD of n = 6 for net work done, and n = 3 for flowability test.
Solid-State Properties and Magnesium Stearate 429

point of contact. Compared with this, compaction the selected samples was done in two steps to determine
involves relatively static particles with relative movement the critical contributors:
of static and dynamic planes in the opposite direction
against a lot of resistance. The particle characteristics 1. Characterization of particle- and bulk-level parameters
required to improve flowability and to decrease frictional (i.e., state of agglomeration, surface area, particle size,
forces during compression are expected to be different. and crystal habit)
Magnesium stearate samples having particles in a 2. Characterization of molecular-level parameters (i.e.,
deagglomerated state will distribute quickly and perform hydration state and crystal structure)
better in the lubrication efficacy test. However, a smaller
Pharmaceutical Development and Technology Downloaded from informahealthcare.com by University of Guelph on 05/04/12

particle size will lead to poorer flow properties. Studies on Particle Size, Shape, and Specific
Surface Area. Because lubrication mechanism is a
Solid-State Characterization surface phenomenon, it can be correlated to properties
such as particle size and specific surface area. The particle
Lubrication efficacy of magnesium stearate has been size and surface area of selected magnesium stearate
associated with factors such as chemical composition and samples I to IV are reported in Table 1. The difference in
its physicochemical properties, such as the specific the particle size and area demonstrates the batch-to-batch
surface area, particle size distribution, moisture content, and manufacturer-to-manufacturer variability in magne-
hydration state, crystal structure, and crystal habit.[13,14] sium stearate. The order of samples in terms of specific
Earlier reports have shown that fatty acid content (stearic surface area was sample I> III >II > IV, and in terms of
and palmitic) do not influence magnesium stearate to an particle size was sample IV > Iffi II ffi III. Smaller particles
extent to alter its lubrication efficacy.[9,14,15] Magnesium tend to accommodate themselves well on other particles
stearate samples with different chemical compositions and form an effective coating, thereby reducing the
have been reported to exhibit the same lubrication ef- frictional forces. Sample III, despite having particle size
For personal use only.

ficacy, and the same batch of the lubricant with different distribution equivalent to I and II, as well as surface area
physicochemical properties has shown different perfor- better than II, did not perform well. This can be attributed
mance patterns, hence suggesting the role of properties to presence of agglomerates as was evident in optical
other than composition in the lubrication efficacy of mag- microscopy. The sample preparation during determination
nesium stearate. of particle size distribution (PSD) involves sonication that
Different reports claim the role of diverse properties breaks the agglomerates into individual particles, and
such as particle size and surface area,[9,14] or moisture thus, does not account for agglomerates present in the
content and crystalline structure[15] or bulk density,[6] to original samples. Effective coating, however, can signif-
influence its lubrication properties. Hence, the study was icantly increase the disintegration and dissolution times,
aimed at comprehensively studying the role of all solid- as well as decrease the hardness of the tablets,[16] because
state levels on the lubrication efficacy of magnesium the film makes it difficult for interparticle bonding to
stearate. Characterization at different solid-state levels of occur.[13] Milling, which leads to a change only in particle

Table 1
Particle size and specific surface area of selected grades of magnesium stearate

Sample of magnesium stearate

Experiment I II III IV

Particle size (mm) d1% > 24.12 d1% > 27.41 d1% > 20.37 d1% > 84.63
d50% > 6.12 d50%> 7.21 d50%> 5.85 d50%> 24.42
d95%> 1.12 d95%> 1.22 d95%> 1.11 d95%> 1.38
BET one-point surface area (m2/g) 5.49 2.75 3.89 1.31
BET specific surface area (m2/g) 6.63 3.24 4.52 1.66
Correlation coefficient 0.9946 0.9965 0.9976 0.987
Monolayer volume,* Vm at STP# (cc/g) 1.5224 0.7445 1.0395 0.3814

Note: From texture analyzer studies, it was found that samples I and II are high-efficiency lubricants in contrast to samples III and
IV that are less effective.
*The larger the surface area, the more volume of gas adsorbed.
#
Standard temperature and pressure.
430 K. P. Rao et al.
Pharmaceutical Development and Technology Downloaded from informahealthcare.com by University of Guelph on 05/04/12
For personal use only.

Figure 5. Optical photomicrographs of different samples of magnesium stearate at a magnification of 630  . (a) Sample I,
(b) sample II, (c) sample III, and (d) sample IV.

size without causing change in any other property Thermal Analysis. DSC and thermogravimetric
(moisture content, hydration state, etc.), improved lubri- analysis (TGA) were performed on magnesium stearate
cation efficacy confirming direct correlation.[13,14] It has samples to characterize the type of hydrate and to study
also been observed that the particle size is directly related the nature of water present because it is stated that
to radial transmission ratio (R) and tensile strength,[17] lamellar dihydrates are more efficient lubricants.[18] None
which are a measure of lubrication efficacy and bond of the samples showed endotherm corresponding to the
formation, respectively. adsorbed water (which would have been converted into
ice at subzero temperatures) when subjected to low
Optical Microscopy. Optical microscopy indicat- temperature DSC indicating nonhygroscopic nature of
ed that samples I and II have smaller particles as magnesium stearate. More than one peak of dehydration
compared with samples III and IV. The shape of crystals in DSC indicated the presence of water in different
was flat and platelike in samples I and II, whereas thermodynamic states.
agglomerates were observed in sample III (Figure 5). In sample I, desolvation in TGA was observed as an
Optical microscopy provides a better reflection of the endotherm in DSC. In TGA (Figure 6a), the weight loss
actual particle size distribution, which gets ‘‘masked’’ was observed up to 107C, indicating that the first three
during particle size distribution determinations using endotherms in DSC correspond to dehydration. The
automated processes involving sonication during sample number of water molecules calculated by TGA was found
preparation, leading to deagglomeration. to be 1.8. Both cleavage of interactions of water with the

Figure 6. DSC and TGA curves of various samples of magnesium stearate. (a) Sample I, (b) sample II, (c) sample III, and
(d) sample IV. The values in parenthesis represent the percentage weight loss, with a standard deviation of ± 0.5. Temperature range
was 10C to 210C for DSC and 35C to 210C for TGA. Sample analysis was performed in pin-holed aluminum pans in DSC and
alumina pans in TGA at a heating rate of 5C/min. All samples contain varying proportions of water molecules, showing the
inconsistency in different samples from different manufacturers.
Pharmaceutical Development and Technology Downloaded from informahealthcare.com by University of Guelph on 05/04/12
For personal use only.

Solid-State Properties and Magnesium Stearate


431
Pharmaceutical Development and Technology Downloaded from informahealthcare.com by University of Guelph on 05/04/12
For personal use only.

432

Figure 6.
K. P. Rao et al.

Continued.
Solid-State Properties and Magnesium Stearate 433

Mg2 + and the evaporation of free water, thus formed, that samples I and IV have their water almost tightly
contribute to the sum of enthalpies of peaks. bound, and prominent water loss starts at 72C. Sample II
Sample II had no sharp thermal events indicating with large bulk of loosely bound water, showed 82%
loosely bound water and less crystallinity. The two peaks weight loss by the time it reached a temperature of 85C.
in DSC can be attributed to water loss because they Sample III behaved differently because the water loss was
coincide with the thermal events of TGA. Water loss was observed until 130C, most of it being lost between 94C
observed till 115C in TGA, and the corresponding to 116C, indicating more tightly bound water compared
thermal events in DSC were found to merge with melting. with other samples (I, II, IV) that showed weight loss only
The prominent peak in sample III shows both dehydration up to 115C.
Pharmaceutical Development and Technology Downloaded from informahealthcare.com by University of Guelph on 05/04/12

and melting. A close observation of this peak (Figure 6c) From the previous results, various possibilities can be
revealed that it starts with a small hump, indicating the predicted for the arrangement of water molecules. They
dehydration of tightly bound water at 82C, followed by can be present as either noncrystalline water (adsorbed
simultaneous dehydration and melting. Single broad water) or as crystalline water (within the crystal). The
endotherm in sample IV (Figure 6d) indicated merging crystalline water can be in different thermodynamic states
of dehydration and melting events. as pure dihydrate or higher hydrates. Furthermore, a
Calculation of water molecules from TGA (Figure 6) mixture of different hydrates may have water at different
indicated that none of the samples are dihydrate in nature, energy levels.
although the loss of water was observed in two stages in
sample I, II, and III. The FTIR studies however, showed FTIR Studies. Table 2 summarizes the results of
that samples IV and I are dihydrates with respect to their FTIR. All magnesium stearate samples showed a carbonyl
crystal structure (discussed later). This anomaly is peak at about 1570 cm 1, indicating the interaction of Mg
explained by considering the magnesium stearate samples atom with the fatty acid chain changing the unionized
as mixtures of different hydrated forms with varying state of acid to an ionized state. This was confirmed in a
For personal use only.

proportions of water molecules. This also explains the previously reported work, by observing a similar peak
reason for the number of water molecules getting related to carbonyl group in the stearic acid sample at
expressed as fractions. Sample I and IV are close to 1700 cm1.[19] Bands between 3600 cm1 and 3100 cm1
dihydrate in terms of their moisture content. TGA showed provide evidence for crystal water. Bands related to fatty
(Figure 6) that all samples except sample IV have water at acid portion showed the presence of a doublet at
two or more different thermodynamic states because the 720 cm1 in samples I and IV, indicating orthorhombic
water loss was observed in more than one step. The or monoclinic structure in them.[20] This particular band is
simultaneous differential analysis (SDTA) curves showed due to the CH2 rocking mode, and its appearance is

Table 2
Summary of Fourier transform infrared results

Sample
Peak (cm 1), threshold—5%,
normalized to 0 – 100 scale Vibration I II III IV

3600 – 3100 Water of 3447 (m)


crystallization 3525 (m)
3447 (m) 3525 (m)
3546 (m) 3567 (m)
35,967 (m) 3546 (m)
3567 (m)
3567 (m)
3300 – 2500 -OH stretching 2957 (s)
2957 (s) 2957 (s) 2957 (s)
2850 (s)
2900 C – H symmetric 2918 (s) 2917 (s) 2917 (s) 2918 (s)
1700 C = O asymmetric — — — —
1616 – 1540 COO asymmetric 1542 (s) 1574 (s) 1574 (s) 1543 (s)
720 CH2 rocking 723 (m) (d) 721 (m) (sig) 722 (m) (sig) 723 (m) (d)

Note: s, strong; m, moderate; sig, singlet; d, doublet. Standard values of bands were taken from Ertel and Carstensen (1988).[15]
434 K. P. Rao et al.

related to the packing of the hydrocarbon chains within at 1542 cm1 of samples IV and I shifted to 1579 and
the crystal.[19] Furthermore, it can be concluded that these 1594 cm1, respectively, after drying. This shift was not
samples are dihydrates because the monoclinic and observed in the other two samples. It indicates that water
orthorhombic packing was shown to be consistent with loss from the higher hydrates (samples I and IV) and
this form in the metallic salts of stearic acid.[21] The TGA transformation to lower hydrates is responsible for this
results also showed that the number of water molecules is shift. Water loss leads to the absence of interaction with
close to 2. The other varieties (samples II and III) can be water, which otherwise satisfies the positive charge on the
concluded to have either hexagonal or triclinic packing Mg2 + ion to some extent. Therefore, water loss leads to
because a singlet is observed at 720 cm 1. Therefore, strengthening of the bond between fatty acid chain and
Pharmaceutical Development and Technology Downloaded from informahealthcare.com by University of Guelph on 05/04/12

they can be expected to have anhydrate or partially Mg2 +. It resulted in the COO asymmetric stretch to shift
hydrated form. to higher wave numbers. Because other samples are
The FTIR analysis of samples dried at 90C under already in a lower hydration state, they did not show
vacuum for 20 hr showed loss of water bands above much change in this band. It was also observed that the
3100 cm1 in all samples except sample III. This sample 720-cm1 band became singlet upon drying in samples I
showed two bands at 3563 and 3468 cm1, indicating the and IV confirming the previous result. It can also be
strong interaction of water with magnesium. Even in concluded that the dehydration is accompanied by the
TGA, the water loss was observed until 130C, indicating change in crystal structure, and thus, crystal water plays
the strongly bound nature of water in this sample. Band an important role in crystal system formation. Thus, from
For personal use only.

Figure 7. XRD pattern of magnesium stearate samples from 2q 5 to 70. (a) Sample I, (b) sample II, (c) sample III, and
(d) sample IV.
Solid-State Properties and Magnesium Stearate 435

Table 3
Correlation between solid-state properties at different levels and lubrication efficacy of magnesium stearate samples

Bulk level Particle level Molecular level

Sample Agglomeration a
PS (mm) a 2
SSA (m /g) Crystal habit Water molecules 720 cm 1 band d spacing

I XX No XX 6.12 XX 6.63 XX Plate XX 1.8 XX Doublet XX 11.7


II XX No XX 7.21 X 3.24 XX Plate  0.6 X Singlet X 10.5
III  Yes XX 5.85 X 4.52  Irregular  1.26 X Singlet X 10.8
IV XX No  24.42  1.66  Irregular XX 1.86 XX Doublet  9.9
Pharmaceutical Development and Technology Downloaded from informahealthcare.com by University of Guelph on 05/04/12

Note: The unshaded cells with XX show favorable properties of magnesium stearate as a lubricant, the dark-shaded cells with 
show the different properties at various levels of solid state that hamper the lubrication efficacy of magnesium stearate, and the light-
shaded cells with X show the intermediate behavior of magnesium stearate.
a
PS (d50%), particle size; SSA, specific surface area.

TGA and FTIR studies, it is evident that samples I and IV particle-level characteristics, which fundamentally influ-
have dihydrate structure, whereas samples II and III have ence the layer-forming tendency of the lubricant. As
mono or still lower hydrate structures. mentioned earlier, deagglomerated particles, smaller
particle size, greater specific surface area, and platelike
X-Ray Diffraction Studies. XRD patterns of the crystal habit are the optimum characteristics that can
four samples were studied to know the difference in attribute good lubrication efficacy. The microscopic
crystal structure and its influence on the lubrication studies revealed that samples I and II have platelike
efficacy. Variation in the composition and proportion of crystal habit. Therefore, they can delaminate during
For personal use only.

the fatty acids makes it difficult to get consistent XRD mixing and compression, which ultimately leads to
pattern. In general, the strong and fine harmonics are reduced friction. Second, lubricants act in the later stages
typical of single well-crystallized form, whereas broad of compression when the contact area between the
reflections are indicative of poorly crystallized form. particles and the die wall increases. At this stage,
Therefore, sample II with only one prominent peak at lubricants having sufficient shearability only can perform
2q =21 can be said to be ‘‘poor in crystalline content.’’ the function. Shearing involves elongation of one of the
Figure 7 shows the XRD pattern of magnesium stearate largest lattice spaces of the crystal. So, crystals having
samples. Samples I and IV indicated a crystal structure as higher lattice space perform better. The order (sample
a mixture of dihydrate and other hydrates, whereas sample I >sample III> sample II > sample IV) of highest lattice
III indicated a partially dehydrated dihydrate.[7,15,17] The spacing (d spacing) values showed the same array. Hence,
amount of moisture content (crystalline water) dictates the higher hydrates were found to improve the lubrication
crystal structure, and it was found to depend on the drying efficacy by delaminating to a greater extent. Sample I was
rate and the pH of the precipitation medium during found to be a dihydrate with greater long spacing
preparation of magnesium stearate.[22] Only a small compared with other samples and showed good lubrica-
difference (0.6 moles of water) in the number of tion efficacy. However, the lubrication efficacy of sample
molecules of water per molecule of magnesium stearate IV was not in agreement with this principle. The reason
is sufficient to convert a poorly crystalline form into a can be attributed to the other solid-state properties of
well-crystallized form. Furthermore, the crystal long different levels that overshadow the molecular level. The
spacing is a key property because it attributes shearability difference in their d spacing can be owed to the different
to the lubricant.[23] Highest d spacing (Table 3) shown by proportions of fatty acids, which ultimately dictates the
sample I and lowest by sample IV correlate well to their crystal structure. These molecular-level characteristics
good and poor lubrication efficacy, respectively. depend on the source of stearic acid, manufacturing, and
storage conditions. Each property has different extent of
contribution toward lubrication efficacy, and it is difficult
Solid-State Levels and Performance of Magnesium to compare the samples, which vary in more than one of
Stearate: A Reminisce the previously mentioned variables.
The extreme performers, samples I and IV, have
Magnesium stearate acts by coating the particles and similar crystal structure. The difference in lubrication ef-
thereby reducing the interactions between moving surfac- ficacy is purely attributed to the difference in specific
es. Particle size, surface area, and the crystal habit are the surface area, indicating that it plays an important role in
436 K. P. Rao et al.

lubrication. The d spacing is more in sample I compared to how the critical balance between the three levels affects
with all other samples, indicating its importance as its ultimate lubrication efficacy. It is this interplay be-
mentioned earlier. Samples II and III are almost similar tween the three levels that decides the exact behavior of
in FTIR studies, particle size, and specific surface area. the sample. With the use of an effective lubricant, it is
Interestingly, although sample III has slightly higher spe- possible to decrease the concentration of hydrophobic
cific surface area and smaller particle size, the lubrication lubricants in the formulation, which would otherwise be
efficacy is less compared with sample II, having less required in higher quantities, and therefore affect the
specific surface area. This can be explained on the basis of biopharmaceutical performance of the dosage form,
the bulk- and particle-level characteristics. At the bulk especially for BCS class II and IV drugs.
Pharmaceutical Development and Technology Downloaded from informahealthcare.com by University of Guelph on 05/04/12

level, sample II had no agglomerates, allowing it to be Our study revealed that all the levels of solid state
adsorbed onto the other excipient particles in the blend. At contribute to, but to a different extent individually, the
the particle level, the platelike crystal habit favored it to performance of the lubricant. Particle-level characteristics
spread uniformly in comparison to sample III, which had seemed to be the most important and overshadow the
agglomerates and irregular crystal habit. Because not molecular-level characteristics in the experimental con-
much difference was observed in the particle size and ditions. Moreover, total work of compression can be used
specific surface area of sample III over sample II, the to screen the lubricants for their efficacy and texture
previously mentioned affects could not be overshadowed, analyser is a sensitive and suitable instrument to evaluate
leading to better performance of sample II over sample III. lubrication efficacy by this method.
Although sample II is a monohydrate with a specific Such studies can be incalculably useful in scale-up of
surface area half the value of sample I, it performed well. both tablets and capsules, where discrepancy at any of the
The reason can be explained in terms of crystal habit and molecular, particle, or bulk level may multiply manyfold,
lack of agglomerates and considerable d spacing. making the processing unpredictable and affect the
It can be said that the inferior performance of sample performance of the dosage form.
For personal use only.

IV is due to the presence of irregular crystal habit, large


particle size (less specific surface area), and smaller
lattice spacing. ACKNOWLEDGMENTS
Therefore, the criterion for good lubrication can be
given in terms of different levels of solid state: We are grateful to Prof. R. K. Ray, Head, Advanced
Center for Material Sciences (ACMS), IIT Kanpur, for his
Bulk level: Absence of agglomerates kind permission to work in ACMS.
Particle level: Smaller particle size, larger specific surface
area, platelike crystal habit
Molecular level: Dihydrate in nature, especially water REFERENCES
should present in two thermodynamic states, larger
d spacing 1. Vitkova, M.; Chalabala, M. The use of some
hydrophobic substances in tablet technology. Acta
Pharm. Hung. 1998, 68 (6), 336– 344.
CONCLUSION 2. Marshall, K. Compression and consolidation of
powdered solids. In The Theory and Practice of
The commercially available magnesium stearate, Industrial Pharmacy; Lachman, L., Liberman, A.,
used as a lubricant in pharmaceutical industry, is subject Kanig, J.L., Eds.; Varghese: Bombay, India, 1976;
to quality variations imposed by source and manufactur- 66– 99.
ing process. The type of hydrate formed due to different 3. Billany, M.R.; Richards, J.H. Batch variation of
manufacturing conditions dictates many solid-state – magnesium stearate and its effect on the dissolution
related properties, which have momentous contributions rate of salicylic acid from solid dosage forms. Drug
to its lubrication performance. In this investigation, an Dev. Ind. Pharm. 1982, 8 (4), 497– 511.
attempt was made to correlate the solid-state properties 4. Banker, G.S. Tablets. In The Theory and Practice of
of magnesium stearate with the lubrication efficacy and Industrial Pharmacy; Lachman, L., Liberman, A.,
to come up with the critical parameters, which affect its Kanig, J.L., Eds.; Varghese: Bombay, India, 1976;
properties as a lubricant. The variations in solid-state 293– 345.
properties (state of agglomeration, surface area, particle 5. Jones, B.E. The filling of powders into two-piece
size, crystal habit, hydration state, and crystal structure) of hard capsules. Int. J. Pharm. 2001, 227 (1), 5 –26.
magnesium stearate provide an appealing explanation, as 6. Baichwal, A.R.; Augsburger, L.L. Development
Solid-State Properties and Magnesium Stearate 437

validation of modified annular shear cell (MASC) to and lubricant properties of magnesium stearate.
study frictional properties of lubricants. Int. J. J. Pharm. Sci. 1988, 77 (7), 625– 629.
Pharm. 1985, 26 (1), 191 –196. 16. Frattini, C.; Simioni, L. Should magnesium stearate
7. Butcher, A.E.; Jones, T.M. Some physical character- be assessed in the formulation of solid dosage forms
istics of magnesium stearate. J. Pharm. Pharmacol. by weight or by surface area? Drug Dev. Ind. Pharm.
1972, 24 (Suppl.), 1P –9P. 1984, 10 (7), 1117 –1130.
8. Bolhuis, G.K.; Lerk, C.F.; Broersma, P. Mixing 17. Bracconi, C.; Andres, C.; Ndiaye, A. Structural
action and evaluation of tablet lubricants in direct properties of magnesium stearate pseudopoly-
compression. Drug Dev. Ind. Pharm. 1980, 6, 15 –33. morphs: effect of temperature. Int. J. Pharm. 2003,
Pharmaceutical Development and Technology Downloaded from informahealthcare.com by University of Guelph on 05/04/12

9. Dansereau, R.; Peck, G.E. Effect of variability in the 262 (1), 109 –124.
physical and chemical properties of magnesium 18. Miller, T.A.; York, P. Pharmaceutical tablet lubri-
stearate on the properties of compressed tablets. cation. Int. J. Pharm. 1988, 41 (1), 1– 19.
Drug Dev. Ind. Pharm. 1987, 13 (6), 975 – 999. 19. Ertel, K.D.; Carstensen, J.T. An examination of
10. Miller, T.A.; York, P. Physical and chemical physical properties of pure magnesium stearate. Int.
characterization of some high pure magnesium J. Pharm. 1988, 42 (1), 171 – 180.
stearate and palmitate powders. Int. J. Pharm. 20. Chapman, D. The 720 cm1 band in the infrared
1985, 23 (1), 55 –67. spectra of crystalline long chain compounds. J.
11. Wada, Y.; Matsubara, T. Pseudopolymorphism and Chem. Soc. 1957, 4489 – 4491.
lubricating properties of magnesium stearate. Pow- 21. Muller, B.W. Die pseudopolymorphie von magne-
der Technol. 1994, 78 (2), 109 – 114. siumsalzen hoherer fettsauren. Arch. Pharm. 1977,
12. Buckley, D.H.; Johnson, R.L. Lubrication with 310, 693 –704.
solids. Chem. Technol. 1972, 2, 302 –310. 22. Rubino, J.T. Electrostatic and non-electrostatic free
13. Barra, J.; Somma, R. Influence of physicochemical energy contributions to acid dissociation constants
For personal use only.

variability of magnesium stearate on its lubricant in cosolvent-water mixtures. Int. J. Pharm. 1988, 42
properties. Drug Dev. Ind. Pharm. 1996, 22 (11), (1), 181– 191.
1105– 1120. 23. Shah, A.C.; Mlodozeniec, M.L. Mechanism of
14. Leinonen, U.I.; Vihervaara, P.A.; Laine, E.S.U. surface lubrication: influence of duration of lubri-
Physical and lubrication properties of magnesium cant-excipient mixing on processing characteristics
stearate. J. Pharm. Sci. 1992, 81 (12), 1194 –1198. of powders and properties of compressed tablets.
15. Ertel, K.D.; Carstensen, J.T. Chemical, physical, J. Pharm. Sci. 1977, 66, 1377 –1378.

Potrebbero piacerti anche