Sunteți pe pagina 1din 35

Therapeutic Function: Antiinflammatory

Chemical Name: m-Benzoylhydratropic acid


Common Name: 2-(3-Benzoyiphenyl)propionic acid
Structural Formula:

Trade Name Manufacturer Country Year Introduced


Profenid Specia France 1973
Zaditen Sandoz Japan 1983

Raw Materials
Ethanol
(3-Benzoylphenyl)acetonitrile
Sulfuric acid
Sodium
Methyl iodide

Manufacturing Process
In an initial step, the sodium derivative of ethyl (3-benzoylphenyl) cyanoacetate is
prepared as follows: (3-benzoylphenyl)acetonitrile (170 9) is dissolved in ethyl carbonate
(900 g). There is added, over a period of 2 hours, a sodium ethoxide solution [prepared
from sodium (17.7 g) and anhydrous ethanol (400 cc)], the reaction mixture being heated
at about 105° to 115°C and ethanol being continuously distilled. A product precipitates.
Toluene (500 cc) is added, and then, after distillation of 50 cc of toluene, the product is
allowed to cool. Diethyl ether (600 cc) is added and the mixture is stirred for 1 hour. The
crystals which form are filtered off and washed with diethyl ether (600 cc) to give the
sodium derivative of ethyl (3-benzoylphenyl)cyanoacetate (131 g). Then, ethyl methyl(3-
benzoylphenyl)cyanoacetate employed as an intermediate material is prepared as follows:
The sodium derivative of ethyl (3-benzoylphenyl)cyanoacetate (131 g) is dissolved in
anhydrous ethanol (2 liters). Methyl iodide (236 g) is added and the mixture is heated
under reflux for 22 hours, and then concentrated to dryness under reduced pressure (10
mm Hg). The residue is taken up in methylene chloride (900 cc) and water (500 cc) and
acidified with 4N hydrochloric acid (10 cc). The methylene chloride solution is decanted,
washed with water (400 cc) and dried over anhydrous sodium sulfate. The methylene
chloride solution is filtered through a column containing alumina (1,500 g). Elution is
effected with methylene chloride (6 liters), and the solvent is evaporated under reduced
pressure (10 mm Hg) to give ethyl methyl(3-benzoylphenyl)cyanoacetate (48 g) in the
form of an oil.
In the final production preparation, a mixture of ethyl methyl(3-
benzoylphenyl)cyanoacetate (48 g), concentrated sulfuric acid (125 cc) and water (125
cc) is heated under reflux under nitrogen for 4 hours, and water (180 cc) is then added.
The reaction mixture is extracted with diethyl ether (300 cc) and the ethereal solution is
extracted with N sodium hydroxide (300 cc). The alkaline solution is treated with
decolorizing charcoal (2 g) and then acidified with concentrated hydrochloric acid (40
cc). An oil separates out, which is extracted with methylene chloride (450 cc), washed
with water (100 cc) and dried over anhydrous sodium sulfate. The product is concentrated
to dryness under reduced pressure (20 mm Hg) to give a brown oil (33.8 g). This oil is
dissolved in benzene (100 cc) and chromatographed through silica (430 g). After elution
with ethyl acetate, there is collected a fraction of 21 liters, which is concentrated to
dryness under reduced pressure (20 mm Hg). The crystalline residue (32.5 g) is
recrystallized from acetonitrile (100 cc) and a product (16.4 g), MP 94°C, is obtained. On
recrystallization from a mixture of benzene (60 cc) and petroleum ether (200 cc), there is
finally obtained 2-(3-benzoylphenyl)propionic acid (13.5 g), MP 94°C.
1256

CRYSTALLIZATION OF KETOPROFEN IN PRESSURE SENSITIVE


ADHESIVE MATRIX. Hoo-Kyun Choi*1, Hye-Chin Pak1, Young-Joo Cho1, Sang-
Chul Shin2 and Jin-Hwan Lee1. 1College of Pharmacy, Chosun University, Kwangju,
Korea; 2College of Pharmacy, Chonnam National University, Kwangju, Korea.

Purpose.

To examine the crystallization of ketoprofen in pressure sensitive adhesive matrix and the
inhibitory effect of various compounds on the crystallization of ketoprofen.

Methods.

The pressure sensitive adhesive matrix containing ketoprofen was prepared by solvent
casting method. Polyisobutylene (PIB) was used as a pressure sensitive adhesive matrix.
The drug and appropriate additives were dissolved in PIB solution and the solution was
casted on silicone coated release liner using a casting knife. The samples were stored at
25° C. They were examined visually and microscopically at specified time intervals.

Results.

Among various compounds tested, polyvinylpyrrolidone (PVP) was found to be the most
effective crystallization inhibitor. The samples containing more than 1.05 % of PVP
exhibited no crystallization up to 3 weeks after the storage at 25° C. In the samples
containing 1.05 % of Poloxamer 407, tween 20, tween 80, Labrasol, Labrafil 2609, span
80, or Transcutol and in the samples with no additive, ketoprofen crystallized within 1
day after the preparation. In case of span 80, Labrafil 2609, and Transcutol, increasing
the amount of each compound up to 3.15% did not change the time for the ketoprofen to
start crystallization. However, the samples containing 3.15 % Poloxamer, tween 20,
tween 80, or Labrasol showed slower crystal formation compared to those containing less
than 3.15 %. When the effect of storage temperature on the crystal formation of
ketoprofen was studied using the samples containing 1.05 % PVP, no drug crystallization
was observed up to 2 weeks after the storage at 4, 25, 30, 50 and 80° C.

Conclusions.

Ketoprofen was crystallized in PIB pressure sensitive adhesive matrix shortly after the
preparation. The extent of crystallization depended on the amount and the kind of
additives used. Among tested additives, PVP was found to be the most effective
crystallization inhibitor.
http://www.aapsj.org/abstracts/AM_1998/1256.html
We claim:

1. A process for accelerating the absorption of ketoprofen in vivo which comprises


combining 1 part by weight of ketoprofen with 1 to 25 parts by weight of an inorganic
buffer substanceof magnesium hydroxide, magnesium oxide or magnesium carbonate
into a buffered administration form and orally administrating the so combined ketoprofen
and buffer.

2. A process according to claim 1, wherein the active compound is ketoprofen in the form
of its enantiomers S(+)- or R(-)- ketoprofen in pure form or as mixtures in a ratio of 1:99
to 99:1.

3. A process according to claim 1, wherein the buffer substance is at least one of


magnesium oxide and magnesium hydroxide.

4. A process according to claim 1, wherein the buffer substance is magnesium oxide.

5. A Process according to claim 1, wherein said combination is in the form of tablets,


capsules, granules, powder mixtures or suspensions.

6. A process for accelerating the absorption of ketoprofen in vivo which comprises


combining ketoprofen with magnesium hydroxide in a buffered administration form and
orally administrating the so combined ketoprofen and magnesium hydroxide.

7. A process according to claim 6, wherein the active compound is ketoprofen in the form
of its enantiomers S(+)- or R(-)- ketoprofen in pure form or as mixtures in a ratio of 1:99
to 99:1.

8. A process according to claim 6, wherein said combination is in the form of tablets,


capsules, granules, powder mixtures or suspensions. Description: BACKGROUND OF
THE INVENTION
The invention relates to the combined use of ketoprofen and special inorganic basic
substances with an improved quality of action.

It is already known that basic substances such as magnesium hydroxide, magnesium


oxide and sodium bicarbonate or mixtures thereofhave an influence on the absorption of
certain active compounds, such as anthranilic acid derivatives, propionic acidderivatives,
acetic acid derivatives, salicylic acid derivatives or salts thereof or pyrazolols or
benzothiazine derivatives (cf. Neuvonen WO 89/07439). In this Application, ketoprofen
is also mentioned as an example of a propionic acid derivative.

It is also known to the expert that the absorption-regulating effect of certain additives
cannot be generalized for all active compounds (cf. D'Arcy et al., Drug Intelligence and
Clinical Pharmacy, 21, 607 (1987)). The granted claims of theabovementioned PCT
Application by Neuvonen were limited to the specific active compounds tolfenamic acid,
mefenamic acid -and ibuprofen, and only magnesium hydroxide and magnesium oxide
are claimed as basic partners in the combination. This confirms thestatement in the
publication by D'Arcy et al. that the action of basic substances such as antacids on the
absorption properties of active compounds cannot be predicted. The later publication by
Neuvonen in Br. J. Clin. Pharmac. 31, 263 (1991) alsoshows, with the aid of two cross-
over studies, that a higher plasma concentration occurs as a result of addition of
magnesium hydroxide only in the case of ibuprofen. In the case of ketoprofen, neither a
significant increase in the rate of absorptionnor an increased extent of absorption was
found.

SUMMARY OF INVENTION

Knowing of this prior art, it was not to be expected that by combining ketoprofen in
racemic form and in the form of its S(+) and R(-) enantiomers with basic auxiliaries such
as magnesium hydroxide and magnesium carbonate, a faster action and asignificant
increase in the maximum plasma level is achieved as compared with unbuffered tablets.
DETAILED DESCRIPTION OF THE DRAWINGS

FIGS. 1 and 2 show the plasma concentration of R(-)- and S(+)- ketoprofens respectively
following administration of buffered verses unbuffered ketoprofen tablets.

FIG. 3 shows the rate of ketoprofen release over time of buffered verses unbuffered
ketoprofen tablets.

DETAILED DESCRIPTION OF THE INVENTION

As shown in FIG. 1 and FIG. 2, a significant increase in maximum plasma level and a
faster action is achieved using the buffered ketoprofen tablets of the invention versus the
unbuffered tablets.

The absorption-accelerating action of magnesium hydroxide and magnesium carbonate


can be seen from the C.sub.max AUC ratio according to Table 1.

Furthermore, the region of the t.sub.max values starts earlier in the case of the buffered
tablets of the combination according to the invention, and shows less scatter than in the
case of the unbuffered tablets. The comparative studies alsoshow that the interindividual
variation in the individual plasma levels is smaller in the case of the buffered
combination tablets according to the invention than in the case of the unbuffered tablets.

The differences found in the maximum plasma concentrations (C.sub.max) of the three
different tablet formulations are significant. The ketoprofen tablet buffered with
magnesium hydroxide in particular leads to higher plasma concentrations of
theketoprofen being achieved earlier than with the unbuffered tablet. In pain indication in
particular, it is therefore advisable to use the buffered tablets with their faster onset of
action compared with the unbuffered tablets.
The inventive step not only lies in overcoming the prejudice known from the literature
that the absorption of ketoprofen cannot be influenced positively by basic substances, but
is also intensified by our own in vitro test. A study of therelease from ketoprofen tablets
(25 mg) in vitro shows that the formulation with the combination of ketoprofen and
magnesium hydroxide has the slowest release, as can be seen from FIG. 3. From these
negative in vitro results, it was not to be expectedthat the combination of ketoprofen +
magnesium hydroxide in particular shows such advantageous absorption properties in
vivo.

Fixed combinations according to the invention which are of particular interest are those
in the form of tablets, effervescent tablets, capsules, granules, powder mixtures,
suspensions, emulsions and drops, which preferably comprise 1 part byweight of
ketoprofen racemate or S(+)- ketoprofen or R(-)- ketoprofen in pure form or as mixtures
in a weight ratio of 1 to 99 to 99 to 1, and comprise 1 to 25 parts by weight of the basic
buffering additive, in particular magnesium hydroxide.

The buffer capacity of the basic partner of the combination in the presentation forms
according to the invention is preferably at least 3 milliequivalents (meq).

Oral administration forms having a low individual scatter of the plasma concentrations,
an increased rate of absorption and a higher maximum plasma concentration are
preferred.

The fixed combinations according to the invention are prepared by customary methods,
for example by mixing and subsequent pressing or by dissolving the individual
components.

EMBODIMENT EXAMPLES

Example 1
The substances of Example 1 are processed to a tablet which releases ketoprofen at a
moderate rate in vitro.

______________________________________ Non-lacquered tablet Ketoprofen


(racemate) 25.0 mg Magnesium hydroxide 150.4 mg Colloidal silicic acid 12.0 mg
Sodium carboxymethyl-starch 7.0 mg Sodium citrate, tertiary 50.0 mg Magnesium
stearate0.6 mg Coating shell HPM cellulose 1.2 mg Polyethylene glycol 4000 0.4 mg
Titanium dioxide 0.4 mg Total weight 147.0 mg
______________________________________

Peparation

The ketoprofen, magnesium hydroxide, sodium carboxymethyl-starch and sodium citrate


are granulated under aqueous conditions and then dried.

The remainder of the constituents (colloidal silicic acid, magnesium stearate) are admixed
to these granules and this mixture is pressed to tablets of 8 mm diameter on suitable tablet
presses.

Example 1a and 1b

Tablets comprising S(+)- and R(-)- ketoprofen are prepared in an analogous manner.

Example 2

The substances of Example 2 are processed to a tablet which releases ketoprofen rapidly
in vitro.

______________________________________ Non-coated tablet Ketoprofen (racemate)


25.0 mg Magnesium carbonate, basic 258.0 mg Sodium carboxymethyl-starch 10.0 mg
Polyvinylpyrrolidone 25 7.4 mg Colloidal silicic acid 2.0 mg Magnesium stearate0.6 mg
Coating shell HPM cellulose 1.8 mg Polyethylene glycol 4000 0.6 mg Titanium dioxide
0.6 mg Total weight 306.0 mg ______________________________________

Preparation

The ketoprofen, magnesium carbonate, sodium carboxymethyl-starch and PVP are


granulated under aqueous conditions and then dried.

The remainder of the constituents (colloidal silicic acid, magnesium stearate) are added to
these granules and this mixture is pressed to tablets of 9 mm diameter on suitable tablet
presses.

Comparison Example 3(without buffer)

The substances of Example 3 are processed to a tablet which releases ketoprofen rapidly
in vitro. The tablet comprises no buffering additive.

______________________________________ Ketoprofen (racemate) 25.0 mg Maize


starch 48.0 mg Avicel 30.0 mg Lactose 32.0 mg Sodium carboxymethylcellulose (Ac-Di-
Sol) 4.3 mg Magnesium stearate 0.7 mg Coating shell HPM cellulose 0.6 mg
Polyethylene glycol 4000 0.2 mg Titanium dioxide 0.2 mg Total weight 141.0 mg
______________________________________

Preparation

The ketoprofen, maize starch, Avicel, lactose and Ac-Di-Sol are granulated under
aqueous conditions and then dried.

These granules are mixed with magnesium stearate and pressed to tablets of 7 mm
diameter on suitable tablet presses.
TABLE 1
_______________________________________________________________________
___ Unbuffered tablet Buffered tablet (Mg(OH).sub.2) Buffered tablet (MgCO.sub.3)
R(-)- S(+)- R(-)- S(+)- R(-)- Parameter ketoprofen ketoprofen ketoprofen ketoprofen
ketoprofen S(+)-ketoprofen
_______________________________________________________________________
___ C.sub.max /AUC (1 h) 0.60 0.61 0.84 0.87 0.74 0.75
_______________________________________________________________________
___
http://www.patentgenius.com/patent/5776505.html
RESEARCH PAPER Year : 2006 | Volume : 68 | Issue : 1 | Page : 76-82 Design
and evaluation of controlled onset extended release multiparticulate systems for
chronotherapeutic delivery of ketoprofen

HN Shivakumar1, Sarasija Suresh2, BG Desai1


1
Department of Pharmaceutical Technology, K. L. E. S's College of Pharmacy,
Rajajinagar 2nd Block, Bangalore-560 010, India
2
Department of Pharmaceutics, Al-Ameen College of Pharmacy, Hosur Road,
Bangalore-560 027, India

Correspondence Address:
H N Shivakumar
Department of Pharmaceutical Technology, K. L. E. S's College of Pharmacy,
Rajajinagar 2nd Block, Bangalore-560 010
India

Source of Support: None, Conflict of Interest: None

DOI: 10.4103/0250-474X.22969

Abstract

An oral controlled onset extended release dosage form intended to approximate the
chronobiology of rheumatoid arthritis is proposed for site-specific release to the colon.
The multiparticulate system consisting of drug-loaded cellulose acetate cores
encapsulated within Eudragit S-100 microcapsules was designed for chronotherapeutic
delivery of ketoprofen. Drug-loaded cellulose acetate cores were prepared by emulsion
solvent evaporation technique in an oily phase at different drug:polymer ratios (1:1, 2:1
and 4:1). These cores were successfully microencapsulated with Eudragit S-100
following the same technique at the core:coat ratio of 1:5. Scanning electron microscopy
(SEM) revealed that the cellulose acetate cores were discrete, uniform and spherical with
a porous and rough surface, whereas the Eudragit microcapsules were discrete and
spherical with a smooth and dense surface. In vitro drug release studies of the Eudragit
microcapsules were performed in different pH conditions following pH-progression
method for a period of 16 h. The release studies indicated that the microcapsules posses
both pH-sensitive and controlled-release properties, showing limited drug release below
pH 7.0 (6.40 to 8.94%), following which the cellulose acetate cores effectively controlled
the drug release for a period of 11 h in pH 7.5. The differential scanning calorimetric and
powder X-ray diffraction studies demonstrated that ketoprofen was present in dissolved
state in the cellulose acetate polymeric matrix, which could explain the controlled drug
release from the cores. The release of ketoprofen from Eudragit microcapsules in pH 7.5
depended on the cellulose acetate levels and was characterized by Higuchi's diffusion
model.

How to cite this article:


Shivakumar HN, Suresh S, Desai BG. Design and evaluation of controlled onset extended
release multiparticulate systems for chronotherapeutic delivery of ketoprofen. Indian J
Pharm Sci 2006;68:76-82

How to cite this URL:


Shivakumar HN, Suresh S, Desai BG. Design and evaluation of controlled onset extended
release multiparticulate systems for chronotherapeutic delivery of ketoprofen. Indian J
Pharm Sci [serial online] 2006 [cited 2010 Jan 27];68:76-82. Available
from: http://www.ijpsonline.com/text.asp?2006/68/1/76/22969

Chronotherapeutics refers to a clinical practice of synchronizing drug delivery in a


manner consistent with the body's circadian rhythm, including disease states, to produce
maximum health benefit and minimum harm[1]. The site-specific delivery of drugs to the
colon has implications in a number of therapeutic areas, which include topical treatment
of colonic disorders such as Crohn's disease, ulcerative colitis, constipation, colorectal
cancer, spastic colon and irritable bowel syndrome. A colonic delivery system would
additionally be valuable when a delay in absorption is therapeutically desirable in
treatment of diseases like rheumatoid arthritis, which are influenced by circadian
rhythms[2]. The disease is known to have the peak symptoms when awaking from
nighttime sleep.

Ketoprofen is a potent non-steroidal anti-inflammatory drug with a short biological which


is prescribed for long-term treatment of musculoskeletal and joint disorders such as
rheumatoid arthritis, osteoarthritis, alkylosing spondilytis and acute gout[3]. The
mechanism of action of ketoprofen is mainly associated to the inhibition of the body's
ability to synthesize prostaglandin. Adverse effects on the gastric mucosa have been
observed when the drug is administered orally. Ketoprofen is rapidly absorbed from the
gastrointestinal tract and reaches high bioavailability (>92%). The drug has been reported
to be transported across the intestinal epithelial cells by trans-cellular passive
diffusion[4].

Colon-specific delivery can be achieved with a suitable mechanism that triggers off the
drug release upon reaching the colon. The physiological change in the pH of the
gastrointestinal tract has been extensively exploited to convey the actives to the colon.
Methods based on pH-sensitive delivery, such as delayed-onset dosage forms, could be a
simple and practical means for colon targeting. Several polymers, particularly Eudragit S-
100[5] and EudragitTMS[6], have been investigated for colonic delivery. These polymers
have been designed to be soluble at pH values higher than 7, keeping in mind the pH
prevalent in the large intestine. As reported, the pH of the colon in normal subjects drops
from 7.5±0.4 in the terminal ileum to 6.4±0.6 in the ascending colon[7]. However, the
major disadvantage of these systems is the possibility of the drug being released in the
terminal ileum rather than the colon. This problem was thought to be solved by utilizing
two polymers: one having pH-sensitive, and the other imparting a controlled-release
property. Eudragit S was used to prevent the drug release till the formulation reaches the
terminal ileum, whereas cellulose acetate avoids the complete release in the ileum and
effectively conveys the drug to the colon.

A multiparticulate system presents several advantages in comparison to single unit forms


in that they exhibit higher colonic residence time, more predictable gastric emptying, and
cause less local irritation[8]. With all these considerations in mind, a multiparticulate
system consisting of drug-loaded cellulose acetate cores encapsulated within Eudragit S-
100 microcapsules was designed for chronotherapeutic delivery of ketoprofen. As a core
forming polymer, cellulose acetate, whose application in the microencapsulation has been
extensively investigated, was selected[9]. With this system, the aim was to minimize drug
release in the upper part of the gastrointestinal tract and target the drug to the colon.

Materials and methods

Ketoprofen was kindly donated by Rhone Poulenc (I) Ltd., Mumbai. Eudragit S-100 was
generously donated by Rohm Pharma, Darmstadt, Germany. Cellulose acetate was
supplied by Rolex Chemicals, Mumbai. The rest of the chemicals of analytical grade,
supplied by S. D. Fine Chemicals, Mumbai, included light liquid paraffin, Span-80,
acetone, n-hexane, and methanol.

Preparation of cellulose acetate cores containing ketoprofen:

Cellulose acetate cores containing ketoprofen were prepared by emulsion-solvent


evaporation technique in an oily phase[10]. Cellulose acetate was dissolved in acetone to
get a homogenous polymer solution (1% w/v). Ketoprofen was dissolved in 10 ml of the
polymer solution at drug:polymer ratio of 1:1, 2:1 and 4:1, and the resulting solution was
added in thin streams to 70 ml of liquid paraffin containing 1% w/w of span-80. About 10
ml of acetone was added to the external phase to produce a stable o/o emulsion. The
system was maintained under constant stirring (1000 rpm) using a variable speed
propeller stirrer (RQ 125 D, Remi Udyog Ltd., Mumbai) for a period of 3 h to allow
complete solvent evaporation. The cellulose acetate cores formed were separated, washed
with n-hexane, and dried for 48 h in a vacuum desiccator.

Microencapsulation of drug-loaded cellulose acetate core:

Cellulose acetate cores containing ketoprofen were encapsulated following the same
technique with Eudragit S-100[10]. The drug-loaded cellulose acetate cores (100 mg)
were suspended in 5 ml of ethanolic solution of Eudragit S-100 (10% w/v) and emulsified
into 70 ml of liquid paraffin containing 1% w/w of span-80. Emulsification was
maintained using a variable-speed propeller stirrer at 1000 rpm to allow complete solvent
evaporation. The microcapsules formed were separated, washed with n-hexane, and dried
for 48 h in a vacuum desiccator.

IR spectra of ketoprofen, cellulose acetate, Eudragit S and the microcapsules were


recorded in a FTIR spectrophotometer (Jasco FTIR, 460 plus) to check the chemical
integrity of the drug in the microcapsules[11].

Scanning electron microscopy (SEM):

Morphology and surface topography of the microparticles were examined by scanning


electron microscopy[12] (SEM-Jeol, JSM-840A, Japan). The samples were mounted on
the SEM sample stab, using a double-sided sticking tape and coated with gold (200A°)
under reduced pressure (0.001 torr) for 5 min to improve the conductivity using an Ion
sputtering device (Jeol, JFC-1100 E, Japan). The coated samples were observed under the
SEM and photomicrographs of suitable magnifications obtained.

Particle-size distribution:

The particle-size distribution of the microparticles was determined using optical


microscopy[12]. The projected diameter of a total of 200 microparticles from each batch
was observed. The size distribution data got were attempted to fit into normal and log
normal distribution, and the equivalent diameter based on surface number basis (dsn) was
computed using Hatch-choate equation[13].

Estimation of drug content:

An accurately weighed quantity of the microparticles was dissolved in acetone. The


acetone was evaporated and the residue left behind was vortexed with 75% methanol for
30 min to extract the drug. The dispersion was filtered and the absorbance of the filtrate
was measured at 258 nm after appropriate dilution in a UV-visible spectrophotometer
(Jasco V-530, Japan). The drug content was estimated in triplicate using a calibration
curve constructed in the same solvent. Polymers did not interfere with the assay at this
wavelength.

Thermal analysis[11]:

Samples of the ketoprofen, cellulose acetate, physical mixtures and the cellulose acetate
cores were taken in a flat-bottomed aluminium pans and heated over a temperature range
of 40-180° at a constant rate of 5°/min with purging of nitrogen (50ml/min) using
alumina as a reference standard in a differential scanning calorimeter (Perkin Elmer DSC,
Pyris-1).

X-ray powder diffraction studies[11]:

The diffraction studies were carried out in a powder X-Ray diffractometer (Philips, PW
1050/37) with a vertical goniometer using Cu K a radiation with Ni filter at a voltage of
40 kV and a current of 20 mA. Powder XRD patterns for ketoprofen, cellulose acetate,
physical mixture and cellulose acetate cores were obtained by scanning from 0 to 50° 2q.

In vitro drug release studies:

Dissolution studies of the Eudragit microcapsules were carried out in triplicate employing
USP XIII dissolution rate test apparatus-1 (Electrolab, TDT-06T) following pH
progression method simulating the gastrointestinal tract conditions[10]. Weighed
quantities of the microcapsules were loaded into the basket of the dissolution apparatus,
and the pH changes were performed, starting with 900 ml of 0.1 N hydrochloric acid for
2 h, mixed phosphate buffer of pH 5.5 for 1 h, phosphate buffer of pH 6.8 for 2 h,
followed by mixed phosphate buffer of pH 7.5 till the end of the test. The temperature of
the dissolution fluid was maintained at 37±0.5° with a stirring speed of 100 rpm. The
samples were withdrawn every hour, filtered through a Millipore filter (0.22 mm), and
assayed spectrophotometrically at 258 nm for the samples of pH 1.2, and at 260 nm for
the rest of the samples. However, the dissolution studies of the cellulose acetate cores
were also performed under the same set of experimental conditions using mixed
phosphate buffer of pH 7.5 as the dissolution fluid mimicking the pH at the end of the
small intestine[7].

Results and discussion

The compositions of different batches of cellulose acetate cores are shown in [Table - 1].
Cellulose acetate cores containing ketoprofen were prepared by emulsion solvent
evaporation technique, where the organic solution containing the drug and cellulose
acetate in acetone was emulsified into an external oil phase of liquid paraffin. A small
amount of acetone added to the external oily phase was known to avoid rapid diffusion of
the organic solvent into the oily phase. This would prevent immediate polymer
precipitation before the organic solution could be dispersed into droplets in the oily phase
leading to formation of a stable emulsion. Span 80 (1% w/w) was used as an emulsifier to
stabilize the o/o emulsion produced.

Photomicrographs of the cellulose acetate cores are shown in [Figure - 1]a. It is vivid
from SEM photomicrographs that the cores were discrete, uniform and spherical with a
porous and rough surface. The rough surface of the cores can be attributed to the rapid
solvent diffusion and quick precipitation of cellulose acetate during the formation of o/o
emulsion. It has been reported that microspheres with a porous and rough surface were
produced by solvent evaporation technique with crystalline polymers[14].

The viscosity of the polymer solution used during microencapsulation is known to


determine the size of the microspheres produced[10]. The viscosity of the organic
solution depended on the polymer concentration in the solution, organic solvent used and
the temperature. Since all the three batches of cellulose acetate cores were prepared using
the organic solution having the same polymer concentration (1% w/v of cellulose
acetate), there was no significant difference in the emulsion globule size. Accordingly,
the cellulose acetate cores of the three batches did not vary significantly in their size. The
size distribution data obtained from optical microscopy, when represented as log-
probability plots, gave straight lines indicating a log-normal distribution in all the three
batches of cellulose acetate cores produced. The surface number diameters (dsn) of the
three batches of drug-loaded cores, as computed using Hatch-choate equation, are
represented in [Table - 1].

The percent drug loading and entrapment efficiency of the three batches of drug loaded
cores are depicted in [Table - 1]. The values of entrapment efficiency were found to
decrease with increase in the initial drug loading, which can be ascribed to better drug
entrapment within the cores with increase in cellulose acetate levels. The drug loss during
the microencapsulation process can possibly be related to the partitioning of the drug to
the oil phase.

The in vitro release profiles of cellulose acetate cores are portrayed in [Figure - 2]. As the
cellulose acetate cores remain protected by the Eudragit S-100 coat below pH 7, the
dissolution tests of the cellulose acetate cores were conducted using phosphate buffer of
pH 7.5 also mimicking the pH prevalent at the end of the small intestine[7]. The
dissolution studies indicated that the cores were characterized by an initial burst effect
during the first hour. The burst effect was reduced with increasing cellulose acetate levels
in the cores, which may be attributed to better drug entrapment within the cores with
increase in cellulose acetate levels. The slow release phase was followed by a controlled
release phase, during which the drug release depended on the cellulose acetate levels in
the cores. The mechanism of drug release was found to be characterized by Higuchi's
diffusion model[15] as plots of amount of drug released versus square root of time were
found to be linear. The values of Higuchi rate constant (KH) were found to range
between 16.1 to 23.3 %h -1/2 with a distinct increasing trend as the cellulose acetate levels
in the cores decreased. The effect of varying cellulose acetate levels in matrix diffusional
systems on drug release has been documented[9].

DSC has been one of the most widely used calorimetric techniques to characterize the
solubility and physical state of drug in the polymeric matrix. [Figure - 3] depicts the DSC
thermograms of ketoprofen, cellulose acetate, physical mixtures and cellulose acetate
cores. The DSC thermogram of ketoprofen exhibited a single sharp endothermic peak at
94.34° corresponding to its melting transition temperature[16]. This peak was also
observed in the thermogram of the physical mixture, even though slightly broadened but
shifted to lower temperature (94.06°). This may be possibly due to fact that presence of
cellulose acetate in the physical mixture depresses the melting point of ketoprofen and
broadens its melting point endotherm. The thermograms of the drug-loaded cellulose
acetate cores showed no such characteristic peak, indicating that the drug was present in
the dissolved state in cellulose acetate polymer matrix.

Powder XRD technique has been extensively utilized along with DSC to study the
physical state of drug in the polymer matrix. Powder XRD patterns for ketoprofen,
cellulose acetate, physical mixture and cellulose acetate cores are shown in [Figure - 4].
The crystalline nature of ketoprofen was clearly demonstrated by its characteristic PXRD
pattern containing well-defined peaks. The PXRD diffractogram of the physical mixture
of the drug and cellulose acetate also exhibited the characteristic diffraction pattern of the
crystalline drug, indicating that the drug was dispersed in cellulose acetate in the physical
mixture, The PXRD spectra of the cores did not reveal any such characteristic PXRD
pattern corresponding to the crystalline drug, confirming the fact that the drug existed in
the dissolved state in the cellulose acetate polymer matrix. These results could explain the
controlled drug release from the cellulose acetate cores.

The second part of the research work was focused on microencapsulation of the cellulose
acetate cores with a pH-sensitive acrylic polymer. The drug-loaded cores were
microencapsulated following the emulsion solvent evaporation technique with Eudragit
S-100 that dissolves at pH of above 7. Eudragit S was selected to protect the cellulose
acetate cores in the upper part of the gastrointestinal tract, avoiding any significant drug
release before reaching the colon. Once the acrylic coat dissolves, it was expected that the
cellulose acetate cores would effectively control the drug release at the target site.

As a part of the research work, a preliminary screening study was undertaken to select a
suitable organic solvent that would dissolve Eudragit S-100 and, at the same time,
maintain the integrity of the cellulose acetate cores. Ethanol was chosen as solvent as it
met the above said criteria; moreover, ethanol diffuses quickly into the external oily
phase, resulting in encapsulation of the drug-loaded cellulose acetate cores. [Table - 2]
depicts the compositions of different batches of Eudragit microcapsules.

SEM revealed that the Eudragit microcapsules were discrete, uniform and spherical, with
a smooth and dense surface. The photomicrographs of the Eudragit microcapsules are
portrayed in [Figure - 1]b. It has been already established that microspheres with a
smooth and dense surface were produced by solvent evaporation technique with
amorphous polymers[14].

As mentioned earlier, the particle size of the microcapsules produced depended on the
viscosity of the polymer solution used. As all the three batches of microcapsules were
produced using the polymer solution having the same Eudragit concentration (10% w/v),
the microcapsules of the three batches did not differ significantly in their particle size.
The particle size distribution data, as determined by optical microscopy when represented
as log-probability plots, gave straight lines indicating a log-normal distribution in all the
three batches of microcapsules produced. The surface number diameters (dsn) of the
three batches of microcapsules, along with their percentage drug loading and
encapsulation efficiency values, are represented in [Table - 2]. The values of percentage
drug loading and encapsulation efficiency portray that emulsion solvent evaporation
technique allows favourable drug encapsulation using Eudragit S 100.

IR Spectrophotometry has been employed as a useful tool to identify the drug excipient
interaction. The IR spectra of ketoprofen and the microcapsules were identical. The
principal IR absorption peaks of ketoprofen at 1698 cm-1 (carboxylic acid carbonyl) and
1655 cm-1 (ketonic carbonyl) appeared in the spectra of ketoprofen as well as the
microcapsules. These observations indicated no chemical interaction between the drug
and other excipients used.

[Figure - 5] portrays the in vitro drug release profiles of Eudragit microcapsules as


determined by pH progression method. The studies showed that the microcapsules
exhibited both pH-sensitive and controlled-release properties. The drug release depended
on the pH of the dissolution media and the cellulose acetate levels in the microcapsules.
A limited drug release was observed from the microcapsules below pH 7 during the first
five hours of dissolution (6.39 to 8.94 %), which can be ascribed to the pH-sensitive
nature of Eudragit S-100 coating. Eudragit S-100 is a pH-sensitive acrylic polymer
having a threshold pH of 7 5. It was observed that once the acrylic coating dissolved at pH
7.5, the cellulose acetate cores effectively controlled the drug release for a period of 11 h.
As revealed by the DSC and PXRD studies, the physical state of ketoprofen in the
cellulose acetate polymeric matrix could explain the controlled release of the drug from
the cellulose acetate cores. Diffusion of the drug through the cellulose acetate polymeric
matrix was the rate-controlling step that could characterize the mechanism of drug
release. The mechanism of drug release from the microcapsules in pH 7.5 was found to
be diffusion controlled and was characterized by Higuchi's diffusion model. The values
of kinetic constant ranged between 15.79 to 22.31% h -1/2, showing a distinct increasing
trend as the cellulose acetate levels in the Eudragit microcapsules decreased.

A multiparticulate system having both pH-sensitive and controlled-release property is


described for chronotherapeutic delivery of ketoprofen. The results collectively prove
that dual coated microcapsules with enteric and controlled release properties can be
successfully developed using double microencapsulation procedure. Bedtime
administration of such a device could improve the anti-inflammatory therapy in the
management of rheumatoid arthritis.

http://www.ijpsonline.com/article.asp?issn=0250-
474X;year=2006;volume=68;issue=1;spage=76;epage=82;aulast=Shivakumar

Indian journal of pharmaceutical sciences


Nanoparticles Containing Ketoprofen and Acrylic Polymers Prepared by an Aerosol Flow
Reactor Method
Hannele Eerikäinen,1,2 Leena Peltonen,3 Janne Raula,1 Jouni Hirvonen,3 and Esko I.
Kauppinen1,4
1
Center for New Materials, Helsinki University of Technology, PO Box 1602, FIN-02044
VTT, Finland
2
Present address: Orion Corporation Orion Pharma, Pharmaceutical Product
Development, PO Box 65, FIN-02101 Espoo, Finland
3
Faculty of Pharmacy and Viikki Drug Discovery Center, University of Helsinki, PO Box
56, FIN-00014 Helsinki, Finland
4
Aerosol Technology Group, VTT Processes, PO Box 1602, FIN-02044 VTT, Finland

Correspondence to:
Esko I. Kauppinen
Tel: +358 9 456 6164
Fax: +358 9 456 7021
Email: Esko.Kauppinen@vtt.fi

Submitted: April 27, 2004; Accepted: September 23, 2004; Published: December 31,
2004

Keywords: nanoparticles, ketoprofen, aerosol, polymer, Eudragit

Abstract

The purpose of this study was to outline the effects of interactions between a model drug
and various acrylic polymers on the physical properties of nanoparticles prepared by an
aerosol flow reactor method. The amount of model drug, ketoprofen, in the nanoparticles
was varied, and the nanoparticles were analyzed for particle size distribution, particle
morphology, thermal properties, IR spectroscopy, and drug release. The nanoparticles
produced were spherical, amorphous, and had a matrix-type structure. Ketoprofen
crystallization was observed when the amount of drug in Eudragit L nanoparticles was
more than 33% (wt/wt). For Eudragit E and Eudragit RS nanoparticles, the drug acted as
an effective plasticizer resulting in lowering of the glass transition of the polymer. Two
factors affected the preparation of nanoparticles by the aerosol flow reactor method,
namely, the solubility of the drug in the polymer matrix and the thermal properties of the
resulting drug-polymer matrix.

Introduction

Drug nanoparticles can be defined as drug-containing particles having size smaller than 1
µm.1,2 These submicron-sized particles consist of the drug and, optionally, a stabilizing or
functional biocompatible polymer. Several applications of nanoparticles have been
proposed, such as tissue targeting in cancer therapy,3 controlled release,4 carrier action for
the delivery of peptides,5,6 and increase in the solubility of drug.7

Previously, a method capable of producing drug-polymer nanoparticles, namely, an


aerosol flow reactor method, has been presented.8,9 This method produces spherical,
amorphous, matrix-type drug-polymer nanospheres directly as dry powder in a 1-step
operation. In the previous study,9 the properties of nanoparticles consisting of an acrylic
polymer, Eudragit L, and drug materials ketoprofen or naproxen were studied. It was
observed that crystallization of the drug in the polymer matrix was the limiting factor for
drug loading. In this study, the polymeric component is varied, while ketoprofen is used
as a model drug.

The polymer nanoparticles prepared by the aerosol flow reactor method have an
amorphous solid solution structure.9 When the polymer glass transition temperature is
above the ambient temperature, the polymeric component is in a glassy state, which
provides mechanical strength to the particles. Therefore, the mechanical hardness and
integrity of the particles can be maintained and coalescence of the particles can be
avoided, which allows the collection as dry powder.

The aim of this study was to evaluate how different polymers and interactions between
the drug and the various polymers affect the physical state of the nanoparticles. Three
acrylic polymers were used in this study. These functional polymers, namely, Eudragit L,
Eudragit RS, and Eudragit E, are widely used in the pharmaceutical industry, and are
accepted for oral use.10 These 3 polymers have different chemical compositions and
functional groups. For the purposes of this study, first, it was expected that the solubility
properties of the nanoparticles could be varied due to different solubilities of the
polymers.10,11 Second, as the functional groups of the polymers are different, interactions
between the polymers and the acidic model drug molecule, ketoprofen, were expected to
be different.

The structural formulas of the polymer materials used are shown in Figure 1. Eudragit L
is a copolymer consisting of methyl methacrylate and methyl methacrylic acid repeating
units in a ratio of 1:1.10 It is soluble when the pH is greater than 6 due to the ionization of
the acid groups; below this pH it is insoluble.11 Eudragit E is a copolymer consisting of a
1:2:1 ratio of methyl methacrylate, dimethylaminoethyl methacrylate, and butyl
methacrylate monomers.10 The tertiary amino groups are ionized at acidic conditions, and
this polymer is soluble when the pH less than 5.11 Eudragit RS is a copolymer consisting
of ethyl acrylate, methyl methacrylate, and trimethylammonioethyl methacrylate chloride
in a ratio of 1:2:0.1.10 This polymer has pH-independent permeability.11
Materials and Methods

Preparation of Particles

Materials

Ketoprofen (2-(3-Benzoylphenyl) propionic acid) was purchased from Sigma (St Louis,
MO) and was used as received. Eudragit L 100 PO, Eudragit RS 100 PO, and Eudragit E
100 were obtained from Röhm (Röhm Pharma, Darmstadt, Germany), and were used as
received.

Preparation of Drug Solution

The drug-polymer solutions were prepared by separately dissolving the polymer and drug
into ethanol (99.6%, Alko Oyj, Rajamäki, Finland) using a magnetic stirrer and
combining the solutions at respective amounts. Total solids concentration of the starting
solution was fixed at 2 g/L. The compositions of prepared particles are shown in Table 1.

Experimental System Set-up

The experimental system set-up for the preparation of nanoparticles has been described in
detail previously.8,9 Briefly, the ethanolic solution containing the drug and the polymer
was atomized using a collision-type air jet atomizer TSI 3076 as the aerosol generator
(TSI Inc Particle Instruments, St Paul, MN). The resulting droplets were suspended into
nitrogen, and the aerosol generated was passed through a heated tubular laminar flow
reactor, which was used to evaporate the solvent from the droplets and to allow particle
formation to complete. The reactor wall temperature used in this study was kept constant
at 80°C and the flow rate of carrier gas was 1.5 L/min. The nanoparticle aerosol was
diluted in a porous tube aerosol diluter with nitrogen (20°C) in a ratio of 1:17 before
collecting the nanoparticles with a Berner-type low-pressure impactor onto aluminum
foil.

Powder Collection

Dry powder samples of particles were collected after diluting the aerosol (ratio 1:17,
dilution gas nitrogen at 20°C) using a Berner-type low-pressure impactor onto aluminum
foil. The impactor was kept at room temperature. The impactor classified the aerosol into
11 stages, and for this study, the dry powder samples were formed by combining the
material deposited on stages 1 to 9. The dry powder samples were stored in a refrigerator
(+2 to +8°C) prior to analyses. Scanning electron microscope (SEM) and transmission
electron microscope (TEM) observations, differential scanning calorimetry (DSC)
analyses, infrared spectroscopy (IR) analyses, and drug release analyses were performed
for these dry powder samples.
Characterization of Particles

Particle Morphology

Particle morphology was analyzed using a field-emission SEM (Leo DSM982 Gemini,
LEO Electron Microscopy Inc, Oberkochen, Germany) using an acceleration voltage of 2
kV. The samples from dry powder particles were prepared by gently dipping a copper
grid (for SEM) or lacey carbon-coated copper grid (for TEM) (Agar Scientific Ltd,
Essex, UK) into the dry nanoparticles and carefully blowing off excess material. The
samples for SEM observations were coated with a thin platinum coating. Particle
morphology and internal structure were further analyzed using a field-emission TEM
(Philips CM200 FEG, FEI Co, Eindhoven, the Netherlands) using an acceleration voltage
of 200 kV.

Particle Size and Size Distribution

Particle size distribution analysis was performed directly from the nanoparticle aerosol
using a TSI scanning mobility particle sizer (SMPS), equipped with a long differential
mobility analyzer (DMA, model 3081; TSI Inc Particle Instruments) and a condensation
particle counter (CPC, model 3022; TSI Inc Particle Instruments). For particle size
measurements, an additional aerosol diluter (1:10, dilution gas nitrogen at 20°C) was
added before the measurements to reduce the particle concentration to a suitable level.
The particle number size distribution measurements were performed 6 times at each
experimental condition to reduce random error, and an average of the 6 measurements
was calculated and used for analysis.

Differential Scanning Calorimetry

The thermal behavior of the particles was analyzed using a DSC instrument (Mettler
Toledo DSC 822e, Mettler Toledo AG, Greifensee, Switzerland) equipped with a Stare
computer program. Approximately 3 mg of sample was accurately weighed into a 40-µL
aluminum pan and sealed with a punched lid. Temperature range of -50°C to 200°C was
scanned using a heating rate of 10°C/min. A nitrogen purge of 50 mL/min was used in
the oven. The samples were heated above their Tg and studied using a microscope (Zeiss
Axioskop, Oberkochen, Germany) equipped with a heating stage (Linkam THMS 600,
Surrey, UK). When the samples were heated above glass transition temperature, Tg, it
was observed that the nanoparticles formed a coalesced drug-polymer matrix and did not
consist of single, separate nanoparticles anymore. Therefore, to characterize the thermal
behavior of the nanoparticles, Tg values were determined in the first heating cycles in
DSC experiments.

Infrared Spectroscopy

Infrared absorption spectra of raw materials and nanoparticles in the wavelength region
4000 cm-1 to 650 cm-1 were recorded using a Fourier transform IR spectrometer
(Spectrum One, PerkinElmer Instruments LLC, Shelton, Connecticut) equipped with a
Universal ATR sampling accessory (PerkinElmer Instruments LLC, Shelton,
Connecticut). Resolution used in the scans was 1 cm-1, and the spectra were averaged
over 3 scans.

Drug Release from Nanoparticles

Drug release tests were performed using a system based on the general drug release
standard for delayed-release (enteric-coated) articles, method A.12 An amount of
nanoparticles corresponding to ~2 mg of ketoprofen was weighed and filled into a size 0
gelatin capsule. The capsule was further girdled with a metal wire to ensure that the
capsule settled down in the vessel.13 Round-bottomed cylindrical glass vessels having a
total volume of ~150 mL were used as release chambers. The solutions were stirred using
a magnetic stirrer at a speed of 50 rpm. The temperature was controlled to 37.0°C ±
0.5°C. In the acid stage, 75 mL of 0.1 N hydrochloric acid was used as the release
medium. Aliquots were withdrawn at predetermined time intervals and immediately
replaced with fresh medium equilibrated at 37°C. After 2 hours, 25 mL of 0.2 M tribasic
sodium phosphate was added to change the pH of the test medium to 6.8, and the test was
continued for a further 4 hours. The amount of the drug released was determined using a
spectrophotometer (Pharmacia LKB Ultrospec III, Pharmacia LKB Biochrom Ltd,
Cambridge, UK) using wavelength of 260 nm. The tests were performed with 2 parallel
runs; the values reported are mean values of the 2 runs. The repeatability of the method
was evaluated by analyzing 6 parallel samples, and it was found that the results are
repeatable. The measured dissolution values have a standard deviation of 6% on average,
while the maximum standard deviation was less than 10%. The highest standard
deviation values were observed immediately after the pH change, probably due to
incomplete mixing and equilibration of the pH in the dissolution vessel.

Results and Discussion

Effect of the Polymer and the Amount of Drug on Particle Size

The atomizer used to spray the nanosized droplets produced a unimodal and lognormal
droplet size distribution. After drying the droplets, the particle size distribution of the
solid nanoparticles reflected the droplet size distribution produced by the atomizer. The
geometric standard deviation of the distributions was less than 2.0 for all the studied
drug-polymer particles, which was in good accordance with the atomizer specifications.14
In Figure 2, particle size distributions of nanoparticles containing 10% (wt/wt) ketoprofen
are shown. The number mean geometric particle size was calculated from the size
distribution curve. In Figure 3, the number mean particle sizes are plotted as a function of
drug amount. As a general trend, the particle size slightly decreased as the amount of
drug was increased in the nanoparticles. It was observed that Eudragit L produced larger
particles than either Eudragit E or Eudragit RS, and that Eudragit RS produced the
smallest particles. This was most likely caused by different viscosities and surface
tensions of the solutions, which affected the atomization and the droplet size.15
Collection of the Nanoparticles

The different polymers used led to different stability of the nanoparticles during
collection. Powders could be collected when the amount of drug was equal to or less than
50% (wt/wt) for Eudragit L nanoparticles, whereas for Eudragit E and Eudragit RS
nanoparticles, nanoparticles containing 33% (wt/wt) or less drug could be collected.
When higher amounts of drug were incorporated to these polymers, the product collected
was tacky and transparent, and seemed not to consist of individual particles.

The collected powders were analyzed by electron microscopy to observe the morphology
of the nanoparticles. The nanoparticles made of polymer Eudragit L and containing 33%
(wt/wt) or less of drug were spherical, had smooth surfaces, and showed no crystallites
(see Figure 4A). When the amount of drug was 50% (wt/wt), some crystallites were also
observed. For Eudragit E, the nanoparticles containing 10% (wt/wt) or less drug were
spherical, separate nanoparticles. When the drug amount was increased to 25% (wt/wt),
the nanoparticles showed coalescence, and separate nanoparticles could not be detected
(see Figure 4C). Similarly to Eudragit E, the Eudragit RS nanoparticles were separate,
distinct nanoparticles when the amount of drug was 10% (wt/wt) or less (see Figure 4B).
Also for these nanoparticles, coalescence and loss of integrity was found when the
amount of drug was 25% (wt/wt), as shown in Figure 4D. The compositions and observed
appearances of the nanoparticles prepared are summarized in Table 1.

TEM observations were performed for the successfully prepared nanoparticles.


Transmission electron microscopy (see Figure 5) showed solid, homogeneous drug-
polymer particles. Grain boundaries or crystals were not detected and, therefore, it was
concluded that these nanoparticles had a matrix-type structure.

Thermal Behavior of the Nanoparticles

To explain the reason for the coalescence of the nanoparticles, the thermal behavior of
the particles was analyzed with DSC. Specifically, the glass transition temperature of the
composite nanoparticles was determined. The Tg values are listed in Table 2. For the
nanoparticles prepared from Eudragit L, the glass transition was slightly lowered as a
function of drug amount (see Table 2). However, the glass transition of all Eudragit L
nanoparticles was clearly above room temperature. When the drug amount was equal to
or less than 33% (wt/wt), the DSC curves showed no signal attributable to melting peak
of the drug (see Figure 6). Therefore, it could be concluded that the drug was
incorporated in these nanoparticles in an amorphous form.16-20 For the nanoparticles
containing 50% (wt/wt) drug, however, an endothermic transition attributable to the
melting of drug crystals was observed at 94°C. Pure ketoprofen showed a distinct crystal
melting peak at 96°C (see Figure 6). For the nanoparticles containing 50% (wt/wt) drug,
the solubility limit of drug in the polymer matrix was exceeded, and the excess drug
formed crystals. The ketoprofen melting peak appeared at a lower value in the
nanoparticles, most likely due to small, imperfect drug crystals formed in the polymer
matrix.21 Also, interaction with the polymer could lead to lowering of the melting
point.18,21,22

On the contrary, for the nanoparticles prepared from Eudragit RS or Eudragit E,


crystallization of drug could not be detected within the composition range studied (see
Figure 6). As no crystals were observed, the amount of ketoprofen was below the
solubility limit of ketoprofen in these polymer matrices. Instead, a significant lowering in
the glass transition temperature of the polymer was observed, as shown in Table 2. The
glass transition temperatures of Eudragit E and Eudragit RS composite nanoparticles
were much lower than of Eudragit L nanoparticles. For Eudragit E and Eudragit RS
nanoparticles, when the drug amount was 25% (wt/wt), the glass transition temperatures
were close to room temperature. The glass transition temperatures measured were 24°C
and 28°C for the Eudragit E and Eudragit RS nanoparticles, respectively. The glass
transition temperatures were close to the collection temperature (room temperature) of
the nanoparticles. Consequently, the nanoparticles were softened and the mechanical
strength was not sustained. Ketoprofen drug molecule acted as an effective plasticizer for
these polymers, lowering the glass transition temperature.23,24

IR Spectroscopy

Infrared spectroscopy was used to study the interactions between the drug and the
polymers. Ketoprofen has a carboxylic acid group, which can interact with the functional
groups of the polymers. The carbonyl peaks in the IR spectra of ketoprofen were
recorded at 1694 cm-1 and 1654 cm-1, and have previously been assigned to dimeric
carboxylic acid carbonyl group and ketonic carbonyl group stretching vibrations,
respectively.25,26

Exemplary IR spectra of the nanoparticles are shown in Figure 7. For clarity and stronger
absorptions arising from ketoprofen, materials containing 33% (wt/wt) ketoprofen are
shown as examples, even though nanoparticle collection was not successful for all these.
Considering first the nanoparticles containing no drug but only polymer (Figure 7,
Eudragit L curve C), the Eudragit L polymer contains both carboxylic acid and ester
groups. Therefore, the IR spectra showed overlapping carbonyl vibrations of the ester
group at 1724 cm-1 and the carboxylic acid at 1710 cm-1.27 However, when ketoprofen
was included in the nanoparticles, this peak was split into 2 peaks as a function of
increasing drug amount (see Figure 7, Eudragit L curve B). The peak at ~1724 cm-1 could
be attributed to stretching vibrations of the ester group carbonyl, similarly to Eudragit E
and Eudragit RS polymers. However, the peak at the lower wavenumber was recorded at
1710 cm-1, 1705 cm-1, 1703 cm-1, 1700 cm-1, 1700 cm-1, and 1699 cm-1 for the
nanoparticles containing 0% (wt/wt), 5% (wt/wt), 10% (wt/wt), 25% (wt/wt), 33%
(wt/wt), and 50% (wt/wt) ketoprofen, respectively. This peak was interpreted as arising
due to the formation of a dimer by the carboxylic acid groups of the polymer and the
drug. Due to the formation of the dimer, the vibration of the carboxylic acid carbonyl was
shifted to lower wavenumbers.

From Figure 7 it can be seen that the Eudragit RS and Eudragit E materials exhibit quite
similar spectra (Figure 7, Eudragit RS curve C and Eudragit E curve C). The strong
stretching vibration of the carbonyl moiety of ester groups could be identified for both the
materials at ~1724 cm-1.28 For the Eudragit E and Eudragit RS nanoparticles containing
ketoprofen, the position of the ester group carbonyl peak at 1724 cm-1 was not changed
(see Figure 7, Eudragit E curve B and Eudragit RS curve B). However, the peak
corresponding to the carboxylic acid group of ketoprofen at 1694 cm-1 was not seen in the
spectra of the composite nanoparticles, whereas the peak arising from the ketone
carbonyl at 1654 cm-1 could be identified. The carboxylic acid group of the ketoprofen
molecule interacted with the polymers, leading to the disruption of the carboxylic acid
dimer of the crystalline ketoprofen. As a result, the carboxylic acid stretching vibration
occurred at higher wavenumbers,25,26 was overlapped by the strong ester vibrations of the
polymer, and could not be detected.

Eudragit E is a polymer containing secondary amino groups capable of accepting a


proton from an acid molecule. It was initially assumed that at least a fraction of the amino
groups of the polymer would be protonated by the acidic drug in the ethanolic solution.
The peaks corresponding to the amino groups have been identified previously29 at 2820
cm-1 and 2770 cm-1. However, any change in the position of these peaks was not observed
when ketoprofen was incorporated in the nanoparticles. Therefore, it was concluded that
ketoprofen drug mainly interacted with the ester groups of the Eudragit E polymer,
similarly to Eudragit RS.

Drug Release

Drug release was evaluated for the nanoparticles prepared from different polymers, and
the results are shown in Figure 8. Nanoparticles prepared from Eudragit L showed an
initial, instant release of ~30% of the drug in the acidic stage of the test. However, the
drug release was slightly slower than of pure ketoprofen. After the pH change, complete
dissolution of the nanoparticles took place rapidly. At the buffer conditions (pH 6.8), the
Eudragit L copolymer was ionized and soluble in the medium, thereby releasing the drug
immediately. Eudragit E polymer, however, is soluble at the acidic conditions. Complete
release of the drug was obtained for Eudragit E nanoparticles in the acidic medium in less
than 15 minutes as a result of polymer dissolution. As the polymer dissolved, also the
drug contained in the nanoparticles was forced into the solution. Drug release from the
nanoparticles was much faster than of pure ketoprofen (see Figure 8). Eudragit RS
nanoparticles showed sustained release of the drug at the acidic conditions, and the drug
release was found to be approximately linear. Similar to Eudragit L nanoparticles, ~30%
of the drug was released initially. Further drug release from the nanoparticle matrix was
controlled by the polymer. When the pH was changed, the amount of the released drug
changed rapidly from ~60% to almost 80%. At the buffer conditions, release of the drug
was faster as the ketoprofen molecules were ionized. Most likely, the ketoprofen
molecules at or close to the surface of the nanoparticles could deprotonate at these
conditions, and had a greater tendency to dissolve.

Two possible mechanisms are proposed for the initial 30% drug release from Eudragit L
and Eudragit RS nanoparticles. First, the nanoparticles can contain a larger a proportion
of drug at the surface of the particles in comparison to the interior of the particles. Such
an uneven distribution could arise from diffusion of the small molecules to the particle
surface during the particle preparation and drying process. When such particles are
immersed in dissolution medium, the drug at the surface is immediately released, and
only further control of drug release is due to the polymers. Second, if the dissolution
medium can penetrate to some extent to the polymer matrix, the drug molecules at and
close to the surface will dissolve.30,31 Theoretically calculated for a 100-nm nanoparticle,
which has a uniform distribution of drug (10% [wt/wt]) and polymer (90% [wt/wt]),
medium should be able to initially penetrate to a 5-nm depth to allow 30% drug release.
As in the previous case, further control of drug release is controlled by the properties of
the polymer.

Conclusion

In this study, it was observed that 2 factors affect the stability of nanoparticles during
collection in the aerosol flow reactor method. First, the interactions between the drug and
the polymer had an effect on the drug loading of the nanoparticles. For Eudragit L
nanoparticles, the solubility of drug in the polymer matrix was the limiting factor for drug
incorporation. When the amount of drug in the polymer matrix was higher than the
solubility limit of drug in the polymer, crystallization of drug was observed. Second, the
thermal properties of the drug-polymer composite nanoparticles affected the stability of
the nanoparticles. For the nanoparticles containing Eudragit RS and Eudragit E, the
thermal properties of the nanoparticles did not allow collection of dry powders. The drug
acted as a plasticizer to the polymers, and the glass transition temperature of the
nanoparticles was lowered close to room temperature. Consequently, the mechanical
strength of the nanoparticles was lost, which led to coalescence of the nanoparticles.

Acknowledgements
The authors are grateful to Mr Marc Donsmark (Donsmark Process Technology,
Fredriksberg, Denmark) for donating the Eudragit materials. The authors wish to thank
Mr Raoul Järvinen for his assistance in building the experimental set-up. Prof Heikki
Tenhu (University of Helsinki, Department of Chemistry) is acknowledged for DSC and
IR analysis equipment time.

References

1. Kreuter J. Nanoparticles. In: Swarbrick J, Boylan J C, eds. Encyclopedia of


Pharmaceutical Technology. Vol. 10. New York, NY: Marcel Dekker; 1994:165-190.

2. Couvreur P, Dubernet C, Puisieux F. Controlled drug delivery with nanoparticles:


current possibilities and future trends. Eur J Pharm Biopharm. 1995;41:2-13.

3. Brigger I, Dubernet C, Couvreur P. Nanoparticles in cancer therapy and diagnosis.


Adv Drug Deliv Rev. 2002;54:631-651.
PubMed DOI

4. Peltonen L, Koistinen P, Karjalainen M, Häkkinen A, Hirvonen J. The effect of


cosolvents on the formulation of nanoparticles from low-molecular-weight poly(l)lactide.
AAPS PharmSciTech. 2002;3:E32.
PubMed DOI

5. Damgé C, Michel C, Aprahamian M, Couvreur P, Devissaguet JP. Nanocapsules as


carriers for oral peptide delivery. J Control Release. 1990;13:233-239.
DOI

6. Damgé C, Vranckx H, Balschmidt P, Couvreur P. Poly(alkyl cyanoacrylate)


nanospheres for oral administration of insulin. J Pharm Sci. 1997;86:1403-1409.
PubMed DOI
7. Chen X, Young TJ, Sarkari M, Williams RO III, Johnston KP. Preparation of
cyclosporine A nanoparticles by evaporative precipitation into aqueous solution. Int J
Pharm. 2002;242:3-14.
PubMed DOI

8. Eerikäinen H, Kauppinen EI. Preparation of polymeric nanoparticles containing


corticosteroid by a novel aerosol flow reactor method. Int J Pharm. 2003;263:69-83.
PubMed DOI

9. Eerikäinen H, Kauppinen EI, Kansikas J. Polymeric drug nanoparticles prepared by


an aerosol flow reactor method. Pharm Res. 2004;21:136-143.
PubMed DOI

10. Shukla AJ. Polymethacrylates. In: Wade A, Weller P J, eds. Handbook of


Pharmaceutical Excipients. 2nd ed. Washington, DC: American Pharmaceutical
Association, Pharmaceutical Press; 1994.

11. Dittgen M, Durrani M, Lehmann K. Acrylic polymers: a review of pharmaceutical


applications. STP Pharma Sci. 1997;7:403-437.

12. US Pharmacopeia XXVII. <724> Drug Release. Rockville, MD: United States
Pharmacopeial Convention; 2003.

13. US Pharmacopeia XXVII. <711> Dissolution. Rockville, MD: United States


Pharmacopeial Convention; 2003.
14. TSI Incorporated. Model 3075/3076 Constant Output Atomizer Instruction Manual.
St Paul, MN: TSI Incorporated; 2000.

15. Lefebvre AH. Atomization and sprays. In: Chigier N, ed. Combustion: An
International Series. New York, NY: Hemisphere Publishing Corporation; 1989.

16. Bodmeier R, Chen H. Preparation and characterization of microspheres containing


the anti-inflammatory agents, indomethacin, ibuprofen, and ketoprofen. J Control
Release. 1989;10:167-175.
DOI

17. Habib MJ, Mesue R. Development of controlled release formulations of ketoprofen


for oral use. Drug Dev Ind Pharm. 1995;21:1463-1472.

18. Dubernet C, Rouland JC, Benoit JP. Ibuprofen-loaded ethylcellulose microspheres:


analysis of the matrix structure by thermal analysis. J Pharm Sci. 1991;80:1029-1033.
PubMed

19. Palmieri GF, Bonacucina G, Di Martino P, Martelli S. Gastro-resistant microspheres


containing ketoprofen. J Microencapsul. 2002;19:111-119.
PubMed DOI

20. Pignatello R, Ferro M, Puglisi G. Preparation of solid dispersions of nonsteroidal


anti-inflammatory drugs with acrylic polymers and studies on mechanisms of drug-
polymer interactions. AAPS PharmSciTech. 2002;3:E10.
PubMed DOI

21. Wunderlich B. Thermal Analysis. San Diego, CA: Academic Press, Inc; 1990.
22. Dubernet C. Thermoanalysis of microspheres. Thermochim Acta. 1995;248:259-269.
DOI

23. Wu C, McGinity JW. Non-traditional plasticization of polymeric films. Int J Pharm.


1999;177:15-27.
PubMed DOI

24. Wu C, McGinity JW. Influence of ibuprofen as a solid-state plasticizer in Eudragit


RS 30 D on the physicochemical properties of coated beads. AAPS PharmSciTech.
2001;2:E24.
DOI

25. Sancin P, Caputo O, Cavallari C, et al. Effects of ultrasound-assisted compaction on


Ketoprofen/Eudragit S100 mixtures. Eur J Pharm Sci. 1999;7:207-213.
PubMed DOI

26. Mura P, Faucci MT, Parrini PL, Furlanetto S, Pinzauti S. Influence of the
preparation method on the physicochemical properties of ketoprofen-cyclodextrin binary
systems. Int J Pharm. 1999;179:117-128.
PubMed DOI

27. Lin S-Y, Liao C-M, Hsiue G-H, Liang R-C. Study of a theophylline-Eudragit L
mixture using a combined system of microscopic Fourier-transform infrared spectroscopy
and differential scanning calorimetry. Thermochim Acta. 1995;254:153-166.
DOI

28. Lin SY, Perng RI. Solid-state interaction studies of drugs/polymers: I.


Indomethacin/Eudragit E, RL or S resins. STP Pharma Sci. 1993;3:465-471.
29. Lin S-Y, Yu H-L, Li M-J. Formation of six-membered cyclic anhydrides by
thermally induced intramolecular ester condensation in Eudragit E film. Polym.
1999;40:3589-3593.
DOI

30. Krause H-J, Schwarz A, Rohdewald P. Polylactic acid nanoparticles, a colloidal


delivery system for lipophilic drugs. Int J Pharm. 1985;27:145-155.
DOI

31. Higuchi T. Mechanism of sustained-action medication. J Pharm Sci. 1963;52:1145-


1149.
PubMed

http://www.aapspharmscitech.org/view.asp?art=pt050468&pdf=yes#ref25

S-ar putea să vă placă și