You are on page 1of 10

G Model

COLSUB-6552; No. of Pages 10 ARTICLE IN PRESS


Colloids and Surfaces B: Biointerfaces xxx (2014) xxx–xxx

Contents lists available at ScienceDirect

Colloids and Surfaces B: Biointerfaces


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

Promising dissolution enhancement effect of soluplus on crystallized


celecoxib obtained through antisolvent precipitation and high
pressure homogenization techniques
Alireza Homayouni a,b , Fatemeh Sadeghi a,b , Jaleh Varshosaz c ,
Hadi Afrasiabi Garekani b,d,∗∗ , Ali Nokhodchi e,f,∗
a
Targeted Drug Delivery Research Center, School of Pharmacy, Mashhad University of Medical Sciences, Mashhad, Iran
b
Department of Pharmaceutics, School of Pharmacy, Mashhad University of Medical Sciences, Mashhad, Iran
c
School of Pharmacy, Esfahan University of Medical Sciences, Esfahan, Iran
d
Pharmaceutical Research Center, School of Pharmacy, Mashhad University of Medical Sciences, Mashhad, Iran
e
Medway School of Pharmacy, University of Kent, Chatham Maritime, ME4 4TB Kent, UK
f
Drug Applied Research Center and Faculty of Pharmacy, Tabriz University of Medical Sciences, Tabriz, Iran

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

Article history: Poor solubility and dissolution of hydrophobic drugs have become a major challenge in pharmaceuti-
Received 24 April 2014 cal development. Drug nanoparticles have been widely accepted to overcome this problem. The aim of
Received in revised form 21 July 2014 this study was to manufacture celecoxib nanoparticles using antisolvent precipitation and high pres-
Accepted 22 July 2014
sure homogenization techniques in the presence of varying concentrations of soluplus® as a hydrophilic
Available online xxx
stabilizer. Antisolvent crystallization followed by freeze drying (CRS-FD) and antisolvent crystallization
followed by high pressure homogenization and freeze drying (HPH-FD) were used to obtain celecoxib
Keywords:
nanoparticles. The obtained nanoparticles were analyzed in terms of particle size, saturation solubility,
Celecoxib nanoparticles
Antisolvent crystallization
morphology (optical and scanning electron microscopy), solid state (DSC, XRPD and FT-IR) and dissolution
High pressure homogenization behavior. The results showed that celecoxib nanoparticle can be obtained when soluplus was added to the
Solid state analysis crystallization medium. In addition, the results showed that the concentration of soluplus and the method
Dissolution used to prepare nanoparticles can control the size and dissolution of celecoxib. Samples obtained in the
presence of 5% soluplus through HPH technique showed an excellent dissolution (90%) within 4 min. It is
interesting to note that celecoxib samples with high crystallinity showed better dissolution than those
celecoxib samples with high amorphous content, although they had the same concentration of soluplus.
DSC and XRPD proved that samples obtained via HPH technique are more crystalline than the samples
obtained through only antisolvent crystallization technique.
© 2014 Elsevier B.V. All rights reserved.

1. Introduction poorly water soluble drugs [4]. A reduction in particle size increases
the effective surface area of particles and this in turn increases
It has been reported that approximately 40% of marketed drugs the dissolution rate according to the Noyes–Whitney equation
and 70% of new drug candidates suffered from poor water solubil- [3]. Moreover, the reduction in the particle size of drugs down
ity [1,2]. The oral bioavailability of these drugs could be improved to nanometer range enhances the saturation solubility due to the
by enhancing the solubility and dissolution rate. Several strategies Ostwald–Freundlich equation which is normally not observed for
have been developed to improve dissolution rate of poorly water micronized particles [5]. The preparation of nanoparticles can be
soluble drugs [3]. Nanoparticle engineering is one of the recent basically classified into two major methods, top-down and bottom-
strategies that has been used to improve the oral bioavailability of up methods. In top-down methods nanoparticles are produced by
breaking down the large particles to smaller particle size using high
energy processes such as wet milling and high pressure homog-
enization [2,6]. In contrast, preparation of nanoparticles through
∗ Corresponding author at: Medway School of Pharmacy, University of Kent,
bottom-up methods starts from molecular level. In this method,
Chatham Maritime, ME4 4TB Kent, UK. Tel.: +44 1634 202947.
∗∗ Corresponding author. nanoparticles can be produced by precipitation/crystallization [7].
E-mail addresses: hadiafrasiabi@yahoo.com (H. Afrasiabi Garekani), In precipitation technique simply the drug is dissolved in an organic
a.nokhodchi@kent.ac.uk (A. Nokhodchi). solvent and then the drug is precipitated by the addition of an

http://dx.doi.org/10.1016/j.colsurfb.2014.07.037
0927-7765/© 2014 Elsevier B.V. All rights reserved.

Please cite this article in press as: A. Homayouni, et al., Promising dissolution enhancement effect of soluplus on crystallized cele-
coxib obtained through antisolvent precipitation and high pressure homogenization techniques, Colloids Surf. B: Biointerfaces (2014),
http://dx.doi.org/10.1016/j.colsurfb.2014.07.037
G Model
COLSUB-6552; No. of Pages 10 ARTICLE IN PRESS
2 A. Homayouni et al. / Colloids and Surfaces B: Biointerfaces xxx (2014) xxx–xxx

antisolvent. One of the main problems of nanoparticles is the adding 10 ml of the methanolic solution of CLX using a syringe at
crystal growth after the precipitation [8]. The presence of stabi- a rate of 10 ml/min into 200 ml of water containing soluplus under
lizer could be more useful to inhibit the crystal growth and also stirring condition (800 rpm at room temperature). The amount of
aggregation of nanoparticles. Moreover, a hydrophilic stabilizer SOL was varied from 0.025 to 0.25% (w/v). The concentration of
could increase the wettability of hydrophobic drugs hence enhanc- methanolic solution was varied from 5 to 9.5% (w/v). Different sam-
ing the water solubility and oral bioavailability [7]. In order to ples of CLX:SOL were prepared to achieve final composition ratios of
control the particle size and prevent aggregation of nanocrystals 95:5, 90:10, 75:25 and 50:50. After the addition of methanolic solu-
in suspension, combination of both techniques (bottom-up and tion, CLX particles start precipitating immediately (the total time
top-down) has recently been used. The Baxter company invented a for stirring after the addition of methanolic solution was 1 min).
combination technology called “NANOEDGE” in which the precip- After 1 min, the obtained suspension was immediately placed in
itated nanocrystals was immediately subjected to high pressure −80 ◦ C and freeze dried with the Heto freeze dryer (Denmark) for
homogenization (HPH) to inhibit the crystal growth and break 48 h. Each crystallization experiment was repeated three times.
the large agglomerates [2,9]. Polyvinyl caprolactam–polyvinyl
acetate–polyethylene glycol (soluplus® ) is a new graft copolymer 2.2.2. Preparation of CLX:SOL freeze dried nanoparticles
with amphiphilic properties that has been introduced by BASF. precipitated by antisolvent crystallization method followed by
This polymer could act as stabilizer and also as solubilizing agent high pressure homogenizer (HPH-FD)
in formulation of poorly water soluble drugs. Soluplus has been The suspensions obtained from the antisolvent crystalliza-
shown excellent solubilizing properties for poorly water soluble tion (CRS) process were further homogenized by a high pressure
drugs [10–12]. Furthermore, it has been reported that this polymer homogenizer (Avestin, Emulsiflex C3) at 500 and 1000 bars for 3 and
was capable of increasing the oral bioavailability of danazol, 5 cycles respectively. The obtained suspension was immediately
fenofibrate and itraconazol [13]. Similar results were reported for freeze dried at −80 ◦ C for 48 h using the freeze dryer as described
bicalutamide nanocrystals prepared by antisolvent crystallization above.
using soluplus [14].
Celecoxib is a selective cyclooxygenase 2 inhibitor used for the 2.2.3. Preparation of physical mixtures of drug carrier
treatment of osteoarthritis, rheumatoid arthritis, and acute pain. For better comparison between various samples, the physical
Furthermore, it is used as a combined therapy with chemotropic mixtures of CLX and SOL with the same ratios used in other methods
agents in cancer treatments [15]. This drug is classified as a BCS mentioned above were also prepared. To this end, the sieved frac-
Class II drug (poor solubility, high permeability) with an aqueous tions (smaller than 250 ␮m) of the both CLX and SOL were mixed
solubility of 1–3 ␮g/ml [16,17]. The poor solubility of celecoxib using mortar and pestle for 20 min to obtain a uniform mixture. The
(CLX) in gastrointestinal tract leads to an incomplete absorp- formulations were kept in glass vials until further studies.
tion and hence poor bioavailability. It has been reported that
the oral bioavailability of CLX capsule is 30% when administrated 2.2.4. Preparation of amorphous CLX
to dogs [15]. Thus, increasing the solubility and dissolution rate Amorphous CLX was prepared by melting the pure CLX at 170 ◦ C
could enhance the oral bioavailability of CLX. Many formulation under nitrogen atmosphere and then rapidly cooling it in an ice bath
approaches have been used to increase the solubility and dissolu- as described by Andrews et al. [20]. The prepared amorphous CLX
tion rate of CLX in last few years [16–28]. These include solid dis- was used in DSC analysis.
persion with PVP, meglumine, HPMC, PVA, PEG, Kollicoat [18–21],
polymer induced crystallization using PVP, HPMC, Cremophor, SDS 2.2.5. Morphological analysis
[23,24,29], cocrystallization using nicotinamide [16], salt formation A scanning electron microscope (LEO 1450 VP, Germany) at an
in combination with crystallization inhibitor [25], complexation acceleration voltage of 20 kV was used to investigate the morpholo-
with ␤-cyclodextrin [26] and nanoparticle formation [15,27,28]. To gies of pure CLX and CLX:SOL samples. All freeze-dried samples
the best of our knowledge there is no systematic study on the poten- were coated with a thin gold-palladium layer by sputter coater
tial use of soluplus (SOL) on the solubility/dissolution enhancement (SC 7620, England). Optical microscope (Olympus BX-60, Japan)
of CLX through making micro/nanocrystals. In other words, the aim was also used to investigate the morphology of nanosuspensions
of the present study was to enhance the dissolution rate of CLX by obtained from CRS and HPH methods before freeze drying.
making micro/nanocrystals of CLX in the presence of soluplus as a
hydrophilic stabilizer using high pressure homogenization (HPH) 2.2.6. Particle size measurement
technique. In addition, a full comprehensive solid state analysis of To measure the size of crystals obtained through the anti-
the resultant micro/nanoparticles was performed to elaborate the solvent crystallization (CRS) and high pressure homogenization
results obtained in the present study. (HPH) techniques, a nano-zetasizer (Malvern, UK) was used. The
nanosuspension was diluted with deionized water to approxi-
2. Materials and methods mately 1 mg/ml for each measurement. The measurements were
run three times to determine the mean and standard deviation of
2.1. Materials Z-average size and zeta potential of the particles.

Celecoxib (CLX) was purchased from Arastoo chemical com- 2.2.7. Solubility measurement
pany, soluplus was donated from BASF (Germany) and sodium The saturation solubilities of pure CLX, physical mixtures
dodecyl sulfate was obtained from Merck (Germany). All other sol- of CLX:SOL and treated CLX-SOL samples obtained via differ-
vents and chemicals were of analytical grade. ent methods were determined. To this end, an excess amount
of samples were added to 15-ml screw-capped glass vials con-
2.2. Methods taining 10 ml double-distilled water. The vials were shaken at
200 rpm in an air bath (25 ◦ C) for 48 h. Preliminary results showed
2.2.1. Preparation of CLX:SOL freeze dried microparticles that 48 h was enough to reach the equilibrium condition. The
precipitated by antisolvent crystallization method (CRS-FD) resulting suspensions were then filtered through a 0.22 ␮m filter
Celecoxib and soluplus were dissolved in methanol and distilled (Biofil MCE membrane). The concentration of CLX was determined
water respectively. Antisolvent crystallization was performed by spectrophotometrically at 253 nm (Cecile 900, USA). The solubility

Please cite this article in press as: A. Homayouni, et al., Promising dissolution enhancement effect of soluplus on crystallized cele-
coxib obtained through antisolvent precipitation and high pressure homogenization techniques, Colloids Surf. B: Biointerfaces (2014),
http://dx.doi.org/10.1016/j.colsurfb.2014.07.037
G Model
COLSUB-6552; No. of Pages 10 ARTICLE IN PRESS
A. Homayouni et al. / Colloids and Surfaces B: Biointerfaces xxx (2014) xxx–xxx 3

Table 1
Mean particle size, zeta-potential, glass transition temperature (Tg ), melting point and enthalpy of fusion (H) of pure CLX, samples prepared by antisolvent crystallization
(CRS) and samples prepared by antisolvent crystallization followed by high pressure homogenization (HPH) in the presence of various soluplus (SOL) ratios (data are the
mean and standard deviations of three determinations).

SampleCLX:SOL Z-average (nm) Zeta-potential 1st peak (◦ C)Tg /H (J/g) 2rd peak (◦ C)/H (J/g)

CLX 5528 ± 265 −23 – 163.39/90.26


CLX CRS – – – 162.91/95.73
95:5 CRS 1248 ± 451 −24 – 160.67/52.07
90:10 CRS 1474 ± 368 −21 – 161.59/35.83
75:25 CRS 1306 ± 344 −16 66.62/0.35 153.50/9.11
50:50 CRS 293 ± 221 −13 68.97/0.42 –
CLX HPH 514 ± 40 −26 – 163.75/94.59
95:5 HPH 872 ± 51 −22 – 162.99/55.78
90:10 HPH 577 ± 62 −17 – 160.67/42.86
75:25 HPH 677 ± 86 −15 – 155.42/11.70
50:50 HPH 439 ± 35 −14 – 161.75/3.47

results are the mean and standard deviation of three determina- Elmer spectrum II (PerkinElmer, Waltham, MA). The obtained disk
tions. (sample-KBr) was scanned against pure KBr disk at wavenumbers
ranging from 450 to 4000 cm−1 with a resolution of 1.0 cm−1 .
2.2.8. Differential scanning calorimetry (DSC)
DSC analysis was conducted to study the thermal behaviors of
the samples. To this end, DSC 822e (Mettler Toledo, Switzerland) 2.2.11. Dissolution studies
equipped with a refrigerated cooling system was used and the A USP dissolution apparatus 2 (paddle method) (Pharmatest,
instrument was calibrated using indium standard. Samples of Germany) was used to monitor the dissolution profiles of CLX
pure celecoxib and CLX:SOL micro/nanoparticle samples (2–3 mg) formulations. Each dissolution vessel was filled with 1000 ml dis-
were placed in aluminum pans sealed with a lid. The crimped tilled water containing 0.25% (w/v) sodium dodecyl sulfate (SDS)
aluminum pans were placed inside the DSC and were scanned equilibrated to 37 ◦ C and the paddles rotated at 50 rpm. An accu-
from 20 to 200 ◦ C at a scanning rate of 10 ◦ C/min under nitro- rate weight of samples equivalent to 40 mg CLX was placed at
gen gas at a flow rate of 80 ml/min. To compare the thermal the top of the dissolution medium and immediately the parti-
behavior of various samples their onset temperatures, melting cles were sank to the bottom of the vessels. Samples were taken
points and endothermic and exothermic enthalpies were calcu- from the vessels at selected time intervals through sintered fil-
lated using the software provided (STARe Ver. 10.00 Mettler Toledo, ter by a peristaltic pump (Alitea, Sweden), and the concentration
Switzerland). of CLX in the samples assayed at 253 nm by a multi-cell changer
spectrophotometer (Shimadzu, Japan) based on a calibration curve
2.2.9. X-ray powder diffraction studies (XRPD) obtained for CLX at this wavelength. The dissolution of each sam-
X-ray powder diffraction patterns of the selected samples were ple was determined in triplicate. Soluplus showed no absorption
collected on a Bruker, D8 Advance, Germany diffractometer, with at 253 nm.
Cu K␣ radiation (1.54056 Å). Each sample was scanned in the range
of 5–40◦ (2) with a step size of 0.05.
2.2.12. Dissolution parameters
2.2.10. Fourier transform infrared spectroscopy (FT-IR) In most cases it is more sophisticated and desirable to consider
FT-IR was used to investigate any changes in crystallized CLX the whole dissolution profile, since the use of a single point such as
on the molecular level during the preparation of CLX:SOL particles. t50% or t90% neglects all other data points. Therefore, the dissolution
Each sample was mixed with KBr at a ratio of 1:5 and was com- efficiency (DE) and mean dissolution time (MDT) were calculated
pressed at a pressure of 7 tonnes for 2 min in a hydraulic press to to reflect the overall dissolution performance of various celecoxib
make disks for FT-IR. The spectrum was obtained using a Perkin formulations. The DE is the area under the dissolution curve (y)

Table 2
Saturation solubility, dissolution efficiency (DE%) and mean dissolution time (MDT) of CLX, it physical mixtures with soluplus (PM), samples prepared by antisolvent crystal-
lization (CRS) and samples prepared by antisolvent crystallization followed by high pressure homogenization (HPH) in the presence of various ratios of celecoxib:soluplus
(data are the mean and standard deviations of three determinations).

SampleCLX:SOL Saturation solubility (␮g/ml) DE%90 min MDT (min)

CLX 2.1 ± 0.7 42.2 ± 0.4 17.4 ± 0.3


Amorphous CLX 2.2 ± 0.5 15.8 ± 0.3 31.8 ± 0.2
95:5 PM 2.2 ± 0.1 64.3 ± 1.4 19.0 ± 0.4
90:10 PM 2.5 ± 0.2 60.0 ± 0.9 23.1 ± 1.8
75:25 PM 2.7 ± 0.2 53.3 ± 0.8 25.9 ± 1.7
50:50 PM 3.2 ± 0.3 53.1 ± 1.8 31.2 ± 2.0
CLX CRS 2.2 ± 0.5 47.1 ± 0.5 17.2 ± 0.3
95:5 CRS 5.9 ± 0.3 91.4 ± 0.6 4.0 ± 0.5
90:10 CRS 6.5 ± 0.2 85.2 ± 0.7 8.7 ± 0.6
75:25 CRS 6.7 ± 0.1 79.6 ± 1.6 14.3 ± 0.4
50:50 CRS 8.2 ± 0.2 88.8 ± 1.8 6.2 ± 0.4
CLX HPH 3.9 ± 0.4 62.7 ± 0.4 11.6 ± 0.3
95:5 HPH 7.9 ± 0.1 95.1 ± 1.8 2.6 ± 1.2
90:10 HPH 8.0 ± 0.2 92.3 ± 3.4 4.7 ± 1.5
75:25 HPH 8.1 ± 0.2 88.8 ± 0.4 4.8 ± 0.2
50:50 HPH 8.1 ± 0.1 86.8 ± 1.3 6.2 ± 0.4

Please cite this article in press as: A. Homayouni, et al., Promising dissolution enhancement effect of soluplus on crystallized cele-
coxib obtained through antisolvent precipitation and high pressure homogenization techniques, Colloids Surf. B: Biointerfaces (2014),
http://dx.doi.org/10.1016/j.colsurfb.2014.07.037
G Model
COLSUB-6552; No. of Pages 10 ARTICLE IN PRESS
4 A. Homayouni et al. / Colloids and Surfaces B: Biointerfaces xxx (2014) xxx–xxx

Fig. 1. DSC traces of pure CLX, freeze dried samples prepared by antisolvent crystallization (CLX-CRS) and freeze dried samples prepared by antisolvent crystallization
followed by high pressure homogenization (CLX-HPH) obtained in the presence of different concentrations of soluplus.

up to a certain time t, expressed as a percentage of the area of the 3. Results and discussions
rectangle described by 100% dissolution in the same time [30].
t 3.1. Morphological analysis
0
y · dt
DE% = × 100
y100 · t Morphological analysis (optical microscopy and SEM) was per-
formed in order to investigate if any changes occurred in crystal
Another parameter that describes the rate of dissolution is the mean
habit of CLX particles during antisolvent crystallization or high
dissolution time (MDT); MDT is the statistical analyses that reflect
pressure homogenization (HPH) process. As it was shown in Figure
to the mean time of dissolution and has the advantage of being
S1-A (refer to supplementary materials), freshly crystallized CLX
applicable to all types of dissolution profiles. MDT is the most likely
without any stabilizer exhibited thin needle shaped crystals aggre-
time for a molecule to be dissolved from a solid dosage form or solid
gated to form star-shaped crystals. When this sample was subjected
particle and can be calculated using the following equations.
to the high pressure homogenization smaller particles with dif-

t̄ · Mi ti + ti+1 ferent morphologies were obtained (Figure S1-B, supplementary
MDT = i t̄ = Mi = (Mi+1 − Mi ) materials). CLX crystals obtained in the presence of 5% soluplus
Mi 2
showed different morphology (Figure S1-C, supplementary mate-
where t̄ is the midpoint of the time period during which the fraction rials). When this sample was subjected to homogenization through
M of the drug has been released from sample [30]. HPH, smaller and separate particles were obtained (Figure S1-D,

Please cite this article in press as: A. Homayouni, et al., Promising dissolution enhancement effect of soluplus on crystallized cele-
coxib obtained through antisolvent precipitation and high pressure homogenization techniques, Colloids Surf. B: Biointerfaces (2014),
http://dx.doi.org/10.1016/j.colsurfb.2014.07.037
G Model
COLSUB-6552; No. of Pages 10 ARTICLE IN PRESS
A. Homayouni et al. / Colloids and Surfaces B: Biointerfaces xxx (2014) xxx–xxx 5

polyvinyl caprolactam–polyvinyl acetate as a hydrophobic portion


and polyethylene glycol as a hydrophilic part. This amphiphilic
polymer could be adsorbed on the precipitated hydrophobic drug
surface, while the hydrophilic part provides stabilization in the
aqueous medium hence inhibiting the crystal growth [10]. Gener-
ally increasing the SOL concentration decreased the mean particle
size of the samples obtained via CRS and HPH methods. One should
be noted that due to the formation of gel like suspension the instru-
ment was not able to provide any reading for CLX crystallized
sample in the absence of soluplus.

3.3. Solubility measurement

Solubility measurement was carried out for pure CLX, physical


mixtures (PM) and all freeze dried (FD) samples. As it is shown in
Table 2 pure CLX showed very low solubility in water (2.1 ␮g/ml)
compared to the treated samples in the presence of SOL. In physi-
cal mixture samples (CLX:SOL), increasing the concentration of SOL
enhanced the saturation solubility of CLX that could be attributed
Fig. 2. XRPD pattern of raw CLX, physical mixture and CRS and HPH samples pre-
to the solubilizing effect of SOL [10]. Comparing the solubility data
pared by antisolvent crystallization (CRS) and samples prepared by antisolvent
crystallization followed by high pressure homogenization (HPH) in the presence of PM and crystallized samples (both CRS and HPH techniques)
of various concentrations of soluplus. demonstrated that antisolvent crystallization process in the pres-
ence of SOL has been more effective in enhancing the solubility
of CLX than its physical mixtures with CLX (Table 2). It seems the
supplementary materials). For better visualization (three dimen- mean particle size of samples plays a critical role in the solubil-
sional), SEM was performed after freeze drying which is discussed ity enhancement of CLX in the absence of SOL (for example the
below. saturation solubility of CLX HPH samples showed twofold higher
SEM image of pure CLX showed rod-shaped crystals (Figure than untreated CLX). According to Freundlich–Ostwald equation
S2-A, supplementary materials) while samples obtained by antisol- decreasing the particle size to nanometer range could increase the
vent crystallization followed by high pressure homogenization and saturation solubility [5]. As it was shown in Table 2, generally HPH
freeze drying (HPH-FD) showed smaller particles with platy mor- process could be more effective than simple antisolvent crystalliza-
phology (Figure S2-B, supplementary materials). Freeze-dried CLX tion technique to increase the saturation solubility of CLX except
samples obtained in the presence of 5% (Figure S2-C, supplementary when high concentration of SOL (50%) was used which there was no
materials) or 10% (Figure S2-D, supplementary materials) soluplus significant difference in the solubility of these two samples (ANOVA
(CLX:SOL CRS-FD) exhibited needle shaped crystals. It appears that test; p > 0.05).
the presence of SOL inhibited the crystal growth and the crys-
tal grew faster only in one dimension during the crystallization. 3.4. Differential scanning calorimetry (DSC)
The results also showed that samples with higher ratio of soluplus
(above 25%) produced a big lump of aggregated particles (figures DSC analysis was performed in order to investigate any interac-
not shown). It is interesting to note that when the above sam- tion between CLX and SOL or any changes in polymorphic form of
ples containing soluplus (5 and 10%) were subjected to HPH the CLX during the crystallization process. Fig. 1 shows the DSC traces
elongated particles became smaller (Figure S2-E, supplementary of CLX obtained via CRS-FD and HPH-FD samples and their melting
materials) or in the case of 10% soluplus these elongated particles points and enthalpy of fusion in Table 1. Pure CRS and HPH sam-
disappeared and different particles with a different morphology ples exhibited an endothermic peak around 163 ◦ C that is similar to
(irregular shape) was obtained (Figures S2-E and S2-F, supple- untreated CLX, and this is in agreement with previously published
mentary materials). This shows that the HPH method can break data [31]. The results reported in Table 1 show that the presence of
down the elongated particles (micrometer size) to smaller particles SOL in freeze dried samples decreased the melting point of CLX and
(nanometer size) to produce platy and rod-shaped nanoparticles the enthalpy of fusion which could be due to the partial dissolution
after freeze drying. of CLX in polymer during heating process. On the other hand in
antisolvent crystallization process, SOL could attach to CLX particle
3.2. Particle size measurement surfaces hence following freeze drying the SOL remained among
CLX particles. This occurrence caused smaller particle size of CLX
Particle size analysis was carried out to investigate the effect as well as fine dispersion of SOL among CLX particles. All of these
of processing method (antisolvent crystallization or high pressure reasons could cause a reduction in melting point and enthalpy of
homogenization) and soluplus concentration on the zeta potential fusion of CLX. Similar results were obtained for antisolvent crys-
and size of the particles. The results in Table 1 showed that HPH tallization of many API in presence of stabilizers [14,24,29,32]. As
process generated smaller particles compared to simple antisol- it was shown in Table 1 CRS samples with 25 and 50% SOL exhib-
vent crystallization (CRS) technique in the presence of soluplus. ited a small endothermic peak at 66.62 and 68.97 ◦ C that could be
The same trend was obtained for the samples containing the attributed to the formation of amorphous CLX (these small peaks
other concentrations of SOL. One exception in these samples was are visible at larger smaller scale in Fig. 1). In previous studies it has
CLX:SOL 50:50 CRS samples which showed smaller particles com- been demonstrated that amorphous CLX showed Tg around 59 ◦ C
pared to the samples subjected to HPH technique (Table 1). This [20,31]. In these samples deviation of endothermic peak (corre-
sample produced nanoparticles with the Z-average of 293 nm. This sponding to Tg of CLX) to higher temperature could be attributed
value is remarkably lower than other CRS samples that could be to the dispersion of CLX in SOL. SOL showed the Tg around 74 ◦ C
due to the inhibitory effect of SOL as a stabilizer during anti- (data not shown). In HPH samples this small endothermic peak
solvent crystallization. Soluplus is a graft copolymer containing did not appear. Moreover as it was shown in Table 1 and Fig. 1 in

Please cite this article in press as: A. Homayouni, et al., Promising dissolution enhancement effect of soluplus on crystallized cele-
coxib obtained through antisolvent precipitation and high pressure homogenization techniques, Colloids Surf. B: Biointerfaces (2014),
http://dx.doi.org/10.1016/j.colsurfb.2014.07.037
G Model
COLSUB-6552; No. of Pages 10 ARTICLE IN PRESS
6 A. Homayouni et al. / Colloids and Surfaces B: Biointerfaces xxx (2014) xxx–xxx

Fig. 3. FT-IR spectra of various CLX samples obtained through antisolvent crystallization techniques (CRS) in the presence of different concentrations of soluplus (SOL).

most HPH samples the melting point and fusion enthalpy of CLX are smaller mean particle size [34]. Moreover, by increasing the ratio
higher than similar CRS samples. These results demonstrated that of SOL in both CRS and HPH samples the intensity of peaks become
the crystalinity of CLX in HPH samples was higher than similar CRS smaller so that in the crystallized samples (CRS and HPH) con-
samples. In other words semi-crystalline and unstable fragile form taining 50% SOL the peaks of CLX completely diminished. This
converted to more stable and crystalline form. This phenomenon indicates that the amorphous nature of the crystallized samples
could be due to the high pressure homogenization process that has in the presence of high concentration of SOL increases. This was
been patented as Nanoedge® by Baxter Inc. [8,9,33]. However in not the case for physical mixtures of CLX:SOL (50:50) as it shows
50% SOL CRS and 50% SOL HPH samples it can be concluded that high crystallinity. The authors believe that in the crystallized sam-
most of CLX converted to amorphous form. ples (CRS or HPH) followed by freeze drying, CLX may disperse as
very fine particles (nm size) within the amorphous structure of
3.5. X-ray powder diffraction studies (XRPD) SOL leading to the observed amorphous structure. These results
were in good agreement with DSC studies. Careful comparison
As it is shown in Fig. 2 X-ray diffraction pattern of untreated between XRPD of samples obtained through CRS and HPH showed
CLX showed sharp peaks at 2 values of 5◦ , 10.5◦ , 16◦ and 21.5◦ that the intensity of peaks in HPH samples is slightly greater
which are diagnostic peaks for crystalline nature of CLX [31] while than the intensity of CRS samples (Fig. 2). For example in 5% SOL
SOL did not show any peak and exhibited amorphous state. CLX- HPH sample the intensity of peaks at 2 of 10.5◦ , 16◦ and 21.5◦
CRS and CLX-HPH samples showed similar pattern diffraction to increased approximately 7–13% compared to 5% SOL CRS sam-
raw CLX but with less intensity which is an indication of less ple. These increases in intensity could be due to the effect of
crystallinity for crystallized CLX. Reduction in intensity of char- high pressure homogenization on crystallinity as described before
acteristics peaks of these samples could also be attributed to the (Table 1).

Please cite this article in press as: A. Homayouni, et al., Promising dissolution enhancement effect of soluplus on crystallized cele-
coxib obtained through antisolvent precipitation and high pressure homogenization techniques, Colloids Surf. B: Biointerfaces (2014),
http://dx.doi.org/10.1016/j.colsurfb.2014.07.037
G Model
COLSUB-6552; No. of Pages 10 ARTICLE IN PRESS
A. Homayouni et al. / Colloids and Surfaces B: Biointerfaces xxx (2014) xxx–xxx 7

Fig. 4. FT-IR spectra of various CLX samples obtained through antisolvent crystallization techniques followed by high pressure homogenization (HPH) in the presence of
different concentrations of soluplus (SOL).

3.6. Fourier transform infrared spectroscopy (FT-IR) C O stretching for tertiary amide at 1643 cm−1 . On the basis of this
information Fig. 3 proves that all CRS samples prepared in the pres-
The FT-IR studies were performed to investigate if any interac- ence of SOL showed both amorphous and crystalline nature, while
tion between CLX and SOL has been occurred in molecular level in HPH samples (according to wavenumbers of NH2 groups) there
during the process of crystallization. The FT-IR spectra of pure CLX, is no shift in the wavenumber of all HPH samples compared to crys-
SOL and all crystallized samples are shown in Fig. 3. The FT-IR spec- talline CLX (Fig. 4). These results demonstrated that amorphous CLX
tra of CLX showed characteristic peaks at 3341 and 3234 cm−1 that particles converted to crystalline state when CRS samples were sub-
attributed to N H stretching vibration of SO2 NH2 group, 1348 and jected to high pressure homogenization process. All these findings
1165 cm−1 for the S O asymmetric and symmetric stretching and were supported by DSC and XRPD data (Figs. 1 and 2).
1230 for C F stretching. These results were in good agreement with
previous data [31]. It has been previously reported that in amor- 3.7. Dissolution studies
phous CLX the doublet peak corresponded to NH2 group shifted
to 3388 and 3264 cm−1 . In addition the wavenumber of asymmet- The dissolution test was performed on all treated, untreated
ric group of S O shifted from 1348 to 1340 cm−1 . These shifts are and physical mixtures of CLX:SOL and their dissolution profiles are
attributed to the intermolecular hydrogen bonding in sulfonamide shown in Fig. 5. The figure shows that in all cases the presence of
group that does not occur in crystalline state of CLX [20,31]. The soluplus improved the dissolution of CLX. This dissolution enhance-
spectra of SOL showed ester carbonyl stretching at 1739 cm−1 and ment was more pronounced when 5% SOL was used regardless of

Please cite this article in press as: A. Homayouni, et al., Promising dissolution enhancement effect of soluplus on crystallized cele-
coxib obtained through antisolvent precipitation and high pressure homogenization techniques, Colloids Surf. B: Biointerfaces (2014),
http://dx.doi.org/10.1016/j.colsurfb.2014.07.037
G Model
COLSUB-6552; No. of Pages 10 ARTICLE IN PRESS
8 A. Homayouni et al. / Colloids and Surfaces B: Biointerfaces xxx (2014) xxx–xxx

Fig. 5. Dissolution profiles for different samples: (A) physical mixtures, mixtures of CLX and soluplus, (B) freeze dried samples prepared by antisolvent crystallization (CRS-
FD), and (C) freeze dried samples prepared by antisolvent crystallization followed by high pressure homogenization (HPH-FD) samples (n = 3, and error bars are standard
deviation).

Please cite this article in press as: A. Homayouni, et al., Promising dissolution enhancement effect of soluplus on crystallized cele-
coxib obtained through antisolvent precipitation and high pressure homogenization techniques, Colloids Surf. B: Biointerfaces (2014),
http://dx.doi.org/10.1016/j.colsurfb.2014.07.037
G Model
COLSUB-6552; No. of Pages 10 ARTICLE IN PRESS
A. Homayouni et al. / Colloids and Surfaces B: Biointerfaces xxx (2014) xxx–xxx 9

the technique used to prepare the samples. This could be due to the crystalline state after high pressure homogenization (XRPD, Fig. 2)
hydrophilic nature of SOL resulting in a better wettability of the could be another reason for the higher dissolution rate of HPH-FD
samples. The results showed that the presence of SOL more than samples compared to their counterpart samples obtained through
5% caused a reduction in the dissolution of the drug from physical CRS. The results of one-way analysis of variance (ANOVA) showed
mixtures of CLX:SOL, but its dissolution still is faster than the dis- significant difference between DE% and MDT of CRS-FD and HPH-FD
solution of pure CLX (Fig. 5A). This may be due to the formation of samples (p < 0.05) with the exception of 50% SOL samples. It can be
a viscose gel around drug particles when the concentration of SOL concluded that CLX samples obtained through HPH technique show
is above the optimum level. All these data indicated that although better dissolution performance compared to the samples obtained
the hydrophilicity of SOL should increase the dissolution rate of via antisolvent precipitation technique.
CLX, the increase in the viscosity of the microenvironment around
drug particles produced by high concentration of SOL should not 4. Conclusion
be ignored.
The dissolution of amorphous CLX was also investigated and It was shown that soluplus is a potential stabilizer to be used
the results showed that the dissolution of amorphous CLX is slower in the crystallization of celecoxib by antisolvent crystallization
compared to crystalline CLX (Fig. 5B). The slower dissolution rate of and high pressure homogenization to produce nanoparticles with
amorphous CLX could be attributed to the devitrification of amor- improved dissolution rate. The concentration of soluplus during
phous CLX when in contact to aqueous environment and hence the crystallization can modulate the drug dissolution and also the
recrystallization to more hydrophobic structure that was reported particle size of celecoxib. Solid state analysis of celecoxib samples
previously by Puri et al. [35]. Similar results have been reported showed that the method used to prepare celecoxib particles had
for capecitabine [36], felodipine [37] and diazepam [38]. In the a big impact on the crystallinity of the samples and the particles
case of CRS-FD samples, all exhibited faster dissolution rate com- obtained through HPH showed higher crystallinity compared to the
pared to untreated CLX which could be due to the presence of samples obtained via a simple antisolvent crystallization technique.
smaller particles for all treated samples (see Fig. 5B and Table 1). In The results showed that the dissolution of celecoxib nanoparticels
CRS-FD samples the 5 and 50% SOL samples showed higher disso- is controlled by particle size, viscosity of the microenvironment
lution compared to the samples obtained in the presence of 10 and around the drug particles and solid state of celecoxib. By controlling
25% SOL. Although at higher concentration of SOL it was expected these three parameters the dissolution rate of celecoxib particles
slower dissolution compared to the samples containing 10 and can be optimized to achieve the highest dissolution rate.
25% SOL (as described before in the case of physical mixtures), the
increase in the dissolution of this sample could be attributed to the
Acknowledgments
small size of the sample (293 nm) which could overcome the nega-
tive effect of the viscosity on the dissolution of CLX. In case of 10 and
Financial support of Vice Chancellor for Research of Mashhad
25% SOL as they showed similar range of particle size then the effect
University of Medical Sciences, Iran (Grant No. 900213) is greatly
of particle size on their dissolution should be the same, therefore,
appreciated. The authors are also thankful to Pasaddak Company,
the negative effect of viscosity on dissolution would be more dom-
Iran, for providing gift sample of Soluplus® .
inant. In addition to the viscosity effect, the effect of amorphous
content of CLX in the samples should not be neglected as it has
Appendix A. Supplementary data
been shown that the dissolution of amorphous CLX is poorer than
the dissolution of crystalline CLX (Fig. 5B). XRPD results showed
Supplementary data associated with this article can be
that the crystallinity of CRS samples containing 5% SOL is higher
found, in the online version, at http://dx.doi.org/10.1016/j.colsurfb.
than CLX samples containing 10 and 25%. Therefore, the dissolu-
2014.07.037.
tion rate and dissolution efficiency (DE) is higher than other CRS
samples. In the case of 50% SOL CRS sample the hydrophilicity of
SOL could overcome the negative effect of amorphous nature. References
The increase in dissolution rate of CRS samples with 50% SOL
[1] E. Merisko-Liversidge, G.G. Liversidge, E.R. Cooper, Eur. J. Pharm. Sci. 18 (2003)
could also be due to very small size of CLX particles (Table 1; 113–120.
293 nm) which could overcome the negative effect of viscosity of [2] R. Shegokar, R.H. Müller, Int. J. Pharm. 399 (2010) 129–139.
high concentration of SOL in the dissolution of CLX (Fig. 5B). The [3] A. Singh, Z.A. Worku, G. Van den Mooter, Expert Opin. Drug Deliv. 8 (2011)
1361–1378.
figure also shows that CLX-CRS sample in the absence of SOL did [4] J. Hu, K.P. Johnston, R.O. Williams 3rd, Drug Dev. Ind. Pharm. 30 (2004) 233–245.
not show any remarkable increase in the dissolution compared to [5] F. Kesisoglou, S. Panmai, Y. Wu, Adv. Drug Deliv. Rev. 59 (2007) 631–644.
the untreated CLX. This indicated that the presence of SOL during [6] L. Gao, D. Zhang, M. Chen, J. Nanopart. Res. 10 (2008) 845–862.
[7] A.A. Thorat, S.V. Dalvi, Chem. Eng. J. 181/182 (2012) 1–34.
the crystallization is essential to improve the dissolution rate of
[8] B. Sinha, R.H. Müller, J.P. Möschwitzer, Int. J. Pharm. 453 (2013) 126–141.
CLX. [9] C.M. Keck, R.H. Muller, Eur. J. Pharm. Biopharm. 62 (2006) 3–16.
Fig. 5C shows that HPH-FD samples obtained in absence or [10] H. Hardung, D. Djuric, S. Ali, Drug Deliv. Technol. 10 (2010) 20–27.
[11] Z.K. Nagy, A. Balogh, B. Vajna, A. Farkas, G. Patyi, A. Kramarics, G. Marosi, J.
the presence of SOL showed higher dissolution rate compared to
Pharm. Sci. 101 (2012) 322–332.
untreated CLX that could be due to the smaller particle size of these [12] N.K. Thakral, A.R. Ray, D. Bar-Shalom, A.H. Eriksson, D.K. Majumdar, AAPS
samples compared to raw CLX (Table 1). This trend was backed up PharmSciTech 13 (2012) 59–66.
by the percentage dissolution efficiency (DE) and mean dissolution [13] M. Linn, E.-M. Collnot, D. Djuric, K. Hempel, E. Fabian, K. Kolter, C.-M. Lehr, Eur.
J. Pharm. Sci. 45 (2012) 336–343.
time (MDT) reported in Table 2. Comparing DE values of samples [14] V.B. Pokharkar, T. Malhi, L. Mandpe, Pharm. Dev. Technol. 18 (2013) 660–666.
obtained via CRS and HPH showed the better performance of parti- [15] A. Dolenc, J. Kristl, S. Baumgartner, O. Planinsek, Int. J. Pharm. 376 (2009)
cles obtained through HPH method (Table 2) particularly when the 204–212.
[16] J.F. Remenar, M.L. Peterson, P.W. Stephens, Z. Zhang, Y. Zimenkov, M.B. Hickey,
concentration of SOL was 0, 5, 10 and 25%. Several studies have been Mol. Pharm. 4 (2007) 386–400.
demonstrated the high efficiency of HPH process for enhancing [17] O.A. Abu-Diak, D.S. Jones, G.P. Andrews, Mol. Pharm. 8 (2011) 1362–1371.
the dissolution rate of several drugs [15,27]. The high dissolution [18] E.A. Fouad, M. El-Badry, G.M. Mahrous, F.K. Alanazi, S.H. Neau, I.A. Alsarra, Drug
Dev. Ind. Pharm. 37 (2011) 1463–1472.
rates obtained for HPH-FD samples compared to CRS-FD samples [19] P. Gupta, V.K. Kakumanu, A.K. Bansal, Pharm. Res. 21 (2004) 1762–1769.
could be attributed to the smaller particle size of HPH-FD sam- [20] G.P. Andrews, O. Abu-Diak, F. Kusmanto, P. Hornsby, Z. Hui, D.S. Jones, J. Pharm.
ples (Table 1). Furthermore, the conversion of amorphous CLX to Pharmacol. 62 (2010) 1580–1590.

Please cite this article in press as: A. Homayouni, et al., Promising dissolution enhancement effect of soluplus on crystallized cele-
coxib obtained through antisolvent precipitation and high pressure homogenization techniques, Colloids Surf. B: Biointerfaces (2014),
http://dx.doi.org/10.1016/j.colsurfb.2014.07.037
G Model
COLSUB-6552; No. of Pages 10 ARTICLE IN PRESS
10 A. Homayouni et al. / Colloids and Surfaces B: Biointerfaces xxx (2014) xxx–xxx

[21] J. Lee, M. Kim, H. Yoon, C. Shim, H. Ko, S. Cho, D. Lee, G. Khang, J. Pharm. Invest. [31] G. Chawla, P. Gupta, R. Thilagavathi, A.K. Chakraborti, A.K. Bansal, Eur. J. Pharm.
43 (2013) 205–213. Sci. 20 (2003) 305–317.
[22] T. Lee, C.W. Zhang, Pharm. Res. 25 (2008) 1563–1571. [32] J. Varshosaz, R. Talari, S.A. Mostafavi, A. Nokhodchi, Powder Technol. 187 (2008)
[23] V.R. Gupta, S. Mutalik, M.M. Patel, G.K. Jani, Acta Pharm. 57 (2007) 173–184. 222–230.
[24] M. Nasr, AAPS PharmSciTech 14 (2013) 719–726. [33] B.E. Rabinow, Nat. Rev. Drug Discov. 3 (2004) 785–796.
[25] H.R. Guzmán, M. Tawa, Z. Zhang, P. Ratanabanangkoon, P. Shaw, C.R. Gard- [34] H. Al-Hamidi, A.A. Edwards, M.A. Mohammad, A. Nokhodchi, Colloids Surf. B:
ner, H. Chen, J.-P. Moreau, Ö. Almarsson, J.F. Remenar, J. Pharm. Sci. 96 (2007) Biointerfaces 81 (2010) 96–109.
2686–2702. [35] V. Puri, A.K. Dantuluri, M. Kumar, N. Karar, A.K. Bansal, Eur. J. Pharm. Sci. 40
[26] M.N. Reddy, T. Rehana, S. Ramakrishna, K.P. Chowdary, P.V. Diwan, AAPS J. 6 (2010) 84–93.
(2004) 68–76. [36] J. Meulenaar, J.H. Beijnen, J.H. Schellens, B. Nuijen, Int. J. Pharm. 441 (2013)
[27] S. Wang, Y. Liu, C. Sun, Y. Hao, T. Jiang, L. Zheng, J. Pharm. Pharm. Sci. 13 (2010) 213–217.
589–606. [37] Z.A. Langham, J. Booth, L.P. Hughes, G.K. Reynolds, S.A.C. Wren, J. Pharm. Sci.
[28] A. Tan, S. Simovic, A.K. Davey, T. Rades, C.A. Prestidge, J. Control. Release 134 101 (2012) 2798–2810.
(2009) 62–70. [38] D.J. van Drooge, W.L. Hinrichs, H.W. Frijlink, J. Control. Release 97 (2004)
[29] H. Lee, J. Lee, J. Cryst. Growth 374 (2013) 37–42. 441–452.
[30] P. Costa, J.M. Sousa Lobo, Eur. J. Pharm. Sci. 13 (2001) 123–133.

Please cite this article in press as: A. Homayouni, et al., Promising dissolution enhancement effect of soluplus on crystallized cele-
coxib obtained through antisolvent precipitation and high pressure homogenization techniques, Colloids Surf. B: Biointerfaces (2014),
http://dx.doi.org/10.1016/j.colsurfb.2014.07.037

You might also like