You are on page 1of 11

European Journal of Pharmaceutical Sciences 20 (2003) 429–438

Interactions of ibuprofen with cationic polysaccharides


in aqueous dispersions and hydrogels
Rheological and diffusional implications
Rosal´ıa Rodr´ıguez, Carmen Alvarez-Lorenzo∗ , Angel Concheiro
Departamento de Farmacia y Tecnolog´ıa Farmacéutica, Facultad de Farmacia, Universidad de Santiago de Compostela,
Santiago de Compostela 15782, Spain
Received 8 July 2003; received in revised form 11 September 2003; accepted 11 September 2003

Abstract

Non-steroidal antiinflammatory drugs, such as ibuprofen, are amphiphilic substances capable of self-association in aqueous solutions and
able to be sorbed onto polymers through hydrophobic and electrostatic bonds. The aim of this work was to analyze the association processes
of sodium ibuprofen with cationic celluloses (Celquat® H-100 (PQ-4) and SC-230M (PQ-10)) and cationic guar gums (Ecopol® 261-S and
14-S) and their repercussions on the properties of the aqueous dispersions and cross-linked hydrogels. The interaction process was studied
in aqueous dispersions through transmittance, surface tension, fluorescence, conductivity, viscosity and oscillatory rheometry measurements.
Below cmc, the drug molecules weakly interact with the polymers through hydrophobic and ionic interactions. Around the cmc (4%), a
notable decrease in the viscosity, and storage and loss moduli of the dispersions (even precipitation in PQ-10 systems) was observed. An
additional increase in drug concentration induced the dispersions to recover their initial properties. Since ibuprofen/polymer cationic groups
ratio were in all cases above 1, these observations indicate that drug self-association induces the polymer to coil around the micelles and, as
the number of micelles increases (more drug concentration) the polymer chains interact with more of them, un- coiling again to some extent.
Polymer (1%) dispersions containing 6% ibuprofen showed drug diffusion coefficients much lower than in water. When a surfactant, sodium
dodecylsulfate, was added to these systems the diffusion coefficients decreased even more, suggesting the formation of new associative
structures. Chemically cross-linked hydrogels made of these cationic polysaccharides absorb consid- erable amounts of ibuprofen (up to 15
g/g) and showed a pH-dependent release process. At acidic pH, drug–polymer affinity is main- tained, preventing drug release. In contrast,
at pH 8 the interactions are broken and the release process is sustained for more than 4 h. In summary, ibuprofen interactions with
cationic polysaccharides strongly determine the performance of their aqueous dispersions and hydrogels.
© 2003 Elsevier B.V. All rights reserved.

Keywords: Amphiphilic drugs; Cationic cellulose; Cationic guar gum; Sodium dodecyl sulfate; Polymer–surfactant interactions

1. Introduction (Pouliquen et al., 2003). Most semi-synthetic polysaccha-


rides have a particularly adequate structure to interact with
Polymer association with complementary additives that amphiphilic molecules, which is interesting for this type of
can rapidly strengthen or induce connections between the applications. Cellulose and guar gum derivatives present a
polymeric chains has been shown as a useful way to obtain glucosidic backbone that may establish hydrophobic inter-
considerable increases in the viscosity of the dispersions— actions, while the presence of hydrophilic or charged groups
avoiding the difficult handling of high concentrations or in their substituents provides the polymer with hydrogen
high molecular weight polymers—or even the creation of bonding capacity or high affinity for oppositely charged
responsive systems whose properties are modulated by molecules. In several papers, the use of surfactants to in-
changes of the affinity between the polymer and the additives duce changes in the conformation of cationic polysaccha-
rides and to promote the formation of aggregates has been

+
Corresponding author. Fax: 34-981-547148. proposed as a way to obtain homogeneous aqueous disper-
E-mail address: ffrusdog@usc.es (C. Alvarez-Lorenzo). sions and to modulate their rheological behavior (Goddard

0928-0987/$ – see front matter © 2003 Elsevier B.V. All rights reserved. doi:10.1016/j.ejps.2003.09.004
430 R. Rodr´ıguez et al. / European Journal of Pharmaceutical Sciences 20 (2003) 429–438

and Hannan, 1977; Hoffman et al., 1996; Chronakis and Coulombic forces. These interactions may induce changes in
Alexandridis, 2001). Additionally, the high bioadhesive ca- the system which can be foreseen as useful to achieve an effi-
pacity (Gruber and Kreeger, 1996) and low toxicity (Annon, cient control of the release process from aqueous dispersions
1988) of hydrophilic cationic polysaccharides make them (Paulsson and Edsman, 2001; Jimenez-Kairuz et al., 2002)
particularly useful for preparing cosmetic preparations and or chemically cross-linked hydrogels (González-Rodr´ıguez
drug delivery systems able to combine long residence times et al., 2002; Rodr´ıguez et al., 2003a).
at the application site with adequate mechanical properties. The aim of this work was to analyze the association pro-
As classical surfactants, many kinds of drugs—i.e., tri- cesses of sodium ibuprofen with cationic celluloses and
cyclic antidepressants, þ-blockers, phenotiazine tranquiliz- cationic guar gums and their repercussions on the properties
ers, antihistamines, non-steroidal antiinflammatory drugs of the aqueous dispersions and cross-linked hydrogels. Two
(NSAIDs) and local anesthetics—are surface-active sub- varieties of cationic hydroxyethyl celluloses and two vari-
stances capable of self-association in aqueous solutions eties of cationic guar gums, differing in molecular weight
(Florence and Attwood, 1998; Taboada et al., 2001) and able and content and distribution of the cationic groups, were
to be sorbed onto polymers through hydrophobic and elec- used to evaluate the influence of polymer molecular struc-
trostatic bonds (Hugerth, 2001; Mura et al., 2003; Yomota ture. Additionally, the effect of the concomitant presence of
and Okada, 2003). The repercussions of these phenomena an anionic surfactant, sodium dodecylsulfate, which may
on the properties of polymer gels depend on the tendency compete with the drug for the binding to the polymer, was
of the drug to aggregate and on the strength of its interac- considered. Since NSAIDs are suitable candidates for incor-
tions with the polymer (Alvarez-Lorenzo and Concheiro, poration into polymer dispersions (physical gels) for topical
2003). In Carbopol® 1342 gels, alprenolol forms, at low application or into hydrogels that release them at the small
drug concentrations, micelle-like aggregates with the poly- intestine, minimizing the adverse gastric drug effects after
mer lipophilic residues, increasing the elastic and viscous oral administration (Dominkus et al., 1996), a preliminary
moduli of the gels. However, as the drug concentration is evaluation of the potential interest of the dispersions and
raised, the consistency decreases, the gel collapses and, hydrogels as drug delivery systems was also carried out.
when alprenolol amino groups neutralize the carboxylic acid
groups of the acrylic polymer, precipitation occurs
(Paulsson and Edsman, 2002). The changes observed in 2. Materials and methods
alprenolol diffusion rate are a consequence of both the
interactions with the polymer and the changes induced in 2.1. Materials
the viscosity of the systems. Interaction of phenothiazines
Polyquaternium-4 (PQ-4) (Celquat® H-100, batch FGS
(chlorpromazine, trifluoromazine, promazine, and promet-
1014) and Polyquaternium-10 (PQ-10) (Celquat® SC-230
hazine) with hyaluronate causes the gels made of this an-
M, batch GFS 1139) were provided by Na- tional Starch
ionic polymer to shrink; the minimum drug concentration
and Chemical Ltd., UK. Ecopol® 261-S (E-261, batch
required for that being proportional to the critical micellar
L10159215) and Ecopol® 14-S (E-14, batch L10159214)
concentration, cmc (Yomota and Okada, 2003). The po-
were from Economy Polymers & Chemi- cals, USA.
tential as periodontal drug delivery systems of non-ionic
Sodium ibuprofen, sodium dodecylsulfate and
cellulose ether–surface-active anesthetic drug gels was high-
ethyleneglycol diglycidylether (EGDE, 50 vol.% in water)
lighted by Scherlund et al. (2000). Lidocaine and prilocaine
were from Sigma–Aldrich Chemical Co., USA. Purified
do not interact with ethylhydroxyethyl cellulose (EHEC) or
water by reverse osmosis (MilliQ®, Millipore Spain) with
hydrophobically modified EHEC (HM-EHEC) but strongly
resistivity >1.82 MK cm was used. All other reactives were
affect their interactions with sodium dodecylsulfate and
of analytical grade.
myristoylcholine bromide. In consequence, the drug loaded
systems notably differed from polymer–surfactant disper- 2.2. Methods
sions in the storage and loss moduli.
Non-steroidal antiinflammatory drugs (NSAIDs), such as 2.2.1. Preparation of the dispersions
sodium ibuprofen, adsorb onto the non-ionic cellulose ethers Aqueous dispersions containing 1.0% (w/w) cationic cel-
non-cooperatively up to the cmc and cooperatively above lulose and a wide range of ibuprofen concentrations (0–8%,
this concentration (Ridell et al., 1999). Near cmc, ibupro- fen w/w) were prepared by dispersing the required amounts of
causes phase separation of EHEC dispersions, because of a each component in 100 ml of water under stirring. The sys-
strong hydrophobic interaction that shrinks the polymer tems were left to stand for 24 h before characterization. All
chains but, above cmc, micelles of ibuprofen solubilize the
studies were carried out at 25 ◦C.
hydrophobic parts of the polymer. The presence of apolar re-
gions and ammonium groups in the cationic polysaccharides 2.2.2. Cloudiness
could enhance the intensity of interactions with the aromatic The cloudiness of dispersions was determined by mea-
ring and the hydrophilic carboxylic group of NSAIDs, mak- suring transmittance at 800 nm (Shimadzu UV-240, Japan)
ing it possible to combine both hydrophobic association and against a blank of cationic polysaccharide dispersion.
R. Rodr´ıguez et al. / European Journal of Pharmaceutical Sciences 20 (2003) 429–438 431

2.2.3. pH magnetic rod. Samples (0.10 ml) were taken from the recip-
The measurements were made with a pH-meter Crison, ient compartment at intervals over a 10-h period, for deter-
model GLP22 (Spain), equipped with a sensor (Ag/AgCl) mination of ibuprofen on the basis of absorption at 273 nm
for viscous media no. 52-21. (Shimadzu UV-240, Kyoto, Japan); in each case, recipient
medium volume was immediately made up with iso-osmotic
2.2.4. Conductivity solution. Diffusion coefficients were estimated by non-linear
Conductance measurements were made in a rhadiometer regression applying the Higuchi (1962) equation
conductivity meter model CDM2e (Denmark) equipped with 1/2

a Crison platinum sensor (Spain). A . π Σ (1)


Q Dt
= 2C0

2.2.5. Surface tension where Q/A is the amount of ibuprofen released per unit area
The measurements were made using the platinum ring at the time t, and C0 is the initial concentration of ibuprofen
method with a Lauda tensiometer TD1 (Germany) applying in the dispersion.
the necessary density corrections.
2.2.10. Hydrogel synthesis
2.2.6. Steady-state fluorescence measurements Hydrogels were synthesized as described previously
Pyrene (Py) emission spectra (λ=350–450 nm) were (Rodr´ıguez et al., 2003a). Briefly, EGDE (2 ml) was added
recorded in a Perkin-Elmer LS50B fluorescence spectropho- to a 3% cationic cellulose dispersion in 0.10 M NaOH
tometer (UK), with the excitation wavelength set to 310 nm medium (10 ml) or to a 2% cationic guar gum solution in
and slits set to 5 and 5, for excitation and emission, respec- 0.05 M KOH medium (10 ml). After stirring for 5 min, the
tively. The samples were prepared by the addition of polymer
mixture was transferred to a test tube of 10.5 mm i.d., and
(0.5% cationic celluloses or 0.1% cationic guar gums) and
hermetically closed. The tubes were kept at 60 ◦C for 6 h, in
ibuprofen (0.5–7%) to a pyrene aqueous solution (10−6 M)
the case of guar gums, and 24 h, in the case of cationic
and stored at 25 ◦C for 24 h. The ratio of the intensity at the celluloses. After cooling, the hydrogels were carefully re-
first (I1 at 373 nm) and the third (I3 at 392 nm) vibronic peaks moved and immersed in ultrapure water for 24 h to swell.
was used as an index of the local hydrophobicity of the The hydrogels were then transferred to HCl 0.1N solution
polymer–surfactant aggregates (Rodr´ıguez et al., 2003b). for 12 h to neutralize the basic medium, and finally kept in
ultrapure water for 1 week, changing the medium every 12 h
2.2.7. Viscosity to allow a complete wash out of non-reacted substances.
Determinations of apparent and specific viscosity were
carried out in Cannon-Fenske capillary viscometers (Afora, 2.2.11. Equilibrium swelling of hydrogels
Spain). After being immersed in ibuprofen solutions (as explained
below), equilibrium diameters d of pieces of cylindrical gels
2.2.8. Viscoelastic behavior were measured using a digital micrometer. The degree of
The rheological behavior of the polymer dispersions with swelling was expressed as
and without ibuprofen was evaluated, in triplicate, in a Rheo- 3

= V
lyst AR-1000N rheometer (TA Instruments, UK) equipped . Σ
Swelling ratio (2)
with an AR2500 data analyzer, fitted with a Peltier temper- V0 = d
ature control, and a 6 cm cone-plate measuring geometry, d0

covered with a solvent trap. Oscillatory shear responses (G± where d0 was the diameter of gel discs upon synthesis
(10.5 mm).
or storage modulus, and G±± or loss modulus) were deter-
The water content of the hydrogel was estimated by the
mined at 0.1 Pa over the frequency range of 0.01–50 rad s−1.
difference between the weight of fully swollen hydrogel
The test conditions were inside the linearity range of the
samples (W), after careful wiping of their surfaces with a soft
viscoelastic properties.
tissue, and the weight of the samples after being dried (W0)
2.2.9. Ibuprofen diffusion assays for 1 week at 40 ◦C

Drug release profiles from 1% polymer dispersions con- (W − W0)


taining 6% ibuprofen (and in some cases SDS at different Q = (3)
W0
concentrations) were obtained in triplicate in Franz-Chien
diffusion cells (Vidra-Foc, Spain), thermostated at 37 ◦C, 2.2.12. Ibuprofen sodium loading and release from
and fitted with cellulose acetate filters (0.45µm pore size, cross-linked hydrogels

Teknokroma, Spain) after previous storage at the same tem- 0.785 cm2. A sample of 2.5 ml was placed in the donor com- partment;
perature for at least 1 h. The area available for diffusion was while 6.0 ml of iso-osmotic NaCl aqueous solu- tion filled the recipient
432
compartment R. Rodr´ıguez
and was stirred with a et al. / European Journal of Pharmaceutical
Pieces ofSciences 20 (2003) 429–438
the cylindrical gels (2–3 mm thickness) were
placed in 10 ml aqueous solutions of 1–8% drug for 3 days
at room temperature. The amount loaded was determined as
the difference between the initial amount of drug and the
amount remaining in the outer solution after the 3 days,
R. Rodr´ıguez et al. / European Journal of Pharmaceutical Sciences 20 (2003) 429–438 433

which was measured spectrophotometrically (λ= 273 nm; PQ-10


Shimadzu UV-240, Japan). Then, each loaded hydrogel was H OH H H
superficially rinsed with water and directly immersed in 50 HO H OH
HO
ml water of pH 3 (HCl aq.), pH 6 (not adjusted), or pH 8 HO
O

(phosphate buffer), and kept at 37 ◦C. The release was also


HO O
H H O
OH Cl-

evaluated when the hydrogels where immersed first in pH H H H O (CH CH O) CH CHOHCH N+(CH )

2 2 n 2 2 3 3
3 medium (for 2 h) then, transfer to pH 6 medium (for 2 h), P Q-4

and subsequently to pH 8 solution. The experiments were


H O (CH2CH2O) nH H H
carried out in triplicate. The amount of ibuprofen released
OH
was measured spectrophotometrically (λ = 273 nm) in pe-
HO H
HO O
HO
HO O

riodically taken samples and again placed in the same ves- OH


H O
sel so that the liquid volume was kept constant. The release
H O ( CH2CH2O) nH
profiles between 10 and 80% were characterized by fitting
the following equation by non-linear regression (Korsmeyer
Cl-
and Peppas, 1981)
Cl-

Mt n
Ibuprofen ONa

= Kt (4)
M∞

where Mt/M ∞ is the fraction of drug released at time t.


E-14 and E-261
To detect the presence of ibuprofen remaining in the hy-
OHO R
drogels after the release tests, IR spectroscopy was used. R = H or CH2CHOHCH2N+(CH3)3Cl- H
HO
Samples of hydrogels used for the loading/release were dried
in an oven at 40 ◦C until constant weight, and powdered in
H
HO
H OH
a mortar. IR spectra were recorded over the range of 400–
H
4000 cm−1, in a Bruker IFS 66 V FT-IR spectrometer H OH HO

(Germany), using the potassium bromide pellet technique. OH O OH O

The characteristic peaks of ibuprofen appear at 2956 cm−1, HO O


HO O
benzene ring, and 1584 and 1409 cm−1, ionized carboxylic H H H H z

H H H H
groups (Higgins et al., 2001).

were, in general, homogeneous. Transmittance was in some cases even


greater than that of the polymer alone dispersions. Precipitation was
3. Results and discussion only observed for the PQ-10 dispersions containing 4–5% ibuprofen
(Fig. 2). The pH of the systems
The structure of the polymers is shown in Fig. 1. PQ-10
and PQ-4 are linear cationic hydroxyethyl celluloses in
which the ammonium groups are bonded to the hydroxy-
ethyl substituent (PQ-10) or directly grafted to the cellulose
backbone while the hydroxyethyl substituent remains un-
altered (PQ-4). On the other hand, cationic guar gums are
formed by mannose units linearly bonded that alternatively
have a lateral ramification of galactopiranose. Some po-
sition 6 hydroxyl groups of galactopiranose are bounded to
a hydroxypropyltrimethylammonium group. In addition to
these general differences, the polymers also differ in
molecular weight and amount of cationic groups as reported
previously (Table 1) (Rodr´ıguez et al., 2001).

3.1. Aqueous dispersions

3.1.1. Study of the interactions in aqueous dispersions


Aqueous dispersions of cationic celluloses and cationic
guar gums prepared with several ibuprofen concentrations
Fig. Rodr´ıguez
434 1. Molecular structures ofR. the et al.celluloses
cationic / European Journal of Pharmaceutical Sciences 20 (2003) 429–438
(polyquaternium-4 and polyquaternium–10), the cationic guar
gums (Ecopol® 261-S and 14-S differ in the content in the
cationic groups), and ibuprofen.

increased linearly from around 7 (for PQ-10 and


PQ-4) or 6 (E-14 and E-261) to 8.5 when
ibuprofen concentration changed from 0 to 8%;
the ibuprofen–cationic polysaccha- ride systems
showing the same pH values as the ibuprofen
sodium solutions. The conductivity of the systems
also in- creased from 650–800 to 10,000–12,000
µS/cm with drug concentration, not showing any
inflection points.
The surface activity of ibuprofen was clearly
seen when the evolution of pyrene fluorescence
and surface tension

Table 1
Molecular weight (MW), intrinsic viscosity in water at 25 ◦C
([η]), and characteristics of substitution of the cationic
celluloses and cationic guar gums (Rodr´ıguez et al., 2001)
Polymer MWa [η] (dl/g) MSHECb %N Ammonium
group/sugar
repeating unit
PQ-4 1400000 14.59 2.71 1.27 1:5.3
PQ-10 1700000 70.09 1.28 1.95 1:2.9
E-14 500000 46.78 – 1.16 1:6.5
E-261 200000 45.85 – 0.91 1:8.5
a Dataprovided by the supplier.
b MSHEC average number of hydroxyethyl substituents =
bonded to each glucopiranose unit.
R. Rodr´ıguez et al. / European Journal of Pharmaceutical Sciences 20 (2003) 429–438 435

120 1000 70

Surface tension (mN/m)


100 800 60
Transmittance (%)

Viscosity (mPa·s)
80 600 50

60 400
40
40 200
30
20 0 P

210 6000 70
E-261
180

Surface tension (mN/m)


5000 60
Transmittance (%)

Viscosity (mPa·s)
150 4000
50
120 3000

2000 40
90
1000
30
60
0
012345678 012345678 012345678
Ibuprofen (%)

Fig. 2. Transmittance, viscosity and surface tension of aqueous dispersions of cationic polysaccharides (1%) containing different proportions of ibuprofen (P:
precipitation of PQ-10 dispersions).

measurements as a function of drug concentration were that of the polymer dispersions in the absence of the drug
recorded. With both techniques, the cmc obtained for (Fig. 2). The effect was less pronounced for PQ-4 systems
ibuprofen in water, 4%, was similar to the value found by that showed, as the E-14 dispersions, a significant increase
Ridell et al. (1999). The minimum observed can be at- in viscosity for lower drug concentrations. In general, the
tributed to some surface-active impurities in the drug. The values of G± and G±± followed a tendency similar to that of
micropolarity of the medium was measured by the changes viscosity and continuously decreased until the drug con-
in the I1/I3 ratio in the emission spectra of pyrene. The I1/I3 centration reached the cmc (4%). An additional increase in
ratio is higher in a polar medium (1.87 in water) than in an drug concentration induced the dispersions to recover their
apolar medium (0.66 in cyclohexane) (Anthony et al., 1998). initial properties, showing G± and G±± values similar to or
Without ibuprofen, cationic polysaccharides showed a I1/I3 even greater than those of polymer alone dispersions (Figs.
ratio around 1.2. In the presence of 5% ibuprofen, this value 3 and 4).
decreased to 0.9, in PQ-4 dispersions, and to 0.70–0.75, in These observations may be explained taking into account
the other cationic polysaccharide systems. The greater I1/I3 the electrostatic and hydrophobic interactions that may be
ratio of PQ-4/ibuprofen dispersions is ex- plained by the established between the drug and the polymers. Even for the
more hydrophilic character of this polymer, which have lowest drug concentration studied (0.5%), the molar ra- tio
more hydroxyethyl groups. The surface tension of the of carboxylic drug groups/ ammonium polymer groups was
ibuprofen–cationic polysaccharide dispersions de- creased, greater than 1; this means that in all systems there are
from 50–60 mN/m, in the presence of the drug to reach a enough ibuprofen molecules to ionically neutralize all
minimum, 30 mN/m, around 4% ibuprofen (Fig. 2). No cationic groups of the polymer. Therefore, although be- low
significant differences were observed in the surface ten- sion cmc the interaction seems to be scarcely significant, some
patterns with respect to that obtained in the absence of drug molecules may act as bridges between different
polymer, which indicates that none of these cationic polymer chains, favoring the formation of a three dimen-
polysaccharides significantly interferes in the micellization sional network of slightly greater consistency. Considering
process of the drug and that the main interactions may hap- the viscosity and transmittance data, the critical aggrega-
pen for drug concentrations around or above cmc. These tion concentration, cac, may be established to be around 2%
observations were confirmed by the changes that occurred in ibuprofen. When the cmc is reached, the drug molecules self-
the transmittance and rheological properties of the dis- associate causing the polymer chains to coil to wrap the
persions. The viscosity of the systems containing ibuprofen micelles. This folding favors intramacromolecular bonds
at a concentration around cmc was considerably lower than and, in consequence, produces a decrease in the consistency
436 R. Rodr´ıguez et al. / European Journal of Pharmaceutical Sciences 20 (2003) 429–438

101

100

10-1
G´´ (Pa)

10-2

10-3

10-4

10-5

101

100
G´ (Pa)

10-1

10-2

10-3

10-4

10-5
0.01 0.1 1 10 100
Angular frequency (rad/s)

Fig. 3. Effect of ibuprofen concentration on the viscoelastic behavior of PQ-


10 dispersions. Fig. 5. Schematic drawing of the effect of ibuprofen on the conformation of
cationic polysaccharides, as the drug concentration increases. In all systems
studied, the molar ratio between anionic groups of ibuprofen and cationic
groups of the polymer was above 1. When the cmc of ibuprofen is reached,
the polymer chains coil around the micelles owing to ionic and hydrophobic
interactions. This causes a strong decrease in the consistency of the system.
Increasing the number of micelles, the polymer chains uncoil and the
system recovers its initial properties.
1
of the systems and, in the case of PQ-10, even precipitation
G´ and G´´ (Pa)

(Fig. 5). A similar effect was described by Hugerth (2001) in


0.1 dispersions of dextran sulfate or carrageenan with amitripty-
line. Above cmc, the presence of more micelles makes the
0.01 aggregates and the hydrophobic regions of the polymers
more soluble. The mixed micelles become the connecting
points for transient networks (Hoffman et al., 1996) and,
0.001 therefore, the viscosity and transparency of the dispersions
rise again.
Compared to common anionic surfactants, such as SDS,
100 in which the cac with cationic polysaccharides is one to two
orders of magnitude lower than the cmc (Anthony et al.,
G´ and G´´ (Pa)

1998; Rodr´ıguez et al., 2003b), ibuprofen sorption is less


10
favored and begins nearer the cmc. The binding of ibupro-
fen molecules among themselves is preferred to the bind- ing
1 to the polymer. Similar results were reported by Ridell et al.
(1999) when the association process of this drug with two
non-ionic cellulose ethers was analyzed. The different
0.1 affinity and effects that SDS and ibuprofen induce in the
012345678 properties of the cationic polysaccharides may be related to
Ibuprofen (%) important structural differences between both amphiphilic
Fig. 4. Storage (G±, solid symbols) and loss (G±±, open symbols) moduli of
molecules. SDS has a greater surface activity and, therefore,
cationic polysaccharide dispersions (1%) containing different proportions a higher tendency to hydrophobically sorb onto the poly-
of ibuprofen (angular frequency 0.9 rad/s). mer is expected. Additionally, its sorption is facilitated by
R. Rodr´ıguez et al. / European Journal of Pharmaceutical Sciences 20 (2003) 429–438 437

the relatively flexible apolar chain, compared to the more Table 2


rigid benzene ring of ibuprofen (Bustamante et al., 2000). Apparent viscosities and ibuprofen diffusion coefficients obtained for cationic
polysaccharide (1%)/SDS dispersions
Finally, the sulfate group has a more acidic character prone
to interact ionically. Therefore, in the case of SDS, sorp- tion Polymer SDS (%) Viscosity (mPa s) D (cm2/min)
begins at concentrations much lower than cmc form- ing PQ-4 0 41.01 (0.26) 34.1 × 10−4 (4.5 × 10−4)
hemimicellar structures with the polymer. In contrast, 0.25 36.81 (0.21) 24.9 × 10−4 (3.5 × 10−4)
ibuprofen first forms micelles into which or around which 0.30 44.01 (0.66) 19.5 × 10−4 (1.2 × 10−4)
the polymer chains distribute owing to electrostatic and hy- PQ-10 0 316.2 (0.91) 11.9 × 10−4 (2.9 × 10−4)
drophobic interactions (Fig. 5). Formation of aggregates be- 0.05 156.3 (0.42) 4.8 × 10−4 (0.6 × 10−4)
tween cationic polysaccharides and ibuprofen micelles is 0.10 147.7 (1.89) 6.3 × 10−4 (0.5 × 10−4)
more favored as the polymer has more hydrophobic residues E-14 0 589.4 (26.2) 5.9 × 10−4 (0.5 × 10−4)
and more cationic groups; i.e., cationic cellulose PQ-10 and 0.01 407.6 (2.47) 5.2 × 10−4 (0.3 × 10−4)
cationic guar gum E-14. This explains that the least relevant 0.05 285.9 (3.69) 6.0 × 10−4 (0.8 × 10−4)
changes in viscosity were observed for PQ-4 dispersions, E-261 0 2129.9 (3.3) 5.3 × 10−4 (0.7 × 10−4)
which is a polymer with the greatest content in hydrophilic 0.01 1029.3 (5.1) 5.2 × 10−4 (0.4 × 10−4)
hydroxyethyl groups. The fact that the hydroxyethyl and the 0.05 955.3 (11.5) 6.5 × 10−4 (0.1 × 10−4)
ammonium groups are separated or bound together justifies Means of three replicates (S.D.). Ibuprofen diffusion coefficient in water was
the different behavior shown by the systems based on the 180 × 10−4 (2.2 × 10−4) cm2/min.
two cationic celluloses when they interact with ibuprofen or
SDS. The presence of the free hydroxyethyl groups in PQ-4
provides enough hydrophilicity to avoid the precipitation of Edsman, 2001). In the case of cationic guar gums, the pro-
the aggregates, even when PQ-4 charges were neutral- ized portions of SDS that can be added without causing phase
and the coiled of the polymer occurred. In PQ-10, the separation are relatively small, and, in consequence, the ef-
ammonium groups are directly bonded to the hydroxyethyl fect on ibuprofen diffusion is scarce.
substituents, which may be included in the aggregates dur-
ing association with the amphiphilic molecules. This leaves 3.2. Chemically cross-linked hydrogels
the more hydrophobic unsubstituted cellulose chain exposed
to the water interface, promoting precipitation (Rodr´ıguez et 3.2.1. Drug loading
al., 2003a). For some applications, it may be interesting to keep the
viscosity and integrity of the polymer-based system
3.1.2. Ibuprofen diffusion studies constant, using the amphiphilic drug–polymer interactions
To carry out the diffusion studies, the dispersions con- to regulate the release process. In that case, chemically
taining 6% ibuprofen were selected since they combined a cross-linked hydrogels obtained by reacting the cationic
therapeutically convenient drug dose (Higgins et al., 2001) polysaccharides with ethyleneglycol diglycidylether may be
with adequate rheological properties. Suitable proportions of useful (Rodr´ıguez et al., 2003b). These hydrogels are
SDS were added to some dispersions in order to lead to the translucent, with a homogeneous surface and high consis-
maximum interaction with the cationic polysaccharides (i.e., tency and elasticity. The volume of the hydrogels when fully
those that caused the greatest increases in viscosity in the swollen in water was several times their volume as freshly
polymer alone dispersions) (Rodr´ıguez et al., 2003a) with the prepared: 7.4 for PQ-4, 9.9 for PQ-10, 3.9 for E-261, and 3.4
aim of establishing the possibilities that the addi- tion of this for E-14. These values corresponded to a hydra- tion level of
surfactant offers to modulate the properties of the systems. 70, 250, 114, and 66 g water/g dry hydrogel, respectively.
All formulations showed sustained release for more than 10 This high water affinity makes the hydrogels potentially
h and the Higuchi equation fitted the diffusion profiles well useful as main components of biocompatible systems with a
(r2 > 0.98); the diffusion coefficients being particularly low high permeability to water and small solutes (Capitan et al.,
in cationic guar gums dispersions owing to their thickening 2000).
capability (Table 2). It is interesting to note that, in the Discs of hydrogels immersed in sodium ibuprofen solu-
presence of ibuprofen, SDS did not cause a significant tions of several concentrations (1–8%) showed a decrease in
increase in the viscosity of the polymer disper- sions, but a their volume of 40%, which is related to neutralization of
decrease. Despite this effect, the diffusion coef- ficients of some ionic charges and to an increase in the ionic strength of
ibuprofen were significantly lower, especially in cationic the medium. The amount of drug sorbed linearly increased
cellulose systems, in the presence of SDS. There- fore, with the concentration of ibuprofen in the outer solution
ibuprofen may remove SDS from the binding sites of the (Fig. 6). The isotherm patterns were similar to the G-type
polymer leading to the formation of new ionic interac- tions model (Duro et al., 1999). This means that the interactions
with the polymer, hydrophobic interactions with SDS, or between the polymer and the drug are relatively weak. Nev-
mixed micelles of ibuprofen/SDS/polymer, making the ertheless, the hydrogels were able to take up a considerably
diffusion through the dispersion more difficult (Paulsson and high amount of drug, up to 15 mg/mg dry hydrogel, except
438 R. Rodr´ıguez et al. / European Journal of Pharmaceutical Sciences 20 (2003) 429–438

16 60
PQ-10
14 PQ-4

Ibuprofen released (%)


50
12
40
10
30
8
6 20
Ibuprofen loaded (g /g dry gel)

4 10
2
0
0 0 60 120 180 240
(A) Time (min)
16
E-14
14 E-261

12
10
8
6
4
2

0
3600 3000 2400 1800 1200 600
0 1 2 3 4 5 6 7 8
(B) Wavenumbers (cm-1)
Ibuprofen (%)

Fig. 7. (A) Ibuprofen release profiles in water from PQ-10 hydrogels loaded
Fig. 6. Ibuprofen sorption isotherms on chemically cross-linked cationic
celluloses and cationic guar gums. in 6% (O), 7% (☐ ), and 8% ( ) ibuprofenO solutions; and (B) from top to
the bottom, FT-IR spectra of freshly prepared PQ-10 hydrogel, of PQ-
10 hydrogel after ibuprofen loading in 6% solution and release in water,
for E-14 hydrogels in which drug sorption was limited to and of pure ibuprofen sodium.
8 mg/mg.
ions. This explains that when the hydrogels were transferred
3.2.2. Drug release
to pH 8 phosphate buffer, all ibuprofen was released. This
Ibuprofen release profiles in water from the hydrogels
allowed to confirm the total amount of ibuprofen initially
loaded in 6, 7, or 8% ibuprofen solutions are shown in Fig.
loaded by the hydrogels.
7A. The release process reached equilibrium after 2 h, when
Fig. 8 shows the release profiles of hydrogels subse-
a 40–50% of the amount of ibuprofen loaded (60, 72, or 80
mg per disc, respectively) still remained in the hydro- gels, quently immersed in media of increasing pH, simulating the
which did not experiment significant changes in volume in vivo situation after oral administration. At pH 3 (diluted
during the process. The release occurred by Fickian diffu- HCl aqueous solution), the hydrogels just released a low
sion (Eq. (4), n = ± r > 0.98). The permanence of the
0.46 0.5, 2

drug was confirmed by IR spectroscopy (Fig. 7B). After 8 h 100


in water, the hydrogels were dried and the characteris- tic pH 3 pH 6 pH 8
bands of ibuprofen (2956 cm−1 benzene ring; 1584 and 1409
Ibuprofen released (%)

80
cm−1 ionized carboxylic groups) were still clearly vis- ible.
Also, the bands corresponding to the carboxylic groups of 60
ibuprofen were slightly moved to lower wave numbers,
suggesting ionic interactions with the polymer. Therefore,
40
only the unbound or weakly bound drug molecules are re-
leased from the hydrogel. The remaining drug molecules are
ionically trapped by the ammonium groups of the polymer 20
network, similarly to that observed by Jimenez-Kairuz et al.
(2002) for lidocaine in carbopol gels. Note that ibuprofen 0
0 60 120 180 240 300 360 420 480 540 600
concentration is, in all cases, clearly below its solubility co-
efficient in water (Cs=20%; Bustamante et al., 2000). An Time (min)
increase in pH and the presence of more ions in the medium Fig. 8. Effect of pH on ibuprofen release rate from PQ-10 (O), PQ-4 (●),
facilitate the diffusion of all ibuprofen, by acting as counter- E-14 (☐ ), and E-261 (■ ).
R. Rodr´ıguez et al. / European Journal of Pharmaceutical Sciences 20 (2003) 429–438 439

proportion of ibuprofen. The lower solubility of the drug in Acknowledgements


this media may contribute to hinder the release process
directly and, indirectly, by promoting hydrophobic interac- This work was financed by the Xunta de Galicia (PGIDT
tions with the polymer network. When they were transferred 00PX120303PR) and the Ministerio de Ciencia y Tec-
to water, there was no release at all; the hydrogels behaved nolog´ıa, Spain (RYC2001-8). The authors also express their
differently from when they were directly immersed in wa- gratitude to the Xunta de Galicia for an equipment grant
ter. Finally, when the pH increased to 8, the release process (DOG 04/06/97) and to National Starch and Chemical Ltd.
began again and finished after 6 h. The release exponents, n, for providing free samples of cationic celluloses.
in pH 8 phosphate buffer were significantly lower than in
water, especially for hydrogels based on the polymers with a
greater content in cationic groups (0.45 for PQ-4, 0.27 for References
PQ-10, 0.29 for E-14, 0.40 for E-261). This finding sug-
gests that, at least phenomenologically, the release process Alvarez-Lorenzo, C., Concheiro, A., 2003. Effects of surfactants on gel
behavior. Design implications for drug delivery systems. Am. J. Drug
is different in both media. Similar results were observed for
Deliv. 1, 77–101.
another NSAID molecule, diclofenac (Rodr´ıguez et al., Annon, 1988. Final report on the safety assessment of polyquaternium-10.
2003b). At pH below the pKa of ibuprofen (4.5; Bustamante J. Am. Coll. Toxicol. 7, 335–351.
et al., 2000), in addition to some ionic interactions, hy- Anthony, O., Marques, C.M., Richetti, P., 1998. Bulk and surface behav-
drophobic interactions are strongly promoted. This stronger ior of cationic guars in solutions of oppositely charged surfactants.
Langmuir 14, 6086–6095.
association is maintained in water and broken when the pH
Bustamante, P., Peña, M.A., Barra, J., 2000. The modified extended Hansen
and ion concentration rise. meted to determine partial solubility parameters of drugs con- taining a
single hydrogen bonding group and their sodium derivatives: benzoic
acid/Na and ibuprofen/Na. Int. J. Pharm. 194, 117–124.
4. Conclusions Capitan, D., Nobile, M.A., Mensitieri, G., Sannino, A., Segre, A.L., 2000.
13C solid state NMR determination of cross-linking degree in

superabsorbing cellulose-based networks. Macromolecules 33, 430–


Cationic cellulose ethers and cationic guar gums interact 437.
hydrophobically and ionically with ibuprofen. The associa- Chronakis, I.S., Alexandridis, P., 2001. Rheological properties of op-
tion phenomena are directly related to the self-aggregation positely charged polyelectrolyte–surfactant mixtures. Effect of poly-
of the drug molecules; dramatically increasing the intensity mer molecular weight and surfactant architecture. Macromolecules 34,
5005–5018.
of the interactions above the cmc. This behavior notably dif-
Dominkus, M., Nicolakis, M., Kotz, R., Wilkinson, F.E., Kaiser, R.R.,
fers from that previously reported for the association pro- Chlud, K., 1996. Comparison of tissue and plasma levels of ibuprofen
cesses with common anionic surfactants, in which the main after oral and topical administration. Arzeneimittel.-Forschung./Drug
changes in the properties of cationic polysaccharide systems Res. 46, 1138–1143.
occur when the charges of the polymer are neutralized by the Duro, R., Souto, C., Gómez-Amoza, J.L., Mart´ınez-Pacheco, R.,
Concheiro, A., 1999. Interfacial adsorption of polymers and surfac-
surfactant. Ibuprofen-polymer interactions cause impor- tant
tants: implications for the properties of disperse systems of pharma-
modifications in the rheological properties of the poly- mer ceutical interest. Drug Dev. Ind. Pharm. 25, 817–829.
dispersions and considerably delay the drug diffusion rate. Florence, A.T., Attwood, D., 1998. Physicochemical Principles of Phar-
Moreover, the addition to the gel of small amounts of SDS, macy, third ed. Macmillan, London, pp. 199–251.
which causes a slight decrease in viscosity, provokes a González-Rodr´ıguez, M.L., Holgado, M.A., Sanchez-Lafuente, C.,
Rabasco, A.M., Fini, A., 2002. Alginate/chitosan particulate systems
reduction in the values of the diffusion coefficients of al-
for sodium diclofenac release. Int. J. Pharm. 232, 225–234.
most 50%. In the presence of the surfactant, new bonds are Goddard, E.D., Hannan, R.B., 1977. Polymer/surfactant interactions. J.
expected between the drug and the surfactant and be- tween Am. Oil Chem. Soc. 54, 561–566.
the mixed micelles of polymer/drug/surfactant. The Gruber, J., Kreeger, R., 1996. Cellulose ethers, cationic. In: Polymeric
information provided by these studies was particularly use- Materials Encyclopedia, vol. 2. CRC Press, New York, pp. 1113– 1118
Higgins, J.D., Gilmor, T.P., Martelucci, S.A., Bruce, R.D., 2001. Ibuprofen.
ful in the design of chemically cross-linked hydrogels able In: Analytical Profiles of Drug Substances and Excipients, vol. 27.
to control the release process of ibuprofen as a function of Academic Press, San Diego, pp. 265–300
the pH of the medium. At pH 3, hydrophobic and ionic bonds Higuchi, W.I., 1962. Analysis of data on the medicaments release from
are strongly promoted and the drug is not released. The ointments. J. Pharm. Sci. 51, 802–804.
association was partially broken in water and totally Hoffman, H., Kästner, U., Donges, R., Ehrler, R., 1996. Gels from mod-
ified hydroxyethylcellulose and ionic surfactants. Polym. Gels Netw. 4,
vanished in pH 8 phosphate buffer. In summary, the inter- 509–526.
action of cationic polysaccharides with anionic amphiphilic Hugerth, A.M., 2001. Micropolarity and microviscosity of amitripty- line
molecules has considerable consequences on the consistency and dextran sulfate/carrageenan-amitriptyline systems: the nature of
and diffusional properties of their aqueous dispersions and polyelectrolyte–drug complexes. J. Pharm. Sci. 90, 1665– 1677.
chemically cross-linked hydrogels, and highlights the need Jimenez-Kairuz, A.F., Allemandi, D.A., Manzo, R.H., 2002. Mechanism of
lidocaine release from carbomer-lidocaine hydrogels. J. Pharm. Sci. 91,
for a profound characterization of the associative phenom- 267–272.
ena as a first step in the development of drug delivery sys-
tems with these components.

You might also like