Вы находитесь на странице: 1из 7

International Journal of Pharmaceutics 481 (2015) 97–103

Contents lists available at ScienceDirect

International Journal of Pharmaceutics


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

Pharmaceutical nanotechnology

Transdermal delivery of curcumin via microemulsion


Amnon C. Sintov *
Department of Biomedical Engineering, Faculty of Engineering Sciences, Ben Gurion University of the Negev, Be'er Sheva 84105, Israel

A R T I C L E I N F O A B S T R A C T

Article history: The objective of this study was to evaluate the transdermal delivery potential of a new curcumin-
Received 5 November 2014 containing microemulsion system. Three series of experiments were carried out to comprehend the
Accepted 1 February 2015 system characteristics: (a) examining the influence of water content on curcumin permeation, (b)
Available online 2 February 2015
studying the effect of curcumin loading on its permeability, and (c) assessing the contribution of the
vesicular nature of the microemulsion on permeability. The skin permeability of curcumin from
Keywords: microemulsions, which contained 5%, 10%, and 20% of water content (1% curcumin), was measured in vitro
Microemulsion
using excised rat skin. It has been shown that the permeability coefficient of CUR in a formulation
Curcumin
Transdermal drug delivery
containing 10% aqueous phase (ME-10) was twofold higher than the values obtained for formulations
Percutaneous penetration with 5% and 20% water (Papp = 0.116  103  0.052  103 vs. 0.043  103  0.022  103 and
0.047  103  0.025  103 cm/h, respectively. A reasonable explanation for this phenomenon may be
the reduction of both droplet size and droplets’ concentration in the microemulsion as the aqueous phase
decreased from 20% to 5%. It has also been shown that a linear correlation exists between the decrease in
droplet size and the increase of curcumin loading in the microemulsion. In addition, it has been
demonstrated that a micellar system, S/O-mix, and a plain solution of curcumin resulted in a significantly
lower curcumin permeation relative to that presented by the microemulsion, Papp = 0.018  103  0.011
 103, 0.005  103  0.002  103, and 0.002  103  0.000  103 cm/h, respectively, vs. 0.110  103
 0.021 103 cm/h for the microemulsion. The enhancement ratio (ER = Jss-ME/Jss-solution) of CUR
permeated via 1% loaded microemulsion was 55.
ã 2015 Elsevier B.V. All rights reserved.

1. Introduction absorption and rapid metabolism, CUR is not chemically stable in


physiological solutions. As reported by Wang et al. (1997), about
Curcumin (CUR; diferuloyl methane), a component of the 90% of CUR decomposed within 90 min when incubated in serum-
golden spice turmeric (Curcuma longa; family-Zingiberaceae), has free phosphate buffer at pH 7.2. In more recent studies, the most
a variety of pharmacological activities. It has, just to name a few, an prevalent product of CUR was identified as the one formed by auto-
anti-inflammatory effect in Crohn’s disease, ulcerative colitis, oxidative transformation, bicyclopentadione, and since it is also a
rheumatoid arthritis, peptic and gastric ulcer, psoriasis, and metabolic product catalyzed by lipoxygenase and cyclooxygenase
vitiligo, as well as anti-tumor, anti-diabetic, anti-oxidation, anti- it has been proposed as a possible mediator of the biological
microbial and anti-viral actions (Anand et al., 2008; Gupta et al., activity related to CUR (Griesser et al., 2011; Gordon and Schneider,
2013). However, due to its low water solubility (maximum 11 ng/ 2012). Both high-rate chemical instability and the low solubility of
ml in phosphate buffer, pH 5), CUR is slightly absorbed into the CUR are a major problem that limits its use in liquid formulations
body with a very low bioavailability (Yang et al., 2007; Tønnesen and makes difficult for researchers to accomplish either in-vitro
et al., 2002; Tomren et al., 2007). Rapid metabolism and excretion evaluation or in vivo bioavailability studies. Several approaches has
(t1/2 = 28.2 and 44.5 min after i.v. and oral curcumin in rats, been developed to improve CUR bioavailability and stability by
respectively) also curtails CUR bioavailability (Yang et al., 2007; loading the agent into liposomes (Li et al., 2005, 2007; Kunwar
Anand et al., 2007; Metzler et al., 2013). In addition to its poor et al., 2006), nanoparticles (Bisht et al., 2007, 2010; Lim et al., 2011;
Chun et al., 2012), and microemulsions (Cui et al., 2009; Liu et al.,
2011; Liu and Chang, 2011; Hu et al., 2012). Further approaches
include derivatives and analogs (Mosley et al., 2007), and
* Correspondence at: Laboratory for Biopharmaceutics, E.D. Bergmann Campus, phospholipid complexes (Liu et al., 2006; Maiti et al., 2007).
Ben-Gurion University of the Negev, P.O. Box 653, Be'er Sheva 84105, Israel.
Tel.: +972 8 647 2709.
The vast majority of published studies aimed at improving
E-mail address: asintov@bgu.ac.il (A.C. Sintov). curcumin bioavailability have been focused on oral or injectable

http://dx.doi.org/10.1016/j.ijpharm.2015.02.005
0378-5173/ ã 2015 Elsevier B.V. All rights reserved.
98 A.C. Sintov / International Journal of Pharmaceutics 481 (2015) 97–103

delivery systems while transdermal delivery of curcumin has 1:3 glyceryl oleate/Labrasol1. The microemulsion formed sponta-
received less attention so far. Patel et al. (2009) have developed a neously at room temperature as a clear monophasic liquid. The
dried polymeric film-type matrix containing curcumin, which its loaded concentration of CUR in the microemulsions was by default
in-vitro permeation followed Higuchi kinetics. In a series of three 1.0 wt% but can be increased to 4.0 wt%. As a convenient method to
publications dealing with microemulsion vehicles, the percutane- represent an increase of water content while decreasing CoS–S
ous delivery of 0.4% curcumin has been studied in terpene levels, ‘water dilution lines’ were drawn from the water apex to the
microemulsions, such as eucalyptol and limonene combined with opposite oil side of the phase diagram. The line was arbitrarily
polysorbate 80, ethanol and water (Liu and Chang, 2011; Liu et al., denoted as the value of the line intersection with the oil scale (e.g.,
2011), and in microemulsions composed of Pluronic F127 DL11 means line representing a surfactant-to-oil ratio of 89:11).
(surfactant), isopropanol (co-surfactant), myristic acid (oil), and Formulations along a dilution line DL11 was prepared to contain
water (Liu and Huang, 2012). Although microemulsions have not 5%, 10%, and 20% (w/w) of aqueous phase. The stability of CUR in
been widely explored for transdermal curcumin they are 10% water-containing microemulsion was tested at 4  C and at
extensively utilized for a variety of pharmaceutical purposes ambient temperature, resulting in decrease of 5.9% and 8.9%,
including transdermal and dermal drug delivery (Kreilgaard et al., respectively, after 7 days. However, no significant decrease in CUR
2000; Alvarez-Figueroa and Blanco-Mendez, 2001; Kreilgaard, level has been noted when including 0.01% a-tocopherol in the
2001; Kreilgaard et al., 2001; Rhee et al., 2001; Lee et al., 2003; formulation.
Sintov and Shapiro, 2004; Hua et al., 2004; Sintov and Botner,
2006; Sintov and Brandys-Sitton, 2006; Liu et al., 2007; Shevach- 2.3. Light scattering
man et al., 2008; Heuschkel et al., 2008; Gannu et al., 2010; Zhao
et al., 2011; Zhang et al., 2011; Zhang and Michniak-Kohn, 2011; The hydrodynamic diameter spectrum of microemulsion nano-
Sintov and Greenberg, 2014). Microemulsions, which are versatile droplets was collected using CGS-3Compact Goniometer System
systems (oil-in-water, water-in-oil, or bi-continuous systems as (ALV GmbH, Langen, Germany). The laser power was 20 mW at the
well as various structures and components), possess high He–Ne laser line (633 nm). Correlograms were calculated by ALV/
solubilizing capacity of diverse compounds and, in particular, is LSE 5003 correlator during 10 s for 20 times at 25  C. Measure-
capable of solubilizing poorly water-soluble drugs. The increase in ments were performed at permanent angle of 60 . The droplet size
solute concentration and the tendency of the vehicle to favor was calculated using the Stokes–Einstein relationship, and the
partitioning into the stratum corneum make microemulsions a analysis was based on regularization method as described by
useful vehicle to enhance transdermal drug permeability. Provencher (1982).
In the present report, we describe an alcohol-free micro-
emulsion system for transdermal drug delivery with respect to its 2.4. Viscosity
contribution to percutaneous permeability of CUR. The micro-
emulsion system, which contains no chemical enhancers, is The viscosity was measured by using a control stress rheometer
composed of water, isopropyl palmitate, propylene carbonate, CSL 50 (TA Instruments Ltd., UK) at 25  C. Cone and plate geometry
and the commonly-used topical nonionic surfactants, PEG- was used.
8 caprylic/capric glycerides (Labrasol1) and glyceryl oleate. We
studied the in vitro skin permeation of the microemulsions after we 2.5. In-vitro skin penetration study
had surmounted the technical difficulties of experimentation
regarding the high-rate of CUR degradability in aqueous solutions. The permeability of CUR through animal skin was determined
By using microemulsions as a vehicle, we have demonstrated an in vitro with a Franz diffusion cell system (Crown Bioscientific, Inc.,
increased solubilization as well as a significantly enhanced Clinton, NJ). The diffusion area was 1.767 cm2 (15 mm diameter
permeability of the water insoluble and poorly absorbable orifice), and the receptor compartment volumes was from 12 ml.
curcumin. The solutions in the water-jacketed cells were thermostated at
37  C and stirred by externally driven, Teflon-coated magnetic
2. Materials and methods bars. Each set of experiments was performed with at least four
diffusion cells (n  4), each containing skin from a different animal.
2.1. Materials All animal procedures were performed in accordance with
protocols reviewed and approved by the Institutional & Use
Curcumin was obtained from Sigma, Rehovot, Israel. Labrasol1 Committee, Ben Gurion University of the Negev, which complies
(PEG-8 caprylic/capric glycerides) was purchased from Gattefosse with the Israeli Law of Human Care and Use of Laboratory Animals.
(Saint-Priest Cedex, France), and glyceryl oleate was obtained from Sprague-Dawley rats (males, 300–350 g, Harlan Laboratories Ltd.,
Croda Europe Ltd. (East Yorkshire, England). Isopropyl palmitate Jerusalem, Israel) were euthanized by aspiration of CO2. The
and propylene carbonate were purchased from Sigma, Rehovot, abdominal hair was clipped carefully and sections of full-thickness
Israel. a-Tocopherol 1000 IU was obtained from BTSA Biotecno- skin were excised from the fresh carcasses of animals and used
logias Aplicadas (Madrid, Spain). High-performance liquid chro- immediately. All skin sections were measured for transepidermal
matography (HPLC) grade water and solvents were obtained from J. water loss (TEWL) before mounted in the diffusion cells or stored
T. Baker (Mallinckrodt Baker, Inc., Phillipsburg, NJ). at lower temperatures until used. TEWL examinations were
performed on skin pieces using Dermalab1 Cortex Technology
2.2. Preparation of microemulsions instrument, (Hadsund, Denemark) and only those pieces that the
TEWL levels were less than 10 g/m2/h were introduced for testing.
The formulation was prepared as previously published (Sintov The skin was placed on the receiver chambers with the stratum
and Botner, 2006; Sintov et al., 2010; Sintov and Greenberg, 2014). corneum facing upwards, and the donor chambers were then
Shortly, water-in-oil (W/O) liquid microemulsions were prepared clamped in place. The excess skin was trimmed off, and the
by mixing Labrasol1 and glyceryl oleate (surfactants), propylene receiver chamber, defined as the side facing the dermis, was filled
carbonate (co-surfactant) and isopropyl palmitate (oil), followed with 70:30 phosphate buffer (pH 7.4) – ethyl alcohol containing
by gently titration of distilled water. The co-surfactant/surfactants a-tocopherol (0.01%). After 15 min of skin washing at 37  C, the
(CoS/S) weight ratio was 2:5, while the surfactants’ ratio was buffer was removed from the cells and the receiver chambers were
A.C. Sintov / International Journal of Pharmaceutics 481 (2015) 97–103 99

refilled with fresh phosphate buffer–ethanol solution. Aliquots X


t1
Qt ¼ Vr Ct þ Vs Ci
(0.5 g each) of CUR-containing microemulsions or surfactant i¼0
mixtures were applied on the skin at time = 0. Samples (2 ml)
where Ct is CUR concentration of the receiver solution at each
were withdrawn from the receiver solution at predetermined time
sampling timet, expressed by a running number (t = 1, 2, 3, ...tn). Ci
intervals, and the cells were replenished up to their marked
is CUR concentration of the ith sample, and Vr and Vs are the
volumes with fresh buffer–ethanol solution each time. Addition of
volumes of the receiver solution and the sample, respectively. Data
buffer solution to the receiver compartment was performed with
were expressed as the cumulative curcumin permeation per unit of
great care to avoid trapping air beneath the dermis. After the 24-h
skin surface area, Qt/S (S = 1.767 cm2). The steady-state fluxes (Jss)
experimental period, each CUR-exposed skin tissue was washed
were calculated by linear regression interpolation of the experi-
carefully in distilled water, wiped and tape-stripped (15) to
mental data at a steady state:
remove the residues of curcumin adsorbed over the stratum
corneum. The tissue was then cut to small pieces and inserted into DQ t
Jss ¼
2-ml vials. The skin pieces in each vial were extracted by 1-ml ðDt  SÞ
ethyl alcohol. The extraction was performed by incubation in a
Apparent permeability coefficients (Papp) were calculated
shaker (750 rpm) for 30 min. The receiver samples and the skin
according to the equation:
extracts were taken into 1.5-ml vials and kept at 20  C until
analyzed by HPLC within two days. J ss
Papp ¼
Cd
2.6. HPLC analysis of samples from receiver solutions and skin extracts where Cd is CUR concentration in the donor compartment (1.0 wt%
or 1.0  104 mg/ml), and it assumed that under sink conditions the
Aliquots of 20 ml from each sample were injected into a HPLC concentration of CUR in the receiver compartment is negligible
system, equipped with a prepacked column (ReproSil-Pur compared to that in the donor compartment.
300 ODS-3, 5 mm, 250 mm  4.6 mm, Dr. Maisch, Germany), which
was constantly maintained at 30  C. The HPLC system (Shimadzu 2.8. Statistical analysis
VP series) consisted of an auto-injector and a diode array detector.
The quantification of CUR was carried out at 425 nm. The samples The statistical differences between the skin permeation curves
were chromatographed using an isocratic mobile phase consisting of the formulations were analyzed employing the two-way ANOVA
of acetonitrile 0.2% acetic acid solution (80:20) at a flow rate of test followed by Bonferroni’s multiple comparison t-test. The
1 ml/min. differences among group means were considered significant when
p values < 0.05. The statistical differences between mean concen-
2.7. Calculation of the in vitro data tration values of curcumin extracted from skin tissues were
assessed by using of Student’s t-test at p < 0.05. Data were
A calibration curve (peak area versus drug concentration) was expressed as mean values (SD).
constructed by running standard drug solutions in ethanol for each
series of chromatographed samples. As a result of the sampling of 3. Results and discussion
large volumes from the receiver solution (and the replacement of
these amounts with equal volumes of buffer), the receiver solution 3.1. Influence of water content
was constantly being diluted. Taking this process into account, the
cumulative drug permeation (Qt) was calculated from the The permeability of curcumin from microemulsions containing
following equation: 5%,10%, and 20% of water (dilution line DL11, CoS/S = 2:5,1% CUR) was

Fig. 1. In vitro percutaneous permeation of curcumin through fresh rat skin after application of curcumin-loaded microemulsions containing various water contents. A
comparison was also made between 5% (diamonds), 10% (squares), and 20% aqueous phase (triangles) (n = 4).
100 A.C. Sintov / International Journal of Pharmaceutics 481 (2015) 97–103

Table 1
Skin permeability parameters of various curcumin-loaded microemulsion (ME)
formulations as measured in vitro by using rat skin (n  4).

Formulation Papp  103 (cm h1) Q24 (mg cm2)


a
ME with 5% aqueous phase (ME-5) 0.043 (0.022) 6.53 (3.08)
ME with10% aqueous phasea (ME-10) 0.116 (0.052) 18.78 (8.68)
ME with 20% aqueous phasea (ME-20) 0.047 (0.025) 7.73 (4.37)
2% curcumin-loaded ME-10 0.233 (0.051) 39.10 (9.68)
3% curcumin-loaded ME-10 0.295 (0.050) 48.38 (8.39)
S/O-mix (surfactants–oil mixture)a 0.005 (0.002) 1.28 (0.49)
Micellar system (oil-free)a 0.018 (0.011) 3.40 (2.00)
1% curcumin in propylene carbonate 0.002 (0.000) 0.69 (0.06)
a
Loaded with 1% curcumin.

measured in vitro using excised rat skin. The cumulative percutane-


ous penetrating amounts of CUR from the three microemulsions are
illustrated in Fig. 1 and the permeation parameters are summarized
Fig. 3. Log–log plot of shear stress vs. shear rate and viscosity vs. shear rate showing
in Table 1. As seen, the cumulative permeating amount of CUR Newtonian fluid behavior related to 5% and 20% water-containing microemulsion,
following 24h-application (Q24) of a microemulsion containing 10% and a graph characteristic of a Bingham plastic that related to 10% water-containing
water fraction was about 3 times higher than those obtained by microemulsion.
application of microemulsions containing 5% or 20% aqueous
phases (Q24 = 18.78  8.68 vs. 6.53  3.08 and 7.73  4.37 mg/cm2,
size is not the only factor dictating the permeability of active
respectively; Student’s t-test, p < 0.05). The permeability coefficient
molecules and because a smaller fraction of inner phase in
of CUR in formulation containing 10% aqueous phase (ME-10) was
microemulsions naturally results in a lower density of the dispersed
twice higher than the values obtained for formulations with 5% and
droplets, it is reasonably conceived that the W/O formulation with
20% water (Papp = 0.116  103  0.052  103 vs. 0.043  103
5% water contains smaller number of droplets per unit volume than
 0.022  103 and 0.047  103  0.025  103 cm/h, respectively;
the formulation with 10% water. This conclusion was evidenced by
Student’s t-test, p < 0.05) (Table 1). The relatively high permeability
viscosity measurements as presented in Fig. 3. Unlike micro-
demonstrated by microemulsion ME-10 may be explained by the
emulsions containing relatively larger or smaller fractions of water
droplet size of the microemulsions. As measured by DLS (Fig. 2), the
phase, which have demonstrated a Newtonian fluid behavior, the
size of the nanodroplets in microemulsion containing 20% water
same microemulsion type with 10% water exhibited a typical
dropped from 5.95  0.21 nm (98% fraction by weight) to
Bingham plastic behavior. This type of fluid is characterized by a
2.53  0.90 nm (46% fraction) and 0.71  0.02 nm (48% fraction)
phenomenon, in which the shear stress (s ) must exceed a certain
when water content decreased to 10% and 5%, respectively. This
yield value, i.e., yield stress (s y), before the fluid deforms and flows
reduction in droplet size was accompanied by the appearance of an
as described by the following equation:
aggregate peak at approximately 150 nm diameter in both micro-
s
emulsions, however, approximately half of the nanodroplets were h ¼ h0 þ _y
much smaller in these formulations than in the microemulsion with g
20% water which contained almost no aggregates. Although when:
microemulsion with 5% water possessed the smallest droplet size,
s  sy
its CUR permeability was lower than that demonstrated by the h0 ¼
g_
microemulsion containing 10% water and was comparable to the
permeability of the microemulsion with 20% water, which where h is the apparent viscosity, h0 is the plastic viscosity, and g_ is
contained the largest droplet size. As it is obvious that the droplet the shear rate.

Fig. 2. Nanodroplet size (average diameter, nm) and their prevalence (percent of total droplets) in microemulsions containing 5%, 10%, and 20% aqueous phase.
A.C. Sintov / International Journal of Pharmaceutics 481 (2015) 97–103 101

The physical behavior of Bingham fluids with a yield stress is


usually explained in terms of an internal structure in three
dimensions which is capable of preventing deformation for values
of the shear stress less than the yield value. This internal structure
could be flocculated particles in concentrated disperse systems or
designed microemulsions that generate high frictional forces
between the moving nanodroplets, forces that contribute to the
yield value.

3.2. Skin curcumin levels

Fig. 4 presents the extent of curcumin accumulation in rat skin


at 7 and 24 h of microemulsion application. In accordance with the
corresponding percutaneous permeation fluxes, skin accumulation
of CUR was significantly higher after administration of 10% water-
containing microemulsion (ME-10) than after 5% water-containing Fig. 4. In vitro percutaneous penetration of curcumin to rat skin after application of
system (ME-5) at either 7 or 24 h. Microemulsion ME-10 resulted in 5% and 10% water-containing microemulsions. Skin retention of curcumin was
twofold higher CUR concentration in the skin after 7 h compared measured after 7 and 24 h microemulsion application (n = 6).
with microemulsion ME-5 (22.5  3.0 vs. 10.5  4.4 mg/g tissue),
and four times higher skin level of CUR after 24 h (331.3  38.4 vs. loaded microemulsion was 55. From this finding, it can be
83.5  11.8 mg/g tissue). summarized that: (a) the mechanism of CUR delivery through
the skin is based on the nature of microemulsion nanostructure
3.3. The effect of microemulsion nanostructure on curcumin rather than on chemical enhancement driven by its ingredients, (b)
permeability this is in agreement with our previous publications (Sintov and
Shapiro, 2004; Sintov and Greenberg, 2014), showing that micro-
Fig. 5 shows the transdermal permeation profile of CUR applied emulsions are superior to micelles and S/O-mixtures in permeating
in a microemulsion (ME-10, 1% CUR concentration) in comparison active substances through the skin, and (c) oil-free micelles
with permeation of CUR in oil-free micellar system, in a mixture of delivered more CUR across the skin than does S/O-mix, which is in
the surfactants and oil (S/O-mix, water-free system), and in a plain agreement with Sintov and Shapiro (2004) for lidocaine (as base),
solution of 1% CUR in propylene carbonate. The CoS/S and the but it is in contrast with Sintov and Greenberg (2014) for caffeine.
surfactants ratios in the microemulsion as well as the micellar The latter argument indicates that unlike hydrophilic and water-
system and the S/O-mix were kept constant. As seen in Fig. 5 and in soluble molecules such as caffeine for which an oil additive is
Table 1, the micellar system, the S/O-mix, and the plain solution crucial, the oil is redundant or less essential for the percutaneous
resulted in a significantly lower CUR permeation relative to that diffusion process of poorly soluble agents such as CUR.
presented by the microemulsion ME-10 (p < 0.05, ANOVA). The
permeability coefficients (Papp) of the micellar system, the S/O- 3.4. Influence of curcumin loading
mix, and the plain solution were 0.018  103  0.011 103,
0.005  103  0.002  103, and 0.002  103  0.000  103 cm/ The in vitro percutaneous permeation profiles of 1–3% CUR
h, respectively, vs. 0.110  103  0.021 103 cm/h for ME-10. The loaded in a microemulsion type ME-10 is illustrated in Fig. 6. As
enhancement ratio (ER = Jss-ME/Jss-solution) of CUR permeated via 1% shown, by increasing CUR concentration in the microemulsion,

Fig. 5. In vitro percutaneous permeation of curcumin using fresh rat skin after its application in a W/O microemulsion (10% aqueous phase), in micellar system, and in a water-
free surfactants and oil mixture (all formulations contained 1% curcumin, n = 4). Data were compared with 1% curcumin solution in propylene carbonate.
102 A.C. Sintov / International Journal of Pharmaceutics 481 (2015) 97–103

Fig. 6. In vitro percutaneous permeation of curcumin through fresh rat skin after application of curcumin-loaded microemulsions at various curcumin concentrations. A
comparison was also made between 1% (diamonds), 2% (squares), and 3% aqueous phase (triangles) (n = 10).
Average droplet size (nm in diameter)

4 3.65 nm 100

3.5 90
80

Percent of total droplets


3
70
2.5 2.36 nm
60
2 50

1.5 40
1.26 nm
0.80 nm 30
1
20
0.5 10
0 0
0.0 1.0 2.0 3.0
Curcumin concentraon (% wt/wt)
Fig. 7. Nanodroplet size (average diameter, nm) and their prevalence (percent of total droplets) in microemulsions containing 0–3% concentrations of curcumin.

its permeability coefficient increased from 0.116  103  0.052 decrease was accompanied by increase in aggregation, the
 103 cm/h in 1% CUR-loaded ME-10 to 0.233  103  0.051 diminution in size of more than 50% of the droplets may explain
 103 cm/h and 0.295  103  0.050  103 cm/h in 2% CUR and the improvement in CUR permeability.
3% CUR-loaded ME-10, respectively (Table 1). It should be noted
that the Papp values for CUR permeation obtained in this study 4. Conclusion
are comparable with those previously published by Liu et al.
(2011), who used transdermal microemulsions composed of 20% From the results presented in this paper, it has been concluded
limonene (oil), 10% water, 35% polysorbate 80 (surfactant) and that the W/O microemulsion vehicle for transdermal CUR
35% ethanol (cosurfactant). There is no statistically significant permeation is advantageous over a micellar system and a
difference between the 2% and 3% CUR-loaded ME-10, however, surfactant-oil mixture, composed of the same proportions of
there is a significant difference between CUR permeability of ingredient. This indicates that the mechanism of CUR delivery is
these high loaded formulations and 1% CUR-loaded micro- based on the nanostructural characteristics of the microemulsion.
emulsion (p < 0.05, ANOVA). It seems that although micro- In addition, it has been demonstrated that one of the mediating
emulsion of the ME-10 type can be loaded up to 4% CUR, the factors of skin permeation and retention is the droplet size of the
effect of loading on CUR permeation through the skin is microemulsion. One more important factor is apparently the
practically expressed from 1% to 2% CUR concentration. Fig. 7 density of the dispersed droplets in the microemulsion, which
presents the effect of CUR loading on droplet size, showing a depends on both droplet size and the volume fraction of the inner
linear correlation between the decrease in droplet size and the phase. Finally, this report joins to a series of studies demonstrating
increase of CUR loading in the microemulsion. Although this the advantage of the microemulsion system in enhancement of
A.C. Sintov / International Journal of Pharmaceutics 481 (2015) 97–103 103

percutaneous and cutaneous permeability of pharmacologically- Li, L., Ahmed, B., Mehta, K., Kurzrock, R., 2007. Liposomal curcumin with and
active agents. without oxaliplatin: effects on cell growth, apoptosis, and angiogenesis in
colorectal cancer. Mol. Cancer Ther. 6 (4), 1276–1782.
Lim, K.J., Bisht, S., Bar, E.E., Maitra, A., Eberhart, C.G., 2011. A polymeric nanoparticle
Acknowledgment formulation of curcumin inhibits growth, clonogenicity and stem-like fraction
in malignant brain tumors. Cancer Biol. Ther. 11 (5), 464–473.
Liu, A., Lou, H., Zhao, L., Fan, P., 2006. Validated LC/MS/MS assay for curcumin and
The author is grateful for the professional assistance of Mr. Igor tetrahydrocurcumin in rat plasma and application to pharmacokinetic study of
Greenberg and Ms. Lillia Shapiro at the Laboratory for Biophar- phospholipid complex of curcumi. J. Pharm. Biomed. Anal. 40 (3), 720–727.
maceutics. Liu, C.H., Chang, F.Y., 2011. Development and characterization of eucalyptol
microemulsions for topic delivery of curcumin. Chem. Pharm. Bull. (Tokyo) 59
(2), 172–178.
References Liu, C.H., Chang, F.Y., Hung, D.K., 2011. Terpene microemulsions for transdermal
curcumin delivery: effects of terpenes and cosurfactants. Colloid Surf. B
Alvarez-Figueroa, M.J., Blanco-Mendez, J., 2001. Transdermal delivery of Biointerfaces 82 (1), 63–70.
methotrexate: iontophoretic delivery from hydrogels and passive delivery from Liu, C.H., Huang, H.Y., 2012. Antimicrobial activity of curcumin-loaded myristic acid
microemulsions. Int. J. Pharm. 215, 57–65. microemulsions against Staphylococcus epidermidis. Chem. Pharm. Bull. 60 (9),
Anand, P., Kunnumakkara, A.B., Newman, R.A., Aggarwal, B.B., 2007. Bioavailability 1118–1124.
of curcumin: problems and promises. Mol. Pharm. 4 (6), 807–818. Liu, H., Wang, Y., Han, F., Yao, H., Li, S., 2007. Gelatin-stabilized microemulsion-based
Anand, P., Thomas, S.G., Kunnumakkara, A.B., Sundaram, C., Harikumar, K.B., Sung, organogels facilitates percutaneous penetration of cyclosporine A in vitro and
B., Tharakan, S.T., Misra, K., Priyadarsini, I.K., Rajasekharan, K.N., Aggarwal, B.B., dermal pharmacokinetics in vivo. J. Pharm. Sci. 96, 3000–3009.
2008. Biological activities of curcumin and its analogues (congeners) made by Maiti, K., Mukherjee, K., Gantait, A., Saha, B.P., Mukherjee, P.K., 2007. Curcumin–
man and mother nature. Biochem. Pharmacol. 76 (11), 1590–1611. phospholipid complex: preparation, therapeutic evaluation and
Bisht, S., Feldmann, G., Soni, S., Ravi, R., Karikar, C., Maitra, A., Maitra, A., 2007. pharmacokinetic study in rats. Int. J. Pharm. 330 (1–2), 155–163.
Polymeric nanoparticle-encapsulated curcumin (nanocurcumin): a novel Metzler, M., Pfeiffer, E., Schulz, S.I., Dempe, J.S., 2013. Curcumin uptake and
strategy for human cancer therapy. J. Nanobiotechnol. 5, 3. metabolism. Biofactors 39 (1), 14–20.
Bisht, S., Mizuma, M., Feldmann, G., Ottenhof, N.A., Hong, S.M., Pramanik, D., Mosley, C.A., Liotta, D.C., Snyder, J.P., 2007. Highly active anticancer curcumin
Chenna, V., Karikari, C., Sharma, R., Goggins, M.G., Rudek, M.A., Ravi, R., Maitra, analogues. Adv. Exp. Med. Biol. 595, 77–103.
A., Maitra, A., 2010. Systemic administration of polymeric nanoparticle- Patel, N.A., Patel, N.J., Patel, R.P., 2009. Design and evaluation of transdermal drug
encapsulated curcumin (NanoCurc) blocks tumor growth and metastases in delivery system for curcumin as an anti-inflammatory drug. Drug Dev. Ind.
preclinical models of pancreatic cancer. Mol. Cancer Ther. 9 (8), 2255–2264. Pharm. 35 (2), 234–242.
Chun, Y.S., Bisht, S., Chenna, V., Pramanik, D., Yoshida, T., Hong, S.M., de Wilde, R.F., Provencher, S.W., 1982. CONTIN: a general purpose constrained regularization
Zhang, Z., Huso, D.L., Zhao, M., Rudek, M.A., Stearns, V., Maitra, A., Sukumar, S., program for inverting noisy linear algebraic and integral equations. Comp. Phys.
2012. Intraductal administration of a polymeric nanoparticle formulation of Commun. 27, 229–242.
curcumin (NanoCurc) significantly attenuates incidence of mammary tumors in Rhee, Y.-S., Choi, J.-G., Park, E.-S., Chi, S.-C., 2001. Transdermal delivery of ketoprofen
a rodent chemical carcinogenesis model: implications for breast cancer using microemulsions. Int. J. Pharm. 228, 161–170.
chemoprevention in at-risk populations. Carcinogenesis 33 (11), 2242–2249. Shevachman, M., Garti, N., Shani, A., Sintov, A.C., 2008. Enhanced percutaneous
Cui, J., Yu, B., Zhao, Y., Zhu, W., Li, H., Lou, H., Zhai, G., 2009. Enhancement of oral permeability of diclofenac using a new U-type dilutable microemulsion. Drug
absorption of curcumin by self-microemulsifying drug delivery systems. Int. J. Dev. Ind. Pharm. 34, 403–412.
Pharm. 371 (1–2), 148–155. Sintov, A.C., Shapiro, L., 2004. New microemulsion vehicle facilitates percutaneous
Gannu, R., Palem, C.R., Yamsani, V.V., Yamsani, S.K., Yamsani, M.R., 2010. Enhanced penetration in vitro and cutaneous drug bioavailability in vivo. J. Control.
bioavailability of lacidipine via microemulsion based transdermal gels: Release 95, 173–183.
formulation optimization ex vivo and in vivo characterization. Int. J. Pharm. 388, Sintov, A.C., Botner, S., 2006. Transdermal drug delivery using microemulsion and
231–241. aqueous systems: influence of skin storage conditions on the in vitro
Gordon, O.N., Schneider, C., 2012. Vanillin and ferulic acid: not the major permeability of diclofenac from aqueous vehicle systems. Int. J. Pharm. 311,
degradation products of curcumin. Trends Mol. Med. 18 (7), 361–363. 55–62.
Griesser, M., Pistis, V., Suzuki, T., Tejera, N., Pratt, D.A., Schneider, C., 2011. Sintov, A.C., Brandys-Sitton, R., 2006. Facilitated skin penetration of lidocaine by
Autoxidative and cyclooxygenase-2 catalyzed transformation of the dietary using a combination of iontophoresis and microemulsion-based formulation.
chemopreventive agent curcumin. J. Biol. Chem. 286 (2), 1114–1124. Int. J. Pharm. 316, 58–67.
Gupta, S.C., Patchva, S., Aggarwal, B.B., 2013. Therapeutic roles of curcumin: lessons Sintov, A.C., Levy, H.V., Botner, S., 2010. Systemic delivery of insulin via the nasal
learned from clinical trials. AAPS J. 15 (1), 195–218. route using a new microemulsion system: in vitro and in vivo studies. J. Control.
Heuschkel, S., Goebel, A., Neubert, R.H.H., 2008. Microemulsions—modern colloidal Release 148, 168–176.
carrier for dermal and transdermal drug delivery. J. Pharm. Sci. 97 (2), 603–631. Sintov, A.C., Greenberg, I., 2014. Comparative percutaneous permeation study using
Hu, L., Jia, Y., Niu, F., Jia, Z., Yang, X., Jiao, K., 2012. Preparation and enhancement of caffeine-loaded microemulsion showing low reliability of the frozen/thawed
oral bioavailability of curcumin using microemulsions vehicle. J. Agric. Food skin models. Int. J. Pharm.. http://dx.doi.org/10.1016/j.ijpharm.2014.05.040.
Chem. 60 (29), 7137–7141. Tomren, M.A., Másson, M., Loftsson, T., Tønnesen, H.H., 2007. Studies on curcumin
Hua, L., Weisan, P., Jiayu, L., Ying, Z., 2004. Preparation evaluation, and NMR and curcuminoids XXXI. Symmetric and asymmetric curcuminoids: stability,
characterization of vinpocetine microemulsion for transdermal delivery. Drug activity and complexation with cyclodextrin. Int. J. Pharm. 338 (1–2), 27–34.
Dev. Ind. Pharm. 30 (6), 657–666. Tønnesen, H.H., Másson, M., Loftsson, T., 2002. Studies of curcumin and
Kreilgaard, M., Pedersen, E.J., Jaroszewski, J.W., 2000. NMR characterization and curcuminoids. XXVII. Cyclodextrin complexation: solubility, chemical and
transdermal drug delivery potential of microemulsion systems. J. Control. photochemical stability. Int. J. Pharm. 244 (1–2), 127–135.
Release 69, 421–433. Wang, Y.J., Pan, M.H., Cheng, A.L., Lin, L.I., Ho, Y.S., Hsieh, C.Y., Lin, J.K., 1997. Stability
Kreilgaard, M., Kemme, M.J.B., Burggraaf, J., Schoemaker, R.C., Cohen, A.F., 2001. of curcumin in buffer solutions and characterization of its degradation products.
Influence of a microemulsion vehicle on cutaneous bioequivalence of a J. Pharm. Biomed. Anal. 15 (12), 1867–1876.
lipophilic model drug assessed by microdialysis and pharmacodynamics. Yang, K.Y., Lin, L.C., Tseng, T.Y., Wang, S.C., Tsai, T.H., 2007. Oral bioavailability of
Pharm. Res. 18, 593–599. curcumin in rat and the herbal analysis from Curcuma longa by LC-MS/MS. J.
Kreilgaard, M., 2001. Dermal pharmacokinetics of microemulsion formulations Chromatogr. B Analyt. Technol. Biomed. Life Sci. 853, 183–189.
determined by in vitro microdialysis. Pharm. Res. 18, 367–373. Zhang, J., Michniak-Kohn, B., 2011. Investigation of microemulsion microstructures
Kunwar, A., Barik, A., Pandey, R., Priyadarsini, K.I., 2006. Transport of liposomal and and their relationship to transdermal permeation of model drugs: ketoprofen
albumin loaded curcumin to living cells: an absorption and fluorescence lidocaine, and caffeine. Int. J. Pharm. 421, 34–44.
spectroscopic study. Biochim. Biophys. Acta 60 (10), 1513–1520. Zhang, Y.T., Zhao, J.H., Zhang, S.J., Zhong, Y.Z., Wang, Z., Liu, Y., Shi, F., Feng, N.P., 2011.
Lee, P.J., Langer, R., Shastri, V.P., 2003. Novel microemulsion enhancer formulation Enhanced transdermal delivery of evodiamine and rutaecarpine using
for simultaneous transdermal delivery of hydrophilic and hydrophobic drugs. microemulsion. Int. J. Nanomed. 6, 2469–2482.
Pharm. Res. 20, 264–269. Zhao, J.H., Ji, L., Wang, H., Chen, Z.Q., Zhang, Y.T., Liu, Y., Feng, N.P., 2011.
Li, L., Braiteh, F.S., Kurzrock, R., 2005. Liposome-encapsulated curcumin: in vitro and Microemulsion-based novel transdermal delivery system of
in vivo effects on proliferation, apoptosis, signaling, and angiogenesis. Cancer tetramethylpyrazine: preparation and evaluation in vitro and in vivo. Int. J.
104 (6), 1322–1331. Nanomed. 6, 1611–1619.

Вам также может понравиться