You are on page 1of 8

NANOMEDICINE

Journal of Controlled Release 141 (2010) 85–92

Contents lists available at ScienceDirect

Journal of Controlled Release


j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / j c o n r e l

Nanoparticles made from novel starch derivatives for transdermal drug delivery
M.J. Santander-Ortega a,b, T. Stauner c, B. Loretz b, J.L. Ortega-Vinuesa a, D. Bastos-González a, G. Wenz c,
U.F. Schaefer b, C.M. Lehr b,⁎
a
Department of Applied Physics, University of Granada, Granada, Spain
b
Department of Biopharmaceutics and Pharmaceutical Technology, Saarland University, Saarbrücken, Germany
c
Department of Organic Macromolecular Chemistry, Saarland University, Saarbrücken, Germany

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

Article history: The goal of this paper was aimed to the formulation of nanoparticles by using two different propyl-starch
Received 14 May 2009 derivatives – referred to as PS-1 and PS-1.45 – with high degrees of substitution: 1.05 and 1.45 respectively. A
Accepted 11 August 2009 simple o/w emulsion diffusion technique, avoiding the use of hazardous solvents such as dichloromethane or
Available online 21 August 2009
dimethyl sulfoxide, was chosen to formulate nanoparticles with both polymers, producing the PS-1 and PS-1.45
nanoparticles. Once the nanoparticles were prepared, a deep physicochemical characterization was carried out,
Keywords:
Starch
including the evaluation of nanoparticles stability and applicability for lyophilization. Depending on this
Nanoparticles information, rules on the formation of PS-1 and PS-1.45 nanoparticles could be developed. Encapsulation and
Physico-chemical Characterization release properties of these nanoparticles were studied, showing high encapsulation efficiency for three tested
Drug delivery drugs (flufenamic acid, testosterone and caffeine); in addition a close to linear release profile was observed for
Skin hydrophobic drugs with a null initial burst effect. Finally, the potential use of these nanoparticles as transdermal
drug delivery systems was also tested, displaying a clear enhancer effect for flufenamic acid.
© 2009 Elsevier B.V. All rights reserved.

1. Introduction Besides, an increasing number of researchers is investigating on


the improvement of percutaneous drug absorption [15]. However, the
In the last years, the use of polysaccharides to prepare nanopar- excellent barrier properties of the skin is a handicap that must be
ticles has increased considerably [1–3]. Polysaccharides possess many overcome [16]. Usually, to reach a high and constant drug flux through
recognition functions, allowing e.g. mucoadhesion or specific receptor the skin a change of the skin barrier function is necessary. However, in
recognition [4], as well as providing neutral coatings with low surface the last years, different authors [17–20] have demonstrated that the
energy, preventing non specific protein adsorption [5]. On the other use of particles as transdermal drug delivery systems (TDDS) enhance
hand, the high amount of hydroxyl groups in the polysaccharide the rate and extent of transport across skin, without compromising
backbone allows the incorporation of different specific ligands to the skin barrier function. Although such systems were undoubtedly
obtain polyfunctional colloidal systems [6]. able to enhance skin penetration and distribution, the mechanism by
In this context, starch has an interesting potential, which is so far which this enhancement was achieved is still unclear [17,21].
relatively unexplored. Starch is a biocompatible, biodegradable, non- The aim of this work was to explore nanoparticulate drug carriers
toxic polymer, abundantly occurring in nature as the major polysac- based on starch derivatives, by using the advantages of hydrophobic
charide storage in higher plants [7,8]. However, despite these starch derivatives. With this in mind, we have selected propyl-starch
properties, some problems can occur. The hydrophilic nature of derivatives instead of the well known acetyl-starch derivatives [22–
starches is a major constraint that seriously limits the development of 24] to formulate our nanoparticles. The inclusion of propyl groups,
starch-based nanoparticles [9]. A good alternative to solve this even with low degree of substitution, would enable good solubility in
problem is the grafting of hydrophobic side chains to the hydrophilic low hazardous organic solvents, such as ethyl acetate. Moreover,
starch backbone [10–14]. However, an important restriction usually propyl-starches may allow a better quality control since degree of
arising from the use of these hydrophobic polysaccharides derivatives substitution can be determined after acidic degradation to the
to prepare nanoparticles, is the necessity to employ organic solvents, respective glucose derivatives by 1H-NMR spectroscopy. Two pro-
such as dichloromethane or dimethyl sulfoxide, with considerable pyl-starch derivatives and one un-modified starch polymer were used
toxicological and other safety risks [10,11]. as basic constituents for the preparation of nanoparticles, represent-
ing two different degrees of substitutions (Ds) 1.05 (PS-1) and 1.45
(PS-1.45) respectively.
⁎ Corresponding author. Tel.: +49 681 302 3039; fax: +49 681 302 4671. Nanoparticles were formulated by emulsification–diffusion tech-
E-mail address: lehr@mx.uni-saarland.de (C.M. Lehr). nique. This method has several advantages such as high yields generally

0168-3659/$ – see front matter © 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.jconrel.2009.08.012
NANOMEDICINE
86 M.J. Santander-Ortega et al. / Journal of Controlled Release 141 (2010) 85–92

obtained, high batch to batch reproducibility and easy scaling up. 2.5. Lyophilization
Moreover, it is possible to control the size and polydispersity of nano-
particles by the control of the oil/water phase ratio [25]. PS-1 and PS-1.45 nanoparticles were lyophilized using a Freeze-drier
Once nanoparticles were formed, a thorough physicochemical char- Alpha 2-4 LSC (Christ, Osterode, Germany) and sucrose or trehalose as
acterization was carried out. Afterwards, the capacity of these cryoprotectant agent. Briefly, different volumes of 10% (w/v) sucrose or
nanoparticles as drug delivery systems was explored by the encapsu- trehalose solution were poured onto different aliquots of nanoparticles
lation and release of three different model drugs, flufenamic acid (FFA), in order to obtain a cryoprotectant range between 0 and 1% (w/v).
testosterone (Test) and caffeine (Caff). Finally, considering the nano- Previously to the lyophilization step, the different samples were frozen
particles characterization results, it was possible to establish the use of at −80 °C for 2 h. Hydrodynamic mean size of these aliquots was
these nanoparticles as TDDS. Hence, the potential application of these measured before and after the lyophilization step.
nanoparticles as TDDS was analyzed by studying their influence on the
permeability of human heat-separated epidermis (HSE). 2.6. Lactat dehydrogenase (LDH) assay

Caco-2 cell line clone C2BBe1 (ATCC No CRL-2102) was used as test
2. Materials and methods
system cultured in DMEM supplemented with 1% Non-essential
amino acids (both from Gibco, Karlsruhe, Germany) and 10% FBS
2.1. Materials
(F7524, Sigma-Aldrich, Taufkirchen, Germany).
Cytotoxicity detection Kit (LDH) from Roche (Mannheim, Ger-
Maize Starch polymer with an amylose content of 25% was kindly
many) was used basically following the instruction manual with
donated by BASF, ethyl acetate (Fluka Chemie GmbH, Buchs, Switzer-
minor deviations. Caco-2 cells were seeded into 24-well plates and
land), polyvinyl alcohol (PVA); Mowiol® 4-88 (Kuraray Specialities
the toxicity assay was performed at the state of a confluent
Europe GmbH, Frankfurt, Germany), Igepal® CA-630 (Octylphenyl-
monolayer. Samples were prepared by dilution of aqueous suspension
polyethylene glycol) (Sigma, St. Louis, MO, USA), cellulose membrane
of starch particles with supplemented cell culture medium at ratios
MW 12–14 kDa (Medicell International Ltd., London, UK) were used
1:3, 1:6, 1:12 and 1:24 resulting in concentrations of 0.0042 to
as obtained from the suppliers. Flufenamic acid, testosterone and
0.033 mg nanoparticles/ml. As negative control cells were incubated
caffeine were purchased from Sigma-Aldrich (St. Louis, MO, USA). All
with cell culture medium and as high control with medium plus 1% of
other solvents and chemicals used were of the highest grade
the detergent Triton X-100. Samples were taken after 4 h and 24 h
commercially available.
incubation time and analysed according to the manufacturer's
protocol. No interaction was found with the detection method and
2.2. Preparation of the nanoparticles the nanoparticles.

Nanoparticles with the different polymers were formulated by a 2.7. MTT assay
simple o/w emulsion diffusion method. Briefly, specific starch
derivative (un-modified starch, PS-1 or PS-1.45) was dissolved in Thiazolyl Blue Tetrazolium Bromide (M5655, Sigma-Aldrich,
ethyl acetate (1 mg/ml), and 1 ml of this organic solution was poured Taufkirchen, Germany) was dissolved in PBS pH 7.4 to yield a final
on 4 ml of an aqueous phase with different percentages (w/v) of PVA concentration of 5 mg/ml for the stock solution.
(0, 0.1, 0.5 and 1). This biphasic system was emulsified with a high Caco-2 cells were seeded into 24-well plates and the toxicity assay
speed homogenizer (Ultra Turrax® Ika®, Staufen, Germany) at was performed at the state of a confluent monolayer by exposure to
14,000 rpm during 15 min. Then, high purified water was added up particle suspension as reported for LDH assay. After the incubation
to 10 ml to force the complete diffusion of the organic solvent to the period of 4 h the particle suspension was removed. Cells were washed
aqueous phase. Finally, the organic solvent was evaporated under once with phosphate buffered saline before fresh medium and 50 µl
vacuum at 35 °C (Rotavapor Büchi®, Labortechnik AG, Flawil, MTT stock solution per well was added. After further 3 h incubation
Switzerland), generating stable nanoparticles. After nanoparticles 500 µl lysis buffer (10% SDS in 0.01 mM HCl) were added. The
preparation, high purified water was added to obtain a colloidal absorbance at λ = 550 nm was analysed in a plate reader. Viability
solution with a final volume of 10 ml. was calculated in comparison to the positive control, untreated cells
as 100% value, and negative control 1%-Triton solution as 0% value.
2.3. Characterization of the nanoparticles
2.8. Encapsulation of flufenamic acid, testosterone and caffeine
Size and ζ-potential of the nanoparticles were analysed by photon
In order to assess the potential use of these nanoparticles as drug
correlation spectroscopy (PCS) using a Nano-ZS (Malvern Instruments,
delivery system three different drugs were chosen to study the
Malvern, UK). For ζ-potential measurements nanoparticles were diluted
encapsulation behaviours of PS-1 and PS-1.45 nanoparticles. These
in NaCl 3 mM. AFM images were obtained using an Atomic Force
drugs present different hydrophobicity (expressed as the logarithm of
Microscopy Nanoscope IV Bioscope™ (Veeco Instruments, Santa
the octanol–water partition coefficient; Log P) and net charge. Exactly,
Barbara, CA, USA). Imaging was done using Taping mode and a silicon
the chosen drugs were: flufenamic acid (Log P = 4.80; pKa = 3.90;
cantilever with a spring constant of approximately 40 N/m and a
Mw = 281.23), testosterone (Log P = 3.47; non-ionizable molecule;
resonance frequency of about 170 kHz. The scan speed applied was
M w = 288.40) and caffeine (Log P = − 0.08; p K b = 10.40;
0.2 Hz.
Mw = 194.20). The preparation method of drug loaded PS-1 and PS-
1.45 nanoparticles was not modified with respect to the un-loaded
2.4. Colloidal stability nanoparticles. The specific molecule and the PS-1 or PS-1.45 polymer
were dissolved in the organic phase (1:1 ratio), following the same
Stability of the nanoparticles was studied as a function of salinity of procedure described in section 2.2. The characterization of these
the medium using NaCl and CaCl2 as aggregating salts. Particle nanoparticles was performed as it was comment previously for the
aggregation was analyzed by photon correlation spectroscopy (PCS) un-loaded nanoparticles. Encapsulation efficiency (EE) was calculated
using a 4700c System (Malvern Instruments). For more details see using a Franz diffusion cell. The exact amount of each drug was
supporting information that can be find in the webpage of this journal. determined by HPLC analysis [19,26].
NANOMEDICINE
M.J. Santander-Ortega et al. / Journal of Controlled Release 141 (2010) 85–92 87

2.12. Quantification of flufenamic acid, testosterone and caffeine

2.12.1. HPLC system


Pump: Dionex P580 pump; Autosampler: Dionex ASJ 100 automated
sample injector; detector: UVD 170 S detector; column oven: Dionex
STH 585 column oven; software: chromeleon 6.50 SP2 build 9.68.
A 125 × 4 mm LiChrospher® 100/RP-18 column (Merck-Hitachi,
Darmstadt, Germany) with a 4 × 4-mm LiChrospher® 100/RP-18 guard
column (5 µm) (Merck-Hitachi, Darmstadt, Germany) was used as
stationary phase for all substances.

2.12.2. Solvents and buffers


Methanol (Lichrosolv, Merck, Darmstadt, Germany); Acetonitril
HPLC grade (Merck, Darmstadt, Germany); Deionised water (Millipore
Milli Q Synthesis, Heidelberg, Germany); McIlvaine buffer (pH 2.2):
1 l contains 20.8 g citric acid, 0.4 g (Na2HPO4 × 2 H2O) and deionised
water; Buffer (pH 2.6): 1 l contains 1.15 ml phosphoric acid, 4.08 g
KH2PO4 anhydrous and deionised water.
Fig. 1. Hydrodynamic size distribution of PS-1 nanoparticles formulated with different
percentages (w/v) of PVA, (dash-dotted line) 0%, (dotted line) 0.1%, (dashed line) 0.5%, 2.12.3. HPLC analytics of the test compounds
(solid line) 1% (n ≥ 3). Flufenamic acid: mobile phase: 80:20 (v/v) methanol: McIlvaine
buffer pH 2.2; retention time: 3.1 min ± 0.2 min; detection wavelength:
284 nm; flow rate: 1.2 ml/min; calibration from 25 to 5000 ng/ml
2.9. In-vitro release of flufenamic acid, testosterone and caffeine (r 2 = 0.999); injection volume: 50 µl.
Testosterone: mobile phase: methanol/water (v/v); 70:30; retention
Release of FFA, testosterone and caffeine from loaded PS-1 and PS-1.45 time: 4.8 min ± 0.2 min; detection wavelength: 250 nm; flow rate:
nanoparticles was studied using Franz diffusion Cells (FdC) type 6G-01- 1.2 ml/min; calibration from 50 to 5000 ng/ml (r2 = 0.999); injection
00-15 (Perme-Gear, Riegelsville, PA, USA). Briefly, a mixture of 0.5 ml of volume: 50 µl.
nanoparticles dispersion and 1.5 ml of phosphate buffer pH 7 (2 mM) was Caffeine: mobile phase: buffer pH 2.6/acetonitrile; 90:10 (v/v);
poured in the donor compartment of FdC, while the receptor compart- retention time: 5.1 ± 0.2 min; detection wavelength: 262 nm; flow rate:
ment was filled with 12 ml of the same buffer solution. In the case of the 1.2 ml/min; calibration from 25 to 10,000 ng/ml (r2 = 0.999); injection
testosterone, the buffer solution was changed due to the low solubility of volume: 50 µl.
this drug. In this case, the medium consisted in phosphate buffer pH 7 Drugs in samples were quantified by the external standard method.
(2 mM) with addition of 2% (v/v) Igepal® CA-630 (a non-ionic poly-
ethylene glycol derivative) and 0.4% (v/v) ethanol. In all cases, donor and 3. Results
receptor compartments were separated by a cellulose membrane with a
molecule cut off of 12–14 kDa. (Medicell Int. Ltd, London, UK.). FdCs were 3.1. Characterization of the nanoparticles
incubated at 32 °C for a period of time of 72 h. Samples of 0.4 ml were
removed from the receptor compartment at regular time intervals up to Propyl-starch derivatives, soluble in organic solvents like ethyl
72 h and replaced with an equal volume of fresh buffer. The concentration acetate, were synthesized by reaction of starch with propyl bromide
of free drug from the receptor was analyzed by HPLC [19,26]. as described in the supporting information that can be found in the
webpage of this journal. Fig. 1 shows the hydrodynamic size dis-
2.10. Skin preparation tribution of nanoparticles formulated with PS-1 polymer and different
amounts of PVA. Similar results were obtained for PS-1.45 polymer
Excised human skin from Caucasian female patients, who had (data not shown). However, in the case of un-modified starch the
undergone abdominal plastic surgery, was used. Permission of the formation of nanoparticle could not be observed.
Ethical Committee of the Caritas-Traegergesellschaft, Trier, Germany As can be observed in Fig. 1, the incorporation of PVA to the
(July 6, 1998). For more details see supporting information that can be formulation improved the nanoparticle hydrodynamic size distribution.
found in the webpage of this journal. A mono-modal size distribution was achieved when the aqueous phase
contains a PVA concentration equal or higher than 0.5% (w/v). The
2.11. Permeation of flufenamic acid, testosterone and caffeine through narrowest size distribution was reached when the aqueous phase
human heat-separated epidermis presented a 1% (w/v) of PVA. Consequently, nanoparticles formulated
with an aqueous phase of 1% (w/v) of PVA (PS-1 and PS-1.45
Permeation studies of the three encapsulated drugs in PS-1 or PS- nanoparticles) were selected for the next studies. Table 1 summarizes
1.45 nanoparticles were carried out using FdC. The whole procedures the mean hydrodynamic size and PDI of these nanoparticles. PS-1
are described in the supporting information that can be found in the nanoparticles exhibited a lower hydrodynamic mean size than PS-1.45.
webpage of this journal. If we consider the reactivity of the hydroxyl starch groups (primary

Table 1
Hydrodynamic mean size (nm), PDI, ζ-potential (mV), ccc (mM) and csc (mM) values in presence of NaCl and CaCl2 of PS-1 and PS-1.45 nanoparticles.

Size (nm) PDI ζ-Pot. ccc (mM) csc (mM) Sizea (nm) PDIa
(mV)
NaCl CaCl2 NaCl CaCl2

PS-1 150.5 ± 3.5 0.12 ± 0.01 − 8.3 ± 0.3 32 6 585 205 180.3 ± 1.1 0.24 ± 0.01
PS-1.45 182.7 ± 6.7 0.08 ± 0.02 − 5.8 ± 0.5 36 9 185 55 185.5 ± 5.4 0.09 ± 0.01
a
After incubation for 25 days at room temperature.
NANOMEDICINE
88 M.J. Santander-Ortega et al. / Journal of Controlled Release 141 (2010) 85–92

alcohol > secondary alcohol), and Ds of both starch derivatives, these


results suggest that the modification of the primary alcohol is enough for
the nanoparticles formation. But, it is plausible that the increase of Ds
modified the behaviour of starch derivative during the emulsification
step, and then, the final characteristic of nanoparticles will be also
modified. This will be discussed in more detail below. As illustrated by
AFM images (Fig. 2a and b), PS-1 and PS-1.45 nanoparticles showed a
spherical shape with a narrow size distribution.
Regarding to the negative ζ-potential values displayed by both
colloidal systems, it is worthy to remark that neither starch
derivatives nor PVA polymers present ionizable chemical groups.
Hence, the low and negative ζ-potential summarized in Table 1 can be
attributed to the specific interaction of the nanoparticles surface with
the ions present in the medium (as ζ-potential measurement were
developed in an aqueous solution of NaCl 3 mM) [27].

Fig. 3. Stability factor as a function of salt concentration for PS-1 (square) and PS-1.45
3.2. Colloidal stability (circle) nanoparticles. NaCl (closed symbols, solid line); CaCl2 (open symbols, dashed
line).
Stability of PS-1 and PS-1.45 nanoparticles as a function of the
electrolyte concentration using NaCl and CaCl2 as aggregating salts was surface hydrophilicity, but also on the nature and concentration of the
determined. Fig. 3 illustrates the value of the Fuchs factor (W) as a hydrated counterions that surround the particles. Then, as Ca2+ is a
function of the salt concentration; the corresponding values of critical more hydrated cation than Na+, the re-stabilization power of the
coagulation concentration (ccc) and critical stabilization concentration former is higher than that of the latter. This explains the lower csc
(csc) are summarized in Table 1. For both nanoparticle systems ccc values obtained for CaCl2 in comparison with NaCl. However, using a
values were lower when CaCl2 was used as aggregating salt. This was given electrolyte, the lower the csc, the higher the hydrophilicity of
due to the greater screening capacity of Ca2+ ions, which significantly the particle surface [30]. Then, the lower csc values found for PS-1.45
reduces the repulsive potential among the particles, favoring the nanoparticles, independently of the employed salt, with respect to PS-
aggregation of the system [28]. Nevertheless, although ccc values 1 nanoparticles, indicate a higher hydrophilic surface of the former.
were slightly lower for PS-1 nanoparticles, it was not possible to find a
clear difference between ccc values of both colloidal systems. Moreover,
the low values of ccc for both systems indicate that only a low amount of 3.3. Lyophilization
PVA must have been incorporated on the nanoparticles surface, that it
was not enough to produce a steric stabilization [29]. A typical limitation of colloidal systems suspended in aqueous
Interesting information can be obtained when the csc values are mediums is their low stability when conserved for a longer time;
analyzed. In this case, differences between both salts and both lyophilization is a good alternative to obtain long-term stable systems
colloidal systems were clearly found. It should be noted that csc values [31]. The lyophilization of PS-1 and PS-1.45 nanoparticles was studied
depends on hydration forces and they do not only depend on the as a function of the type and concentration of cryoprotectant agent.
Fig. 4 displays the hydrodynamic mean size of the reconstituted
systems using trehalose as cryoprotectant, similar results were
observed with sucrose (data not shown). To obtain reconstituted
PS-1 nanoparticles with a similar size that the un-lyophilized, it was
necessary to use a cryoprotectant concentration of 0.5% (w/v).
On other hand, reconstituted PS-1.45 nanoparticles displayed a
similar size that the original nanoparticles. This pattern was totally
independent of the concentration and type of cryoprotectant agent.
Fig. 2c depicts an AFM image of reconstituted PS-1.45 nanoparticles
that were lyophilized without cryoprotectant agent. Comparison of

Fig. 2. AFM images of PS-1 (a), PS-1.45 (b), lyophilized PS-1.45 (c) and FFA loaded PS-1 Fig. 4. Hydrodynamic mean size of reconstituted PS-1 (■) and PS-1.45 (●)
(d) nanoparticles. nanoparticles after lyophilization as a function of cryoprotectant concentration.
NANOMEDICINE
M.J. Santander-Ortega et al. / Journal of Controlled Release 141 (2010) 85–92 89

Table 2
Cytotoxicity of PS-1 and PS-1.45 nanoparticles after 4 h and 24 h incubation.

Concentration MTT assay LDH assay


(mg NP/ml)⁎ 102 (viability, %) (cytotoxicity, %)

4h 4h 24 h

PS-1 3.33 97.84 ± 1.41 − 0.89 ± 1.65 0.35 ± 4.55


1.66 102.20 ± 3.66 − 1.98 ± 0.53 − 0.97 ± 0.36
0.83 104.57 ± 4.24 − 0.89 ± 1.27 − 0.49 ± 1.40
0.42 102.95 ± 3.74 0.40 ± 0.37 0.00 ± 1.15
PS-1.45 3.33 102.16 ± 3.97 − 1.59 ± 1.72 − 1.51 ± 2.72
1.66 112.46 ± 1.34 − 2.02 ± 1.36 − 3.24 ± 1.25
0.83 108.88 ± 2.22 0.62 ± 0.88 − 2.43 ± 1.13
0.42 108.34 ± 3.02 − 1.88 ± 1.18 − 1.04 ± 0.55

Mean viability/toxicity n = 4 ± SD.

Fig. 2c and b shows that the reconstituted nanoparticles presented a


similar appearance to the un-lyophilised nanoparticles. Fig. 5. Release profile of FFA (■), testosterone (□) and caffeine (⊠) from PS-1
nanoparticles.

3.4. Cytotoxicity
PS-1.45 nanoparticles, Fig. 6 presented different release pattern as
Cytotoxicity was analysed using two diverse in vitro tests. LDH assay a function of the encapsulated drug. FFA presented the lowest release
is a test detecting the leakage of cell membrane and is useful in the rate, followed by testosterone. Both drugs showed a negligible burst
investigation of nanoparticle toxicity since the plasma membrane is the effect and a nearly linear profile. Finally, caffeine showed an initial fast
main place of contact between particles and cells. MTT assay monitors release, more pronounced than in the case of PS-1, followed by close
the mitochondrial metabolism of cells as an indicator of their viability. to zero release kinetic.
The results of both tests are consistent and show that propyl-starch
particles up to a concentration of 0.033 mg NPs/ml are not cytotoxic in 3.7. Permeation of flufenamic acid, testosterone and caffeine
the Caco-2 test model (Table 2). The tested samples PS-1 and PS-1.45
differed in their degree of substitution, but toxicological properties of Permeation profiles of FFA are illustrated in Fig. 7a, Papp values
the particles were not affected by the increase of modification. The same calculated from these profiles are summarized in Table 3. Clear
holds true for LDH assay following 24-h incubation. differences were found between the patterns and Papp values of free
and encapsulated molecules. Moreover, Papp values calculated with
3.5. Encapsulation of flufenamic acid, testosterone and caffeine both colloidal systems were significantly different (ANOVA, α < 0.05),
being higher for PS-1.45 nanoparticles. It is worthy to remark the
Table 3 summarizes the main characteristics of PS-1 and PS-1.45 similarity between the transport across a cellulose membrane (release
nanoparticles as a function of the encapsulated molecule. When experiment, Figs. 5 and 6) and across HSE in the permeation
compared to Table 1, it is easy to see that the encapsulation of FFA, experiment (Fig. 7a) for this drug, suggesting that release from the
testosterone or caffeine in PS-1 or PS-1.45 nanoparticles did not alter particles rather than skin permeation is rate limiting.
their original hydrodynamic mean size and PDI. As can be seen in the Regarding testosterone, in Fig. 7b, both the free and the
AFM image (Fig. 2d), encapsulation of FFA did not produce any encapsulated drug displayed very similar permeation profiles, proving
intelligible change in the spherical shape and soft surface of PS-1 that there was no effect of the nanoencapsulation on skin permeation.
nanoparticles; similar results were found for testosterone and caffeine No significant differences (ANOVA, α < 0.05) were found among the
and with PS-1.45 nanoparticles (data not shown). different Papp values (Table 3).
Finally, it is also interesting to note that PS-1 and PS-1.45 Similar results were obtained with caffeine (Fig. 7c); its Papp values
nanoparticles exhibited a high EE for the three tested drugs, >95% indicate that this molecule showed the lowest permeation rate of the
for FFA and testosterone; and >80% for caffeine. three drugs. Consequently, rather large differences were observed
between the release patterns (Figs. 5 and 6) and the permeation
3.6. In-vitro release of flufenamic acid, testosterone and caffeine experiments (Fig. 7c), suggesting that permeation across the skin
barrier remains rate limiting.
Once FFA, testosterone and caffeine were encapsulated in PS-1 and
PS-1.45 nanoparticles, next step was to study their release patterns. For 4. Discussion
PS-1 nanoparticles (Fig. 5), FFA and testosterone presented a sustained
release without any burst effect and a nearly linear profile. On the other Results illustrated in Fig. 1 show the clear effect of PVA on the
hand, the hydrophilic drug caffeine showed a much faster but linear nanoparticle formation. In this case, the well known surfactant
release within the first 10 h before the release reached a plateau phase. properties of this polymer [32,33] improved the nanoparticles

Table 3
Hydrodynamic mean size (nm), PDI, ζ-potential (mV), EE (%) and Papp (10− 6 cm/s) values of PS-1 and PS-1.45 nanoparticles with FFA, testosterone or caffeine encapsulated.

Encap. drug Size (nm) PDI ζ-Pot. (mV) EE (%) Papp (10− 6 cm/s) Papp⁎(10— 6 cm/s)

PS-1 FFA 159.7 ± 1.8 0.14 ± 0.02 − 15.6 ± 1.1 >95 2.43† ± 0.42 0.21 ± 0.05
Test 148.4 ± 0.7 0.10 ± 0.01 − 14.1 ± 3.3 >95 0.40 ± 0.10 0.32 ± 0.07
Caff 158.8 ± 1.5 0.11 ± 0.01 − 16.4 ± 2.3 >80 0.08 ± 0.01 0.09 ± 0.01
PS-1.45 FFA 185.5 ± 3.4 0.06 ± 0.02 − 12.3 ± 1.5 >95 3.11† ± 0.41 0.21 ± 0.05
Test 176.6 ± 6.0 0.11 ± 0.04 − 12.7 ± 0.6 >95 0.43 ± 0.11 0.32 ± 0.07
Caff 183.3 ± 5.4 0.11 ± 0.01 − 10.3 ± 1.2 >80 0.11 ± 0.03 0.09 ± 0.01

Papp⁎, Papp value calculated for a solution of un-encapsulated drug. † Indicates statistically significant differences.
NANOMEDICINE
90 M.J. Santander-Ortega et al. / Journal of Controlled Release 141 (2010) 85–92

Fig. 6. Release profile of FFA (●), testosterone (○) and caffeine (⊗) from PS-1.45
nanoparticles.

formation. In addition, the increase of PVA concentration in the external


aqueous phase gave a size reduction and a lower PDI. These results agree
with those published in the 90s by Scholes et al. [34], who found that the
decrease in size and PDI when PVA concentration was increased can be
explained by the increase of the external aqueous phase viscosity.
Regarding to the low and negative ζ-potential shown by both colloidal
systems, it is plausible that, insofar as neither starch derivative nor PVA
present ionizable groups, this value proceed from the specific
interaction of NaCl with the nanoparticles surface. The specific in-
teraction of the ions with the nanoparticle surface depends on their
polarizability, being higher for the anions. Then the results suggest a
higher accumulation of the Cl− anions than Na+ cations although the
nanoparticles surface must be neutral [27].
The analysis of the colloidal stability reveals additional informa-
tion on the role of each polymer during the particle formation. As
mentioned earlier (section 3.2), both colloidal systems showed a low
stability (see ccc values in Table 1). This is indicative of a low
incorporation of PVA in the interface during the emulsification step
[29]. The amount of adsorbed PVA can be enough to stabilize the
emulsion but it was not enough to give a clear steric stabilization. On
the other hand, the csc values are useful to obtain more information of
the superficial composition of both nanoparticles systems. In this case,
PS-1.45 nanoparticles displayed lower csc values than PS-1 with both
salts (see Table 1). This result indicates that the nanoparticles
formulated with the polymer with higher hydrophobic character
(PS-1.45) displayed a more hydrophilic surface. This is because the
adsorption of molecules onto a substrate is mainly controlled by the
hydrophobic forces [35,36]. Hence, the adsorption through the acetate
moieties of the PVA onto the drops of the organic phase during the
emulsion step must be more favourable in the case of the more
hydrophobic starch, PS-1.45, than for PS-1. Concomitantly, due to the Fig. 7. Permeation profile of FFA (a), Testosterone (b) and Caffeine (c) through HSE.
hydrophilic chains of the PVA, the higher accumulation of this Encapsulated in PS-1 (■) or PS-1.45 (●) nanoparticles and free drug (▲).
molecule gives a higher hydrophilic character to the surface of the
nanoparticles. Regarding to the null effect of the different amounts of
PVA incorporated in both colloidal systems on the ccc magnitude, systems [31]. For this reason, lyophilization is advisable. This is why we
studies carried out previously focused on the adsorption of a hydro- studied the freeze-drying behaviour of PS-1 and PS-1.45 nanoparticles.
philic surfactant on PLGA nanoparticles showed that the csc value It is well known that lyophilization may generate a high stress that could
presented a clear dependence with the amount of surfactant placed destabilize colloidal systems. Hence, the use of cryoprotectant agents to
on the nanoparticles surface, while ccc values were less sensitive to protect the nanoparticles integrity from the freezing stress is often
these changes [29]. This observation suggests that although PS-1 and necessary [31]. For this reason, we selected two different disaccharides,
PS-1.45 nanoparticles presented similar ccc values, the obviously sucrose and trehalose, as cryoprotectant agents, as they are well known
different csc values showed by both systems indicate a clearly dif- for their good cryoprotectant properties [37–39].
ferent surface composition. Hence, stability results indicate that the After lyophilization and redispersation in water, the average size of
Ds of the starch had a clear effect in the final properties of these the nanoparticles was analyzed (see Fig. 4). For PS-1, in order to recover
nanoparticles. the initial properties of the un-lyophilized PS-1 nanoparticles, it was
As can be deduced from the previous lines, the long-term colloidal necessary to add at least a concentration of cryoprotectant agent higher
stability is a serious handicap against the clinical use of nanoparticles than 0.5% (w/v), either sucrose or trehalose. Similar behaviour between
NANOMEDICINE
M.J. Santander-Ortega et al. / Journal of Controlled Release 141 (2010) 85–92 91

both disaccharides was previously published by others authors [40]. testosterone-PS-1.45 interaction. Concomitantly, it is also logic the faster
However, it is easy to find in the literature that trehalose seems to be the release of caffeine as observed for PS-1.45 nanoparticles due to the weaker
preferable cryoprotectant due to its better properties in comparison interaction within the matrix caffeine-PS-1.45 than the caffeine-PS-1.
with other sugars. Among these properties, it is remarkable its higher As a function of the stability results discussed previously, topical
glass transition temperature Tg [41,42]. Crowe et al. [43] found that, for application seems to be most promising route for these nanoparticles
long storage times trehalose is better cryoprotectant than sucrose, due and hence was also addressed in the last section of this study.
to its higher Tg. At this moment it is necessary to remark that PS-1 For FFA, encapsulation in starch nanoparticles caused a significant
nanoparticles were resuspended in purified water just after lyophiliza- (more than 10-fold, see Table 3) increase in its Papp value across human
tion and may be, for this reason, it was not possible to find a clear skin. A similar skin penetration enhancement by polymeric nanopar-
difference between sucrose and trehalose. ticles had been previously observed by us using PLGA [19], and by other
For PS-1.45 nanoparticles, from Figs. 4 and 2c, it is possible to authors with other drugs [17,20,21]. For testosterone, no such
conclude that it was not necessary to use any cryoprotectant to obtain enhancement was observed after nanoencapsulation in starch poly-
reconstituted nanoparticles with similar properties as the original mers, suggesting that transport across the skin barrier remained rate
nanoparticles. Experimental results commented in the previous limiting also for the encapsulated drug. Similar results were obtained
section indicated that both starch derivative nanoparticles present with caffeine, which are in agreement with those previously reported by
different surface composition, being the PS-1.45 nanoparticles Schäfer-Korting et al. [26]. The low Papp values of particle-bound and
enriched in PVA in comparison with the PS-1. Different publications free caffeine can be due to its inherent physicochemical characteristics
have shown that a fraction of this PVA is strongly attached to the [16,46]. Considering Log P and Mw of caffeine the expression calculated
nanoparticle surface [33,44]. Moreover, several authors have demon- by Potts and Guy [46] estimates a Papp ~0.03·10− 6 cm/s for this
strated that this PVA shell can enhance the nanoparticles stability molecule, which is very close to our result.
during freeze-drying, allowing to omit additional cryoprotectants Overall, the skin permeation data for the three drugs suggest that
[31]. Accordingly, our data corroborate that the excess of PVA starch nanoparticles have potential for transdermal drug delivery
presented in the surface of the PS-1.45 nanoparticles, in comparison applications. Although no attempt was made to further investigate the
with PS-1, leads to a better stability. interaction of the nanoparticles with the skin barrier itself, the low drug
To complete the characterization of these systems we have also permeation obtained with testosterone and caffeine and the physico-
evaluated their cytotoxicity. At least in the tested concentration range, chemical characteristics of PS-1 and PS-1.45 nanoparticles hint that
none of the two polymers showed any measurable cytotoxic effect, these colloidal systems themselves did not cross the HSE [16].
neither with the LDH nor with the MTT assay.
Finally, encapsulation and release properties of PS-1 and PS-1.45
nanoparticles were analyzed. Encapsulation of FFA, testosterone or 5. Conclusions
caffeine did not produce a clear effect in the size and PDI of both
colloidal systems, see Table 3 and Fig. 2a–d. On the other hand, In this study, the formulation of nanoparticles by the emulsion
considering the chemical characteristics of the starch derivatives and diffusion technique using propyl-starch derivatives with two different
encapsulated drugs, EE summarized in Table 2 can be classified as a Ds was analyzed. From the colloidal stability, the different surface
function of the hydrophobic character of the encapsulated macro- characteristics were determined, indicating a higher amount of PVA on
molecule, being higher and very similar for FFA and testosterone (Log the nanoparticles surface prepared with the more hydrophobic
P ~5 and ~ 3.5, respectively) and lower for the less hydrophobic drug, polymer, PS-1.45. Superficial differences between PS-1 and PS-1.45
caffeine (Log P ~− 0.1). nanoparticles determined their lyophilization, since PS-1.45, the system
Figs. 5 and 6 depict the release profile of the three drugs from the PS-1 with a PVA enriched surface, recovered better their initial properties.
and PS-1.45 nanoparticles respectively. To simplify the discussion, PS-1 In addition, we explored the potential of nanoparticles formulated
nanoparticles will be firstly analyzed. The three release profiles with propyl-starch derivatives for transdermal drug delivery. Both
illustrated in Fig. 5 can be clearly separated in two different groups. nanoparticle systems, PS-1 and PS-1.45, showed high encapsulation
The first group was composed by FFA and testosterone, while caffeine efficiency for a wide range of chemically different drugs. The drug
forms the second one. As was commented above, FFA and testosterone release was driven by hydrophobic character of both the polymer and
are relatively hydrophobic (Log P ~5 and ~3.5, respectively), while the drug. Both systems showed a close to linear release pattern for
caffeine is much more hydrophilic (Log P ~−0.1) and moreover has a hydrophobic drugs. Finally, PS-1 and PS-1.45 nanoparticles showed a
positive net charge under release conditions (pkb ~10.5). Hence, from remarkable enhancer effect for FFA across the skin barrier.
Fig. 5 can be concluded that the release pattern from PS-1 nanoparticles
was mainly controlled by the hydrophobic interaction between the
encapsulated macromolecule and the nanoparticle matrix. Considering Acknowledgements
the starch derivative structure, without any ionisable groups, it appears
that the degree of drug ionisation did not much affect drug release, The authors wish to thank to Dr. Kostka (Caritas-Krankenhaus
which instead seemed to be mainly governed by drug-matrix hydro- Lebach, Germany) for the supply with human skin samples as well as the
phobic interactions [45]. financial support given by Galenos Fellowship in the Framework of the
Fig. 6 illustrates the release profiles of PS-1.45 nanoparticles. In this EU Project “Towards a European PhD in Advanced Drug Delivery”, Marie
case three drugs displayed three different release patterns. Release Curie Contract MEST-CT-2004-404992; the German Bundesminister-
velocity followed the sequence caffeine>>testosterone>FFA. Consider- ium für Bildung und Forschung (BMBF) program “Nanochance” project
ing the drug characteristics described above, this sequence fits perfectly number 13N9133 and the companies Toroma Organics Ltd, and BASF
with the drug's hydrophilic character, caffeine>>testosterone>FFA, and the Spanish “Ministerio de Ciencia e Innovación” (MICINN) project
independently of the ionic drug state. Hence, for both nanoparticles number MAT2007-66662-C02-01.
suspensions the control of the release pattern can be explained on the
basis of the drug-polymer hydrophobic interactions, which were more
intense for the PS-1.45 case. This can be attributed to the higher Appendix A. Supplementary data
hydrophobic character of PS-1.45 matrix with respect to PS-1 [45]. This
allows discrimination between FFA and testosterone, in spite of their Supplementary data associated with this article can be found, in
similar Log P, being more favourable the FFA-PS-1.45 than the the online version, at doi:10.1016/j.jconrel.2009.08.012.
NANOMEDICINE
92 M.J. Santander-Ortega et al. / Journal of Controlled Release 141 (2010) 85–92

References [25] D. Quintanar-Guerrero, E. Allemann, H. Fessi, E. Doelker, Preparation techniques and


mechanisms of formation of biodegradable nanoparticles from preformed polymers,
Drug Development and Industrial Pharmacy 24 (12) (1998) 1113–1128.
[1] R. Gref, J. Rodrigues, P. Couvreur, Polysaccharides grafted with polyesters: novel [26] M. Schafer-Korting, U. Bock, A. Gamer, A. Haberland, E. Haltner-Ukomadu, M. Kaca, H.
amphiphilic copolymers for biomedical applications, Macromolecules 35 (27) Kamp, M. Kietzmann, H.C. Korting, H.U. Krachter, C.M. Lehr, M. Liebsch, A. Mehling, F.
(2002) 9861–9867. Netzlaff, F. Niedorf, M.K. Rubbelke, U. Schafer, E. Schmidt, S. Schreiber, K.R. Schroder, H.
[2] R. Fernandez-Urrusuno, P. Calvo, C. Remunan-Lopez, J.L. Vila-Jato, M.J. Alonso, Spielmann, A. Vuia, Reconstructed human epidermis for skin absorption testing:
Enhancement of nasal absorption of insulin using chitosan nanoparticles, results of the German prevalidation study, Atla-Alternatives to Laboratory Animals 34
Pharmaceutical Research 16 (10) (1999) 1576–1581. (3) (2006) 283–294.
[3] P. Calvo, C. Remunan-Lopez, J.L. Vila-Jato, M.J. Alonso, Chitosan and chitosan/ [27] T. Lopez-Leon, A. Elaissari, J.L. Ortega-Vinuesa, D. Bastos-Gonzalez, Hofmeister effects
ethylene oxide-propylene oxide block copolymer nanoparticles as novel carriers on poly(NIPAM) microgel particles: macroscopic evidence of ion adsorption and
for proteins and vaccines, Pharmaceutical Research 14 (10) (1997) 1431–1436. changes in water structure, Chemphyschem 8 (1) (2007) 148–156.
[4] J.J. Listinsky, G.P. Siegal, C.M. Listinsky, alpha-L-fucose — a potentially critical [28] M.J. Santander-Ortega, N. Csaba, M.J. Alonso, J.L. Ortega-Vinuesa, D. Bastos-Gonz
molecule in pathologic processes including neoplasia, American Journal of Clinical lez, Stability and physicochemical characteristics of PLGA, PLGA:poloxamer and
Pathology 110 (4) (1998) 425–440. PLGA:poloxamine blend nanoparticles: a comparative study, Colloids and Surfaces
[5] E. Osterberg, K. Bergstrom, K. Holmberg, T.P. Schuman, J.A. Riggs, N.L. Burns, J.M. A: Physicochemical and Engineering Aspects 296 (1–3) (2007) 132–140.
Vanalstine, J.M. Harris, Protein-rejecting ability of surface-bound dextran in end- [29] M.J. Santander-Ortega, A.B. Jodar-Reyes, N. Csaba, D. Bastos-Gonzalez, J.L. Ortega-
on and side-on configurations — comparison to Peg, Journal of Biomedical Vinuesa, Colloidal stability of Pluronic F68-coated PLGA nanoparticles: a variety of
Materials Research 29 (6) (1995) 741–747. stabilisation mechanisms, Journal of Colloid and Interface Science 302 (2) (2006)
[6] C. Lemarchand, R. Gref, P. Couvreur, Polysaccharide-decorated nanoparticles, 522–529.
European Journal of Pharmaceutics and Biopharmaceutics 58 (2) (2004) 327–341. [30] J.A. Molina-Bolivar, J.L. Ortega-Vinuesa, How proteins stabilize colloidal particles
[7] A.P. Marques, R.L. Reis, J.A. Hunt, The biocompatibility of novel starch-based by means of hydration forces, Langmuir 15 (8) (1999) 2644–2653.
polymers and composites: in vitro studies, Biomaterials 23 (6) (2002) 1471–1478. [31] W. Abdelwahed, G. Degobert, S. Stainmesse, H. Fessi, Freeze-drying of nanopar-
[8] M.A. Araujo, A.M. Cunha, M. Mota, Enzymatic degradation of starch-based ticles: formulation, process and storage considerations, Advanced Drug Delivery
thermoplastic compounds used in protheses: identification of the degradation Reviews 58 (15) (2006) 1688–1713.
products in solution, Biomaterials 25 (13) (2004) 2687–2693. [32] J.W. Fong, France Patent FRA 81 116941981.
[9] F. Delval, G. Crini, S. Bertini, N. Morin-Crini, P.M. Badot, J.L. Vebrel, G. Torri, [33] M.F. Zambaux, F. Bonneaux, R. Gref, P. Maincent, E. Dellacherie, M.J. Alonso, P.
Characterization of crosslinked starch materials with spectroscopic techniques, Labrude, C. Vigneron, Influence of experimental parameters on the characteristics
Journal of Applied Polymer Science 93 (6) (2004) 2650–2663. of poly(lactic acid) nanoparticles prepared by a double emulsion method, Journal
[10] J.S. Rodrigues, N.S. Santos-Magalhaes, L.C.B.B. Coelho, P. Couvreur, G. Ponchel, R. of Controlled Release 50 (1–3) (1998) 31–40.
Gref, Novel core (polyester)-shell(polysaccharide) nanoparticles: protein loading [34] P.D. Scholes, A.G.A. Coombes, L. Illum, S.S. Davis, M. Vert, M.C. Davies, The
and surface modification with lectins, Journal of Controlled Release 92 (1–2) preparation of sub-200 nm poly(lactide-co-glycolide) microspheres for site-
(2003) 103–112. specific drug delivery, Journal of Controlled Release 25 (1–2) (1993) 145–153.
[11] S. Hornig, T. Heinze, Nanoscale structures of dextran esters, Carbohydrate [35] T. Tadros, in: T.F. Tadros (Ed.), Colloid Stability: The Role of the Surface Forces, Part
Polymers 68 (2) (2007) 280–286. 1, vol. 1, Wiley-vch Verlag, Weinheim, 2007.
[12] C.K. Simi, T.E. Abraham, Hydrophobic grafted and cross-linked starch nanoparticles [36] M.J. Santander-Ortega, D. Bastos-Gonzalez, J.L. Ortega-Vinuesa, Electrophoretic
for drug delivery, Bioprocess and Biosystems Engineering 30 (3) (2007) 173–180. mobility and colloidal stability of PLGA particles coated with IgG, Colloids and
[13] N. Teramoto, T. Motoyama, R. Yosomiya, M. Shibata, Synthesis and properties of Surfaces B-Biointerfaces 60 (1) (2007) 80–88.
thermoplastic propyl-etherified amylose, European Polymer Journal 38 (7) (2002) [37] M. Chacon, J. Molpeceres, L. Berges, M. Guzman, M.R. Aberturas, Stability and
1365–1369. freeze-drying of cyclosporine loaded poly(D,L lactide-glycolide) carriers, Europe-
[14] C. Lemarchand, P. Couvreur, M. Besnard, D. Costantini, R. Gref, Novel polyester- an Journal of Pharmaceutical Sciences 8 (2) (1999) 99–107.
polysaccharide nanoparticles, Pharmaceutical Research 20 (8) (2003) 1284–1292. [38] D. Quintanar-Guerrero, A. Ganem-Quintanar, E. Allemann, H. Fessi, E. Doelker,
[15] W. Smith, H.I. Maibach, Francis Taylor, Percutaneous Penetration Enhancers, CRC Influence of the stabilizer coating layer on the purification and freeze-drying of
Press, 2006. poly(D,L-lactic acid) nanoparticles prepared by an emulsion-diffusion technique,
[16] G. Cevc, Lipid vesicles and other colloids as drug carriers on the skin, Advanced Journal of Microencapsulation 15 (1) (1998) 107–119.
Drug Delivery Reviews 56 (5) (2004) 675–711. [39] M. Sameti, G. Bohr, M.N.V.R. Kumar, C. Kneuer, U. Bakowsky, M. Nacken, H.
[17] R. Alvarez-Roman, A. Naik, Y.N. Kalia, R.H. Guy, H. Fessi, Enhancement of topical Schmidt, C.M. Lehr, Stabilisation by freeze-drying of cationically modified silica
delivery from biodegradable nanoparticles, Pharmaceutical Research 21 (10) nanoparticles for gene delivery, International Journal of Pharmaceutics 266 (1–2)
(2004) 1818–1825. (2003) 51–60.
[18] J. Luengo, Human Skin Drug Delivery Using Biodegradable PLGA-Nanoparticles, [40] Y.H. Liao, M.B. Brown, A. Quader, G.P. Martin, Protective mechanism of stabilizing
Saarland University, 2007. excipients against dehydration in the freeze-drying of proteins, Pharmaceutical
[19] J. Luengo, B. Weiss, M. Schneider, A. Ehlers, F. Stracke, K. Konig, K.H. Kostka, C.M. Research 19 (12) (2002) 1854–1861.
Lehr, U.F. Schaefer, Influence of nanoencapsulation on human skin transport of [41] J.H. Crowe, F.A. Hoekstra, L.M. Crowe, Anhydrobiosis, Annual Review of Physiology
flufenamic acid, Skin Pharmacology and Physiology 19 (4) (2006) 190–197. 54 (1992) 579–599.
[20] R. Toll, U. Jacobi, H. Richter, J. Lademann, H. Schaefer, U. Blume-Peytavi, [42] A. Simperler, A. Kornherr, R. Chopra, P.A. Bonnet, W. Jones, W.D.S. Motherwell, G.
Penetration profile of microspheres in follicular targeting of terminal hair follicles, Zifferer, Glass transition temperature of glucose, sucrose, and trehalose: an
Journal of Investigative Dermatology 123 (1) (2004) 168–176. experimental and in silico study, Journal of Physical Chemistry B 110 (39) (2006)
[21] R. Alvarez-Roman, A. Naik, Y. Kalia, R.H. Guy, H. Fessi, Skin penetration and 19678–19684.
distribution of polymeric nanoparticles, Journal of Controlled Release 99 (1) (2004) [43] L.M. Crowe, D.S. Reid, J.H. Crowe, Is trehalose special for preserving dry
53–62. biomaterials? Biophysical Journal 71 (4) (1996) 2087–2093.
[22] C. Fringant, J. Desbrieres, M. Rinaudo, Physical properties of acetylated starch- [44] S.K. Sahoo, J. Panyam, S. Prabha, V. Labhasetwar, Residual polyvinyl alcohol associated
based materials: relation with their molecular characteristics, Polymer 37 (13) with poly (D,L-lactide-co-glycolide) nanoparticles affects their physical properties
(1996) 2663–2673. and cellular uptake, Journal of Controlled Release 82 (1) (2002) 105–114.
[23] L. Tuovinen, S. Peltonen, K. Jarvinen, Drug release from starch-acetate films, [45] D.Y. Arifin, L.Y. Lee, C.H. Wang, Mathematical modeling and simulation of drug
Journal of Controlled Release 91 (3) (2003) 345–354. release from microspheres: implications to drug delivery systems, Advanced Drug
[24] L. Tuovinen, E. Ruhanen, T. Kinnarinen, S. Ronkko, J. Pelkonen, A. Urtti, S. Peltonen, Delivery Reviews 58 (12–13) (2006) 1274–1325.
K. Jarvinen, Starch acetate microparticles for drug delivery into retinal pigment [46] R.O. Potts, R.H. Guy, Predicting skin permeability, Pharmaceutical Research 9 (5)
epithelium — in vitro study, Journal of Controlled Release 98 (3) (2004) 407–413. (1992) 663–669.

You might also like