Vous êtes sur la page 1sur 10

International Journal of Pharmaceutics 455 (2013) 219–228

Contents lists available at ScienceDirect

International Journal of Pharmaceutics


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

Pharmaceutical nanotechnology

Chitosan nanoparticles: Preparation, size evolution and stability


Antonio Rampino a , Massimiliano Borgogna a , Paolo Blasi b,∗∗ ,
Barbara Bellich a , Attilio Cesàro a,∗
a
Laboratory of Physical and Macromolecular Chemistry, Dept. Life Sciences, University of Trieste, Via Giorgieri 1, 34127 Trieste, Italy
b
Dept. of Chemistry and Technology of Drugs, University of Perugia, Via del Liceo 1, 06123 Perugia, Italy

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

Article history: Purpose: Characterisation of chitosan-tripolyphosphate nanoparticles is presented with the aim of corre-
Received 2 May 2013 lating particle shape and morphology, size distribution, surface chemistry, and production automatisation
Received in revised form 9 July 2013 with preparation procedure, chitosan molecular weight and loaded protein.
Accepted 10 July 2013
Methods: Nanoparticles were prepared by adding drop wise a tripolyphosphate-pentasodium solution to
Available online 22 July 2013
chitosan solutions under stirring. Trehalose, mannitol and polyethylene-glycol as bioprotectants were
used to prevent particle aggregation and to reduce mechanical stress during freezing and drying pro-
Keywords:
cesses.
Chitosan nanoparticles
Ionotropic gelation
Results: As a novel result, time evolution of the particle size distribution curve showed the presence of a
Nanoparticle ageing bimodal population composed of a fraction of small particles and of a second fraction of larger particles
Freeze-drying attributed to the rearrangement of particles after the addition of tripolyphosphate. Storage for 4 weeks
Spray drying resulted in a slight increase in average size, due to the continuous rearrangement of small particles.
Protein carriers Improvement of nanoparticle stability after lyophilisation and spray-drying was observed in the presence
of all bioprotectants. Trehalose was the best protectant for both methods. Finally, in vivo tests using chick
embryos assessed the biocompatibility of chitosan, tripolyphosphate and the nanoparticles.
Conclusion: The simple ionotropic gelation method with low-MW chitosan was effective in achieving
reproducible nanoparticles with the desired physico-chemical and safety characteristics.
© 2013 Elsevier B.V. All rights reserved.

1. Introduction 2001; Prego et al., 2005) and to modulate drug pharmacokine-


tics, thus improving efficacy and reducing drug toxicity (Kammona
The study of novel pharmaceutical formulations is gradually and Kiparissides, 2012; Ricci et al., 2004; Rosen and Abribat, 2005;
abandoning the parenteral route and is developing medicines Shmulewitz et al., 2006; Zhang et al., 2008).
administered by alternative routes in order to avoid the typical A wide range of materials, such as natural and synthetic
disadvantages of the former. In particular, among the non-invasive polymers, lipids, and surfactants, have been employed to prepare
administration routes, the mucosal route, such as the nasal, pul- drug-containing nanocarriers (Blasi et al., 2007; Duncan, 2003;
monary, oral and vaginal routes, is attracting a great deal of interest, Kumari et al., 2010). Among the materials proposed for mucosal
thanks to its enabling local delivery to target tissues or its deliv- delivery, polysaccharides have received increasing attention
ering the active pharmaceutical ingredient (API) to the systemic because of their outstanding physical and biological properties (Liu
circulation (Andrews et al., 2009). et al., 2008). In particular, cationic polymers such as chitosan and
Micro and nanoparticles are the most promising of the deliv- its derivatives generate particular interest. It is a natural biopoly-
ery systems showing potential for the mucosal delivery of drugs mer consisting of ␤-1 → 4 linked 2-amino-2-deoxy-glucopyranose
and antigens. First described for pharmaceutical applications by (GlcN) and 2-acetamido-2-deoxy-␤-d-glucopyranose (GlcNAc)
Birrenbach and Speiser (1976), nanostructured carriers of appro- residues, manufactured commercially on a large scale by alka-
priate size and surface charge should be able to protect drugs line N-deacetylation of chitin, an abundant biopolymer isolated
from enzymatic degradation (Des Rieux et al., 2006), to improve from the exoskeleton of crustaceans, such as crabs and shrimps
their penetration across the mucosal epithelium (Lamprecht et al., (Tømmeraas et al., 2002). The preparation of chitosan with
controlled properties by means of micro-organism (i.e. fungi) fer-
mentation is of increasing interest (Niederhofer and Müller, 2004;
∗ Corresponding author. Tel.: +39 0405583684; fax: +39 0405583691. Wang et al., 2008). Chitosan offers several advantages for mucosal
∗∗ Corresponding author. Tel.: +39 0755852057; fax: +39 0755855123. delivery, such as low toxicity and good biodegradability (Lee et al.,
E-mail addresses: paolo.blasi@unipg.it (P. Blasi), cesaro@units.it (A. Cesàro). 1995), as well as immunostimulating (Nishimura et al., 1986) and

0378-5173/$ – see front matter © 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.ijpharm.2013.07.034
220 A. Rampino et al. / International Journal of Pharmaceutics 455 (2013) 219–228

high mucoadhesive properties (Lehr et al., 1992). Chitosan has macromolecular properties as reported elsewhere (Rampino,
been used as an excipient in different formulations, such as tablets, 2011). The experiments and the analysis of the polymer conforma-
matrix, micro- and nanoparticles (Dash et al., 2011). tional features (molecular weight, hydrodynamic radius, intrinsic
The cationic amino groups of chitosan complex with metal viscosity and viscosity exponent parameter) were carried out with
anions or small multiple-charged anionic molecules, such as a size exclusion chromatography (SEC) system (SEC3, Viscotek,
sulphates, citrates, and phosphates (Dambies et al., 2001). The for- USA) in the laboratory of UFT – Centre for Environmental Research
mation of particles using ionic gelation is advantageous because and Sustainable Technology, Bremen, Germany. The description of
the process is relatively simple and mild and avoids the use of the SEC3 system experiments and of the analysis of raw data has
organic solvents and high temperatures, thus allowing the success- already been described elsewhere (Weinhold et al., 2009). A full
ful encapsulation of fragile molecules, such as proteins (Al-Qadi report of the methods and analysis of the results of the solution
et al., 2012; Berger et al., 2004; Nasti et al., 2009; Xu and Du, conformational properties of the two chitosan samples and some
2003). In particular, many researchers have explored the potential chitosan derivatives will be reported shortly (Rampino et al., 2013,
pharmaceutical usage of tripolyphosphate (TPP)-chitosan com- in preparation).
plexes after Bodmeier and Pramar reported the preparation of
beads by dropping chitosan into a TPP solution (Bodmeier et al., 2.3. Preparation of the nanoparticles
1989). The prepared hydrogels provide the opportunity to for-
mulate carriers with tunable geometries, positive surface charge, The NPs were prepared following the procedure described by
and, generally, with a good capacity of API encapsulation. How- Calvo et al. (1997). A 2.5 mg/mL chitosan solution was prepared by
ever, nanoparticles (NPs) prepared by ionotropic gelation methods dissolving LMW or VLMW chitosan in a 0.05% (v/v) acetic acid solu-
are usually characterised by aggregation/particle fusion immedi- tion and leaving it under stirring for 24 h. The pH was adjusted to 5.5
ately after preparation or by limited physical/chemical stability with a 0.5 M sodium hydroxide solution and diluted in deionised
when nanoparticle suspensions are stored for an extended period water to the final desired concentrations. TPP was dissolved in
of time. In addition, the technological development of nanocarri- deionised water to a final concentration of 0.25 mg/mL. TPP and
ers is paralleled by the increasing concern about their safety and chitosan solutions were filtered through a 0.45 ␮m membrane (Mil-
possible toxic effect: both the carrier and the materials used must lipore). Then, the TPP solution was added to the chitosan solution
be biocompatible and biodegradable and particular attention is drop wise (0.3 mL/min) at different TPP:chitosan ratios under vig-
placed on the side effects of nanoentities (Aggarwal et al., 2009; orous magnetic stirring at room temperature (Calvo et al., 1997).
Bystrzejewska-Piotrowska et al., 2009). To accomplish this aim, it The resulting suspension was then left to gelify for 30 min.
is worth exploring the possibility of using biopolymeric nanopar- The dynamic evolution of particle formation was investigated by
ticles for oral drug delivery applications (Food Safety Authority of monitoring particle size (as described in Section 2.6.1) over time
Ireland, 2008). on the formulation with the best characteristics. After the com-
The aim of this work was the preparation and characterisa- plete addition of TPP, one set of samples was left under stirring and
tion of chitosan-TPP nanoparticles as mucosal drug carriers using another was removed from the stirrer, and both were left at room
the ionotropic gelation method, as well as their optimisation by temperature. Particle size was measured on freshly prepared parti-
monitoring particle shape and morphology, size distribution, sur- cles, after 2 and 24 h from preparation. Moreover, the possibility of
face charge, and drug loading capacity. Different proteins have a scale-up has been explored increasing both the final volume (10
been encapsulated as model compounds and attempts to scale times) and the component concentration (3 times).
up the production methodology and to produce re-dispersible dry Protein-loaded nanoparticles were prepared as follows: differ-
powders by freeze-drying and spray-drying have been performed ent volumes of the protein stock solution (4 mg/mL) in deionised
as well. Finally, chitosan and chitosan NPs biocompatibility was water were added to the chitosan solution, obtaining a final protein
investigated by chorioallantoic membrane (CAM) assay to fulfil the concentration of 0.2, 0.4, and 0.6 mg/mL. Only LMW chitosan was
current request for nanoparticle safety assessment. employed to prepare nanoparticles loaded with the three differ-
ent proteins: bovine serum albumin (BSA), ovalbumin (OVA) and
2. Materials and methods human insulin (HI). BSA and OVA were added directly to the chi-
tosan solution, while HI was added to the TPP solution and then
2.1. Materials dripped. The dispersion was left under constant stirring for 30 min
at room temperature, in order to allow a homogeneous gelation.
Low molecular weight (LMW) chitosan, tripolyphosphate pen-
tasodium (TPP), albumin from bovine serum (BSA), albumin from 2.4. Nanoparticle yield
chicken egg white (OVA), trehalose from corn starch, mannitol
and polyethylene glycol 10000 (PEG 10000) were purchased from To remove the excess unreacted chitosan, the suspension was
Sigma–Aldrich Co. (St. Louis, Missouri). Very low molecular weight centrifuged and the supernatant was recovered for further anal-
(VLMW) chitosan was prepared by controlled sodium nitrite degra- ysis. The best separation condition, giving the highest amount of
dation, starting from medium molecular weight chitosan (MMW) particle recovery, was 3270 RCF (Relative Centrifugal Force) for 2 h.
(Sigma–Aldrich Co. St. Louis, Mo), as elsewhere reported (Janes and By increasing the speed or time of centrifuging, higher amounts of
Alonso, 2003) with slight modifications. Briefly, a 1% (w/v) NaNO2 precipitate could be obtained, but it was not possible to recover
solution was added to a 1% (w/v) chitosan acid solution (ratio 1:15) nanoparticles even using an ultrasonic bath or an ultrasonic probe.
at room temperature under magnetic stirring. The reaction was Moreover, the addition of a glycerol bed, generally used to prevent
maintained for 1 h. Zinc-free human insulin (HI) was donated by particle aggregation due to centrifuge packing, was not suitable for
Novo Nordisk, (Copenhagen, Denmark). All other chemicals and the complete recovery of nanoparticles.
reagents were of the highest purity grade commercially available. Chitosan quantitative determination in the supernatant was
performed by employing the colorimetric protocol set-up by
2.2. Polymers characterisation Muzzarelli (1998) with very slight modifications. The buffer solu-
tion was prepared with a final pH of 3 and the standard solutions
The two polymer samples, LMW-chitosan and VLMW- of chitosan were prepared starting from a 0.5 mg/mL stock solu-
chitosan have been fully characterised for their chemical and tion containing 0.05% acetic acid. Absorbance was measured at
A. Rampino et al. / International Journal of Pharmaceutics 455 (2013) 219–228 221

575 nm with a UV–visible spectrophotometer (Cary 4E) in triplicate 2.8. Chorioallantoic membrane (CAM) assay
and results were expressed as mean value and standard deviation
(±SD). The in vivo biocompatibility of the raw materials and NPs was
evaluated using chick embryo chorioallantoic membrane assay, a
2.5. Nanoparticle recovery and drying test that has been used in the past and is receiving renewed atten-
tion (Saw et al., 2008; Vargas et al., 2007). Fertilised eggs were first
NP suspensions were centrifuged for 2 h at 3270 RCF and the disinfected with alcohol 70◦ before being placed in the incubator
obtained sediment was resuspended in deionised water, analysed at 38 ◦ C with 60% humidity. At incubation day 3, a window opening
and lyophilised or alternatively spray dried, using trehalose, man- was punctured at the blunt end of the egg and living embryos were
nitol and PEG 10000 as cryoprotectants. Dried particles were selected for the experiment. The opening was then covered with
redispersed and characterised in terms of size, surface charge and a polyethylene film glued with albumen to avoid water loss and
morphology. contamination. At incubation day 6, solid samples (6 mm tablets of
To prepare spray-dried nanoparticles a new kind of spray- raw materials and blank nanoparticles) were applied directly on the
dryer was used, the Nano Spray Dryer B-90 (Buchi, Switzerland). CAM (Ribatti et al., 1996; Schoubben et al., 2013). A Leica WILD M32
The spray cap with a 4.0 ␮m mesh hole size was used in order stereomicroscope (equipped with a WILD PLAN 1X lens), connected
to obtain the smallest droplet size. The operating conditions to a Leica DFC 320 camera system, was used to follow the evolution
were set up as follows: inlet temperature at 120 ◦ C, drying air of any effects on the materials on the CAM. After 24 h, all the eggs
flow rate = 100 L/min, pressure, 50 mbar; pump, 1; spray, 100% were examined again and all the acquired images were compared
(Schoubben et al., 2013). with the previous ones for qualitative acute toxicity evolution.

2.6. Nanoparticle characterisation


3. Results and discussion
2.6.1. Particle size and morphology
NP size and ␨ potential were determined using a Malvern 3.1. Nanoparticles preparation and characterisation
Zetasizer Nano ZS (Malvern Instruments, Worcestershire, UK).
Diluted samples were placed in disposable polystyrene cuvettes Chitosan nanoparticles were prepared by ionotropic gelation
and the scatter intensity was measured at 25 ◦ C. Size values were with the dropwise addition of tripolyphosphate (TPP) to a chitosan
reported as mean hydrodynamic diameter (MHD), standard devi- solution (concentration 2.5 mg/mL). Preliminary experiments were
ation (SD), and polydispersity index (PDI). Chitosan nanoparticles addressed to the selection of the optimal ratios for the formation
were observed by transmission electron microscope (TEM) (Philips of nanoparticles of TPP and of two chitosan samples with different
EM 208 microscope, Eindhoven, Netherlands). All samples for TEM molecular weights. For this purpose, the appearance of opalescence
analysis were prepared by allowing a single drop of nanoparticle was used as an indicator of nanoparticle formation, which was also
suspension to dry overnight at room temperature on a Formvar® confirmed by means of dynamic light scattering. Low molecular
coated 200-mesh copper grid (TAAB Laboratories Equipment Ltd., weight (LMW) chitosan had a weight-average molecular weight
Aldermaston, England). Spray-dried chitosan nanoparticles were (Mw) of 150 kDa, with an intrinsic viscosity [] of 2.37 dL/g and
observed by scanning electron microscopy (SEM) (Philips XL30 a hydrodynamic radius Rh of 15.8 nm in a 0.3 M aqueous acetate
SEM, Eindhoven, Netherlands). The dry powder was placed onto buffer solution at pH 4.5. Very low molecular weight (VLMW) chi-
an aluminium specimen stub covered with a double-sided carbon tosan showed a Mw of 32 kDa, with an intrinsic viscosity of 0.6 dL/g
adhesive disc (Taab, Berkshire, UK) and sputter coated with gold and a hydrodynamic radius of 6.4 nm, in the same solvent.
(20 kV for 3 min). The effect of chitosan Mw and ratio with respect to TPP on par-
ticle characteristics is reported in Table 1. Nanoparticle formation
2.6.2. Fourier transform-infrared (FT-IR) spectroscopy began at chitosan:TPP weight ratio of 6:1. The ability of chitosan
IR spectra (range 5000–600 cm−1 ) were recorded on lyophilised to gel rapidly upon contact with TPP relied on the formation of
samples using a Vertex 70 (Bruker Optics GmbH) spectrophotome- inter- and intramolecular cross-linkages mediated by the anionic
ter (spectral resolution of 4 cm−1 ) equipped with a MIRacleTM molecule (De Campos et al., 2001; Xu and Du, 2003). By increas-
ATR device (Pike Optics) with a single reflection diamond crys- ing the amount of TPP, the nanoparticle suspension became more
tal (1.8 mm spot size) and using a MCT detector (HgCdTe, and more turbid (chitosan NP formation) but particle aggregation
mercury-cadmium-tellurium) cooled with liquid nitrogen. Both occurred rapidly and drastically. When the available quantity of
raw materials and formulations were analysed by attenuated total TPP was high, the dominantly inter- and intramolecular cross-links
reflectance (ATR). The samples were deposited on the top of a dia- were associated with TPP which enabled NPs to form larger par-
mond crystal and secured with a high-pressure clamp. ticles and large flocculating aggregates (chitosan/TPP mass ratio
lower than 4:1). The suspensions formed in the range between
the appearance of the opalescence and flocculation were deeply
2.7. Determination of protein loading efficiency
analysed in terms of particle size and reproducibility. Some of
them gave unpredictable results in terms of particle formation or
The loading efficiency was obtained by determining the concen-
flocculation: this is a common behaviour, as with polysaccharide
tration of the protein in the supernatant using the bicinchoninic
preparations, batch-to-batch variation appears to be an unavoid-
acid (BCA) colorimetric assay (Sigma–Aldrich Co., St. Louis, Mis-
able consequence of the polydispersity of the macromolecules, and
souri). The amount of protein loaded in the NPs was calculated as
has a determinant impact on the physicochemical properties of the
the difference between the target loading and the protein recovered
final product (Morris et al., 2011).
in the supernatant. The supernatant of unloaded nanoparticles was
The interaction at molecular lever between chitosan chains and
used as a blank, in order to subtract the interference of the compo-
TPP was evidenced by ATR FT-IR spectra of chitosan NPs compared
nents of loaded samples with the BCA. Loading efficiency (LE) was
with those of native chitosans (data not shown). The results were
calculated using the following equation:
consistent with previous published data (Dudhani and Kosaraju,
target loading − unloaded protein 2010). A broad peak between 3350 and 3270 cm−1 was attributed
LE(%) = × 100 (1)
target loading to a combination of stretching modes of O H and N H bonds in
222 A. Rampino et al. / International Journal of Pharmaceutics 455 (2013) 219–228

Table 1
Mean hydrodynamic diameter (MHD) and surface charge of VLMW and LMW chitosan-TPP NP prepared at different chitosan:TPP mass ratios.

Chitosan:TPP ratio (w/w) VLMW chitosan NPs LMW chitosan NPs

MHD ± S.D. (nm) PDI ␨-pot ± S.D. (mV) MHD ± S.D. (nm) PDI ␨-pot ± S.D. (mV)

3:1 –* –*
4:1 –* –*
5:1 151 ± 10 0.185 24 ± 4 200 ± 24 0.227 25 ± 3
6:1 164 ± 12 0.198 28 ± 3 193 ± 28 Not reproducible
*
Aggregation took place and no particles were recovered.

chitosan matrix. In the sample of chitosan nanoparticles this band method yield. Here, the amount of chitosan in the supernatant
becomes wider and shifts to lower wavenumbers, indicating an after centrifugation was assayed using a colorimetric determina-
enhancement of the hydrogen bond interactions (Yu et al., 1999). In tion (Fig. 1). By adding TPP to the chitosan solution, a slight change
addition, the 1523 cm−1 peak of the NH2 bending vibration of chi- in the amount of precipitate could be noticed: this could be due to
tosan samples shifted to 1533 cm−1 in the NPs. A similar result has the presence of small nanoparticles that cannot be easily separated
been observed in literature on chitosan-TPP NPs (Xu and Du, 2003) from the suspension by centrifuging because of their light weight.
and on chitosan films treated with phosphate (NaH2 PO4 ); the shift By increasing the addition of TPP, a sudden increment of the precip-
was attributed to the interaction between the amino group and the itate can be observed in correspondence to 3.5 mL of TPP solution
phosphate anion (Knaul et al., 1999). The other peaks observed in (0.875 mg, chitosan to TPP ratio 7.14:1). This volume corresponds to
the sample of nanoparticles were those of the P O stretching at the start of the opalescence in the chitosan solution, signalling the
1201 cm−1 and P O bending at 885 cm−1 of the TPP. formation of aggregates of a dimension comparable to the incident
The results of these preparations and characterisations indicate light. The amount of precipitate increases with the addition of over
that NPs prepared with a chitosan/TPP ratio of 5:1 turned out to 1.25 mg of TPP (5 mL of anionic solution, chitosan:TPP ratio 5:1),
be the best formulations for both chitosan Mws, with an average where larger aggregates are formed: at this point, the amount of
size of 151 nm in the case of VLMW chitosan. The increase of chi- precipitated chitosan was 0.7 mg out of 6.25 mg in the original solu-
tosan Mw to 150 kDa provoked an increase in the average particle tion. This result indicates that the yield of nanoparticles at this stage
size from 151 to 200 nm. These results were consistent with find- is low (11.2%). It is reasonable to assume that some nanoparticles
ings that higher Mw chitosan produced larger nanoparticles (Csaba remain in the suspension and are not separated under the mild cen-
et al., 2009; Luangtana-anan et al., 2005; Tang et al., 2003). The poly- trifuge conditions (3270 RFC for 2 h). However, while increasing TPP
dispersity index (PDI) was low for all the preparations evaluated, amounts and/or centrifugation conditions (speed or time) generate
indicating that a homogeneous dispersion was obtained. The PDI flocculation or particle packing, recovering unreacted chitosan and
changed from 0.185 to 0.227 in correspondence to the increase in adding further TPP are also worth exploring.
chitosan Mw (Table 1); a higher variety of particle size was obtained
using a solution of larger polymer chains. The distribution curve 3.2. Nanoparticles’ evolution over time
indicates the presence of two different particle size populations:
indeed, the bimodal curve consists of a fraction of small particles As previously mentioned, the size distribution curve of freshly
(around 40 nm) and a second population of around 250 nm. The prepared samples is characterised by two populations of parti-
presence of two populations was also confirmed by TEM. Size anal- cles, one around 40 nm and the other around 250 nm. This typical
ysis was performed also on NPs deriving from a scale-up procedure. distribution is mainly observed in freshly prepared samples (Ing
When changing the component concentration, the measurements et al., 2012); if samples were kept under stirring and analysed
indicated that tripling the total concentration gave larger particles, after 6 h, the smallest particle population disappeared. Indeed the
increasing the size from 200 to 273 nm (SD 18), in agreement with bimodal curve became a single Gaussian curve, spread up to a few
previous results (Gan et al., 2005; Hu et al., 2008). When increasing hundreds of nanometers. This phenomenon could be explained by
of 10 times the preparation volume, a smaller variation in dimen-
sion was measured (average size of 237 ± 27 nm).
Regarding the surface charge, both chitosan samples gave pos-
itively charged nanoparticles. When chitosan and TPP were mixed
with each other, they spontaneously formed compact complexes
with an overall positive surface charge, as confirmed by measures
of zeta potential values. In accordance with literature (Fan et al.,
2012; Janes and Alonso, 2003), by increasing TPP concentration,
the zeta potential decreased (Table 1). Analysing the differences
between the two chitosan samples, the same trend as noticed for
particle size, could also be observed as a function of the molecular
mass: compared to VLMW chitosan samples, LMW chitosan sam-
ples with the same chitosan:TPP mass ratio had a slightly higher
zeta potential. Chitosan samples with different Mws should have
different accessibility to cationic groups, even if they have the same
degree of deacetilation: 87% in the studied chitosans. The result
is a higher surface charge for particles formed by polymers with
a longer chain, in accordance with results obtained by Gan et al.
(2005).
Although there is a large number of studies on chitosan nanopar-
ticle formation in relation to the different polymer amounts used, Fig. 1. Quantity of recovered nanoparticles (in mg) as a function of the amount of
very few of these studies incorporated results on the production TPP added (in mg).
A. Rampino et al. / International Journal of Pharmaceutics 455 (2013) 219–228 223

in solution has a random coil conformation (Weinhold et al.,


2009). TEM images provide a chitosan chain dimension of around
20–40 nm, consistent with a hydrodynamic radius of 15.8 nm,
as reported in section 3.1. Moreover, the chitosan concentration
value of 2.5 mg/mL is close to the limit of the critical concentration
of coil overlap (C*). This means that chitosan chains are quite close
to one another, some of them statistically in contact with others.
By adding a dilute solution of TPP (0.25 mg/mL) a double effect
is obtained. On the one hand, a dilution of the solution generates
more randomly distributed chitosan coils; on the other, anionic
TPP starts interacting with few cationic groups on chitosan chains
folding over themselves (mainly through intramolecular links).
The dimensions of the new entities are comparable to the random
coil chain ones because they are the result of few bent chains
grafted together or even a single gelified random coil which has
increased in size due to the presence of TPP (osmotic pressure).
Adding further TPP onto chitosan leads to both the formation
of new particles, supported by the constant size value, and also
their rearrangement towards a more compact particle structure. At
the end of the addition of polyanions, the negatively charged TPP
interacts with the chitosan surface groups leading to the primary
aggregates (charged chitosan monomoles exceed TPP by 10 to 1).
The partially neutralised positive charges of chitosan still present in
the primary aggregates, favour a rearrangement of chains and the
formations of more compact particles with a smaller size. A further
rearrangement of nanoparticles is obtained with time by an effi-
cient sharing of neutralising anionic charges, producing particles
described by the presence of a Gaussian distribution curve instead
of a bimodal one. Although such a phenomenon is often described
in literature to be a function of the effective charge density and
of the flexibility of the polymer, a quantitative elucidation is still
lacking.
TEM images confirm this hypothesis: particles appear as small
and individual, with a diameter of around 30–40 nm; larger parti-
cles are due to the aggregation of single small particles that tend
to fuse together generating a larger entity. This aspect is clearly
shown by subsequent TEM images (Fig. 2). Air-dried samples are
not completely desiccated, especially not nanogels as chitosan-TPP
particles. During TEM analysis, when magnifying a NP aggregate as
much as possible, a very fast fusion (seconds/minutes) of single par-
ticles into one entity was observed: the heat of the electron beam
promoted intermolecular links thanks to the still present aqueous
environment inside the gel-network. In order to confirm that this
behaviour was due to a linking agent, TEM images of chitosan solu-
tion were collected showing small particles, of around 40 nm in
diameter, but with a completely different behaviour: magnifying on
several close entities, their structure did not change and no fusion
occurred even if the ray was kept on them for long time (data not
shown).
The dimensions of freshly formed single particles were clearly
visible in TEM images of samples with paraformaldehyde. This
agent was added immediately after the TPP dribble, there-
fore before the complete rearrangement of chains and particles.
Paraformaldehyde blocks the particles’ interactions by limiting
Fig. 2. (a–b) TEM image of chitosan nanoparticles in two enlargements. (c) The
their mobility. TEM images showed that most particles were single
image b taken after few minutes: the fusion process is clearly visible.
isolated ones and only few of them were grouped together: this
is because there was not enough time for a complete TPP redis-
the fact that the small particles aggregate during stirring to form tribution and particle rearrangement (data not shown). This was
a more homogeneous population of particles. This phenomenon also confirmed by particle size analysis showing the inhibition of
is clearly visible in TEM images, where particles range from few particle growing during ageing.
tens to few hundreds of nanometers (Fig. 2). TEM provides the Size stability as a function of time has been studied for both
size of almost-dehydrated particles, while dynamic light scatter- LMW and VLMW chitosan nanoparticles. All samples presented an
ing measurements yield an ensemble average of the particle size in increase in size without leading to the formation of large aggregates
suspension. (micrometric aggregates) (Fig. 3). As regards the particles prepared
Taking into account dynamic light scattering and TEM results, with LMW chitosan, after 1 month a slight increase in size was mea-
a mechanism of various steps could be presented. Chitosan sured, showing a 2-step behaviour: a growth in size was observable
224 A. Rampino et al. / International Journal of Pharmaceutics 455 (2013) 219–228

Table 3
Mean hydrodynamic diameter (MHD), surface charge and loading efficiency of
protein-loaded chitosan nanoparticles.

Protein amount Loading MHD ± SD ␨-pot ± SD


(␮g/mL) efficiency ± SD (nm) (mV)
(%)

BSA
200 39 ± 9 200 ± 11 27 ± 3
400 40 ± 9 225 ± 21 23 ± 8
600 60 ± 6 224 ± 25 21 ± 6
OVA
200 48 ± 4 195 ± 13 17 ± 1
400 57 ± 5 202 ± 17 17 ± 1
600 76 ± 4 215 ± 13 16 ± 2
HI
200 55 ± 8 330 ± 36 30 ± 4

(Gan et al., 2005). Samples left at room temperature showed a ten-


dency to aggregate after two weeks. The total amount of chitosan
was in excess compared to TPP (charged monomole: TPP ratio 10
Fig. 3. Particle size as a function of storage time: different Mws of chitosan were
used. Square points refer to LMW chitosan NPs and triangle points to VLMW chitosan
to 1), and waiting for a more efficient rearrangement of particles
NPs. Size increase is expressed as 100 × (d − d0 )/d0 (d refers to diameter at time t and led to the production of large aggregates: in fact, the formation of
d0 to the initial diameter). larger nanoparticles is favoured at a higher chitosan concentration
(Kaloti and Bohidar, 2010).
after the first hour, completing the particle formation over the next
3.3. Effect of protein loading on size and surface charge
24 h: size was increased by 5%. Then, after 1 month, the size slightly
increased by 8% in total. On the other hand, particles prepared using
Three different proteins were loaded onto chitosan nanopar-
VLMW chitosan showed a linear increase in diameter size, with a
ticles: bovine serum albumin (BSA), ovalbumin (OVA) and human
growth of 7 and 24% after 7 and 28 days, respectively.
insulin (HI). BSA and OVA, with a pI of 4.8 and 4.7 respectively, were
The size evolution during storage is ascribed to many fac-
added directly to the cationic solution and then TPP was dripped
tors, such as particle aggregation, which provides a more efficient
on. HI (pI = 5.6), on the other hand, was diluted with TPP (pH 7.5)
rearrangement, the interaction of free polymer chains with the
and then added to the cationic solution, because it has a neutral
particle network, that leads to a reorganisation of intermolecu-
net charge in the chitosan solution (pH 5.5) leading to precipita-
lar entanglements, syneresis, and swelling, due to the presence
tion. As shown in Table 3, the presence of the model protein has a
of TPP that generates an inflow of water by osmosis (López-León
small influence on particle size, increasing it by few nanometers.
et al., 2005; Tang et al., 2003). The size change is usually due to
This can be ascribed to the ionic interaction and ionic cross-linking
the balance between the above-mentioned forces. The time span
between negatively charged BSA (or OVA) and positively charged
in which the growing stage finished depended on the size of the
chitosan under the preparation conditions (∼pH 6). This generates
initial nanoparticle, which in turn depended on the molecular
a reduction in the electrical repulsion among biopolymers, where
weight of the chitosan. Larger NPs, prepared with a chitosan of
modifications in the electrical state cause the particle to shrink.
higher molecular weight, have more intermolecular entanglements
Moreover, TPP links intra- and intermolecular chitosan chains giv-
and inter/intra-molecular hydrogen bondings, producing a stable
ing a more compact structure.
nanogel which therefore needs more time to change the size of the
The decrease in the zeta potential after the addition of pro-
nanoparticle.
tein is reasonably due to the partial deposition of the negatively
The aggregation level increased rapidly within the first 2 h and
charged protein on the particle surface, reducing the total net
continued after 24 h. The polydispersity index, on the contrary,
charge. Similarly, in the case of HI, an ionic interaction of chi-
increased after 2 h (Table 2) but decreased after 24 h. Particles left
tosan occurs with TPP and HI. The incorporation of HI increases
without stirring took longer to reach the equilibrium. Without stir-
particle size, probably due to concurrent competition between TPP
ring, the particle size did not increase significantly, growing by 4%
and HI in their interaction with chitosan, thus causing the particle
after 2 h and 5% after 24 h (Table 2). A small particle population
to swell: when HI is added, it is negatively charged and interacts
was still present after 24 h and the polydispersity index was 0.186,
with -NH3 + groups of chitosan, competing with TPP. After parti-
lower than that of the samples kept under stirring. The particle size
cle formation, the HI net charge is next to neutrality as its pI of
population was less variable because the system was slowly looking
5.6 is very close to the pH of the particle suspension (≈6): this
for the best equilibrium and few large nanoparticles were present
generates, on the one hand, a weaker interaction with chitosan (a
swollen particle) and on the other, less influence on the total surface
Table 2 charge (higher zeta potential compared to chitosan/TPP particles)
Effect of time and stirring on particle size. Size growth is expressed as percentage (Table 3).
of size increase.

Hours after Stirring No stirring 3.4. Lyophilisation


preparation

Size PDI Size PDI Lyophilisation is one of the most commonly used technologies
increase ± SD increase ± SD that ensures a long-term conservation of polymeric nanoparticles.
(%) (%) Indeed, nanoparticles stored as an aqueous suspension undergo
0 0 0.282 0 0.198 solubilisation and/or degradation of the polymers, drug leakage,
2 31 ± 1 0.456 4±1 0.195 desorption or degradation. Dried nanoparticles are usually easily
24 77 ± 1 0.341 5±1 0.186
redispersible, although in some cases a complete redispersion after
A. Rampino et al. / International Journal of Pharmaceutics 455 (2013) 219–228 225

Fig. 4. Effect of cryoprotectants on particle size. Each bar represents the average
size and error bars refer to SD. In the absence of cryoprotectants, the freeze-drying
of untreated samples produces collapsed NPs (see text).

lyophilisation may be difficult to achieve, due to the aggregation


or irreversible fusion of nanoparticles. Furthermore, the crystalli-
sation of ice can induce a mechanical stress on nanoparticles,
leading to their destabilisation. Thus, the addition of a suitable
lyo/cryo protective agent before freezing, prevents nanoparticles
from aggregating. It is well established that sugars are employed
for this purpose. The mechanism by which sugars protect particles
resides in the interaction with the solute via hydrogen bonding,
which maintains the solute in a “pseudo hydrated” state during Fig. 5. SEM images of microparticles derived from the sintering of unprotected
the dehydrating step, thus providing protection from damage chitosan nanoparticles after spray-drying in two magnifications.
during dehydration and subsequent rehydration. Such a protective
interaction is made possible by the ability of the sugars to remain
amorphous during freeze-drying.
All the chitosan nanoparticles prepared were characterised by crystals of water and sugar may exert mechanical forces on the
a difficult redispersion after lyophilisation (Dyer et al., 2002; Lee nanoparticles leading to their fusion (Cesàro et al., 2008). There-
et al., 2004), therefore the effect of adding trehalose, mannitol and fore, in order to avoid this situation, it is mandatory that at least
PEG 10000 was investigated (Blasi et al., 2013; Luangtana-anan some sugar molecules remain molecularly dispersed throughout
et al., 2010). Following the addition of cryoprotectants, all samples the amorphous nanoparticle phase (Abdelwahed et al., 2006). It is
could be re-dispersed properly after lyophilisation. NPs contain- suggested that sugars isolate individual particles in the unfrozen
ing different concentrations of trehalose, mannitol and PEG 10000 fraction, thereby preventing aggregation during freezing above Tg .
were evaluated in terms of size, after lyophilisation (Fig. 4). A com- In this case, vetrification is not required for this effect and the spa-
mon behaviour was found, indeed lyophilised nanoparticles were tial separation of particles within the unfrozen fraction is sufficient
more similar to freshly prepared nanoparticles with increasing con- to prevent aggregation (Allison et al., 2000).
centrations of cryoprotectant. From these results, it emerged that a
complete re-dispersion, that is a balance between the average size 3.5. Spray drying
and polydispersion index, was obtained when sugars or PEG were
added to the nanoparticle suspension at a concentration of at least The effect of spray-drying on the size of nanoparticles was eval-
5% (w/v). uated; the three selected cryoprotectants were added to the NP
Comparing the different cryoprotectants, suspensions contain- preparation at the concentration of 0.5% (w/v). This concentration,
ing mannitol or PEG were characterised by a greater increase in size, which gave an important increase in size after lyophilisation, was
especially at low concentrations. Therefore the best results were chosen in order to compare the effect of the two processes. The
obtained with trehalose. This could be related to its peculiar char- effect of the addition of the three selected cryoprotectants was
acteristics, which are low hygroscopicity, the absence of internal investigated before and after spray-drying. The sole addition of cry-
hydrogen bonds thus allowing a more flexible formation of hydro- oprotectants in NP suspension slightly increased the average size
gen bonds with nanoparticles, a very low chemical reactivity and, with respect to blank nanoparticles.
finally, a higher glass transition temperature Tg (Crowe et al., 1996). The drying process causes particle aggregation, an aspect that
It must be taken into account that the crystallisation of cryopro- was clearly visible in the SEM picture of unprotected nanoparticles
tectants and the formation of a eutectic with ice can cause phase where they can be seen to be fused together, generating micro-
separation in the cryo-concentrated portion of the frozen nanopar- particles (Fig. 5). After spray-drying, an increase in size was found
ticle suspension with no opportunity for a stabilising interaction for all cryoprotectants. Trehalose and PEG gave nanoparticles of
with nanoparticles. Individual nanoparticles in the nanoparticle- a larger size but they were still nano-sized and re-dispersible;
rich phase can interact and form aggregates. Moreover, the growing however particles containing PEG were characterised by a high
226 A. Rampino et al. / International Journal of Pharmaceutics 455 (2013) 219–228

Fig. 6. CAM image taken from the top of the egg after deposition of prepared tablet. (a–b) Effect of chitosan and NPs on chick embryo after 24 h from deposition, respectively.
(c) Effect of TPP on chick embryo: after 2 h from deposition, a haemorrhagic event occurs with the death of the embryo.

polydispersity index after spray-drying. On the contrary, nanopar- 4. Conclusions


ticles with mannitol remained in suspension as macroaggregates
and it was not possible to properly redisperse the sample after All the results on the preparation, characterisation and size sta-
spray-drying. A comparison of the spray-drying results with the dif- bility of chitosan-TPP nanoparticles as drug carriers rely on the
ferent excipients showed that trehalose is preferable with respect attempt to modulate the experimental conditions and to identify
to PEG and mannitol, as it was also for lyophilisation. Thus, the appropriate procedures. Three main follow-out are worth to be
trehalose is believed to be a preferable protectant against NP aggre- mentioned:
gation, as it is for other biomolecules in cell bioprotection (Crowe
et al., 1996). (1) The simple ionotropic gelation method was shown to be effec-
The comparison between lyophilised and spray dried particles tive although involved a complicated form of dynamic intra-
treated with the same amount of protector (0.5%, w/v) shows a and inter-molecular complexes. The concentration and time
similar increase of the average size. Since spray drying is faster, dependence of the TPP addition suggest that average particle
and cost-effective, than lyophilisation, this technique is recom- size depends not only on the TPP concentration, but also on
mended for the dehydration of chitosan NP (Schoubben et al., a rearrangement of the small compact particles towards larger
2010). particles. It is also worth noting the beneficial effect of quiescent
conditions, as opposed to the stirring conditions, in providing
3.6. Chorioallantoic membrane assay smaller and homogeneous particles and in delaying the aggre-
gation of the final suspension.
Renewed interest has been shown in the chorioallantoic mem- (2) The exploration of appropriate techniques of particle drying
brane (CAM) assay, as this can be carried out at a much lower (i.e., lyophilisation and spray-drying) was carried out with the
cost with greater efficiency and faster measurements than other aim of achieving both the stability of the nanoparticles’ char-
in vivo assays (Saw et al., 2008; Schoubben et al., 2013; Vargas acteristics and also the best protection of labile drugs, such as
et al., 2007). Thus, the CAM assay was used for studying the bio- therapeutic peptides and proteins. The addition of bioprotec-
compatibility of materials and NPs, analysing the inflammatory tants in the suspension of nanoparticles proved very effective
and neo-angiogenic effect (Fig. 6) after 24 h from the treatment. in reducing surface attraction and maintaining the nanopow-
In the case of the raw material (chitosan), no significant changes der in a dispersed form. The best results were obtained by using
occurred after 24 h of incubation. In fact, all the embryos were trehalose, which has been advocated in many cases, despite
alive and there was no sign of any vascular change in the CAM, the still controversial interpretation of the underlying mech-
e.g. haemorrhaging, neoangiogenesis or the occurrence of vessels anism of its action (Cesàro, 2006; Golovina et al., 2010; Sakurai
devoid of blood flow (ghost vessels). This is in good agreement with et al., 2008). An in-depth physicochemical study of the pref-
similar works reported in literature, confirming that our materials erential distribution (i.e. solvation) of trehalose surrounding
are non-toxic, as expected and described: chitosan is a polymer the chitosan nanoparticles may add some useful results to the
approved for dietary applications in Japan, Italy and Finland and “trehalose story”.
it is approved by the FDA for use in wound dressings (Kean and (3) Finally, out of all the main positive results, emphasis has been
Thanou, 2010). given to the assessment of the safety of nanoparticles for
Moreover, the prepared nanoparticles did not cause any death or biological membranes with in vivo tests. This point is often
inflammatory response. This is an important point because, unlike overlooked in the experiments producing nanoparticles for
raw materials, they are not always predictable: a great deal of atten- human or animal use, despite the great attention nowadays
tion is paid to nanosized carriers, concerning the effect (toxicity) of assigned in Europe through the Nanosafety Cluster activities
the shape/surface structure of the new nanoentities on cells and (http://www.nanosafetycluster.eu/). As an alternative to mam-
tissues. These results were in accordance with the in vitro toxicity malian cells for in vivo tests, the use of chick embryos confirmed
test (data not shown). the biocompatibility of both chitosan and nanostructures, and
The only sample that gave an adverse effect was the TPP tablet encouraged future studies using this simple and direct tech-
(Fig. 6c). This was an expected result, because the tripolyphosphate nique.
salt increases the pH of the chorioallantoic membrane drastically,
generating a haemorrhagic event and the death of the embryo. On Acknowledgment
the other hand, the amount of TPP present in nanoparticles was sig-
nificantly lower, its availability limited by chitosan ionic interaction This work has been partially carried out within the Euro-
and no sign of any damage was noticed. Furthermore, low amounts pean Project FP6 NanoBioPharmaceutics (NMP 026723-2). Antonio
of TPP are commonly used as a food additive to preserve foods such Rampino was the recipient of a grant from MIUR (Rome) during
as red meats, poultry, and seafood, helping them to retain their his Ph.D. studies on “Polysaccharide-based nanoparticles for drug
tenderness and moisture. delivery”.
A. Rampino et al. / International Journal of Pharmaceutics 455 (2013) 219–228 227

The authors wish to thank Dr. G. Birarda (Advanced Technology Ing, L.Y., Zin, N.M., Sarwar, A., Katas, H., 2012. Antifungal activity of chitosan nanopar-
and Nanosciences – TASC INFM, Area Science Park, Trieste) and Miss ticles and correlation with their physical properties. Int. J. Biomater, 9 pp,
http://dx.doi.org/10.1155/2012/632698, Article ID 632698.
F. Virgilio (University of Trieste) for providing help with the ATR- Janes, K.A., Alonso, M.J., 2003. Depolymerized chitosan nanoparticles for pro-
FTIR and Prof. Carlo Cirotto and Lanfranco Barberini for support in tein delivery: preparation and characterization. J. App. Polym. Sci. 88,
the CAM tests. 2769–2776.
Kaloti, M., Bohidar, H.B., 2010. Kinetics of coacervation transition versus nanopar-
ticle formation in chitosan–sodium tripolyphosphate solutions. Colloids Surf. B
Biointerfaces 81, 165–173.
References Kammona, O., Kiparissides, C., 2012. Recent advances in nanocarrier-based mucosal
delivery of biomolecules. J. Control. Release 161, 781–794.
Abdelwahed, W., Degobert, G., Fessi, H., 2006. Investigation of nanocapsules sta- Kean, T., Thanou, M., 2010. Biodegradation, biodistribution and toxicity of chitosan.
bilization by amorphous excipients during freeze-drying and storage. Eur. J. Adv. Drug Deliv. Rev. 62, 3–11.
Pharm. Biopharm. 63, 87–94. Knaul, J.Z., Hudson, S.M., Creber, K.A.M., 1999. Improved mechanical properties of
Aggarwal, P., Hall, J.B., McLeland, C.B., Dobrovolskaia, M.A., McNeil, S.E., 2009. chitosan fibers. J. Appl. Polym. Sci. 72, 1721–1731.
Nanoparticle interaction with plasma proteins as it relates to particle biodis- Kumari, A., Yadav, S.K., Yadav, S.C., 2010. Biodegradable polymeric nanoparticles
tribution and biocompatibility and therapeutic efficacy. Adv. Drug Deliv. Rev. based drug delivery systems. Colloids Surf. B Biointerfaces 75, 1–18.
61, 428–437. Lamprecht, A., Schafer, U., Lehr, C.M., 2001. Size-dependent bioadhesion of micro-
Al-Qadi, S., Grenha, A., Carrión-Recio, D., Seijo, B., Remuñán-López, C., 2012. Microen- and nanoparticulate carriers to the inflamed colonic mucosa. Pharm. Res. 18,
capsulated chitosan nanoparticles for pulmonary protein delivery: in vivo 788–793.
evaluation of insulin-loaded formulations. J. Control. Release 157, 383–390. Lee, D.-W., Powers, K., Baney, R., 2004. Physicochemical properties and blood
Allison, S.D., Molina, M.D.C., Anchordoquy, T.J., 2000. Stabilization of lipid/DNA compatibility of acylated chitosan nanoparticles. Carbohydr. Polym. 58,
complexes during the freezing step of the lyophilization process: the particle 371–377.
isolation hypothesis. Biochim. Biophys. Acta 1468, 127–138. Lee, K.Y., Ha, S.W., Park, W.H., 1995. Blood biocompatibility and biodegrad-
Andrews, G.P., Laverty, T.P., Jones, D.S., 2009. Mucoadhesive polymeric platforms for ability of partially N-acylated chitosan derivatives. Biomaterials 16,
controlled drug delivery. Eur. J. Pharm. Biopharm. 71, 505–518. 1211–1216.
Berger, J., Reist, M., Mayer, J.M., Felt, O., Peppas, N.A., Gurny, R., 2004. Structure Lehr, C.M., Bouwstra, J.A., Schacht, E.H., Junginger, H.E., 1992. In vitro evaluation of
and interactions in covalently and ionically crosslinked chitosan hydrogels for mucoadhesive properties of chitosan and some other natural polymers. Int. J.
biomedical applications. Eur. J. Pharm. Biopharm. 57, 19–34. Pharm. 78, 43–48.
Birrenbach, G., Speiser, P.P., 1976. Polymerized micelles and their use as adjuvants Liu, Z., Jiao, Y., Wang, Y., Zhou, C., Zhang, Z., 2008. Polysaccharides-based nanopar-
in immunology. J. Pharm. Sci. 65, 1763–1766. ticles as drug delivery systems. Adv. Drug Deliv. Rev. 60, 1650–1662.
Blasi, P., Giovagnoli, S., Schoubben, A., Ricci, M., Rossi, C., 2007. Solid lipid López-León, T., Carvalho, E.L.S., Seijo, B., Ortega-Vinuesa, J.L., Bastos-González, D.,
nanoparticles for targeted brain drug delivery. Adv. Drug Deliv. Rev. 59, 2005. Physicochemical characterization of chitosan nanoparticles: electroki-
454–477. netic and stability behavior. J. Colloid Interface Sci. 283, 344–351.
Blasi, P., Schoubben, A., Romano, G.V., Giovagnoli, S., Di Michele, A., Ricci, M., 2013. Luangtana-anan, M., Limmatvapirat, S., Nunthanid, J., Chalongsuk, R., Yamamoto, K.,
Lipid nanoparticles for brain targeting II. Technological characterization. Colloid 2010. Polyethylene glycol on stability of chitosan microparticulate carrier for
Surf. B: Biointerfaces 110, 130–137. protein. AAPS. PharmSciTech. 11, 1376–1382.
Bodmeier, R., Oh, K.H., Pramar, Y., 1989. Preparation and evaluation of drug- Luangtana-anan, M., Opanasopit, P., Ngawhirunpat, T., Nunthanid, J., Sriamornsak,
containing chitosan beads. Drug Dev. Ind. Pharm. 15, 1475–1494. P., Limmatvapirat, S., Lim, L.Y., 2005. Effect of chitosan salts and molar mass
Bystrzejewska-Piotrowska, G., Golimowski, J., Urban, P.L., 2009. Nanoparticles: their on a nanoparticulate carrier for therapeutic protein. Pharm. Dev. Technol. 10,
potential toxicity, waste and environmental management. Waste Manag. 29, 189–196.
2587–2595. Morris, G.A., Castile, J., Smith, A., Adams, G.G., Harding, S.E., 2011. The effect
Calvo, P., Remunan-Lopez, R., Vila-Jato, C.J.L., Alonso, M.J., 1997. Chitosan and chi- of prolonged storage at different temperatures on the particle size distribu-
tosan/ethylene oxide–propylene oxide block copolymer nanoparticles as novel tion of tripolyphosphate (TPP)–chitosan nanoparticles. Carbohydr. Polym. 84,
carriers for proteins and vaccines. Pharm. Res. 14, 1431–1436. 1430–1434.
Cesàro, A., 2006. All dried up. Nat. Mat. 5, 593–594. Muzzarelli, R., 1998. Colorimetric determination of chitosan. Anal. Biochem. 260,
Cesàro, A., De Giacomo, O., Sussich, F., 2008. Water interplay in trehalose polymor- 255–257.
phism. Food Chem. 106, 1318–1328. Nasti, A., Zaki, N.M., De Leonardis, P., Ungphaiboon, S., Sansongsak, P., Rimoli, M.G.,
Crowe, L.M., Reid, D.S., Crowe, J.H., 1996. Is trehalose special for preserving dry Tirelli, N., 2009. Chitosan/TPP and chitosan/TPP-hyaluronic acid nanoparticles:
materials? Biophys. J. 71, 2087–2093. systematic optimisation of the preparative process and preliminary biological
Csaba, N., Köping-Höggård, M., Alonso, M.J., 2009. Ionically crosslinked chi- evaluation. Pharm. Res. 26, 1918–1930.
tosan/tripolyphosphate nanoparticles for oligonucleotide and plasmid DNA Niederhofer, A., Müller, B.W., 2004. A method for direct preparation of chi-
delivery. Int. J. Pharm. 382, 205–214. tosan with low molecular weight from fungi. Eur. J. Pharm. Biopharm. 57,
Dambies, L., Vincent, T., Domard, A., Guibal, E., 2001. Preparation of chitosan gel 101–105.
beads by ionotropic molybdate gelation. Biomacromolecules 2, 1198–1205. Nishimura, K., Ishihara, C., Ukei, S., Tokura, S., Azuma, I., 1986. Stimulation of cytokine
Dash, M., Chiellini, F., Ottenbrite, R.M., Chiellini, E., 2011. Chitosan-A versatile semi- production in mice using deacetylated chitin. Vaccine 4, 151–156.
synthetic polymer in biomedical applications. Prog. Polym. Sci. 36, 981–1014. Prego, C., Garcia, M., Torres, D., Alonso, M.J., 2005. Transmucosal macromolecular
De Campos, A.M., Sanchez, A., Alonso, M.J., 2001. Chitosan nanoparticles: a new vehi- drug delivery. J. Control. Release 101, 151–162.
cle for the improvement of the delivery of drugs to the ocular surface. Application Rampino, A., 2011. Polysaccharide-based nanoparticles for drug delivery. University
to cyclosporin A. Int. J. Pharm. 224, 159–168. of Trieste, Trieste, Italy (Ph.D. Thesis).
Des Rieux, A., Fievez, V., Garinot, M., Schneider, Y.-J., Préat, V., 2006. Nanoparti- Ribatti, D., Vacca, A., Ranieri, G., Sorino, S., Roncali, L., 1996. The chick embryo
cles as potential oral delivery systems of proteins and vaccines: a mechanistic chorioallantoic membrane as an in vivo wound healing model. Pathol. Res. Pract.
approach. J. Control. Release 116, 1–27. 192, 1068–1076.
Dudhani, A.R., Kosaraju, S.L., 2010. Bioadhesive chitosan nanoparticles: preparation Ricci, M., Blasi, P., Giovagnoli, S., Perioli, L., Vescovi, C., Rossi, C., 2004. Leucinostatin-A
and characterization. Carbohydr. Polym. 81, 243–251. loaded nanospheres: characterization and in vivo toxicity and efficacy evalua-
Duncan, R., 2003. The dawning era of polymer therapeutics. Nat. Rev. Drug Discov. tion. Int. J. Pharm. 275, 61–72.
2, 347–360. Rosen, H., Abribat, T., 2005. The rise and rise of drug delivery. Nat. Rev. Drug Discov.
Dyer, A.M., Hinchcliffe, M., Watts, P., Castile, J., Jabbal-Gill, I., Nankervis, R., Smith, 4, 381–385.
A., Illum, L., 2002. Nasal delivery of insulin using novel chitosan based formu- Sakurai, M., Furuki, T., Akao, K., Tanaka, D., Nakahara, Y., Kikawada, T., Watan-
lations: a comparative study in two animal models between simple chitosan abe, M., Okuda, T., 2008. Vitrification is essential for anhydrobiosis in an
formulations and chitosan nanoparticles. Pharm. Res. 19, 998–1008. African chironomid, Polypedilum vanderplanki. Proc. Natl. Acad. Sci. U S A 105,
Fan, W., Yan, W., Xu, Z., Ni, H., 2012. Formation mechanism of monodisperse, low 5093–5098.
molecular weight chitosan nanoparticles by ionic gelation technique. Colloids Saw, C.L.L., Heng, P.W.S., Liew, C.V., 2008. Chick chorioallantoic membrane as an
Surf. B Biointerfaces 90, 21–27. in situ biological membrane for pharmaceutical formulation development: a
Food Safety Authority of Ireland, 2008. The Relevance for Food Safety of Applications rewiew. Drug Dev. Ind. Pharm. 34, 1168–1177.
of Nanotechnology in the Food and Feed Industries. Food Safety Authority of Schoubben, A., Blasi, P., Giovagnoli, S., Rossi, C., Ricci, M., 2010. Development of a
Ireland, Dublin, IE. scalable procedure for fine calcium alginate particle preparation. Chem. Eng. J.
Gan, Q., Wang, T., Cochrane, C., McCarron, P., 2005. Modulation of surface charge, par- 160, 363–369.
ticle size and morphological properties of chitosan-TPP nanoparticles intended Schoubben, A., Blasi, P., Marenzoni, M.L., Barberini, L., Giovagnoli, S., Cirotto, C., Ricci,
for gene delivery. Colloids Surf. B Biointerfaces 44, 65–73. M., 2013. Capreomycin supergenerics for pulmonary tuberculosis treatment:
Golovina, E.A., Golovin, A., Hoekstra, F.A., Faller, R., 2010. Water replacement hypoth- preparation, in vitro, and in vivo characterization. Eur. J. Pharm. Biopharm. 83,
esis in atomic details: effect of trehalose on the structure of single dehydrated 388–395.
POPC bilayers. Langmuir 26, 11118–11126. Shmulewitz, A., Langer, R., Patton, J., 2006. Convergence in biomedical technology.
Hu, B., Pan, C., Sun, Y., Hou, Z., Ye, H., Zeng, X., 2008. Optimization of fabrication Nat. Biotechnol. 24, 277–280.
parameters to produce chitosan-tripolyphosphate nanoparticles for delivery of Tang, E.S.K., Huang, M., Lim, L.Y., 2003. Ultrasonication of chitosan and chitosan
tea catechins. J. Agric. Food Chem. 56, 7451–7458. nanoparticles. Int. J. Pharm. 265, 103–114.
228 A. Rampino et al. / International Journal of Pharmaceutics 455 (2013) 219–228

Tømmeraas, K., Köping-Höggård, M., Vårum, K.M., Christensen, B.E., Artursson, P., characterization of chitosan for sustainable biomedical applications: SAR guided
Smidsrød, O., 2002. Preparation and characterisation of chitosans with oligosac- multi-dimensional analysis. Green Chem. 11, 498–509.
charide branches. Carbohydr. Res. 337, 2455–2462. Xu, Y.M., Du, Y.M., 2003. Effect of molecular structure of chitosan on protein delivery
Vargas, A., Zeisser-Labouèbe, M., Lange, N., Gurny, R., Delie, F., 2007. The chick properties of chitosan nanoparticles. Int. J. Pharm. 250, 215–226.
embryo and its chorioallantoic membrane (CAM) for the in vivo evaluation of Yu, J.H., Du, Y.M., Zheng, H., 1999. Blend films of chitosan-gelatin. J. Wuhan Univ.
drug delivery systems. Adv. Drug Deliv. Rev. 59, 1162–1176. (Nat. Sci. Ed.) 45, 440–444.
Wang, W., Du, Y., Qiu, Y., Wang, X., Hu, Y., Yang, J., Cai, J., Kennedy, J.F., 2008. A Zhang, L., Gu, F.X., Chan, J.M., Wang, A.Z., Langer, R.S., Farokhzad, O.C., 2008. Nanopar-
new green technology for direct production of low molecular weight chitosan. ticles in medicine: therapeutic applications and developments. Clin. Pharmacol.
Carbohydr. Polym. 74, 127–132. Ther. 83, 761–769.
Weinhold, M.X., Sauvageau, J.C.M., Keddig, N., Matzke, M., Tartsch, B., Grun-
wald, I., Kubel, C., Jastorff, B., Thoming, J., 2009. Strategy to improve the

Vous aimerez peut-être aussi