Anda di halaman 1dari 9

Colloids and Surfaces B: Biointerfaces 44 (2005) 65–73

Modulation of surface charge, particle size and morphological properties


of chitosan–TPP nanoparticles intended for gene delivery
Quan Gan a,∗ , Tao Wang a , Colette Cochrane a , Paul McCarron b
a School of Chemical Engineering, Queen’s University Belfast, Belfast BT9 5AG, UK
b School of Pharmacy, Queen’s University Belfast, Belfast BT9 7BL, UK

Received 14 March 2005; received in revised form 6 June 2005; accepted 8 June 2005

Abstract

This work investigates the polyanion initiated gelation process in fabricating chitosan–TPP (tripolyphosphate) nanoparticles in the size
range of 100–250 nm intended to be used as carriers for the delivery of gene or protein macromolecules. It demonstrates that ionic gelation
of cationic chitosan molecules offers a flexible and easily controllable process for systematically and predictably manipulating particle size
and surface charge which are important properties in determining gene transfection efficacy if the nanoparticles are used as non-viral vectors
for gene delivery, or as delivery carriers for protein molecules. Variations in chitosan molecular weight, chitosan concentration, chitosan to
TPP weight ratio and solution pH value were examined systematically for their effects on nanoparticle size, intensity of surface charge, and
tendency of particle aggregation so as to enable speedy fabrication of chitosan nanoparticles with predetermined properties. The chitosan–TPP
nanoparticles exhibited a high positive surface charge across a wide pH range, and the isoelectric point (IEP) of the nanoparticles was found
to be at pH 9.0. Detailed imaging analysis of the particle morphology revealed that the nanoparticles possess typical shapes of polyhedrons
(e.g., pentagon and hexagon), indicating a similar crystallisation mechanism during the particle formation and growth process. This study
demonstrates that systematic design and modulation of the surface charge and particle size of chitosan–TPP nanoparticles can be readily
achieved with the right control of critical processing parameters, especially the chitosan to TPP weight ratio.
© 2005 Elsevier B.V. All rights reserved.

Keywords: Chitosan nanoparticles; Nanoparticle surface charge; Nanoparticle morphology; Ionic gelation; Gene delivery

1. Introduction Furthermore, there is evidence demonstrating that cationic


polymers play an important role in both membrane adhesion
Chitosan is a non-toxic biodegradable polycationic poly- [5] and lysosomal escape [6] of the encapsulated DNA, pro-
mer with low immunogenicity. It has been extensively viding a potential explanation for the superiority of polymer-
investigated for formulating carrier and delivery systems mediated gene transfer relative to naked DNA administration
for therapeutic macrosolutes, particularly genes and protein in many applications. These hybrid DNA–chitosan systems
molecules primarily because positively charged chitosan can can be classified into two categories which differ in their
be easily complexed with negatively charged DNAs and pro- mechanism of formation and morphology: complexes and
teins [1,2]. Chitosan can effectively bind DNA and protect nanospheres.
it from nuclease degradation [3,4]. It has advantages of not Gentle mixing, followed by incubation, of chitosan
necessitating sonication and organic solvents for its prepara- and DNA solutions generated ‘broad’ distributions of
tion, therefore minimizing possible damage to DNA during chitosan–DNA particulate complexes with mean sizes
the complexation process. between 100 and 600 nm, depending on the molecular
weight of the chitosan. Since particle formation was elicited
∗ Corresponding author. Tel.: +44 2890 274463; fax: +44 2890 381753. solely by the tropism of the two oppositely charged macro-
E-mail address: q.gan@qub.ac.uk (Q. Gan). molecules for one another, these particles were termed

0927-7765/$ – see front matter © 2005 Elsevier B.V. All rights reserved.
doi:10.1016/j.colsurfb.2005.06.001
66 Q. Gan et al. / Colloids and Surfaces B: Biointerfaces 44 (2005) 65–73

‘complexes’. The simplicity of chitosan–DNA complexes is tion under mild conditions; homogeneous and adjustable size
both an advantage and a drawback. Though such complexes and a positive surface charge that can be easily modulated
are extremely easy to synthesize, the fact remains that their and a great capacity for the association of peptides, proteins,
transfection efficacy is significantly below that of cationic oligonucleotides, and plasmids [11].
liposomes in vitro and viral vectors in vivo. Aside from Therefore, ability to control and modulate the properties
N/P ratio and chitosan molecular weight, there remain of chitosan–TPP nanoparticles, most importantly particle size
few parameters in the synthesis protocol which can be and density of surface charge, is central in determining gene
modulated in an effort to augment transfection. To address transfection efficiency. It is important that these characteris-
this issue, investigators sought to develop more sophisticated tic properties be predictably produced and easily modulated
DNA-loaded chitosan nanoparticles. in a flexible and reliable nano fabrication process with high
The application of DNA–chitosan nanoparticles has yield and particle stability. It is therefore the focus of this
advanced in vitro DNA transfection research, and data have paper to report on how systematically manipulating process-
been accumulating that shows their usefulness for gene ing parameters in the TPP initiated chitosan gelation to obtain
delivery [7,8]. The therapeutic efficacy of the nanoparticles predictable and optimal nanoparticle properties for desired
could be due to their ability to protect the therapeutic agent applications in relation to gene/protein delivery.
from degradation due to lysosomal enzymes.
Due to their sub-cellular and sub-micron size,
chitosan–TPP nanoparticles can penetrate deep into tissues 2. Materials and experimental methods
through fine capillaries, cross the fenestration present in the
2.1. Materials
epithelial lining (e.g., liver) [9]. This allows efficient delivery
of therapeutic agents to target sites in the body. Also, by Three different molecular weight chitosan, derived from
modulating nanoparticles characteristics, such as enzymatic crab shell, in the form of fibrils flakes were obtained
degradation rate, size and surface charge density, one can from Sigma–Aldrich [Catalogue No. LMW 448869, MMW
control the release of a therapeutic agent from nanoparticles 448877, HMW 419419]. The degree of deacetylation for the
to achieve desired therapeutic level in target tissue for low molecular weight chitosan (LMW Chitosan), medium
required duration for optimal therapeutic efficacy [6]. molecular weight chitosan (MWM chitosan) and high molec-
The major drawback associated with using chitosan as ular weight chitosan (HMW chitosan) is 86.6%, 84.7%,
non-viral gene delivery system is the relatively low transfec- and 82.5%, respectively. Sodium Tripolyphosphate was pur-
tion rate in comparison to viral vectors, even though the later chased from Sigma–Aldrich Chemical Co. Ltd. All other
has its own limitations in patient safety, difficulty in scale- reagents used were of analytical grade.
up production, and possible toxicity, immune responses, and
inflammatory responses. It is understood that transfection 2.2. Purification of chitosan
efficacy of cationic polymers depends primarily on: (i) parti-
cle size, which determines their intracellular uptake, different Since medical applications of animal derived biomateri-
pathways of their uptake, intracellular trafficking and sorting als entail an inherent risk of protein contamination which has
into different intracellular compartments, and (ii) the inten- in recent years aroused great awareness and anxiety among
sity of particle surface charge which influence their ability to the public, drug companies, and the industry regulators, it
efficiently condensate DNA and interact with cell. is of utmost importance to ensure that chitosan intended for
Stable and reproducible chitosan nanoparticles were in medical applications is of the highest purity and free of pro-
early days formulated via chemical cross-linking in water- tein contamination. It is therefore decided to further purify
in-oil emulsion system for entraping and delivering drugs the purchased chitosan materials and examine whether there
[10]. However, the negative effects of cross-linking agents, are changes in chemical as well as physical properties before
e.g., glutaraldehyde, on cell viability and the integrity of and after purification. The origin and purity of purchased chi-
macromolecular drugs led to the development of prepara- tosan material depends on its source, season, and conditions
tion method under mild conditions. Preparation methods by of the chemical deacetylation process, which may vary across
ionically cross-linking cationic chitosan with specific polyan- different suppliers. Further purification process is crucial to
ions were particularly successful as, aside from its com- ensure that the starting chitosan material for nanoparticle
plexation with negatively charged polymers, chitosan has fabrication possesses the highest purity and integrity. In this
the ability to gel spontaneously on contact with multivalent work, purchased chitosan materials were subjected to a vig-
polyanions due to the formation of inter- and intramolecular orous purification process which involved mixing the solid
cross-linkage mediated by these polyanions. Among some chitosan flakes in 1 M NaOH solution, allowing 1 g of chi-
polyanions investigated, tripolyphosphate (TPP) is the most tosan for 10 ml NaOH solution. This solid–liquid mixture was
popular because of its non-toxic property and quick gelling heated and continuously stirred for 2 h at 70 ◦ C, and then fil-
ability. The chitosan–TPP nano system exhibits some attrac- tered using a Buchner funnel. Chitosan was insoluble in the
tive features which render them promising carriers for the caustic solution, and the recovered flakes were washed thor-
delivery of macromolecules. These features include forma- oughly and dried at 40 ◦ C for 12 h.
Q. Gan et al. / Colloids and Surfaces B: Biointerfaces 44 (2005) 65–73 67

The NaOH treated chitosan flakes were dissolved in 0.1 M Netherlands). One drop of dilute chitosan–TPP nanoparti-
acetic acid solution which was filtered using a filter paper cles solution was syringe placed on a carbon film 300 mesh
to remove residues of insoluble particles. One molar NaOH copper grid, allowing to sit until air-dried. The sample was
solution was used to adjust pH value of the filtrate to pH 8.0, stained with 1 M uranyl acetate solution for 5 s at 7 ◦ C before
resulting in purified chitosan in the form of white precipi- viewing on the TEM.
tates. The precipitated chitosan was washed thoroughly using
deionized water, and the product was vacuum-dried at room
temperature for 24 h. The dried samples were used for FT-IR 3. Results and discussion
analysis and preparation of the chitosan–TPP nanoparticles.
3.1. Purification and characterisation of supplied
2.3. Preparation of chitosan–TPP nanoparticles chitosan material

Chitosan solutions of different concentration and molecu- FT-IR was used to identify if there were variations in
lar weight were prepared by dissolving purified chitosan with chemical functional groups present at the surface of chi-
sonication in 1% (w/v) acetic acid solution until the solution tosan samples of different molecular weight, and to deter-
was transparent. Once dissolved, the chitosan solution was mine variations among purified and unpurified chitosan sam-
diluted with deionized water to produce chitosan solutions ples. By comparing the characteristic transmittance spec-
of different concentrations at 0.05%, 0.10%, 0.15%, 0.20%, trum of different chitosan samples, it is possible to ascertain
0.25%, and 0.30% (weight/volume). Tripolyphosphate was changes in the constituent surface functional groups (e.g.,
dissolved in deionized water at the concentration 0.7 mg/ml. NH2 , CH2 NH) during the purification process, an indi-
The chitosan solution was flush mixed with an equal vol- cation of removal of impurities from the purchased chitosan
ume of TPP solution and the formation of chitosan–TPP material.
nanoparticles started spontaneously via the TPP initiated Fig. 1 compares the transmittance spectrum of purified
ionic gelation mechanism. The nanoparticles were formed HMW chitosan with the original supplied materials. The con-
at selected chitosan to TPP weight ratios of 3:1, 4:1, 5:1, 6:1 trasting difference in the spectrum evidently demonstrates
and 7:1. The nanoparticle suspensions were gently stirred for the changes in surface chemistry of the original supplied chi-
60 min at room temperature before being subjected to further tosan material after purification, indicating possible removal
analysis and applications. of impurities, such as protein molecules and pigments. No
attempts were made in this paper to identify what partic-
2.4. FT-IR ular chemical bonds are associated with the spectrum peaks
which require additional verification beyond the scope of this
In FT-IR analysis of both purified and raw chitosan sam- study. The variation in the transmittance spectrum for LMW
ples, transmittance spectra were obtained using a Perkin- and MMW chitosan samples was much less pronounced than
Elmer FT-IR spectrometer (SPECTRUM 1000) fitted with the HMW chitosan before and after purification. Fig. 2 com-
an attenuated total reflectance mode (ATR) cell. The equip- pares the FT-IR transmittance spectrum of purified chitosan
ment was positioned in a laboratory maintained at 25 ± 1 ◦ C. samples of different molecular weight. The figure reveals
A small chitosan sample (7.0–9.0 mg) was placed on a NaCl variations in corresponding peak values of the same func-
plate and subjected to light within the infrared spectrum. The tional groups among different molecular weight samples, but
instrument operated with a resolution of 4 cm−1 and 128 not significant variations in presence of the functional groups
scans were collected for each sample. The IR absorbency themselves.
scans were analysed between 700 and 4000 cm−1 for changes
in the intensity of the sample peaks.

2.5. Measurement of size and zeta potential of


chitosan–TPP nanoparticles

Measurement of physical size, zeta potential and polydis-


persity (size distribution) of the chitosan–TPP nanoparticles
were performed using a 3000HSA Zetasizer (Malvern Instru-
ments, England).

2.6. Morphology observation

The morphological characteristics of the nanoparticles


were examined using a high resolution TEM (Transmission Fig. 1. Representative FT-IR transmittance spectra of purified and unpurified
Electron Microscope) machine (Tecnai F-20, Phillips Co., HMW chitosan samples.
68 Q. Gan et al. / Colloids and Surfaces B: Biointerfaces 44 (2005) 65–73

Fig. 2. Representative FT-IR transmittance spectra of purified HMW, MMW


Fig. 3. Effect of chitosan concentration from 0.05% to 0.30% (w/v) on par-
and LMW chitosan samples.
ticles size with three different chitosan molecular weight. Chitosan to TPP
mass ratio = 5:1, T = 20 ± 1 ◦ C, pH 5.0.

The purification study demonstrates convincingly that fur-


ther purification of supplied chitosan materials is essential The effect of chitosan to TPP weight ratio on particle size
especially with high molecular weight chitosans. was also very prominent (Fig. 4), showing a linear increase of
size with increasing chitosan to TPP weight ratio within the
tested chitosan to TPP ratio range. These linear relationships
3.2. Modulation of particle size provide a simple processing window for manipulating and
optimising the nano size for intended applications.
Particle size is one of the most significant determinant
in mucosal and epithelial tissue uptake of nanoparticles and 3.3. Modulation of particle surface charge
in the intracellular trafficking of the particles [6]. Smaller
size nanoparticles (∼100 nm) demonstrated more than 3- Chitosan has a rigid crystalline structure through inter- and
fold greater arterial uptake compared to larger nanoparticles intra- molecular hydrogen bonding. Chitosan molecules in
(∼275 nm) in an ex vivo canine carotid artery model [12,13] aqueous solutions adopt extended conformation with a more
as the smaller nanoparticles were able to penetrate throughout flexible chain because of the electrostatic charge repulsion
the sub-mucosal layers while the larger size micron-particles between the chains. When chitosan and TPP were mixed with
were predominantly localized in the epithelial lining. Prabha each other in dilute acetic acid, they spontaneously formed
et al. [14] investigated the gene transfection levels of differ- compact nano complexes with an overall positive surface
ent size fractions of PLGA nanoparticles and found that the charge, and the density of the surface charge is reflected by
lower size nanoparticle fraction produced a 27-fold higher measured zeta potential values.
transfection in COS-7 cells and 4-fold higher transfection in
HEK 293 cells for the same dose of nanoparticles. These stud-
ies also suggested that uniform particle size distribution are
important to enhance the nanoparticle-mediated gene expres-
sion.
Chitosan’s ability of quick gelling on contact with polyan-
ions relies on the formation of inter- and intramolecular
cross-linkages mediated by these polyanions. Nanoparticles
are formed immediately upon mixing of TPP and chitosan
solutions as molecular linkages were formed between TPP
phosphates and chitosan amino groups. Size and size distri-
bution of the chitosan–TPP nanoparticles depend largely on
concentration, molecular weight, and conditions of mixing,
i.e., stirring or sonication.
Fig. 3 shows the effect of chitosan concentration on parti-
cles size at three different molecular weight. It demonstrates
that the size of HMW chitosan nanoparticles was mostly
affected by the increased chitosan solution concentration, and Fig. 4. Effect of chitosan to TPP mass ratio from 3:1 to 7:1 on parti-
the increase in size with concentration showed a linear rela- cles size with three different chitosan molecular weight. c = 0.50% (w/v),
tionship within the tested range. T = 20 ±1 ◦ C, pH 5.0.
Q. Gan et al. / Colloids and Surfaces B: Biointerfaces 44 (2005) 65–73 69

Fig. 6. Effect of chitosan concentration on particle zeta potential at three


Fig. 5. Effect of chitosan to TPP mass ratio on particle zeta potential with different chitosan molecular weight. Chitosan to TPP mass ratio = 5:1,
three different chitosan molecular weight. c = 0.50% (w/v), T = 20 ± 1 ◦ C, T = 20 ± 1 ◦ C, pH 5.0.
pH 5.0.

medium, the amine groups will be positively charged, con-


It was reported that the ability of nanoparticles to escape ferring to the polysaccharide a high charge density [19].
the endo-lysosomes was dependent on the surface charge of Therefore, the surface charge density of chitosan molecules
the nanoparticles [15,16]. Nanoparticles which show transi- is strongly dependent on solution pH [20,21], and the ionic
tion in their surface charge from anionic at pH 7 to cationic cross-linking process for the formation of chitosan–TPP
in the acidic endosomal pH (pH 4–5) were found to escape nanoparticles is pH-responsive, providing opportunities to
the endosomal compartment whereas the nanoparticles which modulate the formulation and properties of the chitosan–TPP
remain negatively charged at pH 4–5 were retained mostly nanoparticles.
in the endosomal compartment. Thus, by varying the surface To study the effects of changing environmental pH val-
charge, one could potentially be able to direct the nanoparti- ues on nanoparticle size and zeta potential, chitosan–TPP
cles either to lysosomes or to cytoplasm [6]. The efficacy of nanoparticles formulated with MMW chitosan at fixed con-
nanoparticles as drug carriers is also closely related to their centration of 0.15% (w/v) and different chitosan to TPP mass
interaction, predominantly influenced by surface charge, with ratio between 4:1 and 7:1 were examined at varying chitosan
proteins and enzymes in different body fluids. Calvo et al. [17] solution pH values. The variations in particle size and zeta
analysed the interaction phenomenon between lysozyme, a potential with chitosan solution pH are shown separately in
positively charged enzyme that is highly concentrated in Fig. 7A and B. The nanoparticles formed at solution pH 5.5
mucosas, and poly-␧-caprolactone coated nanoparticles, and had a smaller size but a higher particle zeta potential. Fig. 7B
found that the interaction of lysozyme with the nanoparticles also demonstrated an interesting trough in zeta potential val-
and their consequent degradation was highly dependent on ues at pH 5.0.
their surface charge. The surface charge reversal of nanoparticles in the acidic
The zeta potential of chitosan–TPP nanoparticles solution of endo-lysosomes is proposed as the mechanism
increased linearly with increasing chitosan to TPP weight responsible for the endo-lysosomal escape of the nanoparti-
ratio from 3:1 to 7:1 (Fig. 5). Again, this simple linear rela- cles into cytoplasmic compartment for effective release and
tionship could be easily explored for modulating the particle gene expression [22]. Surface charge reversal occurs when
surface charge density to facilitate the adhesion properties protons or hydronium ions from bulk solution are transferred
and transport properties of the nanoparticles. to nanoparticle surface under acidic conditions, resulting an
The effect of chitosan concentration on zeta potential was increased surface charge density and zeta potential, which
also investigated at a fixed chitosan to TPP weight ratio of would allow stronger electrostatic interactions between the
5:1. The results in Fig. 6 show that, unlike the trend of particle nanoparticles and tissue cells, leading to localized destabili-
size increase with increasing chitosan concentration, the zeta sation of the cell membrane and escape of nanoparticles into
potential decreased with increasing chitosan concentration. cytoplasmic compartment.
Low molecular weight chitosan–TPP nanoparticles pro-
3.4. The effect of solution pH duced at solution pH 5.5 were tested for their size and zeta
potential response to changing pH values of the residing
Chitosan is a weak base polysaccharide, having an aver- solution medium by simply adjusting the solution pH to a pre-
age amino group density of 0.837 per disaccharide unit [18], determined value by titration with either 1 M NaOH or 1 M
and insoluble at neutral and alkaline pH values. In an acidic HCl solutions. Fig. 8A and B showed that both measured par-
70 Q. Gan et al. / Colloids and Surfaces B: Biointerfaces 44 (2005) 65–73

Fig. 7. (A) Effect of solution pH on MMW chitosan–TPP particle size at dif-


ferent chitosan to TPP mass ratio. c = 0.15% (w/v), T = 20 ± 1 ◦ C. (B) Effect
of solution pH value on MMW chitosan–TPP nanoparticles zeta potential.
c = 0.15% (w/v), T = 20 ± 1 ◦ C.
Fig. 8. (A) Responsive particle size change in relation to changing resid-
ing solution pH values from 3.2 to 12.2. LMW chitosan–TPP nanoparticles
ticle size and zeta potential are very sensitive to the changing produced at conditions c = 0.50% (w/v), chitosan to TPP mass ratio = 5:1,
T = 20 ± 1 ◦ C, pH 5.5. (B) Responsive particle zeta potential change in
pH values of the residing aqueous environment, indicating the relation to changing residing solution pH values from 3.2 to 12.2. LMW
surface density of protonised amino groups and the degree chitosan–TPP nanoparticles produced at conditions c = 0.50% (w/v), chi-
of protonisation are reversibly responsive to changing solu- tosan to TPP mass ratio = 5:1, T = 20 ± 1 ◦ C, pH 5.5.
tion pH values. The increase in measured average particle size
could be caused mainly by particle aggregation when solution 3.5. Stability of the chitosan–TPP nanoparticle system
pH value increased, rather than by further growth of the indi-
vidual particle size after initial formation. The sharp increase The chitosan–TPP nanoparticle colloidal system is ther-
in size at pH > 6.0 suggests that the degree of protonisation modynamically unstable, especially at unfavourable solu-
at surface of the particles were reduced, decreasing elec- tion pH conditions and at high particle concentrations,
trostatic repulsion between the particles thereby increasing because of high surface energy associated with the nano
the probability of particle aggregation. The idea of depro- scale dimensions. Fig. 9 shows size growth kinetics of MMW
tonisation of the particle surface was supported by results chitosan–TPP particles at a dilute chitosan concentration
presented in Fig. 8B which shows a continual decrease in 0.15% (w/v). Ionic gelation and growth of the chitosan–TPP
the positive zeta potential well before the pH value reached nanoparticles were completed within the first 60 min with
pH 6.0. Fig. 8B also shows that an isoelectric point for subsequently slight increase in particle size over the next
the chitosan–TPP nanoparticles is located at around pH 9.0. 24 h. No apparent aggregation of particles was observed
Gelling of the nanoparticle colloidal system could easily during this period at constant temperature and solution
occur when the overall particle surface charge is neutral at pH.
the isoelectric point. The gelling mechanism in relation to However when the initial chitosan concentration was
swinging pH values has been investigated for producing smart increased over and above a critical concentration, large aggre-
responsive nanoparticle systems for targeted drug delivery gates formed instantaneously. The large aggregates were
[20,21]. observed using the TEM imaging technique (Section 3.6),
Q. Gan et al. / Colloids and Surfaces B: Biointerfaces 44 (2005) 65–73 71

and the critical aggregation concentration was studied using


Zetasizer measurement which showed a drastic size increase
accompanied by a sudden large reduction in zeta potential at
the critical concentration. The critical chitosan concentration
for the spontaneous formation of aggregates depends on solu-
tion pH and chitosan molecular weight. Tables 1 and 2 show
that the critical chitosan concentration for LMW, MMW and
HMW chitosan is 0.65%, 0.25%, 0.15% (w/v) at pH 4.0, and
1.00%, 0.85%, 0.75% (w/v) at pH 5.0, respectively.

3.6. Morphological characteristics of chitosan–TPP


nanoparticles
Fig. 9. Kinetics of chitosan–TPP nanoparticle size growth. c = 0.15% (w/v),
The morphological characteristics of the LMW chitosan–
chitosan to TPP mass ratio = 5:1, T = 20 ± 1 ◦ C, pH 5.0.
TPP nanoparticles were examined using the TEM technique.
Table 1
Measured particles size and zeta potential at different chitosan molecular weight and concentration
LMW chitosan Size (nm) Zeta (mv) MMW chitosan Size (nm) Zeta (mv) HMW chitosan Size (nm) Zeta (mv)
(%) (w/v) (%) (w/v) (%) (w/v)
0.05 136.2 48.3 0.05 145.3 43.9 0.05 155.0 37.7
0.10 142.3 44.2 0.10 150.5 40.3 0.10 188.9 34.8
0.15 152.9 41.0 0.15 165.2 39.1 0.15 3884.2 17.4
0.20 171.2 39.7 0.20 182.3 37.2 0.20 5964.3 17.2
0.25 190.3 37.3 0.25 2175.4 16.2 0.25 6136.2 16.7
0.30 203.1 35.6 0.30 10016.2 15.2 0.30 7854.2 16.9
0.35 312.7 33.3 0.35 – – 0.35 – –
0.40 429.6 31.6 0.40 – – 0.40 – –
0.45 578.3 30.4 0.45 – – 0.45 – –
0.50 712.4 28.2 0.50 – – 0.50 – –
0.55 888.6 26.9 0.55 – – 0.55 – –
0.60 997.0 25.2 0.60 – – 0.60 – –
0.65 1628.6 14.3 0.65 – – 0.65 – –
Chitosan to TPP mass ratio = 4:1, T = 20 ± 1 ◦ C, pH 4.0.

Table 2
Measured particles size and zeta potential with different chitosan molecular weight and concentration
LMW chitosan Size (nm) Zeta (mv) MMW chitosan Size (nm) Zeta (mv) HMW chitosan Size (nm) Zeta (mv)
(%) (w/v) (%) (w/v) (%) (w/v)
0.05 143.2 49.2 0.05 156.1 46.8 0.05 162.7 41.3
0.10 152.1 46.8 0.10 163.7 44.4 0.10 176.1 37.4
0.15 159.2 45.6 0.15 170.7 40.3 0.15 208.9 36.9
0.20 172.8 44.3 0.20 181.5 39.8 0.20 216.8 35.1
0.25 181.9 42.7 0.25 192.2 37.8 0.25 234.2 34.1
0.30 189.6 40.7 0.30 209.8 35.3 0.30 257.0 33.2
0.35 273.2 39.4 0.35 310.2 33.8 0.35 362.3 32.6
0.40 387.2 37.1 0.40 427.5 32.7 0.40 478.0 31.8
0.45 519.6 36.8 0.45 546.3 32.1 0.45 566.6 31.0
0.50 604.3 34.2 0.50 654.3 30.4 0.50 697.3 29.5
0.55 692.1 33.3 0.55 721.5 29.0 0.55 748.5 27.8
0.60 726.6 32.7 0.60 783.2 28.6 0.60 828.0 26.3
0.65 740.6 32.1 0.65 848.3 27.5 0.65 898.5 25.9
0.70 846.7 30.8 0.70 881.0 25.3 0.70 976.8 23.2
0.75 858.9 30.0 0.75 936.8 25.1 0.75 1654.9 15.9
0.80 893.6 28.5 0.80 988.6 24.8 0.80 2400.5 10.6
0.85 908.6 26.7 0.85 2371.0 15.2 0.85 5116.8 8.0
0.90 938.7 26.2 0.90 2742.7 12.2 0.90 – –
0.95 998.6 24.3 0.95 4276.9 7.6 0.95 – –
1.00 1819.3 13.5 1.00 – – 1.00 – –
1.10 2059.0 11.2 1.10 – – 1.10 – –
1.20 2851.6 7.6 1.20 – – 1.20 – –
Chitosan to TPP mass ratio = 4:1, T = 20 ± 1 ◦ C, pH = 5.0.
72 Q. Gan et al. / Colloids and Surfaces B: Biointerfaces 44 (2005) 65–73

Fig. 10. TEM image of a single LMW chitosan–TPP nanoparticle.

TEM image of single chitosan–TPP nanoparticles (Fig. 10)


reveals that the nanoparticles have a size range between 140 Fig. 12. TEM image of a large aggregate of LMW chitosan–TPP nanopar-
and 250 nm which conforms with the size measurement by ticles.
photon correlation spectroscopy using Zeasizer 3000HAS.
large aggregate formed with many distinctive single nanopar-
Fig. 11 shows the image of an aggregate of four distinc-
ticles, each still possessing a similar nano-metric dimension
tive single particles with clear joining boundaries formed
as being shown in Fig. 10.
alongside the regular geometry of the proximate polyhedron
Fig. 13 shows a TEM image of a chitosan–TPP particle
(pentagon and hexagon) shaped particles. Different to past
incorporating protein molecules of bovine serum albumin
reported works [9,23], the evidence of the formation of poly-
(BSA). 0.03 mg/ml BSA molecules were added to an equal
hydrons, instead of spheres, at nano-metric scale suggests a
volume of LMW chitosan solution (1.5%, w/v) in acidic acid
nucleation through ionic gelation followed by semi-crystal
and gently mixed for 1 h before TPP solution (0.5%, w/v) was
formation and growth.
added to make up chitosan to TPP mass ratio at 5:1. The BSA
As previously described in Section 3.5, the formation of
incorporated nanoparticles have a size range of 300–350 nm,
large nanoparticle aggregates depends on chitosan concen-
tration and solution pH. Fig. 12 shows the TEM image of a

Fig. 11. TEM image of an aggregate of four single LMW chitosan–TPP Fig. 13. TEM image of a LMW chitosan–TPP nanoparticles incorporating
nanoparticles with distinctive polyhedron shapes. BSA molecules.
Q. Gan et al. / Colloids and Surfaces B: Biointerfaces 44 (2005) 65–73 73

doubling the size of chitosan–TPP particles. The BSA incor- [3] Z. Cui, R.J. Mumper, Chitosan-based nanoparticles for topical
porated particles possessed a typical spherical shape and genetic immunization, J. Control. Release 75 (2001) 409–419.
smooth surface, which confirms the findings by Xu and Du [4] L. Illum, I. Jabbal-Gill, M. Hinchcliffe, A.N. Fisher, S.S. Davis,
Chitosan as a novel nasal delivery system for vaccines, Adv. Drug
[23], and Janes et al. [1]. Future work will investigate the Deliv. Rev. 51 (2001) 81–96.
encapsulation mechanism and efficiency of protein molecules [5] K.A. Mislick, J.D. Baldeschwieler, Evidence for the role of proteo-
in the ionic initiated chitosan–TPP nanoparticle formation glycans in cation-mediated gene transfer, Proc. Natl. Acad. Sci. USA
process, and the release kinetics of protein molecules from 93 (1996) 12349–12354.
the particle. [6] J. Panyam, V. Labhasetwar, Biodegradable nanoparticles for drug and
gene delivery to cells and tissue, Adv. Drug Deliv. Rev. 55 (2003)
329–347.
[7] K. Corsi, F. Chellat, L. Yahia, J.C. Fernandes, Mesenchymal stem
4. Conclusion cells, MG63 and HEK293 transfection using chitosan–DNA nanopar-
ticles, Biomaterials 24 (2003) 1255–1264.
The formation of high yield chitosan–TPP nanoparticles [8] K. Romoren, B.J. Thu, O. Evensen, Immersion delivery of plasmid
with predetermined nano-metric size and surface charge den- DNA. II. A study of the potentials of a chitosan based delivery sys-
tem in rainbow trout (Oncorhynchus mykiss) fry, J. Control. Release
sity can be simply manipulated and controlled by varying the 85 (2002) 215–225.
key processing conditions of chitosan concentration, chitosan [9] S.V. Vinagradov, T.K. Bronich, A.V. Kabanov, Nanosized cationic
to TPP weight ratio, and solution pH value. Within the tested hydrogels for drug delivery: preparation, properties and interactions
range of conditions, the increase in particle size and parti- with cells, Adv. Drug Deliv. Rev. 54 (2002) 223–233.
cle zeta potential showed a simple linear relationship with [10] Y. Ohya, M. Shiratani, H. Kobayashi, T. Ouchi, Release behav-
ior of 5-fluorouracil from chitosan-gel nanospheres immobilizing
increasing chitosan to TPP weight ratio, but the zeta poten- 5-fluorouracil coated with polysaccharides and their cell specific
tial at fixed chitosan to TPP ratio showed a linear decrease cytotoxicity, Pure Appl. Chem. A 31 (1994) 629–642.
with increasing chitosan concentration. Solution pH value [11] X.Z. Shu, K.J. Zhu, A new approach to prepare tripolyphos-
and chitosan concentration also had profound influence on the phate/chitosan complex beads for controlled drug delivery, Int. J.
stability of the nanoparticle system, and the critical chitosan Pharm. 201 (2000) 51–58.
[12] C. Song, V. Labhasetwar, X. Cui, T. Underwood, R.J. Levy, Arterial
concentrations for spontaneous formation of large particle uptake of biodegradable nanoparticles for intravascular local drug
aggregates at pH 5 were found to be 0.65%, 0.25%, 0.15% delivery: results with an acute dog model, J. Control. Release 54
(w/v) at pH 4.0, and 1.00%, 0.85%, 0.75% (w/v) at pH 5.0 (1998) 201–211.
for LMW, MMW and HMW chitosan, respectively. The iso- [13] M.P. Desai, V. Labhasetwar, G.L. Amidon, R.J. Levy, Gastrointesti-
electric point of the chitosan–TPP nanoparticles was found nal uptake of biodegradable microparticles: effect of particle size,
Pharm. Res. 13 (1996) 1838–1845.
at around pH 9.0. [14] S. Prabha, W.Z. Zhou, J. Panyam, V. Labhasetwar, Size depen-
Morphological study of the nanoparticles formed under dency of nanoparticles-mediated gene transfection: Studies with
different conditions revealed the formation of regu- fractionnated nanoparticles, Int. J. Pharm. 244 (2002) 105–
larly shaped polyhydron particles, an indication of semi- 115.
crystallisation mechanism during the particle formation and [15] W. Paul, C. Sharma, Chitosan a drug carrier for the 21st century
S.T.P, Pharm. Sci. 10 (2000) 5–22.
growth, suggesting the particles were of compact structure [16] V. Dodane, D. Vilivalam, Pharmaceutical applications of chitosan,
with orderly molecular arrangement, the discovery of which Pharm. Sci. Technol. Today 16 (1998) 246–253.
bears important implications on gene/protein encapsulation [17] P. Calvo, J.L. Vila-Jato, M.J. Alonso, Effect of lysozyme on the
and release mechanisms. stability of polyester nanocapsules and nanoparticles: stabilization
approaches, Biomaterials 18 (1997) 1305–1310.
[18] F.-L. Mi, H.-W. Sung, S.-S. Shyu, Synthesis and characterization of
biodegradable TPP/genipin co-crosslinked chitosan gel beads, Poly-
Acknowledgement mer 44 (2003) 6521–6530.
[19] K.W. Leong, H.Q. Mao, V.L. Truong-Le, DNA-polycation
To the European Social Funding programme for providing nanospheres as non-viral gene delivery vehicles, J. Control. Release
a scholarship to Colette Cochrane for this project. 53 (1998) 183–193.
[20] J.A. Ko, H.J. Park, S.J. Hwang, Preparation and characterization
of chitosan microparticles intended for controlled drug delivery, J.
Pharm. 249 (2002) 165–174.
References [21] X.Z. Shu, K.J. Zhu, Novel PH–sensitive citrate cross-linked chi-
tosan film for controlled drug release, Int. J. Pharm. 212 (2001)
[1] K.A. Janes, P. Calvo, M.J. Alonso, Polysaccharide colloidal particles 19–28.
as delivery systems for macromolecules, Adv. Drug Deliv. Rev. 47 [22] K. Makino, H. Ohshima, T. Kondo, Transfer of protons from bulk
(2001) 83–97. solution to the surface of poly(l-lactide) microcapsules, J. Microen-
[2] S.C.W. Richardson, H.V.J. Kolbe, R. Duncan, Potential of low molec- capsul. 3 (1986) 195–202.
ular mass chitosan as a DNA delivery system: biocompatibility, body [23] Y. Xu, Y. Du, Effect of molecular structure of chitosan on pro-
distribution and ability to complex and protect DNA, Int. J. Pharm. tein delivery properties of chitosan nanoparticles, Int. J. Pharm. 250
178 (1999) 231–243. (2003) 215–226.

Anda mungkin juga menyukai