You are on page 1of 8

European Journal of Pharmaceutical Sciences 46 (2012) 522–529

Contents lists available at SciVerse ScienceDirect

European Journal of Pharmaceutical Sciences


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

A one-step process in preparation of cationic nanoparticles


with poly(lactide-co-glycolide)-containing polyethylenimine gives efficient gene
delivery
Min Da Shau a,1, Mei Fen Shih b,1, Chi Cheng Lin a, I Chuan Chuang a, Wei Chih Hung c, Wim E. Hennink d,
Jong Yuh Cherng c,⇑
a
Department of Biotechnology, Chia-Nan University of Pharmacy and Science, 60 Erh-Jen Rd., Sec. 1, Jen-Te, Taiwan
b
Department of Pharmacy, Chia-Nan University of Pharmacy and Science, 60 Erh-Jen Rd., Sec. 1, Jen-Te, Taiwan
c
Department of Chemistry and Biochemistry, National Chung Cheng University, 168 University Rd., Chia-yi 621, Taiwan
d
Department of Pharmaceutics, Utrecht Institute for Pharmaceutical Sciences, Utrecht University, P.O. Box 80082, 3508 TB Utrecht, The Netherlands

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

Article history: A one-step preparation of nanoparticles with poly(lactide-co-glycolide) (PLGA) pre-modified with poly-
Received 16 September 2011 ethylenimine (PEI) is better in requirements for DNA delivery compared to those prepared in a two-step
Received in revised form 15 March 2012 process (preformed PLGA nanoparticles and subsequently coated with PEI). The particles were prepared
Accepted 4 April 2012
by emulsification of PLGA/ethyl acetate in an aqueous solution of PVA and PEI. DLS, AFM and SEM were
Available online 13 April 2012
used for the size characteristics. The cytotoxicity of PLGA/PEI nanoparticles was detected by MTT assay.
The transfection activity of the particles was measured using pEGFP and pb-gal plasmid DNA. Results
Keywords:
showed that the PLGA/PEI nanoparticles were spherical and non-porous with a size of about 0.2 lm
Cytotoxicity
Transfection
and a small size distribution. These particles had a positive zeta potential demonstrating that PEI was
PLGA attached. Interestingly, the zeta potential of the particles (from one-step procedure) was substantially
PEI higher than that of two-step process and is ascribed to the conjugation of PEI to PLGA via aminolysis.
The PLGA/PEI nanoparticles were able to bind DNA and the formed complexes had a substantially lower
cytotoxicity and a higher transfection activity than PEI polyplexes. In conclusion, given their small size,
stability, low cytotoxicity and good transfection activity, PLGA/PEI–DNA complexes are attractive gene
delivery systems.
Ó 2012 Elsevier B.V. All rights reserved.

1. Introduction categories of nonviral systems for gene delivery (Zhdanov et al.,


2002; Romøren et al., 2004). Cationic polymers, such as polyethy-
The carriers used for gene delivery are commonly classified as lenimine (PEI) (Boussif et al., 1995; Neu et al., 2005), chitosan
viral or non-viral vectors (Zhang and Godbey, 2006; Mintzer and (Borchard, 2001; Kim et al., 2007), poly-L-lysine (PLL) (Kwoh
Simanek, 2009). Viral-DNA particles normally show a relatively et al., 1999) and polyurethane (Tseng et al., 2005; Shau et al.,
high level of transgene expression. However, possible immunoge- 2006; Hung et al., 2009), have extensively been studied in the last
nicity, restrictions to the size of inserted DNA and potential local two decades as gene delivery systems. However, the complexes of
and systemic toxicity limit their universal applications (Thomas cationic polymers and pDNA (‘polyplexes’) have rather low trans-
et al., 2003; Gorecki, 2001). Further, these viral systems have to fection efficiency compared to viral vectors.
be produced in costly facilities (e.g. P2 laboratory). Nonviral-DNA The most frequently studied cationic polymer is PEI. Polypolex-
complexes are a promising alternative for viral nucleic acid deliv- es based on this polymer generally speaking show high transfec-
ery systems because they can more easily pharmaceutically pre- tion activity in many cell lines, but this polymer is rather toxic
pared and have less safety concerns. Also, the systems can be and together with its non-degradability are major drawbacks for
decorated with e.g. PEG coatings, targeting ligand and pH sensitive its in vivo use (Godbey et al., 2001; van de Wetering et al., 1997).
peptides to improve circulation kinetics, cell specificity and endo- Therefore, there is a need for approaches that reduce PEI’s toxicity
somal escape. Cationic lipids and cationic polymers are two major and at simultaneously preserve, or ideally improve, its gene deliv-
ery performance.
⇑ Corresponding author. Tel.: +886 5 2720411x66416; fax: +886 5 2721040.
Poly(lactide-co-glycolide) (PLGA)-based nanoparticles because
E-mail address: jycherng@yahoo.com (J.Y. Cherng).
of their excellent biocompatibility and low toxicity have been
1
Equally contributed to this article. extensively studied as delivery vehicles of both low molecular

0928-0987/$ - see front matter Ó 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.ejps.2012.04.006
M.D. Shau et al. / European Journal of Pharmaceutical Sciences 46 (2012) 522–529 523

weight drugs as well as biotherapeutics such as pharmaceutical particles were freeze-dried without adding cryoprotectant (vacu-
proteins and pDNA (Romero et al., 2010). However, the low DNA umed dried under 60 torr, FD2-6P-M, KingMech, Taiwan) and the
encapsulation efficiency and capacity hamper their gene therapy freeze-dried samples were stored at 4 °C when they were not use
applications (Jong et al., 1997; Ando et al., 1999). Further, the sta- immediately.
bility of entrapped DNA in PLGA matrices is a concern. To illustrate, Comparing to PLGA/PEI nanoparticles (formulation A via one-
it his has been shown that the super coiled DNA is converted into step preparation), a control experiment of ‘‘PLGA + PEI cationic
its less active linear form during preparation and/or release (Perez nanoparticles’’(via two-step process) was carried out by coating
et al., 2001). Importantly, it is difficult to control the release of the of preformed PLGA nanoparticles with PEI. PLGA (54 mg) dissolved
entrapped DNA from PLGA nanoparticles (Perez et al., 2001; Luten in 4 ml ethyl acetate and subsequently 4 ml of an aqueous solution
et al., 2008). As an alternative for entrapment, DNA can be ad- (contained 2% PVA without PEI) was mixed under stirring for
sorbed onto the surface of PLGA nanoparticles. Because PLGA nano- 10 min and an emulsion was formed. The same procedure as above
particles have a negative surface charge, negatively charged pDNA was followed to obtain PLGA nanoparticles. Next, the PLGA nano-
does not adsorb spontaneously onto such particles (Chumakova particles were washed twice with deionized water to remove
et al., 2008). To render PLGA nanoparticles adsorbing for DNA, cat- PVA by centrifugation and mixed with PEI (3.6 mg in 11 ml deion-
ionic surface charges have been introduced (Bivas-Benita et al., ized water, pH around 10.5) under stirring condition for 16 h. Two
2004; Katas et al., 2009). This has been established by incorporat- washing and centrifugation cycles (15,000 rpm; Hermle Z323K,
ing or coating of preformed PLGA nanoparticles with cationic poly- Germany) for 30 min at 4 °C with deionized water were applied
mers such as polylysine (Capan et al., 1999), chitosan (Nafee et al., to remove the excess PEI.
2007) and PEI (Chumakova et al., 2008) or by preparation of PLGA
nanoparticles using cationic surfactants (Basarkar et al., 2007). In
2.3. Preparation of PLGA/PEI–DNA complexes
the present study, we prepared positively charged PLGA nanopar-
ticles in a one-step process by addition of PEI to an aqueous PVA
PLGA/PEI–DNA complexes with various polymer-to-DNA
solution in which a PLGA solution in ethyl acetate was emulsified.
weight ratios were prepared by adding PLGA/PEI nanoparticles dis-
In this way it is aimed that PEI is partly present at the surface to
persion (various amounts in 0.5 ml of plain DMEM) to plasmid
allow DNA binding, but also partly anchored in the PLGA matrix.
DNA solution (a concentration in 0.5 ml plain DMEM for desire
The obtained PLGA/PEI nanoparticles were complexed with pDNA
weight ratios) followed by vortexing for 5 s. The PEI–DNA com-
and their physicochemical characteristics and cytotoxicity as well
plexes were prepared likewise. These polyplex dispersions were
as transfection efficiency were examined and compared with poly-
kept at room temperature at least 30 min. The final DNA concen-
plexes based on PEI.
tration in the complexes was 10 lg/ml for characterization. For
the transfection and cytotoxicity studies, a fixed DNA concentra-
2. Materials and methods tion (5 lg/ml) was applied to form PLGA/PEI–DNA complexes.
The ratios expressed in figures (Figs. 3–7) are defined as the weight
2.1. Materials of PEI (the added amount for nanoparticle preparation) to DNA. For
instance, a polymer-to-DNA weight ratio of 1/1 means that 5 lg
PLGA with 40–75 kDa, (lactide:glycolide 65:35 mol/mol) and b- PEI in PLGA/PEI particles (80 lg; calculated from Table 1B) was
galactosidase were obtained from Sigma, UK. PVA (30–50 kDa, 87– mixed with 5 lg DNA in 1 ml medium to obtain the formation A
89% hydrolyzed) and PEI (branched, 25 kDa) were obtained from complexes. For formulation B, 55 lg of PLGA/PEI particles (corre-
Aldrich, Germany. ONPG (ortho-nitrophenyl b-galactoside) was sponding with 5 lg PEI) was mixed with DNA. For formulations C
bought from Bio Basic Inc., USA. Ethidium bromide and MTT prolif- and D, 30 lg and 17.5 lg of PLGA/PEI particles were mixed with
eration kit were products of Sigma. COS-7 cells were obtained from DNA, respectively.
American Type Culture Collection (ATCC).

Table 1
2.2. Preparation of PLGA/PEI cationic nanoparticles Formulation compositions and the particle size and zeta potential of formed PLGA/PEI
nanoparticles and PLGA/PEI–DNA complexes (n = 3; PDI = polydispersity index).
The one-step preparation of PLGA/PEI cationic nanoparticles (A) Effect of PVA concentration on the size PLGA/PEI nanoparticles
was performed as follows. PLGA nanoparticles have been prepared PVA (%, w/v) 0.5% 0.6% 1% 2%
by an emulsion–diffusion–evaporation technique (o/w) using eth- Particle size (nm) ± SD 520 ± 15 230 ± 2 213 ± 3 210 ± 2
ylacetate as solvent for PLGA and an aqueous PVA solution as con- (B) Formulation compositions of PLGA/PEI nanoparticles
tinuous phase (Ravi Kumar et al., 2004). We added PEI to this PVA Formulation A B C D
solution to render the particles cationically. In detail: PLGA (9 mg) PLGA (mg) 54 36 18 9
was dissolved in 4 ml ethyl acetate; PEI (3.6 mg) was dissolved in PEI (mg) 3.6 3.6 3.6 3.6
4 ml of an aqueous solution of emulsifier (PVA; concentration
0.5%, 0.6%, 1.0% and 2.0% (w/v)). In another series of experiments, (C) The size and zeta potential of PLGA/PEI nanoparticles from formulations A, B, C,
D
the PVA concentration (2% w/v) and PEI amount (3.6 mg) were
Formulation Size (nm) ± SD PDI ± SD Zeta potential (mV) ± SD
fixed whereas the amount of PLGA dissolved in 4 ml ethyl acetate
A 264 ± 5 0.10 ± 0.03 18.3 ± 1.2
was varied (9, 18, 36, 54 mg) was varied. An aqueous PEI/PVA solu-
B 250 ± 6 0.10 ± 0.06 15.1 ± 1.6
tion (pH around 10.5) was mixed with the PLGA solution in ethyl- C 223 ± 2 0.15 ± 0.05 9.8 ± 2.5
acetate under stirring for 10 min and an emulsion was formed by D 210 ± 2 0.20 ± 0.05 7.5 ± 2.9
homogenizing at 15,000 rpm for 10 min. Next, 7 ml deionized (D) The size and zeta potential of PLGA/PEI–DNA complexes from formulations A,
water was added and homogenization at 15,000 rpm was contin- B, C, D (PLGA/PEI-to-DNA ratio = 2/1)
ued for 2 min. Subsequently the emulsion was stirred overnight Formulation Size (nm) ± SD PDI ± SD Zeta potential (mV) ± SD
at 40 °C to remove the organic solvent. Finally, 18 ml deionized A 273 ± 8 0.15 ± 0.03 1.0 ± 0.9
water was added and the nanoparticles were washed by 2-cycle B 262 ± 12 0.20 ± 0.06 1.3 ± 0.8
centrifugation (15,000 rpm; Hermle Z323K, Germany) for 30 min C 234 ± 11 0.25 ± 0.05 2.0 ± 1.3
D 230 ± 13 0.25 ± 0.05 1.5 ± 1.7
at 4 °C with deionized water to remove PVA and excess PEI. The
524 M.D. Shau et al. / European Journal of Pharmaceutical Sciences 46 (2012) 522–529

2.4. Particle size and zeta potential measurements form by mitochondrial succinate dehydrogenase a water-soluble
yellow dye, MTT, into purple colored water-insoluble formazan.
The sizes of PLGA/PEI nanoparticles and PLGA/PEI–DNA com- The cells were maintained in Dulbecco’s modified Eagle’s medium
plexes were determined by DLS using a Malvern Zetasizer Nano (DMEM) supplemented with 10% fetal bovine serum (FBS) in 5%
ZS (Malvern, UK). The measurements were performed with a CO2 incubator at 37 °C. 4  103 cells were seeded/well in 96-well
4 mW, 633 nm He–Ne laser at an angle of 173° and dispersant RI microtiter plates followed by incubation for 18 h. The PLGA/PEI–
(1.33) and viscosity (0.8872 cp) were used for the hydrodynamic DNA complexes (with a fixed DNA concentration at 0.5 lg/well
diameter calculation. The zeta potential of the different PLGA/PEI and various PLGA/PEI nanoparticle–DNA ratios, w/w) in 100 ll
nanoparticles and PLGA/PEI–DNA complexes was determined in supplemented medium were added into wells. After 4 h incuba-
Hepes buffer (20 mM, pH 7) using a Malvern Zetasizer Nano ZS tion, the medium was replaced with fresh medium and the cells
(Malvern, UK). The measured electrophoretic mobility (l) was were incubated for another 40 h. Next, 20 ll MTT (from stock
used to obtain the zeta potential (z) by Henry equation (l = 2ezf(- 5 mg/ml in PBS, pH 7.4) was added to the cells for further 3 h incu-
ka)/3g where e is dielectric constant and g is viscosity). The value bation. Finally, the medium was removed and dimethyl sulfoxide
for Henry’s function f(ka) was taken as 1.5. (DMSO) (100 ll/well) was added to wells. The absorbance of for-
mazan was determined at 570 nm on a microplate reader. Relative
2.5. Gel electrophoresis of PLGA/PEI–DNA complexes cell viability (%) was expressed as a percentage to untreated cells.
Cytotoxicity studies with PEI–DNA polyplexes were performed
The electrophoretic mobility of PLGA/PEI–DNA complexes was using the same procedure as PLGA/PEI–DNA complexes (with a
measured in 0.8% (w/v) agarose gel with 1% tris–Borate–EDTA fixed DNA concentration at 0.5 lg/well and various PEI–DNA ra-
(TBE; pH 8.0). Electrophoresis was done at 100 V for 35 min and tios, w/w).
the gels were subsequently immersed into a 0.5 lg/ml ethidium
bromide solution for another 30 min. DNA bands were visualized
by UV irradiation.
2.10. Transfection studies
2.6. Characterization of PLGA/PEI nanoparticles by AFM and SEM
COS-7 cells were maintained in supplemented DMEM as for
The surface morphology of PLGA/PEI nanoparticles was ana-
cytotoxicity studies. The incubation for the PLGA/PEI–DNA com-
lyzed by SEM and AFM. For SEM, a drop of samples was introduced
plexes with cells in the supplemented medium was 4 h. Then,
onto a carbon strip mounted on an aluminum stage. After air-dry-
the transfection medium was replaced by fresh medium and cells
ing, the samples were shadowed with gold–palladium at 15 mA for
were incubated for another 43 h. Gene expression was character-
100 s to minimize surface charging. Observations on their surface
ized by observation of pEGFP-positive cells under a florescent
morphology were conducted with a Hitachi S-3000N SEM at an
microscope and by quantification of pb-gal activities with ONPG
accelerating voltage of 10 kV. For AFM, the used instrument (Nano-
assay under an ELISA reader (420 nm). Transfection studies with
Scope E, Digital Instruments Inc.) was set in a contact mode using
PEI–DNA polyplexes were performed using the same procedure
n+/n silicon (resistivity 0.01–0.02 X cm) cantilever with a spring
as PLGA/PEI–DNA complexes (with a fixed DNA concentration at
constant of 0.07–0.4 N/m and a resonance frequency of 10–17 kHz.
0.5 lg/well and various PEI–DNA ratios, w/w).
Scanning was performed at a speed of 1.0 Hz with a resolution of
512  512 pixels.

2.7. FT-IR measurements of PLGA/PEI nanoparticles


3. Results and discussion
FTIR spectroscopy of PLGA nanoparticles, PEI, and PLGA/PEI
nanoparticles in KBR pellets was conducted with a Mattson Gale- 3.1. The effect of PVA concentration on the size of PLGA/PEI
rxy Series 5000 spectrometer. Each spectrum was corrected for nanoparticles
the background to obtain the sample vibrational spectrum.
We aimed for the preparation of nanoparticles with a si-
2.8. Quantification of PEI incorporated in PLGA/PEI nanoparticles and ze < 200 nm because such particles have been shown to be capable
on PEI coated PLGA nanoparticles of transfecting cells after DNA complexation (Cherng et al., 1996).
The size of PLGA nanoparticles depends on both the formulation
The amount of PEI entrapped in PLGA nanoparticles (formula- (e.g. type and concentration of surfactant and the PLGA concentra-
tions A–D) and PEI coated PLGA nanoparticles (PLGA + PEI) were tion of the organic phase) and the processing parameters (e.g.
examined by ninhydrin colorimetric assay (Chertok et al., 2010). emulsion method). In this study we particularly focused on the
Briefly, the ninhydrin reagent (500 ll of 0.2% w/v in 0.1 M buffer concentration of the surfactant and the PLGA concentration to tai-
phosphate, pH 9) was added to 200 ll dispersion of 1 mg nanopar- lor the size of the nanoparticles. PVA is a commonly used surfac-
ticle in buffer phosphate. The samples were subsequently heated in tant to stabilize the formed emulsion and prevents the droplets
a boiling water bath for 30 min and then cooled to room tempera- against aggregation during preparation of PLGA micro- and nano-
ture. The absorbance of supernatants was measured at 570 nm on a particles (Sahoo et al., 2002). First the influence of the PVA concen-
UV/Vis spectrophotometer (SP8001, Metertech Inc, Taiwan) and tration on the size of PLGA/PEI was investigated. Table 1 shows
the amine content was quantified using an established calibration that the size of PLGA/PEI nanoparticles sharply decreased from
curve of PEI. 520 to 230 nm when the PVA concentration increased from 0.5%
to 0.6%, respectively. The particle size distribution was rather nar-
2.9. Cytotoxicity evaluation row (PDI value around 0.15 (for 0.6% PVA)). When the PVA further
increased to 2% the particle size only slight decreased (from 230 to
The cytotoxicity of different PLGA/PEI–DNA complexes from 210 nm, Table 1). Therefore, the PVA concentration was fixed at 2%
four different formulations was evaluated in COS-7 cells using an for all subsequent experiments; a concentration also selected in
MTT assay. This assay is based on an ability of living cells to trans- other studies (Prabha and Labhasetwar, 2004; Feczko et al., 2008).
M.D. Shau et al. / European Journal of Pharmaceutical Sciences 46 (2012) 522–529 525

3.2. The effect of PLGA concentration on the size of PLGA/PEI


a 20
nanoparticles
PLGA/conjugated PEI (Formulation A)
The effect of the PLGA concentration on the size of PLGA/PEI (PLGA+PEI) (physicially mixed; the control)
nanoparticles was investigated. It was found that smaller PLGA/ 15

zeta potential (mV)


PEI particles (from 264 to 210 nm) were obtained when the
amount of PLGA in a fixed volume of ethyl acetate decreased (from
54 mg to 9 mg; Table 1). When the amount of PLGA in 4 ml of ethyl
10
acetate was 63 mg, no nanoparticle was obtained after emulsifica-
tion of this solution in an aqueous PVA/PEI phase.
It is quite rational that a higher concentration of PLGA in the or-
ganic solvent due to its increasing viscosity increases the size of 5
PLGA/PEI nanoparticles (Hans and Lowman, 2002). Although the
size of polyplexes around 200 nm is slightly too large for i.v. admin-
istration and prolonged circulation (Verbaan et al., 2004), these par-
ticles might have a good size for other routes of administration (e.g. 0
local administration via injection or aerosol; Park et al., 2008). PLGA/PEI PLGA+PEI

3.3. Zeta potential and PEI content of PLGA/PEI nanoparticles b


The zeta potential of nanoparticles is an important characteris-
tic for their stability. It has been reported that the zeta potential of
PLGA nanoparticles in a neutral buffer is around 45 mV; this neg-
ative charge is due to uncapped end carboxyl groups of PLGA pres-
ent on the surface of the particles (Stolnik et al., 1995). However,
such negatively charged particles are unable to complex negatively
charged pDNA. To render the PLGA nanoparticles with a positive
zeta potential, in the present study, we dissolved PEI in the contin-
uous PVA phase used to emulsify the PLGA/ethyl acetate solution.
The results of Table 1 show that indeed positively charged PLGA
nanoparticles were obtained. It was also found that a decrease in
PLGA concentration was associated with a decrease in the zeta po-
tential of the obtained PLGA/PEI nanoparticles (from +18.3 mV to 1 2 3
+7.5 mV, Table 1). When PLGA preformed nanoparticles were
coated with PEI (expressed as PLGA + PEI), they had a zeta-poten- Fig. 1. Zeta potential (A) and electrophoretic pattern (B) of PLGA/PEI nanoparticles
tial of +5 mV (Fig. 1A). The differences in zeta-potential could be prepared in a one-step process (formulation A, Table 1) and PLGA/PEI nanoparticles
obtained by coating of preformed PLGA nanoparticles with PEI (two-step process).
correlated to the extent of PEI incorporation into/onto obtained
(B) Lane 1 = free DNA, lane 2 = (PLGA + PEI)–DNA complexes (nanoparticles
PLGA/PEI nanoparticles. The encapsulation efficiency (% of the prepared in a two-step process), lane 3 = PLGA/PEI–DNA complexes (nanoparticles
added amount) of PEI incorporated in PLGA/PEI nanoparticles is prepared in a one-step process.
75%, 61%, 39%, and 34% for formulations A, B, C and D, respectively.
Moreover, 20% of the added amount of PEI was adsorbed onto the
surface of the PLGA nanoparticles. This means that with the new
process (PEI present in the PVA aqueous phase during emulsifica-
tion of the PLGA/ethyl acetate solution), particles with a higher
zeta-potential and thus higher PEI-loading were obtained. With
this new process, PEI not only interacts with the negatively
charged surface of PLGA nanoparticles by electrostatic interactions,
but possibly PEI also had reacted with PLGA by aminolysis (Croll
et al., 2004). This reaction results in the formation of amide bonds
connecting PEI and PLGA. Indeed, in the FTIR spectrum of the PLGA/
PEI nanoparticles apart from a NH2-vibration of PEI at 1636 cm 1, a
weak but evident absorbance at 1623 cm 1 of amide bonds (Cha-
kravarthi and Robinson, 2011) is detected (Fig. 2). On the other
hand, the spectrum of the spectrum of the PEI coated PLGA nano-
particles showed a broad absorbance at 1628 cm 1 (the Supple-
mentary figure). IR spectra however only give qualitative changes
but not a precise PEI content in the nanoparticles. The ninhydrin
assay showed that a higher PLGA amount present in the formula-
tion leads to a greater extent of PEI conjugation that in turn re-
sulted in higher zeta potential of obtained PLGA/PEI nanoparticles.

3.4. Complexation and condensation of DNA by PLGA/PEI nanoparticles

Gel electrophoresis was used to study the ability of the cation-


ically charged PLGA/PEI nanoparticles to bind DNA. Fig. 3 shows
that the different PLGA/PEI nanoparticles in formulations A–D Fig. 2. FT IR spectra of PLGA, PLGA/PEI and PEI.
526 M.D. Shau et al. / European Journal of Pharmaceutical Sciences 46 (2012) 522–529

Formulation A Formulation B
1 2 3 4 5 6 1 2 3 4 5 6

Formulation C Formulation D
1 2 3 4 5 6 1 2 3 4 5 6

DNA 8/1 4/1 2/1 1/1 0.5/1 DNA 8/1 4/1 2/1 1/1 0.5/1

Fig. 3. Electrophoretic patterns of PLGA/PEI–DNA formulations. Lane 1: 400 ng DNA and lanes 2–6: PLGA/PEI–DNA complexes were prepared at Hepes (20 mM, pH 7);
formulations A–D refers to Table 1.

Fig. 4a. SEM micrographs of PLGA/PEI nanoparticles of formulations A–D.

were able to quantitatively complex DNA at PLGA/PEI-to-DNA ra- The lower DNA binding capacity of PEI-coated PLGA nanoparticles
tios (w/w) above 2/1. At lower ratios free DNA was detected. In is likely caused by their lower zeta-potential (+5 mV).
Fig. 1B, the PLGA nanoparticles were coated with PEI had much Binding of DNA to PLGA conjugated PEI nanoparticles (PLGA/
lower DNA binding capacity (at a PLGA/PEI-to-DNA weight ratio PEI–DNA complexes) resulted in a slight increase in size (from
of 2/1). Free DNA was observed present, whereas the PLGA/PEI 230 to 273 nm) whereas the zeta potential of these particles was
nanoparticles (from formulation A) showed complete DNA binding. close to neutral. This indicates that the surface of the particles is
M.D. Shau et al. / European Journal of Pharmaceutical Sciences 46 (2012) 522–529 527

Fig. 4b. AFM images of PLGA/PEI nanoparticles (formulation A, upper graph; and formulation D, lower graph).

70
100
CD 60
B
Beta-galactosidase (mU/mg protein)

A
Relative cell viability (%)

80 BC C
PEI
A AB D 50 A
B
A
D B
60
40

PEI CD
40 CD
PEI 30
AB
PEI
20 A
20 CD
PEI PEI B
PEI 10
0 CD C
1/1 2/1 4/1 8/1 AB D PEI
0
PLGA/PEI-to-DNA ratios 1/1 2/1 4/1 8/1
PLGA/PEI-to-DNA ratios
Fig. 5. Cytotoxicity of PLGA/PEI–DNA complexes prepared using different PLGA/PEI
nanoparticles (formulations A–D, Table 1 and different PLGA/PEI-to-DNA ratios, w/ Fig. 6. Transfection efficiency of PLGA/PEI–DNA (pb-gal) complexes (prepared at
w) and PEI–DNA polyplexes analyzed by MTT assay, 47 h after transfection in COS-7 plain DMEM at different PLGA/PEI-to-DNA ratios, w/w) and PEI–DNA polyplexes in
cells in the presence of 10% serum. Results are presented as mean ± SD (n = 3). the presence of 10% serum. The complexes were incubated for 4 h with COS cells
after which the cells were washed and incubated for another 43 h. Protein
coated with DNA as shown in Fig. 3, the binding capability of PLGA/ expression was measured with ONPG assay under an ELISA reader (420 nm); n=3.
PEI nanoparticles from formulations A to D.
zeta potential (e.g. formulations C and D, Table 1) showed some fu-
3.5. SEM and AFM observations sion of particles (Fig. 4A, pointed with arrows), whereas individual,
non-fused/agglomerated, particles were observed in formulation A
The surface morphology of PLGA/PEI nanoparticles was ana- and B. Clearly, a higher zeta potential (e.g. formulation A and B) of
lyzed by SEM and AFM. The PLGA/PEI nanoparticles with lower the PLGA/PEI nanoparticles prevented agglomeration/fusion due to
528 M.D. Shau et al. / European Journal of Pharmaceutical Sciences 46 (2012) 522–529

Fig. 7. Transfection efficiency of PLGA/PEI–DNA (pEGFP) complexes (prepared at plain DMEM at different PLGA/PEI-to-DNA ratios, w/w) and PEI–DNA polyplexes in the
presence of serum (10%). The complexes were incubated for 4 h with COS cells after which the cells were washed and incubated for another 43 h. The efficiency of PLGA/PEI–
DNA (pEGFP) complexes was analyzed by the fluorescence-expressing cells.

higher repulsion forces. AFM was applied to study the PLGA/PEI Fig. 6 shows that the PLGA/PEI–DNA complexes have a substan-
nanoparticles in solution. In line with the SEM observations, AFM tially better transfection activity than PEI polyplexes. Moreover,
micrographs of nanoparticles of formulation A showed that these the gene expression was found dependent to a PLGA/PEI-to-DNA
nanoparticles were more uniform in size than those of formulation ratio. At higher ratios (e.g. w/w = 2/1 or 4/1), PLGA/PEI–DNA com-
D (Fig. 4B). plexes showed a higher transfection efficiency till a ratio (w/w > 8/
1) where a decrease in transfection activity was observed, likely
due to the cytotoxicity of the formulations (Fig. 5) as we previous
3.6. Cytotoxicity of PLGA/PEI–DNA complexes
observed (Cherng et al., 2011).
Fig. 6 also shows that the formulation A and B have higher
Cationic gene delivery polymers are known for their cytotoxic-
transfection efficiency than C and D which might be explained by
ity (Fisher et al., 2003); e.g. PEI induces cell death by membrane
the higher zeta potential of the naked particles which give better
destabilization prior to cellular internalization (Godbey et al.,
DNA retention and thus better intracellular delivery of the nucleic
2001). In line herewith, Fig. 5 shows that PEI–DNA polyplexes
acid.
(fixed amount of DNA = 0.5 lg/well) in the presence of serum
The transfection ability of PLGA/PEI nanoparticles complexed
showed a dose-dependent toxicity (IC50 between 1/1 and 2/1, w/
with DNA encoding for green fluorescent protein (pEGFP) was also
w) towards COS 7 cells. This figure also shows that the PLGA/
examined in COS-7 cells. In line with the results of Fig. 6, the com-
PEI–DNA formulations showed a substantial lower cytotoxicity
plexes made at PLGA/PEI-to-DNA ratios of 2/1 and 4/1 resulted in
(IC 50 was around 8/1, w/w) than PEI–DNA polyplexes in the
the highest fluorescence (Fig. 7) and thus higher protein expres-
examined range (from w/w = 1/1 to 8/1) meaning that less non-
sion. Also, the PLGA/PEI–DNA complexes had a higher transfection
degradable PEI is required for delivering DNA.
activity than the PEI polyplexes.

3.7. Transfection activity of PLGA/PEI–DNA complexes


4. Conclusion
The different PLGA/PEI nanoparticle DNA formulations were
evaluated for their transfection ability in COS-7 cells in the pres- In this study, a substantial advantage PLGA/PEI nanoparticles
ence of serum and compared with that of PEI polyplexes using prepared in a one-step process has been shown over the prepara-
pDNA encoding for b-galactosidase (pb-gal) as reporter genes. tion of PLGA nanoparticles that were coated with PEI in a two-step
M.D. Shau et al. / European Journal of Pharmaceutical Sciences 46 (2012) 522–529 529

process. The nanoparticles prepared using the one-step procedure Jong, Y.S., Jacob, J., Yip, K.P., Gardner, G., Seitelman, E., Whitney, M., Montgomery, S.,
Mathiowitz, E., 1997. Controlled release of plasmid DNA. J. Control. Release 47,
are spherical in morphology, have a small size distribution and
123–134.
have a much better DNA binding capacity than the particles pre- Katas, H., Cevher, E., Alpar, H.O., 2009. Preparation of polyethyleneimine
pared in a two step process. Importantly, they have a relatively incorporated poly(D,L-lactide-co-glycolide) nanoparticles by spontaneous
low cytotoxicity and a substantial better transfection activity of emulsion diffusion method for small interfering RNA delivery. Int. J. Pharm.
369, 144–154.
PEI polyplexes even in the presence of serum. Kim, T.H., Jiang, H.L., Jere, D., Park, I., Cho, M.H., Nah, J.W., Choi, Y.J., Akaike, T., Cho,
C.S., 2007. Chemical modification of chitosan as a gene carrier in vitro and
in vivo. Prog. Polym. Sci. 32, 726–753.
Appendix A. Supplementary material Kwoh, D.Y., Coffin, C.C., Lollo, C.P., Jovenal, J., Banaszczyk, M.G., Mullen, P., Phillips,
A., Amini, A., Fabrycki, J., Bartholomew, R.M., Brostoff, S.W., Carlo, D.J., 1999.
Stabilization of poly-L-lysine/DNA polyplexes for in vivo gene delivery to the
Supplementary data associated with this article can be found, in liver. Biochim. Biophys. Acta 1444, 171–190.
the online version, at http://dx.doi.org/10.1016/j.ejps.2012.04.006. Luten, J., van Nostrum, C.F., De Smedt, S.C., Hennink, W.E., 2008. Biodegradable
polymers as non-viral carriers for plasmid DNA delivery. J. Control. Release 126,
97–110.
References Mintzer, M.A., Simanek, E.E., 2009. Nonviral vectors for gene delivery. Chem. Rev.
109, 259–302.
Ando, S., Putnam, D., Pack, D.W., Langer, R., 1999. PLGA microspheres containing Nafee, N., Taetz, S., Schneider, M., Schaefer, U.F., Lehr, C.M., 2007. Chitosan-coated
plasmid DNA: preservation of supercoiled DNA via cryopreparation and PLGA nanoparticles for DNA/RNA delivery: effect of the formulation parameters
carbohydrate stabilization. J. Pharm. Sci. 88, 126–130. on complexation and transfection of antisense oligonucleotides. Nanomed.
Basarkar, A., Devineni, D., Palaniappan, R., Singh, J., 2007. Preparation, Nanotechnol. Biol. Med. 3, 173–183.
characterization, cytotoxicity and transfection efficiency of poly(DL-lactide-co- Neu, M., Fischer, D., Kissel, T., 2005. Recent advances in rational gene transfer vector
glycolide) and poly(DL-lactic acid) cationic nanoparticles for controlled delivery design based on poly(ethylene imine) and its derivatives. J. Gene Med. 7, 992–
of plasmid DNA. Int. J. Pharm. 343, 247–254. 1009.
Bivas-Benita, M., Romeijn, S., Junginger, H.E., Borchard, G., 2004. PLGA–PEI Park, M.R., Kim, H.W., Hwang, C.S., Han, K.O., Choi, Y.J., Song, S.C., Cho, M.H., Cho,
nanoparticles for gene delivery to pulmonary epithelium. Eur. J. Pharm. C.S., 2008. Highly efficient gene transfer with degradable poly(ester amine)
Biopharm. 58, 1–6. based on poly(ethylene glycol) diacrylate and polyethylenimine in vitro and
Borchard, G., 2001. Chitosans for gene delivery. Adv. Drug Deliv. Rev. 52, 145–150. in vivo. J. Gene Med 10, 198–207.
Boussif, O., Lezoualc’h, F., Zanta, M.A., Mergny, M., Scherman, D., Demeneix, B., Behr, Perez, C., Sanchez, A., Putnam, D., Ting, D., Langer, R., Alonso, M.J., 2001. Poly(lactic
J.P., 1995. A versatile vector for gene and oligonucleotide transfer into cells in acid)-poly(ethylene glycol) nanoparticles as new carriers for the delivery of
culture and in vivo: polyethylenimine. Proc. Natl. Acad. Sci. USA 92, 7297–7301. plasmid DNA. J. Control. Release 75, 211–224.
Capan, Y., Woo, B.H., Gebrekidan, S., Ahmed, S., DeLuca, P.P., 1999. Influence of Prabha, S., Labhasetwar, V., 2004. Critical determinants in PLGA/PLA nanoparticle-
formulation parameters on the characteristics of poly(DL-lactide-co-glycolide) mediated gene expression. Pharm. Res. 21, 354–364.
microspheres containing poly(L-lysine) complexed plasmid DNA. J. Control. Ravi Kumar, M.N.V., Bakowsky, U., Lehr, C.M., 2004. Preparation and
Release 60, 279–286. characterization of cationic PLGA nanospheres as DNA carriers. Biomaterials
Chakravarthi, S.S., Robinson, D.H., 2011. Enhanced cellular association of paclitaxel 25, 1771–1777.
delivered in chitosan-PLGA particles. Int. J. Pharm. 409, 111–120. Romero, G., Estrela-Lopis, I., Zhou, J., Rojas, E., Franco, A., Espinel, C.S., Fernandez,
Cherng, J.Y., van de Wetering, P., Talsma, H., Crommelin, D.J.A., Hennink, W.E., 1996. A.G., Gao, C., Donath, E., Moya, S.E., 2010. Surface engineered poly(lactide-co-
Effect of size and serum proteins on transfection efficiency of poly((2- glycolide) nanoparticles for intracellular delivery: uptake and cytotoxicitys – a
dimethylamino)ethyl methacrylate)-plasmid nanoparticles. Pharm. Res. 13, confocal Raman microscopic study. Biomacromolecules 11, 2993–2999.
1038–1042. Romøren, K., Thu, B.J., Bols, N.C., Evensen, Ø., 2004. Transfection efficiency and
Cherng, J.Y., Hung, W.C., Kao, H.C., 2011. Blending of polyethylenimine with a cytotoxicity of cationic liposomes in salmonid cell lines of hepatocyte and
cationic polyurethane greatly enhances both DNA delivery efficacy and reduces macrophage origin. Biochim. Biophys. Acta 1663, 127–134.
the overall cytotoxicity. Curr. Pharm. Biotechnol. 12, 839–846. Sahoo, S.K., Panyam, J., Prabha, S., Labhasetwar, V., 2002. Residual polyvinyl alcohol
Chertok, B., David, A.E., Yang, V.C., 2010. Polyethyleneimine-modified iron oxide associated with poly (D,L-lactide-co-glycolide) nanoparticles affects their
nanoparticles for brain tumor drug delivery using magnetic targeting and intra- physical properties and cellular uptake. J. Control. Release 82, 105–114.
carotid administration. Biomaterials 31, 6317–6324. Shau, M.D., Tseng, S.J., Yang, T.F., Cherng, J.Y., Chin, W.K., 2006. Effect of molecular
Chumakova, O.V., Liopo, A.V., Andreev, V.G., Cicenaite, I., Evers, B.M., Chakrabarty, S., weight on the transfection efficiency of novel polyurethane as a biodegradable
Pappas, T.C., Esenaliev, R.O., 2008. Composition of PLGA and PEI/DNA gene vector. J. Biomed. Mater. Res. A 77, 736–746.
nanoparticles improves ultrasound-mediated gene delivery in solid tumors Stolnik, S., Garnett, M.C., Davies, M.C., Illum, L., Bousta, M., Vert, M., Davis, S.S., 1995.
in vivo. Cancer Lett. 261, 215–225. The colloidal properties of surfactant-free biodegradable nanospheres from
Croll, T.I., O’Connor, A.J., Stevens, G.W., Cooper-White, J.J., 2004. Controllable surface poly(b-malic acid-co-benzyl malate)s and poly(lactic acid-co-glycolide).
modification of poly(lactic-co-glycolic acid)(PLGA) by hydrolysis or aminolysis Colloids Surf. A – Physicochem. Eng. Aspects 97, 235–245.
I: physical, chemical, and theoretical aspects. Biomacromolecules 5, 463–473. Thomas, C.E., Ehrhardt, A., Kay, M.A., 2003. Progress and problems with the use of
Feczko, T., Toth, J., Gyenis, J., 2008. Comparison of the preparation of PLGA-BSA viral vectors for gene therapy. Nat. Rev. Genet. 4, 346–358.
nano- and microparticles by PVA, poloxamer and PVP. Colloids Surf. A – Tseng, S.J., Tang, S.C., Shau, M.D., Zeng, Y.F., Cherng, J.Y., Shih, M.F., 2005. Structural
Physicochem. Eng. Aspects 319, 188–195. characterization and buffering capacity in relation to the transfection efficiency
Fisher, D., Li, Y.X., Ahlemeyer, B., Krieglstein, J., Kissel, T., 2003. In vitro cytotoxicity of biodegradable polyurethane. Bioconjugate Chem. 16, 1375–1381.
testing of polycations: influence of polymer structure on cell viability and Van de Wetering, P., Cherng, J.Y., Talsma, H., Hennink, W.E., 1997. Relation between
hemolysis. Biomaterials 24, 1121–1131. transfection efficiency and cytotoxicity of poly(2-dimethylamino)ethyl
Godbey, W.T., Wu, K.K., Mikos, A.G., 2001. Poly(ethylenimine)-mediated gene methacrylate/plasmid complexes. J. Control. Release 49, 59–69.
delivery affects endothelial cell function and viability. Biomaterials 2, 471–480. Verbaan, F.J., Oussoren, C., Snel, C.J., Crommelin, D.J.A., Hennink, W.E., Storm, G.,
Gorecki, D.C., 2001. Prospects and problems of gene therapy: an update. Expert 2004. Steric stabilization of poly(2-(dimethylamino)ethyl methacrylate)-based
Opin. Emerg. Drugs 6, 187–198. polyplexes mediates prolonged circulation and tumor targeting in mice. J. Gene
Hans, M.L., Lowman, A.M., 2002. Biodegradable nanoparticles for drug delivery and Med 6, 64–75.
targeting. Curr. Opin. Solid State Mater. Sci. 6, 319–327. Zhang, X., Godbey, W.T., 2006. Viral vectors for gene delivery in tissue engineering.
Hung, W.C., Shau, M.D., Kao, H.C., Shih, M.F., Cherng, J.Y., 2009. The synthesis of Adv. Drug Deliv. Rev. 58, 515–534.
cationic polyurethanes to study the effect of amines and structures on their Zhdanov, R.I., Podobed, O.V., Vlassov, V.V., 2002. Cationic lipid–DNA complexes–
DNA transfection potential. J. Control. Release 133, 68–76. lipoplexes–for gene transfer and therapy. Bioelectrochemistry 58, 53–64.

You might also like