ArticlePDF Available

Structure and organization of phospholipid/polysaccharide nanoparticles

IOP Publishing
Journal of Physics: Condensed Matter
Authors:

Abstract and Figures

In recent years nanoparticles and microparticles composed of polymeric or lipid material have been proposed as drug carriers for improving the efficacy of encapsulated drugs. For the production of these systems different materials have been proposed, among them phospholipids and polysaccharides due to their biocompatibility, biodegradability, low cost and safety. We report here a morphological and structural investigation, performed using cryo-TEM, static light scattering and small angle neutron and x-ray scattering, on phospholipid/saccharide nanoparticles loaded with a lipophilic positively charged drug (tamoxifen citrate) used in breast cancer therapy. The lipid component was soybean lecithin; the saccharide one was chitosan that usually acts as an outer coating increasing vesicle stability. The microscopy and scattering data indicate the presence of two distinct nanoparticle families: uni-lamellar vesicles with average radius 90 Å and multi-lamellar vesicles with average radius 440 Å. In both families the inner core is occupied by the solvent. The presence of tamoxifen gives rise to a multi-lamellar structure of the lipid outer shell. It also induces a positive surface charge into the vesicles, repelling the positively charged chitosan molecules which therefore do not take part in nanoparticle formation.
Content may be subject to copyright.
IOP PUBLISHING JOURNAL OF PHYSICS: CONDENSED MATTER
J. Phys.: Condens. Matter 20 (2008) 104211 (8pp) doi:10.1088/0953-8984/20/10/104211
Structure and organization of
phospholipid/polysaccharide nanoparticles
Y Gerelli1,MTDiBari
1, A Deriu1,LCant`
u2,PColombo
3,
CComo
3, S Motta3,FSonvico
3and R May4
1Dipartimento di Fisica and CNISM, Universit`a degli Studi di Parma and CRS SOFT,
INFM-CNR, Italy
2Dipartimento di Chimica, Biochimica e Biotecnologie per la Medicina—LITA, Universit`adi
Milano, Italy
3Dipartimento Farmaceutico, Universit`a degli Studi di Parma, Italy
4Institut Laue-Langevin, Grenoble, France
E-mail: Antonio.Deriu@fis.unipr.it
Received 16 July 2007, in final form 1 November 2007
Published 19 February 2008
Online at stacks.iop.org/JPhysCM/20/104211
Abstract
In recent years nanoparticles and microparticles composed of polymeric or lipid material have
been proposed as drug carriers for improving the efficacy of encapsulated drugs. For the
production of these systems different materials have been proposed, among them phospholipids
and polysaccharides due to their biocompatibility, biodegradability, low cost and safety. We
report here a morphological and structural investigation, performed using cryo-TEM, static light
scattering and small angle neutron and x-ray scattering, on phospholipid/saccharide
nanoparticles loaded with a lipophilic positively charged drug (tamoxifen citrate) used in breast
cancer therapy. The lipid component was soybean lecithin; the saccharide one was chitosan that
usually acts as an outer coating increasing vesicle stability.
The microscopy and scattering data indicate the presence of two distinct nanoparticle
families: uni-lamellar vesicles with average radius 90 ˚
A and multi-lamellar vesicles with
average radius 440 ˚
A. In both families the inner core is occupied by the solvent. The presence
of tamoxifen gives rise to a multi-lamellar structure of the lipid outer shell. It also induces a
positive surface charge into the vesicles, repelling the positively charged chitosan molecules
which therefore do not take part in nanoparticle formation.
1. Introduction
Nanoscience has been recently indicated as one of the
most promising fields for technological innovation. Among
other applications, nanotechnologies are expected to provide
breakthrough improvements in healthcare and medicine. In
particular therapy, regenerative medicine, diagnosis and drug
delivery [1,2] are expected to benefit from new nanomaterials.
The need of new formulations for biotech drugs, as well
as the opportunity of revaluing old drugs through new
administration routes and innovative therapeutic applications,
have given great impulse to the development of pharmaceutical
nanotechnologies and microtechnologies. Indeed, their
application allows the modification of the biopharmaceutic
properties of drugs without changing their molecular structure;
it is therefore an interesting approach for the administration
of drugs presenting some hindering characteristics, such
as low water solubility, rapid metabolism or elimination,
narrow therapeutic index, toxic side effects. Several colloidal
drug delivery systems have been proposed, as liposomes,
micelles, nanoemulsions and nanoparticles. Polymeric and
lipidic nanosystems improve drug bioavailability, modify
pharmacokinetics or protect the encapsulated drug from
enzymatic degradation. Some liposomal drugs have been
approved or are under evaluation for clinical application,
in particular for cancer treatment [3,4]. Polymeric
nanoparticles built up from biocompatible and biodegradable
materials, such as poly(lactic acid) or poly(lactic-co-glycolic
acid) have been investigated [5]. Synthetic polymers
however are often insoluble in water and require the use
of organic solvents for the manufacture of colloidal drug
carriers. For this reason, great interest has been devoted
0953-8984/08/104211+08$30.00 ©2008 IOP Publishing Ltd Printed in the UK1
J. Phys.: Condens. Matter 20 (2008) 104211 Y Gerelli et al
Figure 1. Cryo-TEM image of nanoparticles containing tamoxifen (panel (a)). The presence of two families (uni- and multi-lamellar vesicles)
is evident, as well as a high overall polydispersity. In the multi-lamellar vesicles, the thickness tcof the region between bilayers is also
polydisperse. For comparison a cryo-TEM micrograph for nanoparticles loaded with a lipophilic non-charged drug (progesterone) is also
shown (panel (b)). Their structure is similar to that of empty nanoparticles (data to be published).
to colloidal preparations obtained from biomaterials like
polysaccharides and lipids that are considered biocompatible,
biodegradable and safe. Recently, we have started a
systematic investigation of the pharmaceutical [6]and
physical–chemical [7] characteristics of new types of
nanosystems obtained by self-assembling of phospholipids
(soybean lecithin) and chitosan. Lecithin is a natural
mix of phospholipids, obtained from soybean, composed
mainly of phosphatidylcholine and phosphatidylethanolamine,
that has been frequently used for liposome and micelle
preparation [8,9]. Chitosan, a cationic polysaccharide
[poly-β-(1,4) glucosamine, (C6H11O4N)n] derived from chitin
by deacetylation, is a versatile biomaterial with favourable
biochemical properties: no toxicity, biocompatibility,
biodegradability, bioadhesion [10,11]. It has been proposed
for a number of clinical and biomedical applications, ranging
form scaffolds for tissue engineering to the production of
colloidal drug carriers obtained by ionotropic gelation of the
polymer with tripolyphosphate [12,13]. Additionally, chitosan
has been investigated for stabilization of microemulsions in
which lecithin was one of the emulsifying agents [14]. In
most of the cases reported in the literature, it turns out that the
saccharide component acts as an outer coating that stabilizes
the lipid vesicle structure. The lecithin/chitosan nanoparticles
obtained with our preparation method are characterized by an
hollow core surrounded by a lipid bilayer with a thin chitosan
outer coating. This structure is observed in nanoparticles
not containing drugs (void nanovectors), as well as in those
loaded with a lipophilic non-charged drug (cryo-TEM image
reported in figure 1; small angle neutron scattering data to
be published). Here we present a structural investigation,
performed by cryo-TEM (cTEM), static light scattering (SLS)
and small angle scattering of neutrons (SANS) and x-rays
(SAXS) on lecithin/chitosan nanoparticles in the presence of
tamoxifen citrate, a lipophilic positively charged drug largely
used in breast cancer therapy. Tamoxifen is a drug with a very
low water solubility and a poor and erratic bioavailability after
oral administration and therefore is expected to benefit from
its formulation in biocompatible nanocarriers. Considering
the lipophilic nature of this drug is it possible to hypothesize
that its molecules will be mostly present into the lipid part
of the investigated nanocapsules. This is a common feature
of different kind of lipid–drug delivery systems [15–17]. A
detailed knowledge of the changes in the nanoparticle structure
due to the presence of the drug is important in order to optimize
the loading efficiency and the kinetics of drug release.
2. Experimental details
2.1. Sample preparation and preliminary characterization
The main component of our system is Lipoid S45 (a commer-
cial lecithin manufactured by Lipoid AG, Switzerland): a mix-
ture of lipids, phospholipids and 30% of fatty acids with an
overall negative charge. The second component is highly pu-
rified chitosan, a polysaccharide obtained by deacetylation of
chitin which is positively charged at acid pH. The third compo-
nent is tamoxifen citrate [18], a lipophilic drug used in breast
cancer therapy, positively charged at acid pH. We started the
preparation from two solutions: a first one composed by 200
mg of lecithin plus 60 mg of TAM all dissolved in 8 ml of
methanol, a second one made up by 10 mg of chitosan di-
luted in 92 ml of water (H2OorD
2O according to custom).
Self-assembled nanoparticles were obtained by rapid injection
(nozzle diameter 0.75 mm, injection rate 40 ml min1), under
mechanical stirring, of the methanol solution into the aqueous
one. The self-assembling process gave rise to a suspension of
drug-loaded nanoparticles with concentration c=0.2% w/w
and pH =2.7[6]. It is worth noticing that TAM is already
admixed with lecithin in the methanol solution while chitosan
interacts whit lecithin only upon injection. For cTEM, SAXS
and SLS measurements, the solvent was a mixture of pure
H2O (92%) and methanol (8%). For the SANS experiments
two buffers were used: a perdeuterated one (D2O 92% and
D-methanol 8%), and a D2O/H2O mixture corresponding to
the index matching of lecithin (73.7% H2O, 18.3% D2O, 8%
H-methanol). In order to determine the mean hydrodynamic
radius and the Z-potential (connected to the charge per unit
surface area) a preliminary characterization was carried out
by dynamic light scattering. Samples containing lecithin and
lecithin/chitosan nanoparticles were also measured for com-
parison. The results are summarized in table 1. The particle
morphology was investigated by cTEM experiments. cTEM is
based on electron transmission microscopy on thin (5000 ˚
A)
vitrified aqueous films [19] obtained by cooling to liquid nitro-
gen temperature. The limiting factor, when using electrons as
probes, is the small difference between the electronic density
of amphiphilic molecules and that of the surrounding water; the
experimental resolution can not therefore be higher than 40–
50 ˚
A. This corresponds to a typical bilayer thickness, therefore
the internal bilayer structure can not be observed. Figure 1
shows cTEM micrographs for a typical suspension of TAM-
loaded nanoparticles (panel (a)), and nanoparticles loaded with
2
J. Phys.: Condens. Matter 20 (2008) 104211 Y Gerelli et al
Table 1. Comparison between average hydrodynamic radius (RH)
and surface potential of lecithin vesicles (LEC), chitosan/lecithin
nanoparticles (NCL) and those with TAM.
System RH
(˚
A)
Polydispersity
index (%)
Z-potential
(mV)
LEC 350 ±20 20 48 ±2
NCL 1140 ±20 20 45 ±2
NCL+TAM 670 ±20 20 45 ±3
progesterone, a non-charged lipophilic drug (panel (b)). The
difference in structure is very evident: nanoparticles with pro-
gesterone are characterized by an hollow core surrounded by a
lipid bilayer with a thin chitosan outer coating; this structure
is very close to that observed in empty nanoparticles obtained
with the same preparation method. In nanoparticles with TAM
a multi-layered structure is clearly visible.
2.2. Static light scattering
The measurements were performed with a 677 nm incident
wavelength covering a range of momentum Qfrom 3.0×104
to 2.4×103˚
A1. Sample suspensions for SLS were diluted to
a concentration c=0.04% (w/w), adding water and 0.0001%
of HCl in order to keep a constant pH. This was essential to
avoid multiple scattering contributions.
2.3. Small angle neutron scattering
SANS experiments were performed at the Institut Laue-
Langevin, ILL (Grenoble) using the D11 diffractometer. Three
different wavelength, λ, and sample-to-detector distance, D,
configurations were adopted (λ=5˚
A, D=5.5m;λ=
5˚
A, D=20.5m;λ=10 ˚
A, D=34 m) in order to
cover a Q-range of about two decades from 9.4×104to
0.1˚
A1. Sample holders were standard quartz flat cells (1 mm
thick). Standard corrections, cell subtraction and normalization
to absolute scattering units were performed using ILL SANS
routines.
2.4. Small angle x-ray scattering
SAXS data were collected at the ID02 beamline at the
European Synchrotron radiation Facility, ESRF (Grenoble).
Scattered x-rays (λ=1˚
A) were collected using a charge-
coupled device (CCD). The sample-to-detector distance, D was
1.2 m; this configuration allows to explore a Q-range from
8×103to 4 ×101˚
A1. Calibration was performed using
a silver behenate standard. Collected images were processed
and corrected with the ID02 software; accurate background
subtraction was performed by us at a later stage.
3. Data analysis
3.1. SLS and SANS
AsimplelowQGuinier analysis was not feasible for both
SANS and SLS data due to the high overall nanoparticle
polydispersity (see figure 1). We therefore decided to assume
Figure 2. SANS and SLS profiles for the investigated nanoparticle
solutions. SANS measurements in deuterated solvent ()arein
excellent agreement with the SLS ones (). SANS data at the
lecithin matching point are also reported ().
an analytical model for the scattering cross-section d(Q)/d
suitable for the whole Q-range explored. Nanoparticle
suspensions for the SLS experiments required a dilution higher
than that used in SANS measurements (50 times). In order
to properly match SLS and SANS data, we performed a
preliminary SANS measure at the SLS concentration (0.04%,
data not shown); these low concentration SANS data set can
be perfectly superimposed to the higher concentration ones by
re-scaling them according to their dilution ratio. This indicates
that the overall size and structure of the nanoparticles are not
appreciably influenced by dilution. The SANS curves were
expressed in absolute scattering units and the SLS data were
then normalized to the SANS ones obtaining a continuous
curve covering almost three decades in Q(figure 2). The
extremely good agreement of the SANS and SLS data in the
overlap region has to be remarked. For a dilute solution of
homogeneous and identical particles the scattered intensity
I(Q)is proportional to
d(Q)
d=n
Vρ 2
p|F(Q)|2(1)
where n/Vis the particle number density, ρ pρpρsis
the contrast between the scattering length density, SLD, of the
particles (ρp) and of the solvent (ρs)[20]. In order to take into
account the nanoparticle polydispersity the form factor F(Q)
has been weighted according to a Schultz distribution [21,22],
D(σ, R), with polydispersity index σand centroid R0.The
differential scattering cross-section becomes therefore
d(Q)
d=φ
Vρ 2
p+∞
0
dRD(σ, R0)|F(Q,R)|2.(2)
Here the n/Vin equation (1) is replaced by φ/V,where
Vis the mean particle volume according to the adopted
distribution (for a Schultz distribution R3=2+1)
(2σ2+1)R3)andφis the particle volume fraction. The ratio
φ/Vis then equal to the average particle number density.
Guided by the morphological cTEM analysis, we adopted a
model that takes into account both uni-lamellar(ULV) vesicles,
3
J. Phys.: Condens. Matter 20 (2008) 104211 Y Gerelli et al
and multi-lamellar ones (MLV). The resulting differential
cross-section is then the sum of two contributions shown in
equation (3) and explained in detail in the following sections.
AQ-independent background is also added to account for the
incoherent scattering contribution from the hydrogens present
in the sample
d(Q)
d=p×φ
Vu
0
Duu,Ru
0)F2
u(Q,R)dR
+(1p)φ
Vm
0dND
mm,N)
0
dtDc,tc)F2
m(Q,R(N,t))+bkg.(3)
The subscripts u and m indicate the type of particles involved
in each scattering contribution (ULV and MLV, respectively)
and pis the relative population of ULV particles.
3.1.1. ULV particles. The model takes into account
a polydisperse core (mean radius Ru
0), described by a
Schultz distribution (Duu,Ru
0)), surrounded by a shell with
thickness tu
B[23]. The total particle radius, Ru
0+tu
B,has
therefore the same polydispersity index (σu) as the core. The
resulting form factor is
Fu(Q,R)=3ρu
BV(R+tu
B)j1(Q(R+tu
B))
Q(R+tu
B)
V(R)j1(QR)
QR +3ρ0V(R)j1(QR)
QR (4)
where j1is the spherical Bessel function of the first kind. The
parameters of the ULV model are therefore the particle number
density p×φ
Vu, the polydispersity index σu, the inner radius Ru
0,
the lipid bilayer thickness tu
Band the two contrasts, ρ0and
ρ u
Bfor core and bilayer respectively.
3.1.2. MLV particles. In agreement with the morphological
cTEM results, we assumed the contrast profile schematized
in figure 3. It originates from an hollow core with radius
Rm
0and contrast ρ m
0, a first bilayer with thickness tm
Band
contrast ρ m
Bfollowed by a repetition of N−1 units, each
composed by a solvent layer (thickness tcand contrast ρm
0)
plus a lipid bilayer. The particle polydispersity is due to
the variable number of bilayers (described by the distribution
Dmm,N)) and to the polydispersity of the inter-layer
thickness, tc, described by a further Schultz distribution
Dc,tc). The two above distributions (over Nand over
tc) play a totally different role. The first one affects the
external radius value and gives rise to a smearing of the SANS
curve mostly near the Guinier region. The latter is needed
to account for the width of the broad peak centred at Q
7×102˚
A1(see figure 2) and does not affect the mean total
radius Rtot. The form factor for the MLV particles can then
be expressed as
Fm(Q,N,t)=4πρm
0Rm
0
0
dRR
2sin(QR)
QR
+4πρm
BRm
0+tm
B
Rm
0
dRR
2sin(QR)
QR
Figure 3. Radial SANS contrast profile adopted for multi-lamellar
vesicle population (MLV).
+4π
N1
j=0ρ m
0Rc
j
R0
j
dRR
2sin(QR)
QR
+ρ m
BR0
j+1
Rc
j
dRR
2sin(QR)
QR (5)
where the integration limits are
R0
j=Rm
0+(j+1)tm
B+jt
outer radius of the (j+1)th bilayer
Rc
j=Rm
0+(j+1)(tm
B+t)
outer radius of the (j+1)th solvent layer.
Equation (5) can be explicitly expressed in terms of spherical
Bessel functions of the first kind, j1, recalling that
4πB
A
dxx
2sin(Qx)
Qx
=34
3πB3j1(QB)
QB 4
3πA3j1(QA)
QA .(6)
For this model the parameters are: two polydispersity indices,
σmand σc, the number of bilayers, N, the core radius Rm
0,the
thickness of the lipid layer and of the inter-layer (tm
Band tc
respectively) and the contrasts ρ m
0and ρ m
B. During the fit
runs they were not all kept free at the same time. As explained
in the following section, the results of the cTEM analysis and
of first test of the model enabled us to fix some of them and to
select well defined limits for other ones.
3.2. SAXS
SAXS experiments have been performed to exploit the
different x-ray contrasts between lipid, saccharide and drug
components of the nanoparticles and to extend the SANS upper
Q-limit (0.1) to 0.4˚
A1. In this way one can also investigate
the internal structure of the lipid bilayers. In order to do this,
we write the bilayer form factor as
F(Q,d)=+d/2
d/2
dxρ ( x)cos(Qx)(7)
where d=2(th+tt)is the bilayer thickness and ρ(x)
is the local electronic density across the bilayer. The x-
dependent contrast ρ (x)is described by a simplified head–
tail model (see figure 4)[24], implying two uniform scattering
4
J. Phys.: Condens. Matter 20 (2008) 104211 Y Gerelli et al
Figure 4. Electron density profile for the adopted head–tail model
(——) and of the solvent (······). The head and tail thicknesses, th
and tt, and the total bilayer thickness, d, are also indicated.
length densities for the tail and for the head respectively. A
more detailed description of ρ ( x)is not justified since the
composition of the adopted commercial lipid (Lipoid S45) is
partially unknown (see section 2.1). The SAXS intensity can
therefore be expressed as
I(Q)=A
Q2F2(Q,d,ρ(x)). (8)
Fitting parameters are: the scaling factor Aand the thickness
thand tt(see figure 4).TheroutineusedfortheSAXS,SANS
and SLS data is based upon an least square minimization from
the MINUIT package [25].
4. Results and discussion
4.1. Nanoparticle structure
The cTEM micrographs (panel (a)) in figure 1show the
presence of two clearly different nanoparticle populations.
Smaller nanoparticles (ULV, typical size about 20 nm) have a
vesicular structure with a large hollow core and a relatively thin
outer shell attributable to a single bilayer structure. From the
micrographs one can infer a relatively large size polydispersity.
Larger nanoparticles (MLV, typical size above 80 nm) have
a clear multi-layer structure. They are as well polydisperse
both in size and in number of layers. For MLV particles
we have adopted the multi-layer model described previously
and depicted in figure 3. Electronic contrast between water
and saccharides is ten times bigger than that between water
and lipids; cTEM can therefore provide information on the
arrangement of chitosan. The cTEM micrographs suggest that
chitosan is almost absent from the nanoparticles. This can be
understood in terms of the nanocapsule preparation method:
due to hydrophobic attraction, the lipophilic cationic TAM
and the amphiphilic anionic lecithin, already admixed into the
methanol solution, form mixed aggregates upon injection into
the water/chitosan solution. The resulting positive charge of
the resulting nanoparticles (see Z-potential values in table 1)
inhibits surface adhesion of the positively charged chitosan
molecules. Because of the acidity of the final solution (pH =
2.7), chitosan molecules in solution do not aggregate and keep
elongated chain shapes [26].
Figure 5. SANS data: () nanoparticles in deuterated solvent,
() nanoparticles in lipoid-matched solvent, SLS () data. The
continuous lines are the fits on both data sets to equation (9). The
contributionsofULV(----)andMLV(
······) populations are also
shown for nanoparticles in deuterated solvent.
Figure 5shows the two SANS curves in absolute units for
nanoparticle suspensions with concentration c=0.2% (w/w)
in both perdeuterated and lipoid-matched solvents. In the same
figure the SLS data for a similar sample with c=0.04% are
reported after re-scaling to the SANS data. In the overlap
region the agreement between the two curves is remarkable.
It has also to be noted that a very small amount of much larger
particles is likely to be present in the solution, as suggested by
the discrepancy at very low Qbetween experimental SLS data
and the fit. Their presence cannot be appreciated by SANS,
restricted to higher Q, and they have therefore been neglected
in the adopted model.
The two curves shown in figure 5(SANS+SLS data for
the sample in deuterated solvent and SANS data for the one in a
solvent at the lipoid matching point) have been simultaneously
fitted adopting the two-family model described in section 3.1.1
(see equation (3)). Preliminary fits, performed varying all the
free parameters of the model, lead to the conclusion that the
scattering contrasts ρ0and ρ m
0are always very close to
zero (<1014 ˚
A2). This indicates that the core regions of
both families and the inter-layer region of MLV nanoparticles
are occupied by the solvent; these two contrasts were then set
to zero in the subsequent analysis. Under this zero-contrast
hypothesis the simultaneous fit of the data in figure 5implies a
system of two equations which differ only for the SLDs of the
two solvents:
d(Q)
d=16π2φ
Q2[pu
Bρs)2fu(Q)
+(1p)(ρm
Bρs)2fm(Q)]+bkg
d(Q)
d=16π2φ
Q2[pu
Bρs)2fu(Q)
+(1p)(ρm
Bρs)2fm(Q)]+bkg
(9)
where, apart from the contrast factors, fu(Q)and fm(Q)
represent the form factors of the ULV and MLV families,
5
J. Phys.: Condens. Matter 20 (2008) 104211 Y Gerelli et al
respectively,
fu(Q)=1
Vu
0
dRD
uu,Ru
0)R
0
drr sin(Qr)
2
fm(Q)=1
Vm
0dND
mm,N)
0dtDc,tc)
N1
j=0Rm
0+j(t+tm
B)+tm
B
Rm
0+j(t+tm
B)
drr sin(Qr)
2.
It has to be noted that equation (9) contains the product
of the scattering contrast and the particle number density
for both ULVs and MLVs. Nevertheless, since we work in
absolute units and the SLD of both solventsare known, we can
unambiguously determine the values of pand of the contrasts
ρ m
Band ρ u
B. From the cTEM micrographs the thickness
parameters, tb,tm
band tccould be estimated; we could
therefore put narrow limits to their range of variation, namely
from 35 to 60 ˚
A. cTEM provided also an initial guess for the
relative population number, pand for the two polydispersity
indices σmand σu. The curves resulting from the overall fit
are shown as continuous lines in figure 5together with the
contributions of the ULV and MLV families. The results are
summarized in table 2: they confirm that both ULV and MLV
families are highly polydisperse. The values obtained for the
lipid bilayer thicknesses tu
Band tm
B(50 ˚
A) are in agreement
with those typical for extruded pure lipid vesicles [24,27,28].
The SLD values indicate that in ULV particles the lipid bilayer
is mostly composed by Lipoid (ρlipoid ρs=(0.59 ±0.06)×
106˚
A2as measured by index matching, ρtam =1.5×
106˚
A2calculated),the amount of drug is zero within the
experimental accuracy (0.08) in this case. On the other hand,
from the SLD values obtained for MLV bilayers, it is possible
to conclude that the drug/lipid ratio (w/w) is 0.20 ±0.09.
Overall, MLV nanoparticles are characterized by an average
bilayer number N=5 with individualthickness tm
bseparated
by inter-layer regions, with thickness tc, that are occupied by
the solvent. This repetition leads to the broad peak observed in
figure 5at Q7×102˚
A1, its broadening being mainly due
to the width of the Dc,tc)distribution. Indeed the average
thickness of the multi-layer repeating unit ttot =tc+tm
b=
90 ±24 ˚
A corresponds to Q=2π/ttot =(7±2)×102˚
A1.
The average bilayer number N, together with Rm
0,tm
b,and
tcdetermine the mean external nanoparticle radius Rtot
according to
Rtot=Rm
0+tm
B+[N−1] tc+tm
B=440 ˚
A.
The radius polydispersity index σtot =48% is determined by
the distribution in the number of bilayers N.
4.2. Drug encapsulation
On the basis of the obtained SLD values, we can confirm that
TAM is located within the lipid bilayers; this is not surprising
if we consider the lipophilic nature of this drug. As explained
before, TAM is mostly present into the MLV nanoparticles,
with a 0.2–1 TAM to lecithin ratio (w/w). This ratio can
Figure 6. SANS curves for nanoparticles in deuterated ()and
lecithin-matched solvents (). In both data sets a flat incoherent
background has been subtracted. The latter data are also shown after
re-scaling by a factor [ρ m
B(deuterated)/ρm
B(matched)]2=1500
().
Table 2. Fitting parameters obtained by equation (3) applied to
SANS and SLS data.
ULV MLV
Core radius R0(˚
A) 40 ±220±7
Polydispersity σ(%) 50 —
Inter-layer thickness tc(˚
A) 37 ±1
Polydispersity σ(%) 23
Bilayer thickness tB(˚
A) 50 ±553±2
Bilayer SLD ρB(×106˚
A2)0.59 ±0.05 0.78 ±0.08
Population in % 38 62
be compared to the TAM/LEC one in the starting preparation
(see section 2.1): 0.34–1. Taking into account the volume
fraction of MLV nanoparticles, the effective loading efficiency
(encapsulated drug/drug present in the starting solution) is
(60 ±10)%. This is in agreement with the loading capacity
derived from the measured concentration of non-encapsulated
drug extracted from the final solution by ultracentrifugation
which is 60% [29]. This result is also confirmed by the
contrast values resulting from the fit of the data described in
section 4.1. After the flat incoherent background subtraction,
the data at the lipoid matching point, are significantly different
from zero only below Q0.025 ˚
A1.Therethe
scattering signal is dominated by the MLV contribution; in
fact ρ m
B(matched)0.16 and ρ u
B(matched)0.040
(see table 2); the scattered intensity is proportional to (ρ)2
therefore the scattering signal arising from ULVs is negligible.
The ratio between the scattering signals from the sample in
deuterated solvent and from the one at the lipoid matching
point is then: [ρ m
B(deuterated)/ρm
B(matched)]2=1500.
This result is shown in figure 6which shows a good
superposition between data in deuterated solvent and the ones
in the lipoid-matched solvent, once they are re-scaled by the
above ratio. We have to remark that this comparison can be
performed only below Q=0.025 ˚
A1i.e. in the region where
the signal from the lecithin-matched samples, after subtraction
of the flat incoherent background, is still appreciably = 0.
6
J. Phys.: Condens. Matter 20 (2008) 104211 Y Gerelli et al
Figure 7. SAXS () data together with the fit according to
equation (8)(——). The position expected for the low intensity peak
due to the multi-lamellar periodicity is indicated.
Table 3. Fitting parameters obtained by equation (7)appliedtoID02
data.
Zone
SLD
(×106˚
A2)
Thickness
(˚
A)
Head 11 ±111±2a
Tail 9 ±113.0±0.5a
aTotal bilayer thickness
d=2×(th+tt)=48 ±5˚
A.
4.3. Bilayer thickness
The SAXS analysis, performed in the range 8 ×103–4 ×
101˚
A1, enabled us to investigate in more detail the structure
of the lipid bilayer (see figure 7). In comparing the SANS
and SAXS data one has to note that the low intensity peak at
Q0.07 ˚
A1observed in the SANS profile and due to the
multi-lamellar periodicity is almost not visible in the SAXS
curve. This fact is just accidental: the peak occurs close to
the first minimum of the bilayer form factor in the x-ray data
(Q0.05 ˚
A1), while in the neutron case this minimum is
expected at higher Q-values outside the range we investigated.
Owing to the partially unspecified composition of Lipoid
S45, we have adopted a simplified description of the bilayer
form factor leading to equation (7). The parameter values
obtained from the fit are shown in table 3. The total bilayer
thickness, d48 ˚
A agrees well with the one obtained from
the SANS measurements and with literature values for the main
lipid components of our lecithin [24,27,28].
5. Conclusion
The adopted nanoparticle preparation methodology (self-
assembling following rapid injection of a lecithin+TAM
methanol solution into a diluted aqueous solution of chitosan)
gives rise to a polydisperse population characterized by two
distinct families: uni-lamellar vesicles (ULV) with average
radius 90 ˚
A and multi-lamellar vesicles (MLV) with average
radius 440 ˚
A. In both families the inner core is occupied by the
solvent. Tamoxifen is almost absent in ULV particles, while
it is present in the MLV ones which have a loading efficiency
around 60%. TAM molecules are located within each bilayer
as expected for lipophilic molecules.
The presence of tamoxifen changes structure of the
lipid outer shell from the uni-lamellar one observed in ULV
particles, to a multi-lamellar with an average number of
bilayers N=5 and a polydispersity index σm=48%.
The thickness of a single lipid bilayer is similar in ULVs and
in MLVs with a value close to that of pure DMPC [30], the
main lipid component of lecithin. It seems therefore not to be
affect by the presence of a charged drug and of a not negligible
amount of fatty acids.
In lipid-based nanoparticles, chitosan acts usually as
an outer coating that increases the vesicle stability. The
introduction of a charged drug like tamoxifen induces a
positive surface charge in the vesicles that repels the positively
charged chitosan molecules which therefore do not take
part in nanoparticle formation. Even in the absence of
chitosan, the loaded nanoparticles show a stability sufficient
for pharmaceutical applications [29]; this is probably due to
the multi-lamellar assembly induced by the charged drug.
The choice of a multi-component commercial lipid
mixture like Lipoid S45 was determined by the need of
studying systems close to the ones used in pharmacological
applications. The presence of 30% of fatty acids in Lipoid
S45 can in principle additionally complicate the interpretation;
we remark however that the matched solvent used in this
experiment refers precisely to Lipoid S45 therefore, as
concerns the scattering length densities, the presence of fatty
acids is to some extent taken into account. The present
analysis shows that careful morphological and structural
characterization by complementary techniques like cTEM,
SLS, SANS and SAXS can provided enough information to
describe the main structural features of the particles and the
effect of drug–particle interactions even when commercial
components are adopted. In the next future we plan to
extend this work towards nanoparticles in which commercial
lecithin is replaced by a better-controlled binary lipid mixture.
This should enable us to investigate in greater detail the
lipid–solvent–drug interactions, and to better tailor these self-
assembled systems for specific applications.
Acknowledgments
Many thanks to Mr S Barbieri for valuable help in sample
preparation and for helpful discussions. The authors are
grateful to Dr G¨oran Karlsson (University of Uppsala) for
the cryo-TEM measurements. We also thank Dr E DiCola
for her valuable technical assistance on the ID02 beamline
at ESRF. Financial support by the Ministry of University
and Research (PRIN founding framework) is gratefully
acknowledged. Y Gerelli also acknowledges NMI3 for
supporting his participation in ECNS 2007.
References
[1] Sonvico F, Dubernet C, Colombo P and Couvreur P 2005
Curr. Pharm. Des. 11 2091–105
[2] Hughes G A 2005 Nanomedicine 122–30
7
J. Phys.: Condens. Matter 20 (2008) 104211 Y Gerelli et al
[3] Torchilin V P 2005 Nat. Rev. Drug Discovery 4145–60
[4] Kim K Y 2007 Nanomedicine 3103–10
[5] Csaba N, Sanchez A and Alonso M J 2006 J. Control. Release
113 164–72
[6] Sonvico F, Cagnani A, Rossi A, Motta S, DiBari M T,
Cavatorta F, Alonso M J, Deriu A and Colombo P 2006 Int.
J. Pharm. 324 67–73
[7] Sonvico F, DiBari M T, Bove L, Deriu A, Cavatorta F and
Albanese G 2006 Physica B385 725–7
[8] Batzri S and Korn E 1973 Biochim. Biophys. Acta 298 1015–9
[9] Felt O, Buri P and Gurny R 1999 Drug Dev. Ind. Pharm. 24
979–93
[10] Illum L 1998 Pharm. Res. 15 1326
[11] Singla A W and Chawla M 2001 J. Pharm. Pharmacol.
53 1047
[12] Agnihotri S, Mallikarjuna N and Aminabhavi T 2004
J. Control. Release 100 5–28
[13] Janes K, Calvo P and Alonso M 2001 Adv. Drug Delivery Rev.
47 83–97
[14] Calvo P, Remunan-Lopez C, Vila-Jato J and Alonso M 1997
J. Appl. Polym. Sci. 63 125–32
[15] Reddy L H, Vivek K, Bakshi H and Murthy R S 2006 Pharm.
Dev. Technol. 11 167–77
[16] Bond`ı M L, Craparo E F, Giammona G, Cervello M,
Azzolina A, Diana P, Martorana A and Cirrincione G 2007
Drug Delivery 14 61–7
[17] Riske K E, Cho S H, Lee H B, Jeong S Y and Yuk S H 2003
J. Microencapsulation 20 489–96
[18] PubChem 2004 http://pubchem.ncbi.nlm.nih.gov/summary/
summary.cgi?sid=17389724&viewopt=Deposited
[19] Almgren M, Edwards K and Karlsson G 2000 Colloids Surf. A
174 3–21
[20] Sears V F 1992 Neutron News 326–37
[21] Lee J H, Agarwal V, Bose A, Payne G F and Raghavan S R
2006 Phys.Rev.Lett.96 048102–1
[22] Kiselev M A, Wartewig S, Janich M, Lesieur P, Kiselev A M,
Ollivon M and Neubert R 2003 Chem. Phys. Lipids
123 31–44
[23] Bartlett P and Ottewill R H 1992 J. Chem. Phys. 96 3306–18
[24] Riske K A, Amaral L Q and Lamy-Freund M T 2001 Biochim.
Biophys. Acta 1511 297–308
[25] Cern program library: http://cernlib.web.cern.ch/cernlib/
[26] Guo H F, Lin H L and Yu T L 2002 J. Macromol. Sci. A
39 837–52
[27] Kuˇcerka N, Pencer J, Sachs J N, Nagle J F and Katsaras J 2007
Langmuir 23 1292–9
[28] Ristori S, Oberdisse J, Grillo I, Donati A and Spalla O 2005
Biophys. J. 88 535–47
[29] Como C, Sonvico F, Rossi A, Corradini E and Colombo P 2006
AAPS J. S2 8W4163
[30] Zemlyanaya E V, Kiselev M A and Aswal V K 2004 Preprint
0411029v2 [physics.chemph]
8
... In addition, chitosan, a cationic polysaccharide of natural origin, has been recently indicated as one of the most promising excipients for veterinary applications (10)(11)(12). Chitosan presents a series of appealing properties as nasal and vaccine excipient, providing mucoadhesion, enhanced mucosal permeability, control of the antigen release and adjuvant effects (13,14). ...
Article
Full-text available
Nasal vaccination has been shown to provide optimal protection against respiratory pathogens. However, mucosal vaccination requires the implementation of specific immunization strategies to improve its effectiveness. Nanotechnology appears a key approach to improve the effectiveness of mucosal vaccines, since several nanomaterials provide mucoadhesion, enhance mucosal permeability, control antigen release and possess adjuvant properties. Mycoplasma hyopneumoniae is the main causative agent of enzootic pneumonia in pigs, a respiratory disease responsible for considerable economic losses in the pig farming worldwide. The present work developed, characterized, and tested in vivo an innovative dry powder nasal vaccine, obtained from the deposition on a solid carrier of an inactivated antigen and a chitosan-coated nanoemulsion, as an adjuvant. The nanoemulsion was obtained through a low-energy emulsification technique, a method that allowed to achieve nano droplets in the order of 200 nm. The oil phase selected was alpha-tocopherol, sunflower oil, and poly(ethylene glycol) hydroxystearate used as non-ionic tensioactive. The aqueous phase contained chitosan, which provides a positive charge to the emulsion, conferring mucoadhesive properties and favoring interactions with inactivated M. hyopneumoniae. Finally, the nanoemulsion was layered with a mild and scalable process onto a suitable solid carrier (i.e., lactose, mannitol, or calcium carbonate) to be transformed into a solid dosage form for administration as dry powder. In the experimental study, the nasal vaccine formulation with calcium carbonate was administered to piglets and compared to intramuscular administration of a commercial vaccine and of the dry powder without antigen, aimed at evaluating the ability of IN vaccination to elicit an in vivo local immune response and a systemic immune response. Intranasal vaccination was characterized by a significantly higher immune response in the nasal mucosa at 7 days post-vaccination, elicited comparable levels of Mycoplasma-specific IFN-γ secreting cells and comparable, if not higher, responsiveness of B cells expressing IgA and IgG in peripheral blood mononuclear cells, with those detected upon a conventional intramuscular immunization. In conclusion, this study illustrates a simple and effective strategy for the development of a dry powder vaccine formulation for nasal administration which could be used as alternative to current parenteral commercial vaccines.
... With increasing Z the fraction of unilamellar vesicles v u markedly decreases from 40 to less than 20%, but remains significant even at the highest Z value investigated. Nevertheless, combination of the increase of N and the decrease of v u results in a change of the average number of bilayers N with surfactant concentration that is more pronounced than just the change of N. Similarly low values of N were reported for chitosan/lipid multilayer vesicles [50,28,51]. The presence of additional unilamellar vesicles is needed to properly describe the scattering intensity close to the correlation peak (see Fig. S4 in the supporting information). ...
Article
Hypothesis Membrane undulations are known to strongly affect the stability of uni- and multilamellar vesicles formed by surfactants or phospholipids. Herein, based on the same arguments, we hypothesise that the properties of polyelectrolyte mediated surfactant multilamellar vesicles, in particular the multiplicity – i.e. the number of layers forming the vesicle – depend on the dynamics of the membrane. Experiments Small-angle neutron scattering (SANS) and neutron spin-echo (NSE) were used to probe the structure and the dynamics of the multilayered vesicles formed in mixtures of the biopolymer chitosan and oppositely charged alkyl ether carboxylates. The neutron scattering data are complemented by static and dynamic light scattering experiments. Experiments were performed in polyelectrolyte excess conditions, and at a pH close to the pKa of the surfactant. Findings The structural investigation shows very clearly that multilayered surfactant/polyelectrolyte vesicles are formed in the investigated mixtures. Only 3 to 5 layers form, on average, one vesicle, as similarly found in mixtures of chitosan and phospholipid vesicles. NSE shows that the surfactant membrane becomes stiffer upon complexation with chitosan, and that the fluctuation of the layers is strongly coupled in time and space. Such strong coupling and the increase in overall stiffness is associated with a high entropic cost. Accordingly, the combined SANS and NSE study points out that the low multiplicity found in multilayered vesicles involving the rigid polysaccharide chitosan arises from the strongly coupled dynamics of the membrane layers.
... The self-assembly method is based on the binding between negatively charged phospholipids and positively charged chitosan, resulting in the formation of nanostructures with good stability, proving their biodegradability and biocompatibility, very good mucosal adhesion, and insignificant cytotoxicity. This delivery system improved their access to drugs and extended their release times into the body [14,15]. Several recent studies have investigated the efficiency of this novel nanocarrier in delivering different drugs via various routes, such as oral and transmucosal routes of administration [16][17][18][19]. ...
Article
A novel, safe and efficient method was developed to encapsulate a blend of essential oils (EOs) into biodegradable nanoparticles (NPs). The biodegradable and biocompatible nanoparticles were made from chitosan (CH) and lecithin (LE) . The quality of the essential oils was verified using gas chromatography/mass spectrometry (GC/MS). The synthesis of nanoparticles included emulsification, followed by sonication, homogenization, and extrusion. Transmission electron microscopy (TEM) indicated that the nanoparticles were spherical in shape with sizes ranging from 25 to 70 nm, while dynamic light scattering (DLS) showed high negative zeta potentials. The stability of the final formula was evaluated in gastric and intestinal fluids. The chitosan/lecithin encapsulated EOs exhibited promising antimicrobial activity against the multi-drug resistant bacteria Salmonella typhi.
... SVT-LCNs were formed by the electrostatic self-assembly of lecithin and chitosan and have a multilayered structure alternating chitosan aqueous-rich layers and phospholipid bilayers, as previously reported (Gerelli et al., 2008b;Gerelli et al., 2008a). In order to improve drug encapsulation efficiency, nanoparticles were produced with the addition of pharmaceutical grade oils to help drug encapsulation in the lipophilic domain of LCNs. ...
Article
Full-text available
Nasal delivery has been indicated as one of the most interesting alternative routes for the brain delivery of neuroprotective drugs. Nanocarriers have emerged as a promising strategy for the delivery of neurotherapeutics across the nasal epithelia. In this work, hybrid lecithin/chitosan nanoparticles (LCNs) were proposed as a drug delivery platform for the nasal administration of simvastatin (SVT) for the treatment of neuroinflammatory diseases. The impact of SVT nanoencapsulation on its transport across the nasal epithelium was investigated, as well as the efficacy of SVT-LCNs in suppressing cytokines release in a cellular model of neuroinflammation. Drug release studies were performed in simulated nasal fluids to investigate SVT release from the nanoparticles under conditions mimicking the physiological environment present in the nasal cavity. It was observed that interaction of nanoparticles with a simulated nasal mucus decreased nanoparticle drug release and/or slowed drug diffusion. On the other hand, it was demonstrated that two antibacterial enzymes commonly present in the nasal secretions, lysozyme and phospholipase A2, promoted drug release from the nanocarrier. Indeed, an enzyme-triggered drug release was observed even in the presence of mucus, with a 5-fold increase in drug release from LCNs. Moreover, chitosan-coated nanoparticles enhanced SVT permeation across a human cell model of the nasal epithelium (×11). The nanoformulation pharmacological activity was assessed using an accepted model of microglia, obtained by activating the human macrophage cell line THP-1 with the Escherichia coli–derived lipopolysaccharide (LPS) as the pro-inflammatory stimulus. SVT-LCNs were demonstrated to suppress the pro-inflammatory signaling more efficiently than the simple drug solution (−75% for IL-6 and −27% for TNF-α vs. −47% and −15% at 10 µM concentration for SVT-LCNs and SVT solution, respectively). Moreover, neither cellular toxicity nor pro-inflammatory responses were evidenced for the treatment with the blank nanoparticles even after 36 h of incubation, indicating a good biocompatibility of the nanomedicine components in vitro. Due to their biocompatibility and ability to promote drug release and absorption at the biointerface, hybrid LCNs appear to be an ideal carrier for achieving nose-to-brain delivery of poorly water-soluble drugs such as SVT.
... 43 SVT-loaded hybrid lecithin/ chitosan nanoparticles (SVT-LCNs) were obtained by a spontaneous self-assembly process, involving the electrostatic interaction of lecithin, a negative phospholipid, with chitosan, a positively charged polysaccharide. 44,45 These nanosystems combine the versatility of phospholipid-based nanocarriers with the penetration-enhancing properties of positively charged polysaccharide chitosan and have been demonstrated to be promptly biodegradable. 22,46 Lipid-core PCL nanocapsules were obtained using either a nonionic surfactant, i.e., polysorbate 80 (PCL_P80), 29 or a negatively charged polysaccharide-based surfactant, i.e., sodium caproyl hyaluronate. ...
Article
Full-text available
Nanoparticles are promising mediators to enable nasal systemic and brain delivery of active compounds. However, the possibility of reaching therapeutically relevant levels of exogenous molecules in the body is strongly reliant on the ability of the nanoparticles to overcome biological barriers. In this work, three paradigmatic nanoformulations vehiculating the poorly soluble model drug simvastatin were addressed: (i) hybrid lecithin/chitosan nanoparticles (LCNs), (ii) polymeric poly-ε-caprolactone nanocapsules stabilized with the nonionic surfactant polysorbate 80 (PCL_P80), and (iii) polymeric poly-ε-caprolactone nanocapsules stabilized with a polysaccharide-based surfactant, i.e., sodium caproyl hyaluronate (PCL_SCH). The three nanosystems were investigated for their physicochemical and structural properties and for their impact on the biopharmaceutical aspects critical for nasal and nose-to-brain delivery: biocompatibility, drug release, mucoadhesion, and permeation across the nasal mucosa. All three nanoformulations were highly reproducible, with small particle size (∼200 nm), narrow size distribution (polydispersity index (PI) < 0.2), and high drug encapsulation efficiency (>97%). Nanoparticle composition, surface charge, and internal structure (multilayered, core–shell or raspberry-like, as assessed by small-angle neutron scattering, SANS) were demonstrated to have an impact on both the drug-release profile and, strikingly, its behavior at the biological interface. The interaction with the mucus layer and the kinetics and extent of transport of the drug across the excised animal nasal epithelium were modulated by nanoparticle structure and surface. In fact, all of the produced nanoparticles improved simvastatin transport across the epithelial barrier of the nasal cavity as compared to a traditional formulation. Interestingly, however, the permeation enhancement was achieved via two distinct pathways: (a) enhanced mucoadhesion for hybrid LCN accompanied by fast mucosal permeation of the model drug, or (b) mucopenetration and an improved uptake and potential transport of whole PCL_P80 and PCL_SCH nanocapsules with delayed boost of permeation across the nasal mucosa. The correlation between nanoparticle structure and its biopharmaceutical properties appears to be a pivotal point for the development of novel platforms suitable for systemic and brain delivery of pharmaceutical compounds via intranasal administration.
... Going back to solution behaviour, the morphology of DPPC liposomes was first analysed by SANS (Supporting Information, Figure S11). Data fitting using the vesicular model described in [90] demonstrates the presence of both uni-and multi-lamellar vesicles with average radii of 11 and 42 nm, respectively. Further detail on the fitting procedure is given in Supporting Information. ...
Article
Hypotheses. Bile salts (BS) are biosurfactants released into the small intestine, which play key and contrasting roles in lipid digestion: they adsorb at interfaces and promote the adsorption of digestive enzymes onto fat droplets, while they also remove lipolysis products from that interface, solubilising them into mixed micelles. Small architectural variations on their chemical structure, specifically their bile acid moiety, are hypothesised to underlie these conflicting functionalities, which should be reflected in different aggregation and solubilisation behaviour. Experiments. The micellisation of two BS, sodium taurocholate (NaTC) and sodium taurodeoxycholate (NaTDC), which differ by one hydroxyl group on the bile acid moiety, was assessed by pyrene fluorescence spectroscopy, and the morphology of aggregates formed in the absence and presence of fatty acids (FA) and monoacylglycerols (MAG) – typical lipolysis products – was resolved by small-angle X-ray/neutron scattering (SAXS, SANS) and molecular dynamics simulations. The solubilisation by BS of triacylglycerol-incorporating liposomes – mimicking ingested lipids – was studied by neutron reflectometry and SANS. Findings. Our results demonstrate that BS micelles exhibit an ellipsoidal shape. NaTDC displays a lower critical micellar concentration and forms larger and more spherical aggregates than NaTC. Similar observations were made for BS micelles mixed with FA and MAG. Structural studies with liposomes show that the addition of BS induces their solubilisation into mixed micelles, with NaTDC displaying a higher solubilising capacity.
... In the last decades, remarkable progress has been witnessed in the research and development of biocompatible and biodegradable polymers for drug delivery and in particular for their use in micro/nanoparticle manufacturing [1]. Biodegradable polymers such as poly(glycolide) (PGA), poly(lactic acid) (PLA), poly(lactide-co-glycolide) (PLGA), poly(ε-caprolactone) (PCL), poly(3-hydroxybutyrate) (PHB), poly(2-hydroxybutyrate-co-3-hydroxyvalerate) (PHBV), gelatine, chitosan (CHI), and alginate (ALG) are among popular polymers being used for this purpose [2][3][4][5]. ...
Article
Full-text available
A.A.A.); isa_majid@usm.my (M.I.A.M.) Abstract: Polyhydroxyalkanoate (PHA) copolymers show a relatively higher in vivo degradation rate compared to other PHAs, thus, they receive a great deal of attention for a wide range of medical applications. Nanoparticles (NPs) loaded with poorly water-soluble anticancer drug docetaxel (DCX) were produced using poly(3-hydroxybutyrate-co-4-hydroxybutyrate), P(3HB-co-4HB), copolymers biosynthesised from Cupriavidus malaysiensis USMAA1020 isolated from the Malaysian environment. Three copolymers with different molar proportions of 4-hydroxybutirate (4HB) were used: 16% (PHB16), 30% (PHB30) and 70% (PHB70) 4HB-containing P(3HB-co-4HB). Blank and DCX-loaded nanoparticles were then characterized for their size and size distribution, surface charge, encapsulation efficiency and drug release. Preformulation studies showed that an optimised formulation could be achieved through the emulsification/solvent evaporation method using PHB70 with the addition of 1.0% PVA, as stabilizer and 0.03% VitE-TPGS, as surfactant. DCX-loaded PHB70 nanoparticles (DCX-PHB70) gave the desired particle size distribution in terms of average particle size around 150 nm and narrow particle size distribution (polydispersity index (PDI) below 0.100). The encapsulation efficiency result showed that at 30% w/w drug-to-polymer ratio: DCX-PHB16 NPs were able to encapsulate up to 42% of DCX; DCX-PHB30 NPs encapsulated up to 46% of DCX and DCX-PHB70 NPs encapsulated up to 50% of DCX within the nanoparticle system. Approximately 60% of DCX was released from the DCX-PHB70 NPs within 7 days for 5%, 10% and 20% of drug-to-polymer ratio while for the 30% and 40% drug-to-polymer ratios, an almost complete drug release (98%) after 7 days of incubation was observed.
Article
Rubropunctatin, a metabolite isolated from the fungi of the genus Monascus, is a natural lead compound applied for the suppression of tumors with good anti-cancer activity. However, its poor aqueous solubility has limited its further clinical development and utilization. Lecithin and chitosan are excellently biocompatible and biodegradable natural materials, which have been approved by the FDA as drug carrier. Here, we report for the first time the construction of a lecithin/chitosan nanoparticle drug carrier of the Monascus pigment rubropunctatin by electrostatic self-assembly between lecithin and chitosan. The nanoparticles are near-spherical with a size 110-120 nm. They are soluble in water and possess excellent homogenization capacity and dispersibility. Our in vitro drug release assay showed a sustained release of rubropunctatin. CCK-8 assays revealed that lecithin/chitosan nanoparticles loaded with rubropunctatin (RCP-NPs) had significantly enhanced cytotoxicity against mouse mammary cancer 4T1 cells. The flow cytometry results revealed that RCP-NPs significantly boosted cellular uptake and apoptosis. The tumor-bearing mice models we developed indicated that RCP-NPs effectively inhibited tumor growth. Our present findings suggest that lecithin/chitosan nanoparticle drug carriers improve the anti-tumor effect of the Monascus pigment rubropunctatin.
Chapter
The most useful properties of food, i.e. the ones that are detected through look, touch and taste, are a manifestation of the food’s structure. Studies about how this structure develops or can be manipulated during food production and processing are a vital part of research in food science. This book provides the status of research on food structure and how it develops through the interplay between processing routes and formulation elements. It covers food structure development across a range of food settings and consider how this alters in order to design food with specific functionalities and performance. Food structure has to be considered across a range of length scales and the book includes a section focusing on analytical and theoretical approaches that can be taken to analyse/characterise food structure from the nano- to the macro-scale. The book concludes by outlining the main challenges arising within the field and the opportunities that these create in terms of establishing or growing future research activities. Edited and written by world class contributors, this book brings the literature up-to-date by detailing how the technology and applications have moved on over the past 10 years. It serves as a reference for researchers in food science and chemistry, food processing and food texture and structure.
Article
The present study deals with the encapsulation and brain-specific targeting of Lycopene (LYC) as Polysorbate-80 (P-80) coated Phosphatidylserine-chitosan self-assembled nanoparticles (P-80-LYC-PSCNP), with an approach to reduce oxidative stress which is rationalized as a major cause of neurological disease. P-80-LYC-PSCNP was formulated by using the injection technique for self-assembly of Phosphatidylserine and chitosan, applying P-80 coating for surface modification and were assessed for their morphology and physicochemical attributes. The efficiency of LYC liberated from P-80-LYC-PSCNP was determined in Streptozotocin-induced Oxidative stress Model (SOSM); cognitively by behavioral despair-test and biochemically by enzymatic assays at 5 mg/kg LYC-equivalent dose. P-80-LYC-PSCNP were formulated as stable (ζ= 3.89 mV) spherical, regular (≈220 nm), and homogenous entities. It demonstrated a total 76.89% LYC in-vitro release during 28 hrs and shown efficacy for brain-specific delivery with a relative bioavailability of 20.4, Cmax 4.6 times, and 2.5 times swifter than normal LYC. After 28 days, it shortened immobility time by 72.57±8.12 secs (P<0.01) against STZ-control (134.58±13.85 sec, P<0.001). It also improved the antioxidant enzymatic functioning of CAT, SOD, and GPx by 7.90±0.45 Δ240/min/mgP (P<0.05), 32.50±1.70 U/mgP (P<0.01), and 59.40±3.70 U/mgP (P<0.01) respectively against STZ-control which showed 3.05±0.80 Δ240/min/mgP (P<0.05), 15.20±3.10 U/mgP (P<0.05), and 25.15±3.05 U/mgP at 5 mg/kg LYC-equivalent dose respectively. P-80-LYC-PSCNP was successfully formulated, analyzed and had delivered Lycopene (5 mg/kg) by trespassing the BBB with improved pharmacokinetics. It alleviated the experimental Streptozotocin-induced Oxidative stress in turn improving behavioral and cognitive anomalies, while enhancing the antioxidant-related activity of enzymes.
Article
Full-text available
Recently, we have started a systematic study of the structure and dynamics of nano- and microparticles of interest as highly biocompatible drug carriers. For these particles, that are composed of polymeric and lipid material, a detailed understanding of the particle–solvent interactions is of key importance in order to tailor their characteristics for delivering drugs with specific chemical properties. Here we report results of elastic neutron scattering (ENS) investigations on lecithin/chitosan nanoparticles. They were first prepared by autoassembling the two components in aqueous solution; the samples were then freeze-dried and re-hydrated in a D2O atmosphere. The experiments were performed in the temperature range of 20–50K using the backscattering spectrometer IN13 at ILL (Grenoble, France). The comparison of samples in the dry state with similar ones at an hydration level of about 0.3–0.4 (g D2O/g hydrated sample), indicates that the presence of an outer chitosan ‘‘coating’’ reduces the mean square fluctuations of the hydrogens in the lipid component, leading thus to a stiffer nanoparticle structure.
Article
Full-text available
Cryo-transmission electron microscopy, c-TEM, has during the last 10 years contributed significantly to the understanding of the numerous, and often complex, structures formed by amphiphilic molecules in dilute aqueous solutions. In particular, the method has evolved as an important tool for the investigation of liposomes. In this review, we discuss and show examples of how the technique has been utilised to gain new information on the form and structure of liposomes, as well as on the morphological changes taking place upon encapsulation of drugs or interactions with surfactants, DNA and other polyelectrolytes. Several examples where c-TEM has been successfully employed to visualise related amphiphilic structures, such as thread-like micelles and particles of reversed phases, are also presented. In addition, we discuss recent developments concerning sample preparation and interpretation of images, as well as possible artefacts and their origin.
Article
Chitosan has been investigated as an excipient in the pharmaceutical industry, to be used in direct tablet compression, as a tablet disintegrant, for the production of controlled release solid dosage forms or for the improvement of drug dissolution. Chitosan has, compared to traditional excipients, been shown to have superior characteristics and especially flexibility in its use. Furthermore, chitosan has been used for production of controlled release implant systems for delivery of hormones over extended periods of time. Lately, the transmucosal absorption promoting characteristics of chitosan has been exploited especially for nasal and oral delivery of polar drugs to include peptides and proteins and for vaccine delivery. These properties, together with the very safe toxicity profile, makes chitosan an exciting and promising excipient for the pharmaceutical industry for present and future applications.
Article
We have studied the freezing of a binary mixture of colloidal poly(methyl methacrylate) spheres of size ratio 0.31 and composition AB4 (here A refers to the larger spheres). When suspended in a suitable liquid these particles interact via a steeply repulsive (approximately hard sphere) potential. The structure of the colloidal crystals formed in this binary system has been established from a combination of small-angle neutron and light scattering measurements. We find that there is an almost complete size separation on freezing. The crystalline phase contains almost exclusively large spheres while the smaller spheres are excluded from the crystal into a coexisting binary fluid. This observation is in agreement with recent density functional calculations for the freezing of hard sphere mixtures.
Article
The application of thermal neutron scattering to the study of the structure and dynamics of condensed matter requires a knowledge of the scattering lengths and the corresponding scattering and absorption cross sections of the elements. Ln some cases, values for the individual isotopes are needed as well. This information is required to obtain an absolute normalization ofthe scatteredneutron distributions, tocalculate unit-cell structure factors in neutron crystallography, and to correct for effects such as absorption, self-shielding, extinction, multiple scattering, incoherent scattering, and detector efficiency.
Article
Dilute solution properties of chitosan in propionic acid aqueous solutions with various pH values were studied using intrinsic viscosity, static light scattering (SLS), and dynamic light scattering (DLS). We demonstrated that the particle sizes (intrinsic viscosity [η], average hydrodynamic radius ⟨Rh⟩, and average radius of gyration ⟨RG⟩) of chitosan molecules in dilute propionic acid/water solutions increased with decreasing pH value. SLS data also demonstrated that second virial coefficient (A2) increased with decreasing pH value suggesting that solubility of chitosan in water increased with increasing propionic acid concentration. Differential refractive index increment (dn/dC), depolarization ratio (ρv), and ratio of ⟨RG⟩/⟨ Rh⟩ increased with decreasing pH value indicating the increment of chitosan molecular chain anisotropy with increasing propionic acid concentration. Increasing propionic acid concentration in chitosan aqueous solution caused an increase of −NH3 group on chitosan molecules leading to an increase in intra-molecular electrostatic charge repulsion and chain expansion of chitosan molecules. Thus, the molecular size, A2, ρν, dn/dC, and ⟨RG⟩/⟨Rh⟩ of chitosan increased with increasing propionic acid concentration.
Article
Hydrophilic nanoparticulate carriers have important potential applications for the administration of therapeutic molecules. The recently developed hydrophobic-hydrophilic carriers require the use of organic solvents for their preparation and have a limited protein-loading capacity. To address these limitations a new approach for the preparation of nanoparticles made solely of hydrophilic polymers is presented. The preparation technique, based on an ionic gelation process, is extremely mild and involves the mixture of two aqueous phases at room temperature. One phase contains the polysaccharide chitosan (CS) and a diblock copolymer of ethylene oxide and propylene oxide (PEO-PPO) and, the other, contains the polyanion sodium tripolyphosphate (TPP). Size (200–1000 nm) and zeta potential (between +20 mV and +60 mV) of nanoparticles can be conveniently modulated by varying the ratio CS/PEO-PPO. Furthermore, using bovine serum albumin (BSA) as a model protein it was shown that these new nanoparticles have a great protein loading capacity (entrapment efficiency up to 80% of the protein) and provide a continuous release of the entrapped protein for up to 1 week. © 1997 John Wiley & Sons, Inc.
Article
Single bilayer liposomes, indistinguishable from those obtained by sonication, can be prepared by injecting an thanolic solution of phospholipid into water. The dilute suspension is easily concentrated by ultrafiltration and the mild conditions allow neither degradation nor oxidation of the phospholipid.
Article
The aim of this review is to give an insight into the many potential applications of chitosan as a pharmaceutical drug carrier. The first part of this review concerns the principal uses of chitosan as an excipient in oral formulations (particularly as a direct tableting agent) and as a vehicle for parenteral drug delivery devices. The use of chitosan to manufacture sustained-release systems deliverable by other routes (nasal, ophthalmic, transdermal, and implantable devices) is discussed in the second part.
Article
Mucosal delivery of complex molecules such as peptides, proteins, oligonucleotides, and plasmids is one of the most intensively studied subjects. The use of colloidal carriers made of hydrophilic polysaccharides, i.e. chitosan, has arisen as a promising alternative for improving the transport of such macromolecules across biological surfaces. This article reviews the approaches which have aimed to associate macromolecules to chitosan in the form of colloidal structures and analyzes the evidence of their efficacy in improving the transport of the associated molecule through mucosae and epithelia. Chitosan has been shown to form colloidal particles and entrap macromolecules through a number of mechanisms, including ionic crosslinking, desolvation, or ionic complexation, though some of these systems have been realized only in conjunction with DNA molecules. An alternative involving the chemical modification of chitosan has also been useful for the association of macromolecules to self-assemblies and vesicles. To date, the in vivo efficacy of these chitosan-based colloidal carriers has been reported for two different applications: while DNA-chitosan hybrid nanospheres were found to be acceptable transfection carriers, ionically crosslinked chitosan nanoparticles appeared to be efficient vehicles for the transport of peptides across the nasal mucosa. The potential applications and future prospects of these new systems for mucosal delivery of macromolecules are highlighted at the end of the chapter.