ArticlePDF AvailableLiterature Review

Biodegradable polymeric nanocarriers for pulmonary drug delivery

Authors:

Abstract and Figures

Pulmonary drug delivery is attractive for both local and systemic drug delivery as a non-invasive route that provides a large surface area, thin epithelial barrier, high blood flow and the avoidance of first-pass metabolism. Nanoparticles can be designed to have several advantages for controlled and targeted drug delivery, including controlled deposition, sustained release, reduced dosing frequency, as well as an appropriate size for avoiding alveolar macrophage clearance or promoting transepithelial transport. This review focuses on the development and application of biodegradable polymers to nanocarrier-based strategies for the delivery of drugs, peptides, proteins, genes, siRNA and vaccines by the pulmonary route. The selection of natural or synthetic materials is important in designing particles or nanoparticle clusters with the desired characteristics, such as biocompatibility, size, charge, drug release and polymer degradation rate.
Content may be subject to copyright.
Review
10.1517/17425240802141568 © 2008 Informa UK Ltd ISSN 1742-5247 1
1
5
10
15
20
25
30
35
40
45
50
54
Biodegradablepolymeric
nanocarriersforpulmonary
drugdelivery
Erik Rytting, Juliane Nguyen, Xiaoying Wang & Thomas Kissel
Philipps-Universität Marburg, Institut für Pharmazeutische Technologie & Biopharmazie,
Ketzerbach 63, D-35032 Marburg, Germany
Background: Pulmonary drug delivery is attractive for both local and
systemic drug delivery as a non-invasive route that provides a large surface
area, thin epithelial barrier, high blood flow and the avoidance of first-pass
metabolism. Objective: Nanoparticles can be designed to have several
advantages for controlled and targeted drug delivery, including controlled
deposition, sustained release, reduced dosing frequency, as well as an
appropriate size for avoiding alveolar macrophage clearance or promoting
transepithelial transport. Methods: This review focuses on the development
and appli cation of biodegradable polymers to nanocarrier-based strategies
for the delivery of drugs, peptides, proteins, genes, siRNA and vaccines by
the pulmonary route. Results/conclusion: The selection of natural or
synthetic materials is important in designing particles or nanoparticle
clusters with the desired characteristics, such as biocompatibility, size,
charge, drug release and polymer degradation rate.
Keywords: biodegradable polymers, gene and siRNA delivery, nanoparticles, peptides,
proteins, pulmonary drug delivery
Expert Opin. Drug Deliv. (2008) 5(6):1-11
1. Introduction
Pulmonary drug delivery is attractive for several reasons. It is the obvious choice
for the local administration of drugs to treat disease locally within the lung, but
there are several advantages of employing the pulmonary route to achieve systemic
delivery of therapeutics.
Diseases that can be targeted with local pulmonary administration include
chronic obstructive pulmonary disease (COPD), asthma, cystic fibrosis, infectious
diseases, tuberculosis and lung cancer [1]. Delivering drugs via the lungs
also provides a non-invasive route of delivery for targeting the systemic circulation,
as the lungs provide a large surface area, a thin epithelial barrier, high blood
flow, and less enzymatic activity compared to other areas in the body [2-4].
First-pass metabolism can be avoided by pulmonary administration, which can be
especially useful for biopharmaceuticals, which are often extensively degraded
following oral delivery [4,5].
1.1 Pulmonaryphysiology
After inhalation through the nose or mouth, air first enters the trachea, which
divides into two main bronchi. Bifurcations continue through 23 stages before
the alveolar sacs are reached. Several changes in the cellular milieu occur when
moving from the bronchial regions to the alveolar regions deep in the lung,
including differences in epithelial cell type, airway thickness and lining fluid [6].
The airway epithelium contains ciliated cells, secretory cells and basal cells; the
alveolar epithelium is mainly comprised of Type I and Type II alveolar cells.
1. Introduction
2. Biodegradable
nanocarrier materials
3. Pulmonary delivery of
active agents
4. Conclusions
5. Expert opinion
Biodegradablepolymericnanocarriersforpulmonarydrugdelivery
2 ExpertOpin.DrugDeliv.(2008) 5(6)
110
115
120
125
130
135
140
145
150
155
160
164
55
60
65
70
75
80
85
90
95
100
105
109
Narrower airways result in narrower diffusion distances. The
total distance spanning the air–blood barrier decreases from
about 30 – 40 μm in the bronchial region to approximately
500 nm in the alveolar region [4,6]. The mucus layer lining
the upper airways is about 5 – 10 μm thick, whereas the
lung surfactant lining fluid secreted by the Type II alveolar
cells has a layer thickness of only 50 – 80 nm [6].
Barriers to particle deposition include the physical
defense of the oropharyngeal region and the bronchial tree.
Other barriers to particle and drug transport in the airways
include the mucus layer, the alveolar lining fluid, epithelial
cells, basement membrane, the mucociliary escalator,
macrophage clearance and proteolytic degradation [4,5]. Tight
junctions in the distal airways are not as tight as the tight
junctions in the bronchial and alveolar epithelia, suggesting that
targeting these distal airways may result in higher bioavail-
abilities in the systemic circulation [7]. Alveolobronchial
clearance is slower in the peripheral regions, and macrophage
clearance is reported to be minimal for particles less
than 260 nm [3,5]. Metabolism in the lung occurs by
peptidases and Phase I enzymes; these metabolizing enzymes
are found mostly in Clara cells, alveolar Type II cells
and alveolar macrophages [3,6].
1.2 Nanocarriersforpulmonarydelivery
Nanoparticles have gained increasing attention for pulmonary
drug delivery, due to their advantages for targeted deposition,
bioadhesion, sustained release and reduced dosing frequency
for improving convenience to the patient [2]. Some incentives
for using nanoparticles for the controlled delivery of drugs,
peptides, proteins, genes, siRNA and vaccines in the lung
include having an appropriate size for avoiding alveolar
macrophage clearance and promoting transepithelial transport.
Nanocarriers used for pulmonary applications also include
liposomes, solid lipid nanoparticles and nanotubes, but since
this review is limited to polymeric nanocarriers, the reader is
referred to other excellent manuscripts of interest [8-10]. Lung
targeting following intravenous nanoparticle administration
also falls outside the scope of this review, but this has been
discussed previously [11]. Due to their small size, most
nanoparticles would be exhaled, but these multifunctional
particles can be delivered to the lung by nebulization or by
the incorporation of the nanoparticles into larger particles
with an appropriate aerodynamic diameter by flocculation [12],
spray drying, or other means.
Regional deposition of particles delivered to the lung
depends on several factors, including particle properties such
as aerodynamic diameter, charge, surface properties and
hygroscopicity, as well as temperature, breathing pattern and
the timing of the aerosol pulse injection within the breathing
cycle [4]. The aerodynamic diameter (dae) is a function
of size, shape and density. Porous particles will have a
smaller dae than their physical diameter would suggest, and
for non-spherical particles the dae is mostly dependent upon
the short axis and the magnitude of the aspect ratio [5].
Shape does not only affect deposition by its influence
on the dae, but other factors too, as particle fibers are not
as easily cleared by alveolar macrophages compared to
spherical particles [13].
The aerodynamic diameter (dae) has an important
influence on particle destination. The optimal size for
deposition in the deep lung for systemic delivery is
approximately 1 3 μm [5]. Particles larger than 5 – 10 μm
result in oropharyngeal deposition, and are more likely to
be swallowed than to reach the lung. Particles smaller
than 1 μm will likely be exhaled. For particles between
1 and 5 μm, the smaller particles generally reach the deeper
parts of the lung, and the larger particles land in the upper
airways [3]. Particles around 1 2 μm have a higher chance
of crossing the air–blood barrier, and particles smaller
than 150 nm encounter delayed lung clearance, increased
protein interactions and more transepithelial transport
compared to larger particles [13]. Particle size may also
affect particle degradation and drug release rates.
Surface charge is another important property to consider
in particle design. Low surface energy is needed to
avoid particle agglomeration [5,13]. Electrostatic interactions
are also possible between the alveolar wall and oppositely
charged particles, but this depends on hydrophobicity
and humidity [13].
2. Biodegradablenanocarriermaterials
The design and synthesis of biodegradable polymeric
materials that will provide the appropriate nanocarrier chara-
cteristics for temporal and spatial distribution of drug in
the lung has been pursued extensively. Nanocarrier targeting
to the lung tissue based on particle size and surface charge
is an important aspect for material selection and design,
but the release of the active agent will also depend on
its distribution in the nanocarriers and the degradation
rate of the polymer.
A number of synthetic and natural polymers have been
utilized in formulating biodegradable nanoparticles [14].
Synthetic polymers have the advantage of sustaining the
release of the encapsulated therapeutic agent over a period
of days to several weeks. Natural polymers have a compara-
tively short duration of drug release. Polymers used for the
formulation of nanoparticles include natural polymers such
as albumin, gelatin, alginate, collagen, cyclodextrin and chitosan;
synthetic polymers used for pulmonary applications include
poly(lactide-co-glycolide) (PLGA) copolymers, polyacrylates
and polyanhydrides. These polymers, together with novel
polymers representing modifications to PLGA, are outlined
below. The applications of these polymers to pulmonary
drug delivery are then described in the following section.
2.1 PLAandPLGA
Polylactides (PLA) and poly (D,L-lactide-co-glycolide)
(PLGA) have been the most extensively investigated for drug
Rytting,Nguyen,Wang&Kissel
ExpertOpin.DrugDeliv.(2008) 5(6) 3
165
170
175
180
185
190
195
200
205
210
215
219
220
225
230
235
240
245
250
255
260
265
270
274
delivery [14]. As polyesters in nature, these polymers undergo
hydrolysis upon implantation into the body, forming
biologically compatible moieties (lactic acid and glycolic
acid) that are removed from the body by the citric acid
cycle. The degradation products are formed slowly and do
not affect normal cell function. These safe, non-toxic
polymers are currently being widely used in drug delivery
systems and tissue engineering research.
More details regarding the biocompatibility and bio-
degradation of PLA and PLGA in drug delivery applications
are presented in a review by Anderson and Shive [15]. The
release of drug in the PLGA matrix is controlled by diffusion
of the drug through the matrix and by degradation of
the polymer [15]. PLGA’s degradation rate is affected by the
copolymer composition and the molecular weight, and the
release of drug can thus be varied from weeks to months [16].
Several methods such as solvent evaporation, nanoprecipitation
and multiple emulsions allow formulations of small particles
on an industrial scale [14,17].
2.2 PhysicalmodicationofPLGA
Modifications to the particle surface and size can be used to
avoid particle clearance, increase circulation time, improve
transport across biological barriers, or to prolong residence
time at the site of absorption. Prolonging drug presence can
be manipulated by using mucoadhesive materials, such as
the biodegradable polysaccharide chitosan. Yamamoto et al.
produced PLGA nanoparticles surface-modified with chitosan
and encapsulating the peptide elcatonin [18]. The surface-
modified nanoparticles delivered to the lungs of guinea pigs
resulted in prolonged effects compared to the unmodified nano-
particles, and the chitosan-modified nanoparticles were elimi-
nated more slowly than the unmodified version, suggesting that
nanoparticle retention was responsible for the sustained effects.
2.3 ChemicalmodicationofPLGA
Despite its many advantages, PLGA also has some inherent
shortcomings. The lack of hydrophilic and functional
groups leads to challenges regarding drug encapsulation
and stability during storage. Other challenges include
polyphasic release patterns, low encapsulation efficiency
and high burst release [19,20]. A major challenge of nano-
particle delivery to the lungs is formulation instability due
to particle–particle interactions.
A strategy to overcome some of the issues associated with
the use of PLGA for pulmonary delivery of nanocarriers was
to create a polymer that is more hydrophilic in nature than
PLGA and to introduce functional groups for improved drug–
polymer interactions within the nanoparticles. For example,
block copolymers of hydrophilic poly(ethylene glycol) (PEG)
with PLGA have been reported to show accelerated drug
release [21,22]. Star and comb-shaped PLA or PLGA could be
synthesized with multifunctional initiators, such as glycerol,
pentaerythritol, amino-propanediol, poly(vinyl alcohol) and
dextran [23-28]. A primary feature of these materials is that
they have high molecular weights but relatively short PLA or
PLGA chains, and more hydroxyl end groups, which leads to
increased hydrophilicity and faster degradation rates compared
to linear PLA or PLGA of similar molecular weight.
Polyelectrolytes with functional groups in the backbone,
such as amine and sulfonic acid groups, were introduced
into brush-like graft PLGA. These modifications affect the
colloidal stability of carrier systems by imparting positive or
negative surface charges and increasing protein or drug
loading of carriers by electrostatic interactions [29-32]. These
functional groups also accelerate the degradation rate by
enhancing the hydrophilic character of the polyester [29,31,32].
Compared with amphiphilic block copolymers, the
amphiphilic graft copolymers have multi-grafted hydrophobic/
hydrophilic branches along a hydrophilic/hydrophobic polymer
backbone. Therefore, the properties of nanoparticles can
be easily varied by simply adjusting the graft density and
side chain length of the branches.
2.3.1 PVA-PLGA
Dailey et al. reported the synthesis of a series of poly(vinyl
alcohol) (PVA)-based branched polyesters with PLGA side
chains (PVA-g-PLGA, Figure 1A) [33]. The PVA provides a
hydrophilic basis for the copolymer, while the degree of the
copolymer hydrophobicity could be varied according to
the length of the PLGA side chains grafted onto the PVA.
These copolymers exhibited a lower burst effect coupled
with a linear infusion-like release profile of proteins, which
could be controlled by the structure and molecular weight
of the copolymer. Also, in contrast to the bulk erosion
observed for PLGA, the PVA-g-PLGA copolymers exhibited
a surface erosion biodegradation mechanism [34,35]. Further
developments of this type of copolymer could satisfy the
requirements of different drugs and proteins delivered by
the pulmonary route.
PVAs have shown good protein compatibility, muco-
adhesive properties and better temperature stability during
bulk polymerization with lactide and glycolide. PVAs with
molecular weights less than 15,000 g/mol will be eliminated
from the body by renal excretion [36]. Biodegradation occurs
by surface erosion, and the biocompatibility is comparable
to that of linear PLGA [37,38].
Uncharged and charged PVA-PLGA having side chain
lengths higher than 10 have the potential for the formation
of microparticles and nanoparticles, and water soluble
polyesters with PLGA side chain lengths smaller than three
are capable of forming nanocomplexes with oppositely
charged proteins [15,27,35].
2.3.2 SB-PVA-PLGA
Varying amounts of sulfobutyl (SB) groups were attached to
the backbone to create SB-PVA-g-PLGA polymers with an
increasingly negative surface charge (Figure 1B) [39]. The
SB-PVA-g-PLGA polymer allows the preparation of
nanoparticles that exhibit a core-corona structure with the
Biodegradablepolymericnanocarriersforpulmonarydrugdelivery
4 ExpertOpin.DrugDeliv.(2008) 5(6)
282
285
288
275
281
negatively charged hydrophilic sulfobutyl groups oriented
towards the outer aqueous phase, providing for the rearrange-
ment of the hydrophilic backbone to the particle surface.
Jung et al. demonstrated the preparation of particles with
this polymer having diameters between 100 and 500 nm [40].
These colloidal carriers can be prepared without the use
of additional surfactants. This is of extreme importance for
pulmonary application, as the inhalation of synthetic surfac-
tants may disturb the surface tension of the pulmonary lining
fluid and lead to impaired lung function or inflammation.
2.3.3 Amine-modified PVA-PLGA
Different amino groups, such as dimethylaminopropylamine
(DMAPA), diethylaminoethylamine (DEAEA) and
PVA-g-PLGA
Sulfobutyl-PVA-g-PLGA
A.
B.
C.
Amine-PVA-PLGA R = C3H6, C2H4; R′′ = C2H5, CH3
P(VS-VA)-g-PLGA
D.
OH S (O
*
*
O
O
O
O
OH
O
O
O
O O
O-Na+
)(
x))(
y
)(
n
z
)
m
(
N
R′′
R′′
R
O OH (O
*
O
O
O
O
OH
O
O
O
*
O
NH
O
)((w)(
x)(
y
)(
n
)
z
)
m
*
OOH (O
*
O
O
O
O
OH
O
O
OO
)(
(x)(
y)
z
)
m
)(
n
*
OH O(O
*
O
O
O
O
OH
O
O
O
S OO
OH
)(
(x)(
y
)(
n
)
z
)
m
Figure1.ThechemicalstructureofaseriesofPVA-basedbranchedPLGAs.
Rytting,Nguyen,Wang&Kissel
ExpertOpin.DrugDeliv.(2008) 5(6) 5
289
295
300
305
310
315
320
325
330
335
340
343
344
350
355
360
365
370
375
380
385
390
395
398
diethylaminopropylamine (DEAPA) were attached to the
backbone to create polymers with increasingly positive
surface charges (Figure 1C) [41]. Wittmar et al. developed
this class of biodegradable amine-substituted PVA
polymers, onto which hydrophobic PLGA chains were
grafted. The amphiphilic properties of diethylaminopropyl
amine-poly(vinyl alcohol)-grafted-poly(lactide-co-glycolide)
(DEAPA-PVA-g-PLGA) make it a versatile polymer for
pulmonary drug delivery [29,42]. Using a modified solvent
displacement method, nanoparticles can be generated
from DEAPA-PVA-PLGA without high shear forces.
Furthermore, degradation times can be tailored by the
degree of amine substitution to range from a few days
to several weeks, which is a vital aspect for long-term
pulmonary application. In contrast, the long degradation
rates of commercially available PLGA are a critical factor
for its use in lung delivery [43].
Coating DEAPA-PVA-g-PLGA nanoparticles with carboxy-
methylcellulose can prevent aggregation of the particles
during nebulization; Dailey et al. reported the preparation
of such particles in the size range of 76 – 256 nm [42].
Such aggregation occurs with more hydrophobic polymers.
The degradation rate decreased with increasing amounts
of CMC. DEAPA-PVA-g-PLGA nanoparticles were also
shown to be taken up into alveolar epithelial cells (A549)
to a low extent [42].
2.3.4 P(VS-VA)-PLGA
To control precisely the number of functional groups,
P(VS-VA)-g-PLGA was recently developed (Figure 1D).
The polyelectrolyte backbones were obtained by the radical
copolymerization between vinyl acetate and vinyl sulfonic
acid sodium salt and the subsequent hydrolysis. The obtained
poly(vinyl sulfonic-co-vinyl alcohol) (P[VS-VA]) backbones
were grafted with PLGA by ring-opening melt polymerization
using SnOct2 as catalyst through the hydroxyl group [44].
It was demonstrated that the degree of sulfonic acid sub-
stitution and the side length of PLGA can be easily controlled
by the feed ratio. Surface characterization studies showed
that, as observed for SB-PVA-g-PLGA, nanoparticles prepared
from these polymers exhibited a core-corona structure
with the negatively charged, hydrophilic sulfonic groups
oriented towards the outer aqueous phase [45]. Nanoparticles
prepared from this novel polymer class were reported
to range in size from 120 151 nm [45].
2.4 Otherpolymers
Polyanhydrides and polyacrylates have also been recently
investigated for nanocarrier-based pulmonary drug delivery
applications. Fliegel et al. describe a novel class of
biodegradable poly(ether-anhydride) polymers designed
for pulmonary drug delivery [46]. These polymers are
composed of sebacic acid and poly(ethylene glycol) (PEG)
in various ratios. By the addition of 10% PEG, the fraction
of particles deposited in the lower stages of a model
lung could be increased, most likely due to minimized
aggregation from surface roughness. Zhang et al. have
used polybutylcyanoacrylate to prepare nanoparticles for
intratracheal delivery [47].
In summary, chemical modifications to PLGA add
several benefits in the selection of a suitable material for
nanocarriers in the lung. The introduction of positive or
negative charges can enhance the encapsulation efficiency
and release profile of oppositely-charged drugs, proteins, or
genetic material. Adjustments to the polymer structure can
alter the balance between hydrophobic and hydrophilic
groups, which can in turn affect drug loading, release,
nanoparticle orientation, particle size and surface charge.
Furthermore, certain functional groups will affect the
polymer degradation rate, which not only affects the release
of active agent but is also of concern when one considers
the possible accumulation of polymer within the lung
following repeated doses. The design of fast-degrading
polymers, such as DEAPA-PVA-g-PLGA, overcomes some
of the challenges associated with polymer accumulation
due to slower degrading PLGA particles.
3. Pulmonarydeliveryofactiveagents
The previous section introduced several polymers that have
been used and designed for pulmonary delivery of nano-
carriers; this section will present some of the specific appli-
cations of these materials in the delivery of active agents,
including drugs, peptides, proteins, DNA, siRNA and vaccines.
3.1 Drugdelivery
As mentioned above, PLGA exhibits a triphasic drug and
protein release kinetic with an initial burst effect, which
is governed by diffusion kinetics, followed by a lag
phase and a secondary burst phase [48,49]. By varying
the PLGA chain lengths, the proportion of lactic to
glycolic acid and the molecular weight, drug release profiles
can be influenced [15].
Dutt et al. encapsulated isoniazid and rifampicin into
PLGA microparticles and investigated their release profile
from different formulations, demonstrating that isoniazid
shows a sustained release of up to 3 days from porous
microparticles and of up to 6 days from non-porous
microparticles. By hardening the PLGA microparticles, a
sustained release carrier system of up to 7 weeks in vitro
and in vivo could be achieved. In a murine model
one dose of PLGA microparticles was able to clear bacteria
from the lungs and liver more effectively as compared to
a daily administration of free drug [50,51].
For the treatment of tuberculosis, recombinant Myco-
bacterium tuberculosis antigen 85B (Ag85B) was encapsulated
by spray-drying into PLGA-microspheres [52]. With a median
diameter of 3 – 4 μm, these microspheres were suitable
for targeting macrophages and for aerosol delivery to the
lung. PLGA-rAg85B microspheres were able to stimulate an
Biodegradablepolymericnanocarriersforpulmonarydrugdelivery
6 ExpertOpin.DrugDeliv.(2008) 5(6)
454
460
465
470
475
480
485
490
495
500
505
508
399
405
410
415
420
425
430
435
440
445
450
453
antigen response that was two orders of magnitude higher
than that observed with the pure rAG85B. In another study,
PLGA microspheres, prepared using emulsion/solvent
evaporation, were loaded with rifampicin and delivered
to guinea pigs, which were infected with Mycobacterium
tuberculosis [53]. Compared to nebulized rifampicin
suspension, the aerosolized rifampicin-loaded PLGA micro-
spheres were able to reduce most measures of tuberculosis
infection. Encapsulation of three anti-tubercular drugs
rifampicin, isoniazid, and pyrazinamide into PLGA
nanoparticles achieved sustained therapeutic drug levels for
6 – 8 days in the plasma, and for up to 11 days in the
lungs. The drug-loaded nanoparticles were prepared by the
multiple emulsion technique and nebulized after vacuum
drying. A significantly prolonged elimination half-life was
observed compared to the orally administered drug and
no tubercle bacilli could be detected in the lungs after
five doses of treatment [54].
A second widely used class of biodegradable polymers
for pulmonary delivery is chitosan and its derivatives.
The degradation of chitosan has already been tested in
several studies and it has been shown that chitosans are
depolymerized enzymatically by lysozyme, albeit with a very
slow rate [55-57]. Like lactoferrin or peroxide, lysozymes are
present within the lung mucus and lysozyme is the most
abundant antimicrobial polypeptide in respiratory tract
secretions [58,59]. Learoyd et al. investigated the influence of
chitosan molecular weight on the drug release of terbutaline
sulfate spray powders using low, medium and high molecular
weight chitosan. With increasing molecular weights, the
drug release profile changed from a burst release to a
sustained drug release profile over 2 – 4 h. The microparticles
generated displayed a median diameter of 1 2.5 μm
and were therefore suitable for inhalation [60].
Corrigan et al. investigated the influence of the preparation
media on the morphology and characteristics of chitosan
microparticles prepared by spray-drying. As the degree of
acetylation of chitosans affects its physicochemical properties
(i.e., viscosity, degradability and solubility), spray drying
was performed in hydrochloric acid or acetic acid. It was
observed that the presence of acetic acid leads to increased
acetylation of chitosan during spray-drying. When loading
chitosan microparticles with salbutamol by spray-drying,
a high respirable fraction was achieved when aerosolized
into a twin impinger. However, the burst release of the
drug in less than 5 min requires further optimization for
future pulmonary delivery [61].
3.2 Peptideandproteindelivery
With its large alveolar surface area, thin epithelial barrier
and low proteolytic activity compared to other administration
routes, the lung represents an attractive route for the delivery
of macromolecules, such as proteins. Due to their extreme
sensitivity, the design of sophisticated drug carriers is
required to overcome the many barriers of the lungs.
Amidi et al. generated insulin-loaded microparticles by
spray-drying using N-trimethyl chitosan [62]. In all formu-
lations the secondary and tertiary structure of insulin could
be preserved. Even after 1-year storage at 4°C, the particle
characteristics and insulin structure remained unchanged
and intact. Grenha et al. developed a microparticulate
carrier system for insulin-loaded chitosan nanoparticles.
Using mannitol and lactose as excipients, the insulin-
loaded chitosan nanoparticles were microencapsulated
by spray-drying, yielding particles with Ferret diameters of
2 4 μm. In vitro studies showed that approximately
75 – 80% of the encapsulated insulin could be released from
the nanoparticle-loaded microspheres within 15 min [63].
PLGA nanospheres coated with chitosan for pulmonary
delivery of the peptide elcatonin have been mentioned
previously as an example of the advantages of physical
modifications to PLGA nanocarriers [18].
Kawashima et al. dosed PLGA nanoparticles prepared
with insulin to guinea pig lungs and demonstrated a
significant reduction in blood glucose level, with a prolonged
effect over 48 h compared to insulin solution [64].
Insulin-loaded nanoparticles using a different polymer,
poly(butyl cyanoacrylate), delivered to the lungs of rats,
were shown by Zhang et al. to extend the duration of
a hypoglycemic effect over 20 h [47].
3.3 Gene,siRNAandvaccinedelivery
An ideal gene delivery system should show high transfection
levels, be non-toxic and biodegradable for long-term appli-
cation. Polyethylenimine (PEI) is one of the most effective
cationic compounds for plasmid delivery into mammalian
cells [65,66]. The cationic groups of carriers such as PEI
can form complexes with oppositely-charged genetic cargo.
However, the high toxicity and lack of biodegradability
of PEI limits its potential for pulmonary application.
To overcome these drawbacks of PEI and at the same
time maintain its high transfection efficiency, Thomas et al.
developed biodegradable PEIs composed of a linear 423 Da
PEI and a branched 1.8 kDa PEI [67]. These two low
molecular weight PEIs, which have been shown to be
less toxic than its high molecular weight counterpart [68],
are crosslinked with bi- and oligo-functional acrylates
to obtain biodegradable high molecular weight PEIs.
Creating a combinatorial library of vectors, it was
shown that the optimal vector in vivo was the mixed PEI
crosslinked with propylene glycol glycerolate diacrylate. It
combined the highest transfection efficiency (186 times
higher than the physical mixture of the parental PEIs)
with low toxicity, whereas the commercially available 22-kDa
PEI caused 50% mortality.
Another group of biodegradable polymers used for gene
delivery are chitosans, which are considered to be non-toxic.
One of their major drawbacks, however, is the modest
transfection efficiency in vivo and in vitro [69,70]. Further
investigations are necessary to clarify the mechanisms of
Rytting,Nguyen,Wang&Kissel
ExpertOpin.DrugDeliv.(2008) 5(6) 7
509
515
520
525
530
535
540
545
550
555
560
563
564
570
575
580
585
590
595
600
605
610
615
618
uptake and transfection efficiency. Köping-Höggard et al.
administered chitosan-pDNA polyplexes to the lungs of mice,
showing that the high molecular weight chitosan is well
tolerable to mice and is able to promote gene delivery into
the lungs. However, PEI remained a superior pDNA vector
for pulmonary gene delivery [71]. After further optimization,
a comparable luciferase gene expression to that of PEI was
achieved when administered to mouse lungs. The molecular
weight of chitosans in this study ranged from 1.2 – 10 kDa.
Physicochemical studies showed that low molecular weight
chitosans are able to release pDNA in the presence of the
model anion heparin, leading to a higher transfection
efficiency compared to the more stable high molecular
weight chitosan polyplexes [72].
Another approach used to improve the bioactivity of
chitosan oligomer polyplexes was to introduce a trisaccharide
branch that targets cell surface lectins [73]. Lectins
are reported to be exposed on airway epithelial cells
and have the ability to bind sugar residues [74,75]. The trans-
fection efficiency of the trisaccharide-substituted chitosans
was significantly higher in human liver hepatocytes (HepG2)
and a human bronchial epithelial cell line (16HBE14o-)
than unmodified chitosan and PEI 25 kDa. Luciferase
gene expression in mouse lungs was fourfold higher for
trisaccharide-substituted chitosans than unmodified chitosan
but unfortunately no comparison to PEI 25 kDa in vivo
was shown.
Chitosan has also been used for siRNA delivery to
human lung carcinoma cells (H1299). With a chitosan/
siRNA formulation containing sucrose as lyoprotectant, a
70% knock down was achieved [76]. Howard et al. prepared
polyplexes between siRNA and chitosan [77]. Effective
knockdown, both in vitro and in vivo, was observed. In
the H1299 human lung carcinoma cell line and in murine
peritoneal macrophages, knockdowns of 77.9 and 89.3%,
respectively, were achieved. With a 40% reduction in
EGFP fluorescence in bronchial epithelial cells of transgenic
EGFP mice, chitosan/siRNA polyplexes showed successful
RNA interference.
Chitosan has also been reported to display therapeutic
potential in the case of respiratory syncytial virus (RSV).
RSV causes bronchiolitis and pneumonia and is also a severe
risk factor for asthma. Treatment of rats prior to RSV
infection with chitosan/siRNA-polyplexes containing a
siRNA that interferes against the RSV-NS1 gene (siNSI),
reduced the virus titers in the lung. SiNS1-treated rats
showed less inflammation and hyperresponsiveness compared
to the control [78].
Pulmonary DNA vaccination represents a non-invasive
and less painful administration route for immunization. The
opportunity to combine the genetic information of various
antigen epitopes and cytokines, easy production and the
high stability of plasmid DNA compared to recombinant
proteins and pathogens make it an attractive class of vaccines.
Many pulmonary pathogens, such as M. tuberculosis bacillus,
respiratory syncytial virus (RSV) and severe acute respiratory
syndrome corona virus (SARS) could all be treated
once a suitable vaccination has been developed [79-81].
Bivas-Benita et al. prepared poly(D,L-lactide-co-glycolide)
(PLGA) nanoparticles and coated them with polyethylenimine
(PEI). These PLGA-PEI nanoparticles were loaded with
DNA and uptake into the endo-lysosomal compartment
of the human airway submucosal epithelial cell line,
Calu-3, was detected [82].
N-trimethyl chitosan and dextran microparticles were
investigated for pulmonary delivery of diphtheria toxoid.
The microparticles were prepared by drying an aqueous solution
of polymer and diphtheria toxoid using a supercritical fluid
(SCF) spraying process. In contrast to dextran microparticles,
only the N-trimethyl chitosan microparticles led to a detectable
secretion of IgA when administered to the lungs [82].
4. Conclusions
Pulmonary drug delivery is attractive for both local and
systemic drug delivery as a non-invasive route that provides
a large surface area, a thin epithelial barrier, high blood flow
and the avoidance of first-pass metabolism. Nanoparticles
have several advantages for controlled drug delivery by the
pulmonary route, including sustained release, reduced dosing
frequency, as well as being an appropriate size for avoiding
alveolar macrophage clearance or promoting transepithelial
transport. Particles or nanoparticle clusters with aerodynamic
diameters between 1 and 5 μm have the highest probability
of successful lung deposition. The selection of natural or
synthetic biodegradable polymeric materials for nanocarriers
is important in order to design particles with the desired
characteristics. Biocompatibility, size, charge and drug release
rates must all be considered, but in order to avoid accumu-
lation of polymeric materials following repeated dosing, the
polymer degradation rate is crucial.
5. Expertopinion
Although significant progress has been made in recent years
relating to the design of biodegradable nanocarrier strategies
for the delivery of drugs, peptides, proteins, genes, siRNA
and vaccines, the future of pulmonary delivery strategies is
expected to be influenced critically by the outcome of
ongoing discussions.
An issue that remains surrounded by considerable debate
is the question whether the lung should be used as an entry
port for systemic drug administration. In this context, the
safety of the nanocarriers and a lack of inflammatory and
immunogenic potential need to be demonstrated under
chronic treatment conditions. Such studies have not been
presented with drug-loaded nanocarriers, but will be
necessary during future clinical trials. Such issues are not
limited to pulmonary drug delivery, but are also important
in oral and intravenous administration.
Biodegradablepolymericnanocarriersforpulmonarydrugdelivery
8 ExpertOpin.DrugDeliv.(2008) 5(6)
647
650
655
660
665
670
674
619
625
630
635
640
646
A second area where a general lack of information can be
recognized is the interface between nano-objects and lung
tissue/cells. There are still many questions to be answered
with regard to the fate of nanocarriers in the lung. For
example, what physical and chemical characteristics affect
clearance by alveolar macrophages? Which particle properties
affect cellular internalization and transport across the
pulmonary epithelium? Furthermore, predictive correlations
between in vitro, ex vivo and in vivo models are necessary in
preparation for clinical trials of nanocarrier-based drug
delivery systems in the lung.
The third area where more fundamental information
needs to be generated is related to the aforementioned topics
but addresses the aspects of biomaterials used for pulmonary
delivery systems. The residence time of nanocarriers in lung
tissue, their degradation mechanisms and the clearance of
degradation products will ultimately affect the safety and
biocompatibility of such delivery devices. The accumulation
of carrier materials within the lung, including polymer and
its degradation products, may bring about long-term concerns
that outweigh the benefits of therapy with polymeric carriers.
More polymers with short half-lifes would clearly be
desirable, as this would be more suitable for repeated admini-
stration. A consensus about testing strategies, including both
in vitro and ex vivo models, has yet to be reached.
The fourth area where advances would be desirable relates
to the design of nanocarriers, especially the incorporation of
sensitive therapeutic agents, such as proteins, p-DNA and
siRNA, which require particle production methods avoiding
high shear stress. Important progress regarding particle
stabilization is necessary to prolong the shelf life of
nanoparticles intended for pulmonary delivery. Reconstitution
of dried nanosuspensions for nebulization or packaging of
nanoparticles into adequately-sized particle clusters for
inhalation of a dry powder are important steps to deliver the
desired doses to the desired regional targets within the lung.
As diffusion distances in nanocarriers are shorter by
definition, control over drug loading and release under
in vitro as well as in vivo conditions remains a challenge. In
addition to new technologies, polymer design may also help
to address such problems, as we have shown in some of the
examples presented above.
Pulmonary drug delivery is a fascinating area of research
which needs input from various disciplines ranging from
medical sciences to aerosol physics. As these interdisciplinary
research activities continue in the area of biodegradable
nanocarriers for pulmonary drug delivery, one can expect
significant advancements in the future that will extend hope
to healthcare professionals and patients alike.
Acknowledgements
The authors wish to thank the German Ministry for
Education and Research (BMBF) for a nanotechnology
science award 13N8889 and Boehringer Ingelheim
Pharma GmbH for technical support.
Bibliography
Papers of special note have been highlighted
as either of interest (•) or of considerable
interest (••) to readers.
1. Pison U, Welte T, Giersig M, Groneberg DA.
Nanomedicine for respiratory diseases.
Eur J Pharmacol 2006;533:341-50
2. Sung JC, Pulliam BL, Edwards DA.
Nanoparticles for drug delivery to the
lungs. Trends Biotechnol 2007;25:563-70
3. Sakagami M. In vivo, in vitro and ex vivo
models to assess pulmonary absorption
and disposition of inhaled therapeutics for
systemic delivery. Adv Drug Deliv Rev
2006;58:1030-60
4. Scheuch G, Kohlhaeufl MJ, Brand P,
Siekmeier R. Clinical perspectives
on pulmonary systemic and
macromolecular delivery. Adv Drug
Deliv Rev 2006;58:996-1008
•• Anexcellentresourceofpractical
informationpertainingtothelungs
asarouteofadministration.
5. Shoyele SA, Cawthorne S. Particle
engineering techniques for inhaled
biopharmaceuticals. Adv Drug Deliv Rev
2006;58:1009-29
• Anextensivesourceofinformation
relatingtothecellularcomponentsof
theair–bloodbarrier.
6. Steimer A, Haltner E, Lehr CM. Cell
culture models of the respiratory tract
relevant to pulmonary drug delivery.
J Aerosol Med 2005;18:137-82
7. Patton JS, Byron PR. Inhaling medicines:
delivering drugs to the body through the
lungs. Nat Rev Drug Discov 2007;6:67-74
• Manytermsareclearlydenedfor
thenoviceintheeld.
8. Vyas SP, Khatri K. Liposome-based
drug delivery to alveolar macrophages.
Expert Opin Drug Deliv 2007;4:95-9
9. Jaspart S, Bertholet P, Piel G, et al. Solid
lipid microparticles as a sustained release
system for pulmonary drug delivery.
Eur J Pharm Biopharm 2007;65:47-56
10. Xiang QY, Wang MT, Chen F, et al.
Lung-targeting delivery of dexamethasone
acetate loaded solid lipid nanoparticles.
Arch Pharm Res 2007;30:519-25
11. Azarmi S, Roa WH, Lobenberg R.
Targeted delivery of nanoparticles for the
treatment of lung diseases. Adv Drug
Deliv Rev 2008; In press
•• Includesadiscussionofpulmonary
targetingaftersystemicdelivery,
whichisbeyondthescopeofthis
currentreview.
12. Shi L, Plumley CJ, Berkland C.
Biodegradable nanoparticle flocculates
for dry powder aerosol formulation.
Langmuir 2007;23:10897-901
• Apromisingstrategyforpulmonary
nanoparticledelivery.
13. Chow AH, Tong HH, Chattopadhyay P,
Shekunov BY. Particle engineering for
pulmonary drug delivery. Pharm Res
2007;24:411-37
14. Panyam J, Labhasetwar V. Biodegradable
nanoparticles for drug and gene delivery
to cells and tissue. Adv Drug Deliv Rev
2003;55:329-47
15. Shive MS, Anderson JM. Biodegradation
and biocompatibility of PLA and PLGA
microspheres. Adv Drug Deliv Rev
1997;28:5-24
Rytting,Nguyen,Wang&Kissel
ExpertOpin.DrugDeliv.(2008) 5(6) 9
16. Lin SY, Chen KS, Teng HH, Li MJ.
In vitro degradation and dissolution
behaviours of microspheres prepared by
three low molecular weight polyesters.
J Microencapsul 2000;17:577-86
17. Berkland C, King M, Cox A, et al. Precise
control of PLG microsphere size provides
enhanced control of drug release rate.
J Control Release 2002;82:137-47
18. Yamamoto H, Kuno Y, Sugimoto S, et al.
Surface-modified PLGA nanosphere with
chitosan improved pulmonary delivery of
calcitonin by mucoadhesion and opening
of the intercellular tight junctions.
J Control Release 2005;102:373-81
19. Sanders LM, Kent JS, McRae GI,
et al. Controlled release of a luteinizing
hormone-releasing hormone analogue
from poly(d,l-lactide-co-glycolide)
microspheres. J Pharm Sci 1984;73:1294-7
20. Miyajima M, Koshika A, Okada J,
Ikeda M. Effect of polymer/basic
drug interactions on the two-stage
diffusion-controlled release from
a poly(L-lactic acid) matrix.
J Control Release 1999;61:295-304
21. Youxin L, Kissel T. Synthesis and properties
of biodegradable ABA triblock copolymers
consisting of poly(-lactic acid) or poly
(-lactic-co-glycolic acid) A-blocks attached
to central poly (oxyethylene) B-blocks.
J Control Release 1993;27:247-57
22. Kissel T, Li Y, Unger F. ABA-triblock
copolymers from biodegradable polyester
A-blocks and hydrophilic poly(ethylene
oxide) B-blocks as a candidate for
in situ forming hydrogel delivery systems
for proteins. Adv Drug Deliv Rev
2002;54:99-134
23. Arvanitoyannis I, Nakayama A,
Kawasaki N, Yamamoto N. Novel
star-shaped polylactide with glycerol using
stannous octoate or tetraphenyl tin as
catalyst. 1. Synthesis, characterization
and study of their biodegradability.
Polymer 1995;36:2947-56
24. Kim SH, Han YK, Ahn KD, et al.
Preparation of star-shaped polylactide
with pentaerythritol and stannous octoate.
Makromol Chem 1993;194:3229-36
25. Dong CM, Qiu KY, Gu ZW, Feng XD.
Synthesis of star-shaped poly (d, l-lactic
acid-alt-glycolic acid) with multifunctional
initiator and SnOct2 catalyst. Polymer
2001;42:6891-6
26. Arvanitoyannis I, Nakayama A,
Kawasaki N, Yamamoto N. Novel
polylactides with aminopropanediol or
aminohydroxymethylpropanediol using
stannous octoate as catalyst-synthesis,
characterization and study of their
biodegradability. 2. Polymer
1995;36:2271-9
27. Breitenbach A, Kissel T. Biodegradable
comb polyesters. Part 1. Synthesis,
characterization and structural
analysis of poly (lactide) and poly
(lactide-co-glycolide) grafted onto
water-soluble poly (vinyl alcohol)
as backbone. Polymer 1998;39:3261-71
28. Youxin L, Nothnagel J, Kissel T.
Biodegradable brush-like graft polymers
from poly (d,l-lactide) or poly
(d,l-lactide-co-glycolide) and
charge-modified, hydrophilic dextrans
as backbone – synthesis, characterization
and in vitro degradation properties.
Polymer 1997;38:6197-206
29. Oster CG, Wittmar M, Unger F, et al.
Design of amine-modified graft polyesters
for effective gene delivery using
DNA-loaded nanoparticles. Pharm Res
2004;21:927-31
30. Simon M, Wittmar M, Bakowsky U,
Kissel T. Self-assembling nanocomplexes
from insulin and water-soluble
branched polyesters, poly [(vinyl-3-
(diethylamino)-propylcarbamate-co-(vinyl
acetate)-co-(vinyl alcohol)]-graft-poly
(l-lactic acid): a novel carrier for
transmucosal delivery of peptides.
Bioconjug Chem 2004;15:841-9
31. Jung T, Kamm W, Breitenbach A,
et al. Loading of tetanus toxoid to
biodegradable nanoparticles from
branched poly (sulfobutyl-polyvinyl
alcohol)-g-(lactide-co-glycolide)
nanoparticles by protein adsorption:
a mechanistic study. Pharm Res
2002;19:1105-13
32. Li Z, Huang L. Sustained delivery
and expression of plasmid DNA
based on biodegradable polyester,
poly(,-lactide-co-4-hydroxy-proline).
J Control Release 2004;98:437-46
33. Dailey LA, Wittmar M, Kissel T. The
role of branched polyesters and their
modifications in the development
of modern drug delivery vehicles.
J Control Release 2005;101:137-49
•• Amorethoroughreviewofsome
ofthenovelpolymersdiscussedin
Section2ofthisarticle.
34. Breitenbach A, Pistel KF, Kissel T.
Biodegradable comb polyesters. Part II.
Erosion and release properties of poly(vinyl
alcohol)-g-poly(lactic-co-glycolic acid).
Polymer 2000;41:4781-92
35. Frauke Pistel K, Breitenbach A,
Zange-Volland R, Kissel T. Brush-like
branched biodegradable polyesters,
part III: Protein release from microspheres
of poly(vinyl alcohol)-graft-poly
(,-lactic-co-glycolic acid). J Control Release
2001;73:7-20
36. Yamaoka T, Tabata Y, Ikada Y. Comparison
of body distribution of poly(vinyl alcohol)
with other water-soluble polymers
after intravenous administration.
J Pharm Pharmacol 1995;47:479-86
37. Breitenbach A, Li YX, Kissel T. Branched
biodegradable polyesters for parenteral
drug delivery systems. J Control Release
2000;64:167-78
38. Li Y, Nothnagel J, Kissel T.
Biodegradable brush-like graft
polymers from poly(-lactide) or
poly(-lactide-co-glycolide) and
charge-modified, hydrophilic dextrans
as backbone – synthesis, characterization
and in vitro degradation properties.
Polymer 1997;38:6197-206
39. Li Y, Volland C, Kissel T.
Biodegradable brush-like graft
polymers from poly(D,L-lactide) or
poly(D,L-lactide-coglycolide) and
charge-modified, hydrophilic dextrans
as backbone – in vitro degradation
and controlled releases of hydrophilic
macromolecules. Polymer 1998;39:3087-97
40. Jung T, Breitenbach A, Kissel T.
Sulfobutylated poly(vinyl alcohol)-graft-
poly(lactide-co-glycolide)s facilitate
the preparation of small negatively
charged biodegradable nanospheres.
J Control Release 2000;67:157-69
41. Dailey LA, Jekel N, Fink L, et al.
Investigation of the proinflammatory
potential of biodegradable nanoparticle
drug delivery systems in the lung.
Toxicol Appl Pharmacol 2006;215:100-8
42. Dailey LA, Kleemann E, Wittmar M,
et al. Surfactant-free, biodegradable
nanoparticles for aerosol therapy
based on the branched polyesters,
DEAPA-PVAL-g-PLGA. Pharm Res
2003;20:2011-20
43. Shoyele SA, Slowey A. Prospects
of formulating proteins/peptides as
Biodegradablepolymericnanocarriersforpulmonarydrugdelivery
10 ExpertOpin.DrugDeliv.(2008) 5(6)
aerosols for pulmonary drug delivery.
Int J Pharm 2006;314:1-8
44. Kissel T, Wang XY. Graft copolymers
as drug delivery systems.
EP07116143; 2007
45. Wang X, Xie X, Cai C, et al.
Biodegradable branched polyesters
poly(vinyl sulfonate-co-vinyl alcohol)-
graft-P(D,L-lactic-co-glycolic acid) as
negatively-charged polyelectrolyte
platforms for drug delivery: synthesis
and characterization. Macromolecules
2008;41:2791-9
46. Fiegel J, Fu J, Hanes J.
Poly(ether-anhydride) dry powder aerosols
for sustained drug delivery in the lungs.
J Control Release 2004;96:411-23
47. Zhang Q, Shen Z, Nagai T. Prolonged
hypoglycemic effect of insulin-loaded
polybutylcyanoacrylate nanoparticles after
pulmonary administration to normal rats.
Int J Pharm 2001;218:75-80
48. Sanders LM, Kell BA, McRae GI,
Whitehead GW. Prolonged
controlled-release of nafarelin,
a luteinizing hormone-releasing
hormone analogue, from biodegradable
polymeric implants: influence of
composition and molecular weight
of polymer. J Pharm Sci 1986;75:356-60
49. Zolnik BS, Burgess DJ. Effect of
acidic pH on PLGA microsphere
degradation and release.
J Control Release 2007;122:338-44
50. Dutt M, Khuller GK. Chemotherapy of
Mycobacterium tuberculosis infections
in mice with a combination of isoniazid
and rifampicin entrapped in Poly
(DL-lactide-co-glycolide) microparticles.
J Antimicrob Chemother 2001;47:829-35
51. Dutt M, Khuller GK. Sustained release
of isoniazid from a single injectable
dose of poly (DL-lactide-co-glycolide)
microparticles as a therapeutic
approach towards tuberculosis. Int J
Antimicrob Agents 2001;17:115-22
52. Lu D, Garcia-Contreras L, Xu D, et al.
Poly (lactide-co-glycolide) microspheres in
respirable sizes enhance an in vitro T cell
response to recombinant Mycobacterium
tuberculosis antigen 85B. Pharm Res
2007;24:1834-43
53. Garcia-Contreras L, Sethuraman V,
Kazantseva M, et al. Evaluation of
dosing regimen of respirable rifampicin
biodegradable microspheres in the
treatment of tuberculosis in the
guinea pig. J Antimicrob Chemother
2006;58:980-6
54. Pandey R, Sharma A, Zahoor A, et al.
Poly (DL-lactide-co-glycolide)
nanoparticle-based inhalable sustained
drug delivery system for experimental
tuberculosis. J Antimicrob Chemother
2003;52:981-6
55. Lee KY, Ha WS, Park WH. Blood
compatibility and biodegradability of
partially N-acylated chitosan derivatives.
Biomaterials 1995;16:1211-6
56. Varum KM, Myhr MM, Hjerde RJ,
Smidsrod O. In vitro degradation
rates of partially N-acetylated chitosans
in human serum. Carbohydr Res
1997;299:99-101
57. Tomihata K, Ikada Y. In vitro and in vivo
degradation of films of chitin and its
deacetylated derivatives. Biomaterials
1997;18:567-75
58. Nicod LP. Lung defences: an overview.
Eur Respir Rev 2005;14:45-50
59. Skerrett SJ. Lysozyme in pulmonary host
defense: new tricks for an old dog. Am J
Respir Crit Care Med 2004;169:435-6
60. Learoyd TP, Burrows JL, French E,
Seville PC. Chitosan-based spray-dried
respirable powders for sustained delivery of
terbutaline sulfate. Eur J Pharm Biopharm
2008;68:224-34
61. Corrigan DO, Healy AM, Corrigan OI.
Preparation and release of salbutamol from
chitosan and chitosan co-spray dried
compacts and multiparticulates. Eur J
Pharm Biopharm 2006;62:295-305
62. Amidi M, Pellikaan HC, de Boer AH,
et al. Preparation and physicochemical
characterization of supercritically dried
insulin-loaded microparticles for
pulmonary delivery. Eur J Pharm Biopharm
2008;68:191-200
63. Grenha A, Seijo B, Remunan-Lopez C.
Microencapsulated chitosan nanoparticles
for lung protein delivery. Eur J Pharm Sci
2005;25:427-37
64. Kawashima Y, Yamamoto H, Takeuchi H,
et al. Pulmonary delivery of insulin
with nebulized DL-lactide/glycolide
copolymer (PLGA) nanospheres to prolong
hypoglycemic effect. J Control Release
1999;62:279-87
65. Lungwitz U, Breunig M, Blunk T,
Gopferich A. Polyethylenimine-based
non-viral gene delivery systems.
Eur J Pharm Biopharm 2005;60:247-66
66. Zou SM, Erbacher P, Remy JS, Behr JP.
Systemic linear polyethylenimine
(L-PEI)-mediated gene delivery in the
mouse. J Gene Med 2000;2:128-34
67. Thomas M, Lu JJ, Zhang C, et al.
Identification of novel superior
polycationic vectors for gene delivery
by high-throughput synthesis and
screening of a combinatorial library.
Pharm Res 2007;24:1564-71
68. Lv H, Zhang S, Wang B, et al. Toxicity
of cationic lipids and cationic polymers
in gene delivery. J Control Release
2006;114:100-9
• Addressesimportantaspectsregarding
biocompatibilityofcarriermaterials.
69. Germershaus O, Mao S, Sitterberg J,
et al. Gene delivery using
chitosan, trimethyl chitosan or
polyethylenglycol-graft-trimethyl
chitosan block copolymers: establishment
of structure-activity relationships in vitro.
J Control Release 2008;125:145-54
70. MacLaughlin FC, Mumper RJ, Wang J,
et al. Chitosan and depolymerized chitosan
oligomers as condensing carriers for
in vivo plasmid delivery. J Control Release
1998;56:259-72
71. Koping-Hoggard M, Tubulekas I, Guan H,
et al. Chitosan as a nonviral gene delivery
system. Structure–property relationships
and characteristics compared with
polyethylenimine in vitro and after lung
administration in vivo. Gene Ther
2001;8:1108-21
72. Koping-Hoggard M, Varum KM, Issa M,
et al. Improved chitosan-mediated gene
delivery based on easily dissociated chitosan
polyplexes of highly defined chitosan
oligomers. Gene Ther 2004;11:1441-52
73. Issa MM, Koping-Hoggard M,
Tommeraas K, et al. Targeted gene
delivery with trisaccharide-substituted
chitosan oligomers in vitro and after lung
administration in vivo. J Control Release
2006;115:103-12
74. Fajac I, Briand P, Monsigny M, Midoux P.
Sugar-mediated uptake of glycosylated
polylysines and gene transfer into normal
and cystic fibrosis airway epithelial cells.
Hum Gene Ther 1999;10:395-406
75. Weis WI, Drickamer K. Structural basis
of lectin–carbohydrate recognition.
Ann Rev Biochem 1996;65:441-73
76. Andersen MO, Howard KA, Paludan SR,
et al. Delivery of siRNA from lyophilized
Rytting,Nguyen,Wang&Kissel
ExpertOpin.DrugDeliv.(2008) 5(6) 11
polymeric surfaces. Biomaterials
2008;29:506-12
77. Howard KA, Rahbek UL, Liu X, et al.
RNA interference in vitro and in vivo
using a novel chitosan/siRNA nanoparticle
system. Mol Ther 2006;14:476-84
78. Kong X, Zhang W, Lockey RF, et al.
Respiratory syncytial virus infection in
Fischer 344 rats is attenuated by short
interfering RNA against the RSV-NS1
gene. Genet Vaccines Ther 2007;5:4
79. Palese P, Garcia-Sastre A. Influenza
vaccines: present and future.
J Clin Invest 2002;110:9-13
80. Orme IM, McMurray DN, Belisle JT.
Tuberculosis vaccine development:
recent progress. Trends Microbiol
2001;9:115-8
81. Bivas-Benita M, Ottenhoff THM,
Junginger HE, Borchard G. Pulmonary
DNA vaccination: Concepts, possibilities
and perspectives. J Control Release
2005;107:1-29
82. Bivas-Benita M, Romeijn S, Junginger HE,
Borchard G. PLGA-PEI nanoparticles for
gene delivery to pulmonary epithelium.
Eur J Pharm Biopharm 2004;58:1-6
Afliation
Erik Rytting1 PhD, Juliane Nguyen2,
Xiaoying Wang3 PhD & Thomas Kissel†4 PhD
Author for correspondence
1Postdoctoral Research Associate
Philipps-Universität Marburg,
Institut für Pharmazeutische
Technologie & Biopharmazie,
Ketzerbach 63, D-35032 Marburg, Germany
2PhD candidate
Philipps-Universität Marburg,
Institut für Pharmazeutische
Technologie & Biopharmazie,
Ketzerbach 63, D-35032 Marburg, Germany
3Postdoctoral Research Associate
Philipps-Universität Marburg,
Institut für Pharmazeutische
Technologie & Biopharmazie,
Ketzerbach 63, D-35032 Marburg, Germany
4Professor and Head
Philipps-Universität Marburg,
Institut für Pharmazeutische
Technologie & Biopharmazie,
Ketzerbach 63, D-35032 Marburg, Germany
Tel: +49 6421 2825881;
Fax: +49 6421 2827016;
E-mail: kissel@staff.uni-marburg.de
... Nanoparticles have various advantages for treatment of respiratory diseases such as prolonged drug release, cellspecific targeted drug delivery and degradation within an acceptable period of time. [30,92] Glucagon DPIs loaded with PLGA nanospheres (GLG/NS) for chronic glucagon replacement therapy in patients with total pancreatectomy was carried out by Onoue et al. [93] Following the intratracheal administration in rats, evaluation regarding the hyperglycemic effect of a respirable powder formulation (GLG/NS-RP) was carried out, along with an evaluation on the pharmacokinetic behavior of glucagon via EIA (enzyme immunoassay) kit. A biphasic pattern with an initial burst and slow diffusion was observed in the drug-release behavior of GLG/NS. ...
Article
Nanoparticles (NPs) have played a pivotal role in various biomedical applications, spanning from sensing to drug delivery, imaging and anti-viral therapy. The therapeutic utilisation of NPs in clinical trials was established in the early 1990s. Silver nanoparticles (AgNPs) possess anti-microbial, anti-cancer and anti-viral properties, which make them a possible anti-viral drug to combat the COVID-19 virus. Free radicals and reactive oxygen species are produced by AgNPs, which causes apoptosis induction and prevents viral contamination. The shape and size of AgNPs can influence their interactions and biological activities. Therefore, it is recommended that silver nanoparticles (AgNPs) be used as a valuable tool in the management of COVID-19 pandemic. These nanoparticles possess strong anti-microbial properties, allowing them to penetrate and destroy microbial cells. Additionally, the toxicity level of nanoparticles depends on the administered dose, and surface modifications are necessary to reduce toxicity, preventing direct interaction between metal surfaces and cells. By utilising silver nanoparticles, drugs can be targeted to specific areas in the body. For example, in the case of COVID-19, anti-viral drugs can be stimulated as nanoparticles in the lungs to accelerate disease recovery. Nanoparticle-based systems have the capability to transport drugs and treat specific body parts. This review offers an examination of silver nanoparticle-based drug delivery systems for combatting COVID-19, with the objective of boosting the bioavailability of existing medications, decreasing their toxicity and raising their efficiency.
Article
Idiopathic pulmonary fibrosis (IPF) is a chronic and progressive interstitial lung disease that has been well-reported in the medical literature. Its incidence has risen, particularly in light of the recent COVID-19 pandemic. Conventionally, IPF is treated with antifibrotic drugs—pirfenidone and nintedanib—along with other drugs for symptomatic treatments, including corticosteroids, immunosuppressants, and bronchodilators based on individual requirements. Several drugs and biologicals such as fluorofenidone, thymoquinone, amikacin, paclitaxel nifuroxazide, STAT3, and siRNA have recently been evaluated for IPF treatment that reduces collagen formation and cell proliferation in the lung. There has been a great deal of research into various treatment options for pulmonary fibrosis using advanced delivery systems such as liposomal-based nanocarriers, chitosan nanoparticles, PLGA nanoparticles, solid lipid nanocarriers, and other nanoformulations such as metal nanoparticles, nanocrystals, cubosomes, magnetic nanospheres, and polymeric micelles. Several clinical trials are also ongoing for advanced IPF treatments. This article elaborates on the pathophysiology of IPF, its risk factors, and different advanced drug delivery systems for treating IPF. Although extensive preclinical data is available for these delivery systems, the clinical performance and scale-up studies would decide their commercial translation.
Chapter
In recent years, there has been a lot of interest in using nanotechnology in medicine to diagnose and cure various infectious diseases. Although the precise mechanism of action is frequently challenged, many nanosized materials have been extensively explored for this purpose, using their innately unique features. Their functionality has also been enhanced by the addition of various coatings, such as those that increase chemical stability and stealth properties or contain targeting agents that lessen off-target effects while producing effective "smart" nanoplatforms for the early detection, treatment, and possibly resolution of diseases. This book aims to compile research and review articles that concentrate on the synthesis, characterization, and applications of nanomaterials, such as nanoparticles, with a focus on investigating novel ways to improve their properties and create new efficient diagnostic, therapeutic, or theranostic approaches with a clear understanding of how they work.
Article
Pulmonary drug delivery is an efficient way to deliver drugs directly to the site of action i.e., lungs or to the blood circulation with minimum systemic effects. Recent emergence of coronavirus disease-2019 (COVID-19) and expeditious development of nanoparticle-based vaccines have recently reignited considerable interest in designing inhalable nanoparticle-based drug delivery systems as next-generation respiratory therapeutics. In this review, dry powder inhaler (DPI) formulations based on polymeric nanoparticles have been reviewed. It comprehensively describes various biodegradable and biocompatible synthetic (e.g. polylactic-glycolic acid, polyethylene glycol, polyethylenimine, and polycaprolactone) and natural (e.g. alginate, chitosan, dextran, hyaluronic acid, and gelatin) polymers used in formulating nanoparticle-based dry powder inhalers. This review covers the most recent drugs encapsulated in synthetic-based and natural-based biocompatible nanocarriers used in DPIs, providing latest approaches for treating various respiratory and pulmonary disorders. DPIs comprised of biocompatible biodegradable polymeric nanocarriers exhibited favorable particle properties and aerodynamic properties. In addition, these polymeric nanocarriers are chemically versatile in being either synthetic or natural. This chemical versatility enables versatility in the various types of drugs that can be incorporated into DPIs provided that the needed particle properties and aerodynamic properties are maintained.
Patent
Full-text available
The present invention relates to graft copolymers of polyvinyl sulfonic-co-vinyl alcohol)-g-poly(lactide-co-glycolide) (P(VS-VA)-g-PLGA) with negatively charged electrolyte properties, their method of preparation and their use. These negatively charged graft copolymers are suitable for effective loading and sustained-release of especially positively charged drugs, proteins and peptides, and drug-loaded particles from these grafted co-polymers are especially useful as parenteral or mucosal drug delivery systems for pharmaceutical applications.
Article
Full-text available
Several nonviral vectors including linear polyethylenimine(L-PEI) confer a pronounced lung tropism to plasmid DNA when injected into the mouse tail vein in a nonionic solution. We have optimized this route by injecting 50 µg DNA with excess L-PEI (PEI nitrogen/DNA phosphate=10) in a large volume of 5% glucose (0.4 ml). In these conditions, 1–5% of lung cells were transfected (corresponding to 2 ng luciferase/mg protein), the other organs remaining essentially refractory to transfection (1–10 pg luciferase/mg protein).β-Galactosidase histochemistry confirmed alveolar cells, including pneumocytes, to be the main target, thus leading to the puzzling observation that the lung microvasculature must be permeable to cationic L-PEI/DNA particles of ca 60 nm. A smaller injected volume, premixing of the complexes with autologous mouse serum, as well as removal of excess free L-PEI, all severely decreased transgene expression in the lung. Arterial or portal vein delivery did not increase transgene expression in other organs. These observations suggest that effective lung transfection primarily depends on the injection conditions: the large nonionic glucose bolus prevents aggregation as well as mixing of the cationic complexes and excess free L-PEI with blood. This may favour vascular leakage in the region where the vasculature is dense and fragile, i.e. around the lung alveoli. Cationic particles can thus reach the epithelium from the basolateral side where their receptors (heparan sulphate proteoglycans) are abundant. Copyright
Article
Particulate nanocarriers have been praised for their advantageous drug delivery properties in the lung, such as avoidance of macrophage clearance mechanisms and long residence times. However, instilled non-biodegradable polystyrene narospheres with small diameters and thus large surface areas have been shown to induce pulmonary inflammation. This study examines the potential of biodegradable polymeric nanoparticles composed of poly(lactic-co-glycolic acid) (PLGA) and the novel PLGA derivative, diethylaminopropylamine polyvinyl alcohol-grafted-poly(lactic-co-glycolic acid) (DEAPA-PVAL-g-PLGA), to provoke inflammatory responses in the murine lung after intratracheal instillation. Lactate dehydrogenase (LDH) release, protein concentration, MIP-2 mRNA induction, and polymorphonucleocyte (PMN) recruitment in the bronchial alveolar lavage fluid (BALF) were used to evaluate an inflammatory response in Balb-C mice. Two sizes of polystyrene (PS) nanospheres (diameters: 75 nm and 220 mn) were included in the study for comparison. All nanoparticle suspensions were instilled at concentrations of 1 mu g/gl and 2.5 mu g/gl, representative of an estimated "therapeutic dose" and a concentrated "dose" of particles. In all experiments, the 75 nm PS particles exhibited elevated responses for the inflammatory markers investigated. In contrast, biodegradable particles of comparable hydrodynamic diameter showed a significantly lower inflammatory response. The most marked differences were observed in the extent of PMN recruitment. While the 75 mn and 220 nm PS nanospheres exhibited 41 and 74% PMN within the total BALF cell population after 24 h, respectively, PMN recruiting in lungs instilled with both types of biodegradable particles did not exceed values of the negative isotonic glucose control. In conclusion, evidence suggests that biodegradable polymeric nanoparticles designed for pulmonary drug delivery may not induce the same inflammatory response as non-biodegradable polystyrene particles of comparable size. (c) 2006 Elsevier Inc. All rights reserved.
Article
Brush-like branched polyesters, obtained by grafting poly(lactic-co-glycolic acid), PLGA, onto water-soluble poly(vinyl alcohol) (PVAL) backbones, were investigated regarding their utility for the microencapsulation of proteins. Poly(vinyl alcohol)-graft-poly(lactic-co-glycolic acid), PVAL-g-PLGA, offers additional degrees of freedom to manipulate properties such as e.g. molecular weight, glass transition temperature and hydrophilicity. PLGA chain length was varied at a constant molecular weight (Mw) of the PVAL backbone and secondly Mw of the PVAL backbone was varied keeping the PLGA chain lengths constant. The most striking feature of these polymers is their high Mw. Microencapsulation of hydrophilic macromolecules, such as bovine serum albumin, ovalbumin, cytochrome c and FITC-dextran using a w/o/w double emulsion technique was investigated. Surface morphology, particle size, encapsulation efficiencies and protein release profiles were characterized as well. Microencapsulation of model compounds was feasible at temperatures of 0–4°C with yields typically in the range of 60–85% and encapsulation efficiencies of 70–90%. Both, encapsulation efficiency and initial protein release (drug burst) were strongly affected by the glass transition temperature, Tg, of the polymer in contact with water, whereas the in vitro protein release profile depended on the PVAL-g-PLGA structure and composition. In contrast to PLGA, protein release patterns were mostly continuous with lower initial drug bursts. Shorter PLGA chains increased drug release in the erosion phase, whereas initial pore diffusion was affected by the Mw of PVAL backbone. Release profiles from 2 to 12 weeks could be attained by modification of composition and molecular weight of PVAL-g-PLGA and merit further investigations under in vivo conditions. The in vitro cytotoxicity of PVAL-g-PLGA is comparable to PLGA and therefore, this new class of biodegradable polyesters has considerable potential for parenteral drug delivery systems.
Article
Biodegradable, negatively charged, branched polyesters, namely, poly(vinyl sulfonate-covinyl alcohol)-graft-poly(d,l-lactic-coglycolic acid), abbreviated as P(VS-VA)-g-PLGA, were synthesized by ring-opening polymerization using sulfonate-modified poly(vinyl alcohol) backbones as a platform for cationic drug delivery systems. The structures of sulfonate-modified backbones and the graft polyesters were characterized by NMR, FT-IR, GPC-MALLS, DSC, and viscosity measurements. By controlling the degree of sulfonate substitution of the backbone and the feed ratio of the backbone to the monomer, graft polyesters with different degrees of sulfonate substitution and different branch lengths were obtained. The random copolymer structure of PLGA and the successful grafting of PLGA to the P(VS-VA) backbone were confirmed by COSY NMR experiments. In vitro degradation studies demonstrated that the increased sulfonate substitution leads to a faster degradation rate; half-lives as fast as 8 days were observed. Nanoparticles were prepared from these amphiphilic graft copolymers by a solvent displacement technique and were characterized by particle size, polydispersity, zeta potential, and SEM. These novel biodegradable polyesters are promising candidates as negatively charged polyelectrolyte platforms for cationic drug delivery systems.
Article
The ring-opening polymerization of L-lactide with stannous octoate was investigated in the presence of pentaerythritol. By this way it was possible to prepare higher molecular weight star-shaped poly(L-lactide)s compared with the linear ones obtained by stannous octoate only. The weight-average molecular weights of linear and star poly(L-lactide)s were measured by light scattering analysis from hexafluoro-2-propanol solution, and the respective Mark-Houwink equations were derived. The second virial coefficient and the intrinsic viscosity of the star polymers were lower than those of the linear ones, which confirms the star-shaped architecture.
Article
d,l-3-methylglycolide (MG) was successfully polymerized with multifunctional initiator (trimethylolpropane (TMP) or pentaerythritol (PTOL)) and stannous octoate (SnOct2) catalyst in bulk at 110°C. The effects of molar ratios of monomer to initiator, monomer to catalyst and monomer conversion on the molecular weight of polymer were studied. For the homopolymerization of MG with TMP initiator and SnOct2 catalyst, the molecular weight of polymer increases from 6840 to 35 010 with the molar ratio of monomer to initiator (45–450), and the molecular weight distribution is from 1.15 to 1.35. The results indicate that in the homopolymerization of MG, the molecular weight of polymer is proportional to the molar ratio of monomer to initiator and the monomer conversion. The molar ratio of monomer to catalyst has no influence on the molecular weight of polymer at least within the range of 500–4000. 1H NMR spectra of the resulting polymers obtained from the homopolymerization of MG show that the homopolymerization of MG with TMP or PTOL initiator and SnOct2 catalyst produced two types of three-arm or four-arm star-shaped polymers. The bulk ring-opening homopolymerization of MG proceeds through a “coordination–insertion” mechanism and follows the selective acyl–oxygen bond cleavage reaction. 13C NMR spectroscopy indicates that the obtained poly(d,l-lactic acid-co-glycolic acid) (50:50, in molar ratio; d,l-PLGA50) has an alternating structures of lactyl and glycolyl units.
Article
Two series of novel polymers were synthesized by reacting L-lactide (LLA) with aminohydroxymethylpropanediol (AHPD) or aminopropanediol (APD) using stannous octoate as catalyst. Nuclear magnetic resonance, Fourier-transform infra-red spectroscopy, gel permeation chromatography (g.p.c.), differential scanning calorimetry (d.s.c.) and thermogravimetric analysis were employed for the characterization of these polymers. The above-mentioned polymers were semicrystalline (d.s.c. traces) when the molar ratio LLA/AHPD or LLA/APD was higher than 32/1. However, at molar ratios LLA/AHPD or LLA/APD below 32/1, the synthesized polymers were amorphous. D.s.c. and g.p.c. measurements show that, the higher the LLA content in the polymer, the higher the glass transition temperature Tg, melting temperature Tm, enthalpy of melting ΔHm and number-average molecular weight Mn. The potential biodegradability of these polymers was tested with the aid of enzymatic and alkali hydrolysis. Total organic carbon and g.p.c. measurements showed that high AHPD or APD contents impart to polymers a higher susceptibility to degradation.
Article
Parenteral delivery systems for proteins and peptides based on aliphatic polyesters are currently the subject of intensive research efforts. Linear polyesters of lactic acid and glycolic acid present significant problems with respect to the modulation of release properties of higher molecular weight substances, such as proteins. The molecular architecture of poly(lactic acid) (PLA) and its copolymers with glycolic acid (PLG), can be modified by the introduction of a hydrophilic charge-containing backbone, e.g. dextran sulfate sodium (DSS) or diethylaminoethyl dextran chloride (DEAED). The resulting branched molecular structure should offer additional possibilities to manipulate the degradation and release properties of parenteral delivery systems. The degradation of the graft PLG is accelerated significantly by the nonlinear structure, which contains many short biodegradable branches attached to a hydrophilic backbone molecule. The mechanism of the polymer degradation is influenced by the backbone substances. In the case of DEAED as backbone, the predominant chain scission of the graft polyester occurs in a random hydrolytic ester cleavage, similar to PLG. By contrast, a nonrandom chain scission in the vicinity of the branching points of the backbone was found for DSS-PLG. The erosion of the graft PLG proceeds more rapidly in the centre of the devices than at the surface. In contrast to linear PLG, the release of FITC-dextran and bovine serum albumin (BSA) from the microspheres prepared from the graft PLG is continuous. Graft PLGs offer additional possibilities for adjusting the release of proteins and peptides from biodegradable parenteral delivery systems.