SlideShare a Scribd company logo
1 of 21
Download to read offline
1611ISSN 2041-5990Therapeutic Delivery (2011) 2(12), 1611–163110.4155/TDE.11.131 © 2011 Future Science Ltd
Review
Special Focus: Oral Delivery of Macromolecules
Biopharmaceuticals, especially proteins, have
become increasingly popular and relevant in
the treatment of a number of disease states
due to their high specificity and activity at low
concentrations in comparison to traditional
small-molecule therapies [1–6]. However, most
proteins are still delivered parenterally rather
than orally [1,2,5,7]. This is due to a number of
factors including their structural complexity and
sensitivity to degradation, as well as the numer-
ous barriers to oral absorption [1–3,7–10]. When
administered orally, the bioavailability of most
proteins is virtually zero [5], exceptions being
the hydrophobic immunosuppressive cyclospo-
rin A [10,11] and the analogue of vasopressin, des-
mopressin acetate [1,12]. It is thought that non-
parenteral administration will increase patient
compliance, which may be particularly pertinent
in the case for patients who have a phobia of
needles [1,5,10].
Increasing the oral absorption of proteins,
therefore, has long been sought [1,3,5–7,9,10,13] and
a number of oral protein technologies are already
undergoing in vivo testing with some promis-
ing results [5,6]. These include Eligen®
 [14,15],
BioOral®
  [16], HIM2  [1,7,17], Oraldel®
  [18],
AI-401 [19], Macrulin®
 [20] and Orasome®
 [6].
The oral delivery of insulin is the main driver
for the development of these technologies given
its high (and increasing) level of clinical use in
comparison to other proteins [4,5,10].
It is thought that oral insulin delivery will
be advantageous not only in potential improve-
ments in patient compliance, but also avoid-
ing the peripheral hyperinsulinemic effects of
parenteral delivery and replicating endogenous
insulin hepatic delivery  [4,10,21–25] as well as
potentially providing a more sustained control
over blood glucose levels compared with current
parenteral delivery [26]. Insulin has a plasma half-
life of 6 min, and is cleared within 10–15 min,
which means that any delivery system that pro-
vides controlled release would be advantageous
in terms of maintaining glycemic control for
extended periods [4]. However, it is also possible
that the delivery of proteins to non-endogenous
sites may also have unwanted effects such as
insulin-induced gastroparesis [27].
None of these delivery systems are yet widely
available to the public [6]. This is due to a number
of factors including: lack of reproducible biologi-
cal effects and in vivo toxicity; toxic effects of cer-
tain excipients on the protein and patient; produc-
tion and formulation costs; deleterious effects of
formulation on protein conformation/structure;
and the two main barriers to oral protein delivery
of facilitating protein absorption and protecting
them from degradation in the gastrointestinal
(GI) tract [1,6,7,27].
Barriers to oral protein delivery
Due to the mainly hydrophilic nature of proteins
and their large size and relatively high-molecular
weights, oral absorption is difficult to achieve [6,7].
Most proteins are classified as Class III drugs
(Biopharmaceutics Classification System), in that
they are water soluble, but poorly soluble in lipid
and so poorly absorbed (i.e., they have low log P
values) [4,28]. (The exception being cyclosporin A,
which is Biopharmaceutics Classification System
Class IV.) For proteins to be absorbed orally they
need to be taken up by specific receptors in the GI
The oral delivery of proteins using
interpolymer polyelectrolyte complexes
In spite of the numerous barriers inherent in the oral delivery of therapeutically active proteins, research into the
development of functional protein-delivery systems is still intense. The effectiveness of such oral protein-delivery
systems depend on their ability to protect the incorporated protein from proteolytic degradation in the GI tract
and enhance its intestinal absorption without significantly compromising the bioactivity of the protein. Among these
delivery systems are polyelectrolyte complexes (PECs) which are composed of polyelectrolyte polymers complexed
with a protein via coulombic and other interactions. This review will focus on the current status of PECs with a
particular emphasis on the potential and limitations of multi- or inter-polymer PECs used to facilitate oral
protein delivery.
Colin Thompson* &
Chidinma Ibie
Robert Gordon University, School of
Pharmacy and Life Sciences, Aberdeen,
AB10 1FR, UK
*Author for correspondence:
Tel.: +44 1224 262 561
Fax: +44 1224 262 555
E-mail: c.thompson@rgu.ac.uk
For reprint orders, please contact reprints@future-science.com
Review |Thompson & Ibie
Therapeutic Delivery (2011) 2(12)1612 future science group
epithelium (e.g., IGF-I [29] and insulin [4,27,30]) or
via endocytosis [4,31]. However, uptake via these
routes does not lead to pharmacological effects
in vivo due to simultaneous protein metabolism.
Therefore attempts have been made to facilitate
absorption and protect against metabolism by
use of different delivery systems. Such delivery
systems have a number of barriers to overcome
in terms of the GI tract including:
n	Mucus layer on epithelial cells;
n	Unstirred water layer on epithelial cells;
n	Low permeability of cell membranes and GI
tract to macromolecules;
n	The presence of efflux pumps in cell
membranes;
n	Tight junctions between epithelial cells;
n	Chemical and enzymatic degradation/metab­
olism of proteins in GI fluid and on the brush
borders and inside epithelial cells.
The mucus layer on epithelial cells is continuously
replenished every 12–24 h [32] and is a physical
diffusion barrier to drug absorption, particularly
macromolecules [1,5,33,34]. It can act to prevent
absorption due to the adhesive nature of mucin
(the major component of mucus) causing pro-
teins to be trapped in the mucus layer and there-
fore preventing their uptake/passage between
cells  [1,35–37]. Amphiphilic structures such as
proteins may also interact with mucin via non-
specific hydrophobic interactions such as those
seen with bacteria and surfactants [34], further
reducing their ability to be absorbed. Cationic
proteins may also electrostatically interact with
the negatively charged groups on the side chains
of mucin [34,38], limiting their uptake.
The presence of an unstirred water layer on
epithelial cells is less significant in absorption,
but still can play a limiting role, particularly for
hydrophobic drugs. However, in the case of most
proteins they will readily pass across this layer
due to their hydrophilic nature and their absorp-
tion is mainly limited by their inability to diffuse
through lipid cell membranes [1].
There are many potential mechanisms of
drug absorption via either the paracellular or
transcellular routes (Figure 1) [1,3,4,34,39].
The paracellular route is attractive for proteins
due to the lack of enzymatic activity in the inter-
cellular spaces [1,4,8,9]. However, access to intercel-
lular spaces is limited by tight junctions between
cells. The integrity of these tight junctions is
maintained by a number of transmembrane
(occludin and claudin) and cytoplasmic proteins
(a family of zona occluding proteins, e.g., ZO-1)
and when these junctions are closed only very
small molecules, for example, water, may pass
through them [4,10,40]. Also found within the tight
junctions are cadherin glycoproteins, which rely
on maintenance of physiological concentrations of
calcium to preserve tight-junction integrity [4,10].
Therefore, paracellular transport is not normally
an option for macromolecules. As stated above, the
junctions found between cells only allow for the
passage of very small hydrophilic molecules [4,5].
Molecules of 10–50 Å have been shown to pass
between cells, although not usually more than
11.5 Å, due to the gap between cells being less
than a nanometer [3,10,34]. There is also a limit in
molecular weight of 200 g/mol via the paracellu-
lar route, much smaller than most peptides [1,7,39].
Therefore, for macromolecules to have a chance
to pass through tight junctions, they need to be
opened to quite a significant extent [4,10].
Similarly, although a large number of small-
drug molecules are absorbed via transcellular
passive diffusion, this is not possible for pro-
teins given their hydrophilicity, low log P values,
high-molecular weights and charged nature,
which mean they have very low absorption dif-
fusion coefficients [1,3,4,7,35,41]. The presence of
efflux pumps (e.g., P-glycoprotein [P-gp] and
multidrug resistance protein) in GI epithelial
cell membranes may also result in any protein
that is absorbed into cells being removed before
transport to the systemic circulation can be com-
pleted [1,3,4,9,39,42]. However, this tends to be more
of a problem with small molecules [39]. An excep-
tion to this being cyclosporin A, which has been
shown to be a substrate for P-gp [1,9,42].
Peyer’s patches are also be involved in trans­
cellular absorption. These are lymphoid nodules
found primarily in the ileum of the small intes-
tines and are covered by a different type of epi-
thelial cells than the rest of the GI tract, called
microfold or M cells [4]. The patches are predom­
inantly involved in mediating immune response
to pathogens in the GI tract and transporting
absorbed substances to the lymph nodes and
spleen. However, their structure (underdeveloped
villi and mucus layer) allows for nano-sized mate-
rials to easily pass through, usually of 200 nm
or less in diameter [4,43,44], while materials of
3000–5000 nm or more in diameter are trapped
within the patches [4,45]. However, they const­
itute a limited part of the GI tract epithelium as
a whole, which limits their role in absorption of
therapeutic compounds [4,43,45].
The oral delivery of proteins using interpolymer polyelectrolyte complexes | Review
www.future-science.com 1613future science group
Carrier-mediated uptake of some proteins is
feasible due to a number of therapeutic proteins
being of endogenous origin. These proteins can
act as substrates for receptors found in the GI
tract and, in doing so, are absorbed into epithe-
lial cells [46,47]. However, this is only the case
for a limited number of proteins and these pro-
teins are subsequently enzymatically degraded
within cells, prior to passage into the systemic
circulation [3,4].
Further protein breakdown can occur in the
GI fluid and this is primarily achieved by endo-
peptidases: pepsin (stomach), trypsin, a-chymo-
trypsin and elastase (intestines) [1,4,5,7,8,10,48–50].
Enzymatic degradation can take a number of
forms with the hydrolysis of the peptide bonds
within proteins being the most common, as well
as oxidation, deamidation and phosphoryla-
tion [1,9]. Each enzyme has its own target site(s)
on an individual protein (Table  1)  [7,10,48,50],
which makes simultaneous protection against
all GI enzymes difficult to achieve.
A further site for protein breakdown is in the
brush-border membrane of the GI tract, where
degradation can occur by exopeptidases such as
the amino- and carboxy-peptidases [3,4,8,9,51].
Changes in pH can also play a role in protein
degradation as well as stability and conforma-
tion [1]. The low pH in the stomach (pH 1–3)
can induce degradation of peptide bonds as it is
crucial in the activation of pepsin from pepsino-
gen [7,9]. In addition, the change in pH along
the GI tract can have profound effects on pro-
tein conformation, hence, their activity as well
as susceptibility to enzymatic breakdown [1]. As
proteins are amphoteric, their degree of ioniza-
tion and so aqueous solubility will vary consider-
ably with pH and their charge will alter between
anionic and cationic around their isoelectric
point (pI). Changes in charge on proteins will
alter their ability to form ionic and hydrogen
bonds, thus altering their secondary, tertiary and
quaternary structures. This change in confor-
mation can render them inactive and also more
susceptible to degradation, as further enzymatic
target sites may become exposed [1,3,52,53].
As is evident, oral administration presents
numerous challenges as a route for protein
Basolateral (B) side
Paracellular permeation (passive)
Passive transcellular
transport
Carrier-mediated
efflux
Carried-mediated
uptake
Tight junction
Absorptive direction (A to B)
Excretive direction (B to A)
Concentration gradient
Carrier-mediated
uptake
Carrier-mediated
efflux
Passive transcellular
transport
ABL permeation (passive)
Apical (A) side Intestinal membrane
Figure 1. GI tract epithelium physiology and transport routes across the GI tract.
ABL: Aqueous boundary layer.
Reproduced with permission from [116]. © Nature Publishing Group.
Review |Thompson & Ibie
Therapeutic Delivery (2011) 2(12)1614 future science group
delivery. Therefore a number of different technol-
ogies to improve oral bioavailability of proteins
have been developed.
Technologies to facilitate oral
protein delivery
Early steps in improving oral absorption involved
co-administration/co-formulation of proteins
with protease inhibitors, such as sodium glyco-
cholate, aprotinin, bacitracin, soybean trypsin
inhibitor, camostat mesilate [54,55] and absorp-
tion enhancers such as bile salts, salicylates, sur-
factants and chelating agents [56,57]. However,
these were found to have deleterious effects on
the GI tract wall, on normal digestion and the
pancreas itself [1,3–7,9,10].
Subsequently, encapsulation of proteins in
particulate structures of various sizes and mor-
phologies has become one of the most widely
used approaches to facilitate oral delivery [43,58–
62]. This concept is usually aimed at modifying
the pharmacokinetic profile and physico­chemical
characteristics of the resultant construct, as well
as providing an outer polymeric membrane for
the entrapped protein, which can serve as a
protective coat from proteolytic enzymes [61,62].
Particulate carriers having different functional-
ities can be tailor-made by careful selection of
parent polymers that possess desirable proper-
ties such as enzyme inhibition, mucoadhesion,
absorption-enhancing or pH/environmentally
sensitive properties [63–69]. These delivery sys-
tems may also be functionalized by attachment of
various receptor-recognizable/targeting ligands
to preformed particles  [64,70,71]. Particulate
carriers including liposomes [3,6,72], micropar-
ticles [3,6,7,72], and polymeric colloids including
polyelectrolyte nanoparticulates  [3,52,53,73–79],
have been developed with varying degrees
of success.
For oral protein delivery, the use of liposomes
(tiny, vesicular self-assemblies consisting of
hydrophobic and hydrophilic domains formed
when phospholipids are dispersed in aqueous
media) is challenging due to the inability of lipid-
based systems to entrap sufficient quantities of
hydrophilic drugs resulting in problems such as
low GI tract stability, low drug-loading capacity
and leakage of entrapped drug [3,6,72,80].
Moreover, the fabrication of micropart­
icles often requires the use of organic solvents,
high temperatures/pressures and hydropho-
bic polymers, which can denature proteins. In
addition, their use in oral protein delivery has
led to highly variable levels of protein in the
circulation [1,7,81–84].
In contrast, polymeric colloidal systems are
thought to be advantageous, as they can be more
quickly and completely absorbed due to their
large surface areas and nano-range sizes, as well
as protect proteins from degradation and improve
their in vivo stability [3,6,52,75,85,86]. However,
there can be a lack of control of protein release
and levels of protein loading can be low [3,53].
Polyelectrolyte complexes (PECs) are nor-
mally nanosized particulates that are composed
of polyelectrolyte polymers and a protein. Their
formation is driven by coulombic interactions,
that is, electrostatic interactions and hydrogen
bonding, as well as hydrophobic interactions in
some cases (Figure 2) [52,76,78,82,86,87]. They can
also involve prior copolymerization of oppo-
sitely charged polymers, which can then inter-
act electrostatically with each other and with a
protein [65,66].
PECs are able to form spontaneously (self
assemble) under benign conditions and can have
high levels of protein loading, which is advanta-
geous in comparison to some of the other delivery
technologies mentioned above [78,82,84,86,88–90].
Key Terms
Polyelectrolyte complex:
Polyelectrolyte
polymer–protein complex
whose formation is primarily
driven by coulombic forces.
Chitosan: Cationic
polysaccharide derived from
crustaceans used in a large
number of polyelectrolyte
complexes.
Thiomers: Polymers that
contain moieties with
thiol groups.
Caco-2: Colorectal cancer cell
line, one of the most common
in vitro cell models for the GI
tract.
Quaternary ammonium
moiety: Quaternary amine,
commonly produced by the
trimethylation of the primary
amines of chitosan.
Table 1. Various GI tract enzymes and their respective protein target sites.
Enzyme Target sites
Endopeptidases Nonterminal residues
Trypsin After arginine, lysine unless followed by proline
a-chymotrypsin After phenylalanine, tyrosine and tryptophan unless followed by proline; also
after asparagine, leucine, methionine and histidine
Pepsin Before leucine, phenylalanine, tyrosine and tryptophan unless preceded by
proline; can also cleave at many other sites
Elastase After glycine, alanine, serine and valine unless followed by proline
Exopeptidases Terminal residues
Carboxypeptidase A Aromatic/aliphatic hydrocarbon chain residues at the carboxy-terminal
Carboxypeptidase B Cationic residues at carboxy-terminal
Aminopeptidase Residues at amino-terminal
The oral delivery of proteins using interpolymer polyelectrolyte complexes | Review
www.future-science.com 1615future science group
In addition, they can also fulfill the two main
criteria for oral protein-delivery systems in that
they have the potential to both facilitate absorp-
tion and offer a degree of protection against
degradation in the gut.
„„ Chitosan PECs
Chitosan is a weakly cationic polysaccha-
ride derived from the shells of crustaceans.
It has been used as a mucoadhesive, release-
rate controlling excipient in a number of for-
mulations due to its nontoxic, biocompatible
nature [22,23,78,82–84,88–91]. It has also been shown
to promote tight-junction opening, facilitating
paracellular transport of active molecules [78,89,91].
It can do this by electrostatic interactions with
tight-junction proteins to reversibly alter their
morphology and increase the permeability of
intercellular spaces [78,83,92,93]. Its highly positive
charge can also facilitate cell uptake due to inter-
actions with anionic cell membranes to a greater
extent than anionic and non-ionic polymers [23].
However, its weakly basic nature means that it
is fairly insoluble above pH 6.5 (pKa
 = 6.5) [22,82–
84,88–91]. Therefore its ability to gel, to facili-
tate junction opening and adhere to mucus
membranes is limited in the intestines and
colon  [2,23,78,89,92]. To overcome these prob-
lems, derivatives of chitosan have been devel-
oped, which have pH-independent charges; for
example, trimethylated chitosan (TMC), as well
as those that form stronger (of up to 140-fold),
covalent mucoadhesive bonds (e.g., thiolated chi-
tosan [thiomers]), in the intestines/colon than
other forms of chitosan [2,78,89,91–95].
Chitosan and its derivatives have been shown
to offer some protection against enzymatic break-
down, particularly thiolated versions, which are
vital for the function of some enzymes [78,94,95].
They are able to do this as thiol groups can chelate
metal cations, such as zinc, which are responsible
for enzyme function.
These modified chitosans are also capable of
facilitating protein absorption via both paracellu-
larandtranscellularroutes.Thiolatedchitosanhas
been shown to increase membrane permeability,
possible due to an interaction with protein tyro-
sine phosphatases (PTPs). PTPs de­phosphorylate
the tight-junction protein occludin causing tight
junctions to close. It is postulated that by forming
disulfides with PTPs, thiolated chitosan is able
to prevent junction closing for longer by prevent-
ing dephosphorylation [65,95]. The mechanism by
which thiomers first open tight junctions is as
yet unclear. However, it is likely in part due to
chelation of metal cations, such as calcium, which
are responsible for maintenance of tight-junction
integrity [65,66]. Nonetheless, the limited inter-
cellular space available for paracellular transport
means that protein transport via this route can
be negligible for macromolecules regardless of the
extent of junction opening [92].
Thiolated chitosan may even be able to inhibit
P-gp efflux pumps [95] although, to date, this has
only been shown with model molecules (e.g.,
rhodamine 123, which is 380 g/mol) and not a
protein [96,97].
TMC modified by covalent attachment of
poly(ethylene glycol)(PEG)-ylated has been
shown to facilitate adsorptive endocytosis of insu-
lin in a Caco-2 cell model [92]. Uptake was shown
to linearly increase with polymer concentration
andincubationtimeshowingthatthepolymerwas
able to promote absorption for up to 4 h in vitro.
Indeed, insulin uptake increased from approxi-
mately 20 µg/mg protein (insulin only control)
to 60 µg/mg protein when insulin was complexed
with PEGylated TMC (100,000 g/mol) at physio­
logical pH. The promotion of cellular uptake is
thought to be due to electrostatic interaction
between the quaternary ammonium moieties
on TMC and cell membranes, which facilitates
endocytosis [92].
Another way to overcome the problems inher-
ent with using chitosan alone in its native state in
oral delivery systems has been to combine it with
other polyelectrolytes to form interpolymer PECs.
„„ Interpolymer PECs
PECs of polyanions, such as alginate, and
proteins, have also been produced. However,
they suffer from similar problems to chitosan-­
based PECs, in particular pH-dependent
ionization and so solubility and mucoad­
hesion  [2,4,22,24–26,30,66,83,84,88,90,98–103].
Therefore, modified forms, such as thiomers,
Self-assembly
<200 nm
Polyelectrolyte complex micelle
–
–
–
–
–
–
–
– –
–
+
+
+
+
+
+
+
+
Figure 2. Formation of a polyelectrolyte complex.
Reproduced with permission from [117]. © Springer.
Review |Thompson & Ibie
Therapeutic Delivery (2011) 2(12)1616 future science group
interpolymer complexes of polyanions and poly-
cations have been produced to try and mitigate
the problems of using polyanions or polycations
alone [2,84,90,103].
Electrostatically induced intermolecular com-
plexation occurring between polycationic and
polyanionic molecules present in aqueous solu-
tion leads to the formation of discrete interpoly-
mer complexes, which have been shown to be
suitable delivery systems for oral proteins [104,105].
Interpolymer complexes are often prepared by
mixing aqueous solutions containing oppositely
charged polymers at a specific pH resulting in
the formation of interpolymer PECs exhibit-
ing varied morphologies and functionalities as
dictated by their chemical composition [106,107].
However, tailoring the use of these interpoly-
electrolyte structures to surmounting the bar-
riers inherent in oral protein delivery involves
careful selection of complementary polycations
and polyanions, which interact optimally pro-
viding a robust network suitable for the incor-
poration and protection of proteins intended for
oral administration. The concept of interpolymer
complexation also conveniently creates a platform
where unique properties of different polymers
that have been shown to be capable of improving
oral protein bioavailability can be imparted into
a single delivery system, thereby potentiating the
efficacy of the resultant construct. Drug release
from the interpolyelectrolyte network is usually
mediated by decomplexation of the PECs at pH
conditions that reduce the electrostatic forces
between the interacting polyelectrolytes leading
to dissociation of the complex and release of the
incorporated protein.
Invariably, many research groups have
exploited the versatility of chitosan by complex-
ing it/its derivatives with various polyanions of
which the majority are anionic polysaccharides
of natural origin such as glucomannans and
mucopolysaccharides as well as other natural and
synthetic polymers [2,22,30,66,83,84,87,88,90].
As well as forming interpolymer PECs by
mixing chitosan and polyanions, these can also
be formed after prior copolymerization of poly-
electrolytes. This can counteract some of the
potential problems associated with PECs that rely
solely on coulombic interactions. These include
counterion effects and the high solubility of chi-
tosan in the stomach and of polyanions in the
lower GI tract, which can result in partial PEC
collapse and subsequent protein dumping, limit-
ing protection from degradation and promotion
of absorption [2,24–26,65,66,78,82,89,99–102,108,109].
Alginate–chitosan PECs
Alginate is an anionic polysaccharide extracted
from bacteria or algae and is made of residues of
mannuronicandguluronicacidofvaryingratios.It
exists as salt forms due to cooperative binding with
cations, for example, sodium alginate. Alginates
form gels when mixed with divalent cations (at
certain stoichiometric ratios), such as calcium,
due to their exchange with sodium, which then
results in inter-chain associations being formed
(dimerization of the alginate) primarily between
guluronic residues [2,22,24,30,83,100,108]. Alginates are
already used in a number of foods as a thicken-
ing agent/stabilizer and have been shown to be
both biocompatible and biodegradable. They are
approved for human use by the US FDA.
In terms of their use as oral delivery vehicles,
alginates have been shown to form strong mucoad-
hesive bonds when compared with other polymers
suchaschitosan [2,22,30,83].However,alginateshave
a pH-solubility problem in that although the gels
they form release very little protein in the stomach,
once they reach the lower GI tract they can rapidly
dissolve preventing controlled release of protein
as well as providing very limited protection from
degradation. In addition, protein electrostatic
interaction with alginates can cause detrimental
changes in protein conformation and loading of
proteins can be low [2]. Therefore, attempts have
been made to use alginate in concert with other
polymers to overcome these problems.
One such combination involves the use of algi-
nate with chitosan, which forms electrostatic inter-
actions between the carboxyl groups of alginate
and amino groups of chitosan [2,22,24,30,83,100,108].
This predominantly occurs between carboxyls
on mannuronic rather than guluronic residues of
alginate and aminos on chitosan, probably due to
the lower pKa
of mannuronic residues (3.38 com-
pared with 3.65 for guluronic) [110]. The potential
advantage of this combination is that it allows the
formation of a pH-sensitive gel network, which
can complex/load a protein for delivery. This
network does not entirely dissolve in the stom-
ach or intestines due to the presence of alginate
and chitosan, respectively. It can still swell in the
intestines, as alginate does on its own, but due to
the presence of chitosan does not completely dis-
solve there and so may allow for mucoadhesion
and controlled release to take place.
These PECs are produced by first forming
nuclei of alginate complexed with calcium and a
protein followed by addition of chitosan to form
PEC nanoparticles via crosslinking between chito-
san and alginate, as well as between each polymer
Key Term
Interpolymer complex:
Polyelectrolyte complex
consisting of two or more
oppositely charged polymers.
The oral delivery of proteins using interpolymer polyelectrolyte complexes | Review
www.future-science.com 1617future science group
and the protein (Figure 3) [22,30,83]. Pregelation
with calcium is thought to increase PEC stability
by providing an additional level of crosslinking
between alginate chains. Crucially these PECs
have been shown, via derivatized Fourier trans-
form IR and circular dichronism spectra, to have
no significant effect on insulin secondary struc-
ture and that insulin is stable in the PECs for up
to one month, suggesting that the protein is still
in its native conformation after complexation [22].
However, the limitation of these systems is that
polymer–polymer and polymer–insulin interac-
tions are also still pH dependent with insulin
association efficiency falling from 81 to 60% of
total insulin used in nanoparticle manufacture
when PEC solution pH was increased from 4.7
to 5.2 due to a reduction in charge density of
chitosan and insulin (pI = 5.3) [22]. This can lead
to low protein loading and cause burst release at
both low and high pH [2,7]. Indeed it was found
that alginate–chitosan PECs showed burst release
of insulin between 40–60% within 5 min in sim-
ulated gastric (pH 1.2) and simulated intestinal
fluid (pH 6.8) [22]. This was due to the variations
in charge density on chitosan, alginate and insu-
lin at these pH values resulting in weak interac-
tions between the three components of the PEC
and so uncontrolled release of insulin.
In addition to this, some precipitation of alg­
inate and chitosan can still occur over the pH
range in the GI tract resulting in large increases
in particle size, which in turn could limit cellular
interactions and uptake due to the reduction in
PEC surface area. Sarmento and colleagues found
that decreasing pH from 4.7 to 4.2 resulted in
particle size increasing from 797 to 2858 nm,
as the reduction in pH reduces the ionization of
alginate (pKa
between 3.3 and 3.7, depending on
ratio of mannuronic to guluronic residues) [87].
Regardless of these problems, however, algi-
nate–chitosan PECs have been shown to be effec-
tive at reducing blood glucose levels in diabetic
rats in vivo [30]. Oral doses of 50 or 100 IU/kg
contained in alginate–chitosan PECs were shown
to gradually reduce glycemia over a 14 h period
to approximately 60% of initial values. This
O
HO
HO
O
O
O
O
O
O
HO
HO
HO
HO
HO
HO
HO
O
O
O
O
O
O
O
O
O
-
-
-
-
-
-
-
-
O
O
O
O
O
HO
HO
OH
OH
O
OO
NH+
OHO
OH
HO
OH
HO
HO
n
n
HO
HO
HOHO
Guluronic residue
Calcium
crosslink
Deprotonated alginate
Electrostatic
interaction
Mannuronic residue
Protonated
chitosan
O
O
OH
O
OH
OH
OH
OH
OH
O
OO
O
O
O
O
O O
O
O
O
O
O
Ca2+
OH
OHOHOH
+
H3
N
+
H3
N
+
HN
HO
NH3
+
NH3
+
Figure 3. Polyelectrolyte complex of calcium-crosslinked alginate complexed with
chitosan.
Review |Thompson & Ibie
Therapeutic Delivery (2011) 2(12)1618 future science group
compared favorably with a 2.5 IU/kg dose of
noncomplexed insulin given subcutaneously
(sc.), which reduced glycemia by more than 50%
within 2 h, but which returned to near initial
blood glucose levels after 8 h. It was also differ-
ent from an oral insulin solution and a physical
mixture of insulin and empty nanoparticles (both
50 IU/kg), which resulted in a max­imum reduc-
tion in glycemia of only 20% after 6 h, which then
returned to near pretreatment levels rapidly [30].
This indicated that although some insulin
would have been released in the stomach and
overall control of release was probably poor (as
well as possible precipitation/dissolution of the
polymer complex), the PECs allowed insulin
to pass through the GI epithelium intact and
produce a hypoglycemic effect. This effect was
delayed compared with sc. insulin, as it would be
expected that the PECs adhere to the GI mucosa
and produce a prolonged release of insulin. As
well as this, the time taken for GI transit, trans-
port of insulin through the various barriers of
the GI tract epithelium, then the circulation
and finally to the liver before it could exert its
affect would take far longer than transit to the
circulation after sc. administration [4].
These results are promising in the search for
an oral insulin-delivery system in that it would
appear that the alginate–chitosan PECs are able
to protect insulin and facilitate its absorption to
some extent. It would also suggest a prolonged
control of blood glucose levels would be pos-
sible, which would reduce patient dosing as well
as potentially provide for a reduction in diabetic
complications.
However, unsurprisingly, the doses required to
elicit this effect are high compared with sc. doses
(a dose of 25 IU/kg given as an oral nanopar-
ticle produced a much reduced effect [maxi-
mum reduction of 25% glycemia after 8 h]).
Bioavailability of the 50 and 100 IU/kg doses
were 5.1 (equivalent to 7% of administered dose)
and 9.5 ng/ml.h, respectively, which although
much higher than oral insulin administered non-
complexed (2.1 ng/ml.h), could still be improved
in order to reduce dose sizes. It should also be
borne in mind that these relatively high doses are
required due to loss of insulin during GI transit
and transport across the GI epithelium due to
chemical and enzymatic breakdown. Therefore,
further work on these systems’ ability to protect
insulin from degradation could be investigated.
Additionally, although several uptake mecha-
nisms were postulated (absorptive endocytosis,
paracellular transport, Peyer’s patches) the exact
routes and mechanisms of uptake have not yet
been determined. If the mechanisms by which
these PECs facilitate insulin uptake could be fully
elucidated, it could indicate any further changes
to the PECs structure, which may aid uptake and
also minimize insulin loss during delivery.
Further alginate–chitosan PECs were devel-
oped by El-Sherbiny and colleagues involving a
carboxymethylchitosan(CMC)–sodium ­acrylate
copolymer, which could be complexed with cal-
cium-crosslinked alginate and a PEGylated form
of chitosan [99–102]. By grafting sodium acrylate
to the CMC, the number of carboxylic groups
within the PEC structure were increased. This
was done in order to limit the potential for swell-
ing, and therefore protein release, at gastric pH.
This is possible as the carboxylic groups can form
hydrogen bonds with each other and the hydrox-
yls of CMC as well as form strong ionic interac-
tions with calcium ions thereby minimizing the
ability of the PECs to swell at low pHs. The loss
of a large proportion of the amine groups in the
structure of chitosan due to carboxymethylation
and grafting of sodium acrylate further limits
ionization at low pH and therefore swelling due
to repulsion between protonated amino groups.
By coating sodium acrylate–CMC–alginate with
PEGylated chitosan, this should offer ever greater
protection against premature protein release in
the stomach by forming strong electrostatic inter-
actions with CMC–sodium acrylate and alginate
(Figure 4) [100].
It was found that swelling was very low (less
than 5% increase in weight of PECs after 2 h) in
simulated gastric fluid (pH 2.1), but fairly exten-
sive (up to 40% after a further 8 h) in simulated
intestinal fluid (pH 7.4). Release of bovine serum
albumin (BSA) from these PECs was monitored
in the same buffers and was found to be lower
with those PECs coated with PEGylated chitosan
than those with no coating in simulated gastric
fluid (~10 compared with ~25% release after 2 h)
and fairly well controlled once transferred to sim-
ulated intestinal fluid (70–85% after a further
8 h). PECs coated with PEGylated chitosan also
had greater loading capacities than noncoated
forms (63–83% encapsulation for coated PECs
as compared with 31–47% for noncoated) [100].
From this, it can be seen that the additional
hydrogel coating allowed for the formation of
strong electrostatic interactions between it and
the other polymers in the PEC thereby gaining
greater control over swelling and release, but
also allowed for greater interaction with a model
protein to allow for higher rates of encapsulation.
The oral delivery of proteins using interpolymer polyelectrolyte complexes | Review
www.future-science.com 1619future science group
This work appears to also be very promising
in terms of controlled intestinal release. The for-
mulation may also have beneficial mucoadhe-
sive, absorption-enhancing and enzyme-inhib-
iting properties due to the number of carboxylic
groups present as well as the presence of PEG;
however these have yet to be determined.
Further multipolymer complexes for insulin
delivery have been developed by Woitiski et al.
again involving the use of calcium crosslinked
alginate nuclei complexed with chitosan [108].
This PEC was further modified by addition of
another polyanion, dextran sulfate, to try and
aid protein encapsulation and control release by
increasing interpolymer and polymer–insulin
electrostatic interactions (as with the use of multi-
ple polyelectrolytes by [100]). Attempts to improve
steric stabilization of PECs were also made by
hydrogen bonding of PEG or hydrophobic
bonding of poloxamer 188 (Pluronic®
) (a tri-
block copolymer of poly(ethylene oxide) and
poly[propylene oxide]) to chitosan and alginate.
PECs were coated with albumin in order for it to
act as a competitive target for proteolysis to limit
the degradation of insulin (Figure 5) [24,25,108],
which is a fairly novel and interesting mechanism
to reduce insulin degradation.
Importantly, both an in vitro western blot bio-
activity assay and in vivo sc. injection of insulin
(collected after in vitro release from PECs) into
diabetic rats demonstrated that insulin activ-
ity was retained after formulation into these
PECs [25,108]. However, a problem with using
another protein in this formulation was that it
also acted as a competitive binder to the polymers
in the PEC, which reduced insulin encapsula-
tion from 90% at 0.25% albumin to 75% at 1%
albumin [108].
COO-
C=O C=O C=O
O-
Cs-g-PEG
(Thin layer coating)
Alg–CMCs-g-AAs
(Negative core of
hydrogel microparticles)
A B
+
NH3
HOOC
COOH
COOH
COOH
COOH
COOH
COOH
HOOC
COOH
HOOC
OH
OH
+
NH3
NH3
+
HOOC
COOH
HO
-
OOC
COO-
COO-
COO-
COO-
COO-
COO-
COO-
COO-
-
OOC
-
OOC
-
OOC
O-
O-
+
NH3
+
NH3
+
NH3
Ionic crosslinks (with Ca2+
)
Hydrogen bonds
Repulsive interactions
CMCs-g-AAs
Na alginate
Cs-g-PEG
SIF
(pH 7.4)
SGF
(pH 2.1)
Figure 4. The different types of interactions in the developed alginate–CMCs-g-AAs hydrogel microspheres coated with
Cs-g-PEG at both pH 2.1 (SGF) and pH 7.4 (SIF).
CMC: Carboxymethylchitosan; PEG: Poly(ethylene glycol); SGF: Simulated gastric fluid; SIF: Simulated intestinal fluid.
Reproduced with permission from [100]. © Elsevier.
Review |Thompson & Ibie
Therapeutic Delivery (2011) 2(12)1620 future science group
Additionally, although some formulations
limited release in simulated gastric fluid (to at
best ~12% after 2 h), on transferring PECs to
simulated intestinal fluid very rapid bursts (of
up to another 40% of loaded insulin) of release
were seen with all formulations. This would sug-
gest a similar problem as found with the stan-
dard alginate–chitosan PECs discussed above,
which could also not control release on increases
in pH due to polymer dissolution/precipitation,
or swelling owing to electrostatic repulsion
between ionized carboxyl groups [2,22,30,83].
It may be that the additional binding of dex-
tran sulfate as well as PEG/poloxamer and albu-
min to alginate–chitosan complexes reduces,
rather than improves, their stability at intestinal
pH values. These additional polymers and pro-
teins may limit the number of binding sites avail-
able for insulin in the PECs and result in weaker
polymer–protein interactions as well as weaker
interpolymer interactions limiting control over
protein release.
Further to this it was found that the PECs
produced by Woitiski and colleagues were still
able to protect insulin from pepsin degradation
in vitro, compared with a noncomplexed insu-
lin control (18 and 77% degraded, respectively)
probably due to the lack of swelling of the PECs
at gastric pH; the fact that the majority of the
insulin is entrapped within PEC cores and not
on the surface; and that albumin acted as a
sacrificial target for the enzyme [25].
Furthermore, it was found that these PECs
(using a dose of 50 IU/kg) were able to reduce
blood glucose in diabetic rats gradually over 12 h
to a peak reduction of 60% of initial values and
were able to maintain levels to approximately
50% of initial values for a further 12 h and
overall bioavailability was 13% (almost double
that achieved by Sarmento et al. with the same
dose [30]). This is a promising finding as it would
indicate these PECs are able to deliver larger
amounts of bioactive insulin through the GI
epithelium and into the circulation over a long
period of time and crucially maintain low glyce-
mia for up to at least 24 h after administration.
This (together with Sarmento and colleagues
in vitro release and in vivo glycemia data dis-
cussed above) would also suggest that the release
of insulin in vivo is quite different than in vitro,
given that hypoglycemia lasted for up to 24 h
in vivo, but insulin release in vitro was complete
within 180 min. This implied poor in vitro–
in vivo correlation would indicate that the use
of simulated gastric and intestinal fluid in these
kinds of studies has limited use at best, particu-
larly given the other substances, such as food,
numerous enzymes and differences in buffer
capacity and ionic strength of GI fluid compared
with those used on the bench.
Supplementary work with these PECs found
that the presence of albumin also aided insu-
lin transport across both in  vitro cell layers
(Caco-2 alone and co-cultured with HT-29)
and ex vivo rat intestinal mucosa. The perme-
ability co­efficient for albumin-covered PECs
was 28.4  ×  10-6
compared with 19.38 and
7.35 × 10-6
 cm����������������������������������/s for albumin-free PECs and insu-
lin alone, respectively, in ex vivo rat intestinal
mucosa. This was possibly achieved by albumin
acting as a stabilizer for insulin (i.e., it aids pas-
sage across the mucus layer as well as limiting
intracellular and extracellular enzymatic break-
down). However the data for albumin-free PECs
may be more instructive in terms of determin-
ing permeability in vivo as the albumin layer
should be degraded in the stomach and by gas-
tric and intestinal enzymes prior to initiation of
absorption of insulin in the intestines.
Transepithelial electrical resistance (TEER)
values of co-cultured Caco-2 and HT-29 cells
also dropped on incubation with PECs. However,
this was only by approximately 20% of initial
values, which would suggest that transport was
predominantly via transcellular routes [26]. The
Alginate
Dextran sulfate
Poloxamer
Albumin
Chitosan
Calcium
Insulin
Figure 5. Multilayered nanoparticles encapsulating insulin. Nanoparticles
are formed by alginate and dextran sulfate nucleating around calcium ions and
interacting with poloxamer, further stabilized with chitosan and subsequently
coated with albumin.
Reproduced with permission from [25] © Elsevier.
The oral delivery of proteins using interpolymer polyelectrolyte complexes | Review
www.future-science.com 1621future science group
exact mechanisms and transporters involved in
uptake and translocation through cells are also
to be elucidated.
Poly(aspartic acid)–chitosan PECs
Poly(aspartic acid) is also a poly(carboxylic acid)
like alginate. However, it is synthetic rather
than derived from natural resources. It is pro-
duced by polycondensation of aspartic acid to
poly(succinimide) followed by hydrolytic ring-
opening. As it is based on an amino acid, the
polymer formed has a protein-like structure and
is water soluble and biodegradable [84,90].
Shu and colleagues formed PECs between
chitosan and thiolated polyaspartic acid with
a view to produce pH-sensitive swellable
nanoparticles for the oral delivery of insulin
(Figure 6) [90]. The potential advantage of these
PECs is that they may be more stable than other
chitosan–polyanion-based PEC. Previously, the
electrostatic interactions used to form PECs
based on chitosan and other polyanions were
weaker or stronger in different parts of the GI
tract due to the variation in pH, depending upon
their pKa
, and were generally unstable in vivo
due to salting-out effects [22,30,78,83,87,109].
However, the use of a thiolated polymer in
this case has not only allowed for electrostatic
interactions to take place, but for disulfide-bond
crosslinking, which increased the strength of
interactions and overall nanoparticle stability
at low pH (pH 1.2–4) and in increasing salt
concentrations in vitro.
Additionally, in vitro release studies dem-
onstrated that PECs, which relied solely upon
electrostatic interactions, released insulin very
quickly (~90% within 30  min) at pH  1.2,
while chemically crosslinked particles showed a
more controlled release at pH 1.2 (~40% after
120 min). Although after adjustment of the pH
to 6.8, the remaining insulin from crosslinked
nanoparticles was very rapidly released (~90%
after a further 30 min).
The rapid release in pH 6.8 is surprising,
given that disulfide-bond formation should be
increased at a pH above 5 due to the increasing
oxidation of the thiol groups on the PECs [32,63,94–
97,103]. The opposite appears to have happened
whereby the interpolymer interactions have
become so weak that the nanoparticles can no
longer retain insulin at increasing pHs. It may be
that the use of chitosan has resulted in a degree
of precipitation of the formulation at pH 6.8
given that it would be fairly insoluble at that pH.
As the disulfides do not form between chitosan
and aspartic acid, but only between aspartic
acid chains this would not solve the problem
of the lack of solubility of chitosan at higher
pH. This together with the loss of electrostatic
interaction between chitosan and aspartic acid
could have caused the burst release. A variation
on this formulation was also examined where a
low-molecular weight (6000 g/mol), water-solu-
ble chitosan was used instead and poly(aspartic
acid) was used without thiolation. In this study,
the polyelectrolytes were co-mixed with PEG
in order to provide additional sites for inter-
polymer interactions and, therefore, form more
stable PECs [84]. PEG was also used in order to
promote mucoadhesion by improving complex
chain flexibility and mobility as well as maintain
the secondary and tertiary conformation of BSA
when encapsulated.
The addition of PEG may have resulted in
interpolymer interaction between its ether group
and the carboxylic-acid group of the aspartic
acid as well as between the primary amines on
chitosan and the hydroxyl on PEG. As a result of
this, the interpolymer network was strengthened
and more compact nanoparticles were formed
(from a radius of approximately 175 nm with
no PEG to 125 nm at 30 mg/ml PEG). These
nanoparticles were also stable along a range of
salt concentrations with the particle radius not
exceeding 300 nm up to 0.025 mol/l NaCl.
This was in contrast to the previous study where
particle radius increased up to 400 nm at the
same salt concentration [90]. However it is dif-
ficult to attribute this increased stability solely
to the presence of PEG, given the number of
other variables that changed between the two
studies [84,90].
Regardless of the apparent increase in par-
ticle stability (uncontrolled), pH-dependent
release was again found. In vitro BSA release was
approximately 20% at pH 2.5, 60% at pH 1.2
and 80% at pH 7.4 (after 240 min). Again, it
can be seen that noncontrolled release occurred
at a higher pH, which would not be beneficial
in vivo. The reduced protonation of chitosan at
pH 7.4 would again, regardless of its increased
water solubility at low-molecular weight, result
in very loose and diffuse PECs at this pH causing
rapid and uncontrolled protein release.
In order to improve upon this delivery sys-
tem, quaternization and/or thiolation of chitosan
may also be required in order that electrostatic
forces between polymers are strengthened and
are pH independent as well as potentially allow-
ing the formation of disulfides between chitosan
Review |Thompson & Ibie
Therapeutic Delivery (2011) 2(12)1622 future science group
and (thiolated) aspartic acid in order to further
strengthen the interpolymer network. In addi-
tion, conjugation of PEG to either polymer may
also facilitate greater complex stability, solubil-
ity, biocompatibility and mucoadhesion than the
weaker hydrogen bonding seen above between
PEG and chitosan [82,88,89,91,92]. Indeed, it has
been postulated that the counterion effects of
salt can be counteracted to a limited extent
by steric hindrance of quaternary ammonium
salt groups, grafted PEG and disulfide-bond
formation [82,89].
However, it must also be borne in mind that
the ability to control release of proteins is also
dictated by their charge which will also change
with pH. The ability of a protein to form elec-
trostatic interactions with a polymer will alter
drastically around their pI, as the charge on the
polymer goes from cationic to anionic or vice
versa [92]. Therefore any system that is to pro-
duce controlled release in the GI tract and avoid
release in the stomach, must not rely solely upon
electrostatic interactions between the polymer
and protein.
Poly(methacrylic acid) copolymer PECs
Poly(methacrylic acid) (PMMA) is another weak
acid, which has been used to complex with pro-
teins. However, as with other polyanions men-
tioned above, when used alone it is unable to
Electrostatic
interaction
HO
HO
HO
HO
HO HO
HO
HO
+
H3
N
+
H3
N
+
H3
N +
H3
N
+
HN
H2
H2 H2
H2
H2
CH2
CH2
CH2
CH2
CH2
CH2
H2
NH
+
OH
OH
OH
OH
OH
OHOH
O
O O
O
O O
CC
C C C
C = O
NH
NH
C = O
C
O
S
S
O
O
O
O
- - -
- - -
-
-
O
O O O O
O O
O O
O
O
O
O
O
O
O
n
x y
n
yx
H
N
H
N
H
N
H
N
H
N
H
N
O O
O
O
O
OH
Disulfide
crosslink
Ionised
poly(aspartic acid)
Protonated
chitosanNH3
+
NH3
+
Figure 6. Polyelectrolyte complex of disulfide crosslinked thiolated poly(aspartic acid)
complexed with chitosan.
The oral delivery of proteins using interpolymer polyelectrolyte complexes | Review
www.future-science.com 1623future science group
eff­iciently complex with proteins at the pH range
of the GI tract. Therefore attempts have been
made to produce more stable complexes by copo-
lymerizing with PEG. This increases complex
stability as the PEG ether groups and carboxyl
groups of PMMA can form hydrogen bonds to
produce an interpolymer complex [13,111–113].
PMMA-based nanoparticles have been pro-
duced by copolymerizing with an amine mod­
ified PEG, which facilitated interpolymer ionic
crosslinking [13]. Individual PMMA–PEG par-
ticles were approximately 800 nm in diameter.
However, a high degree of aggregation occurred
at gastric pH resulting in the form­
ation of
micron-sized particulates. At intestinal/physi-
ological pH, above the pKa
of PMMA, increased
interparticle repulsion occurred resulting in
smaller, discrete particles. This is very impor-
tant as smaller particles are more likely to inti-
mately interact with the GI tract epithelium in
terms of promoting both mucoadhesion and
absorption [13,88].
PMMA–PEG particles were able to demon-
strate ex vivo mucoadhesion for up to 3 h due to
both the predominantly anionic charge on the
particles at pH 7 and possibly the ability of PEG
to entangle with mucus glycoproteins. However,
longer lasting mucoadhesion would be desirable
to set up and maintain a protein concentration
gradient across the GI tract mucosa into the cir-
culation and so further particle size reduction
could be beneficial.
BSA was loaded into these particles with
encapsulation efficiencies of up to 80%. In vitro
release from these particles showed a very slow
and incomplete release at pH 1.2 (less than 10%
after 3 h) and relatively rapid release at pH 7.4
(~90% after 3 h). This pH-dependent release
could be attributed to the increase in swelling of
the particles at elevated pH due to the increased
ionization of carboxyl groups on PMMA caus-
ing repulsion between PMMA chains above its
pKa
 [13]. Again, this is not ideal as although the
protein would not be released in the stomach, the
rapid release in the intestines would be unlikely
to result in the maintenance of therapeutic levels
and would most likely result in the requirement
of repeated dosing. The same study also utilized
the PMMA–PEG PECs to load insulin alone and
precomplexed to carboxymethyl-b-cyclodextrin
(Cb-CD). The Cb-CD was used to improve
insulin stability in the delivery system, by pre-
venting insulin hydrophobic intermolecular
interactions which can lead to aggregation and
a reduction in activity [21].
Encapsulation efficiency was high at
approximately 78% (insulin alone) and 73%
(insulin–Cb-CD) and insulin was shown to
still be biologically active after encapsulation
using an ELISA. However, insulin release from
the particles was very rapid at pH 7.4 (in effect,
dose dumping) with approximately 70% release
within 60 min regardless whether insulin was
encapsulated alone or complexed to Cb-CD.
The more rapid release of insulin compared
with BSA may in large part be due to its much
smaller-molecular weight compared with BSA
(5800 compared with 66,000 g/mol). The pore
size of the swollen particles may be such that
insulin can pass easily through them, while the
larger BSA requires more time in order to be able
to diffuse through the interpolymer complex.
Therefore, more precise control over swelling
(and hence pore size) is required for delivery of
smaller proteins.
One way to achieve this may be by increas-
ing the molecular weight of the PEG used.
Higher-molecular weight PEG could result in
greater interpolymer interactions, thereby reduc-
ing swelling and therefore pore size. However,
this reduction in pore size could also limit insu-
lin loading. Increasing PEG molecular weight
in PECs from 400 to 1000 has been shown to
reduce in vitro insulin release (from 41 to 39%
after 2 h) and also reduced insulin encapsulation
(from 38 to 32%) [112]. A reduction in insulin
loading may be necessary to provide greater con-
trol of insulin release. However, the reductions
in release seen here are quite small and so other
options need to be considered, perhaps including
the use of considerably higher molecular weight
PEGs, for example, ≥10000 g/mol.
Other work with PMMA–PEG copolymers
has involved the incorporation of chitosan
into PECs. Sajeesh and Sharma copolymerized
PMMA, chitosan and polyether (a copolymer
of PEG and polypropylene glycol) [21]. In this
case the electrostatic interaction was between the
primary aminos on chitosan and the carboxyl
groups on PMMA while polyether formed hydro-
gen bonds with PMMA. The choice of polyether
rather than PEG alone was due to the potential
for polyethers to promote absorption across epi-
thelial barriers via a surfactant-like effect on cell
membranes (as with the Pluronics®
) [21]. This
PEC was complexed with insulin either alone or
after insulin was again complexed with a hydro-
philic derivative of b-cyclodextrin (2-hydroxy-
propoyl b-cyclodextrin [HPbCD]) to increase
its stability in the formulation.
Review |Thompson & Ibie
Therapeutic Delivery (2011) 2(12)1624 future science group
Encapsulation efficiency of insulin was approx-
imately 85% regardless of whether insulin was
used alone or as part of a CD complex indicating
a lack of interaction between HPbCD and the
interpolymer complex. This is borne out by the
similarity in in vitro release from particles, which
carried insulin alone or as part of a HPbCD com-
plex; pH-dependent release was demonstrated
with approximately 15% release within 2 h at
pH 1.2, and 75–80% release at pH 7.4 over the
same time regardless of the presence of HPbCD.
The lack of control of release at pH 7.4 had
still not been overcome by these further formula-
tion changes, and would suggest that while the
use of a CD ensures that insulin remains active
when formulated it does nothing to aid control
of insulin release as it cannot interact with the
polymers present in the PEC.
This study was also able to show that these
PECs can show some concentration-depen-
dent trypsin inhibitory activity when using
N-a-benzoyl-l-arginine ethyl ester (BAEE)
and casein as low- and high-molecular weight
substrates, respectively. Greater protection was
offered to casein than BAEE over a PEC concen-
tration range of 0.25–1% (w/v), which may be
due to differences in substrate molecular weight
(24,000 g/mol for casein compared with approx­
imately 300 g/mol for BAEE). As the studies were
conducted at pH 7.6, the PECs would have been
swollen and so release of BAEE would have been
even more rapid than insulin, while casein release
could have been much slower. This difference in
free and encapsulated substrate could have meant
that as more casein was encapsulated for longer,
then a physical barrier to trypsin binding, as well
as any potential chelation of metal ions by the
carboxyl groups of PMMA, would have provided
greater protection to it than BAEE. However,
the protective effects seen are modest and only
reached at most to approximately 40% inhibition
of trypsin at 1%(w/v) PEC concentration.
The lack of a substantial inhibitory effect
may be due to the lack of ‘free’ carboxyls on the
PMMA after copolymerization. Electrostatic
interactions and hydrogen bonding within the
PEC between polymers and also the encapsu-
lated protein maintain the PEC structure in
solution. Therefore it may be that very few car-
boxyls were not already interacting with other
PEC components and so very few of them were
available to chelate metal ions.
This would seem to be confirmed by a simi-
lar study, which showed that carbopol 974 had a
greater inhibitory effect on trypsin activity against
casein than PMMA–chitosan–PEG PECs; 80%
inhibition compared with 50% inhibition at a
concentration of 1%(w/v), respectively  [88].
Carbopols are weakly crosslinked hydrogels of
acrylic acid and alkyl ethers and, as such, may
have a fewer number of carboxyls involved in
crosslinking than these PMMA-based PECs.
However, it would be useful to compare PMMA
alone with crosslinked versions to confirm this.
In addition, attempts to reduce the degree of
interaction between these polymers, perhaps by
varying the molar ratios used for polymerization
or molecular weights of the polymers, would
provide greater insight into this. However, a
reduction in interaction between polymers used
in these systems would probably result in even
faster release of encapsulated proteins due to fur-
ther increases in pore size on swelling or perhaps
collapse of PECs altogether on increases in pH.
Further studies in this area have been con-
ducted to try and improve control of protein
release and enzyme inhibition of PMMA–chi-
tosan–PEG nanoparticles by conjugation of cys-
teine to the surface of these PECs [65,66]. As with
other thiomers mentioned above, the presence
of these thiol groups should confer additional
properties to the PECs including covalent muco-
adhesive bond formation and greater chelation
of metal cations to facilitate both tight-junction
opening and enzyme deactivation.
These PMMA–chitosan–PEG thiomers are
different in that the surface of the PECs is thio-
lated and not the interior. To date, most thiomers
have relied on disulfide crosslinking within PECs
to form stable particles as the pH rises to five
and above [32,63,94–97,103]. However, as mentioned
above, this reduces the number of ‘free’ thiol
groups available for mucoadhesion and enzyme
inhibition. One way to improve these properties
may be to attach thiol-­containing moieties to the
surface of PECs instead, which could maximize
the ‘free’ thiol content of the formulation. It
has been found, however, that disulfides do still
form between surface thiol groups. Sajeesh et al.
found that thiol content of thiolated PMMA–
chitosan–PEG PECs increased from 506 to
1214 µmol/100 mg of PECs after treatment with
sodium borohydride to cleave disulfides [65].
The potential to form disulfides between par-
ticles should also be considered, which would
promote particle aggregation, particularly at ele-
vated pHs, such as in the lower GI tract. If this
were to occur, intimate contact between particles
and the GI mucosa could be reduced as well as
the ability to promote mucoadhesion, enzyme
The oral delivery of proteins using interpolymer polyelectrolyte complexes | Review
www.future-science.com 1625future science group
inhibition and mucosa permeability, thus, limit-
ing the ability to both protect a loaded protein
and facilitate its absorption. Limiting disulfide
formation may be achieved by the presence of
other functional groups or ligands, which could
increase steric hindrance between thiols, limit-
ing their ability to conjugate, such as quaternary
ammonium moieties or PEG [82,89].
Although no electron microscopy was per-
formed on the thiolated PMMA–chitosan–PEG
PECs in this study, particles would appear to be
aggregated at pH 7, given that sizes increased with
the number of thiol groups grafted to the particle
surface. PECs with no thiol groups registered a
hydrodynamic diameter of 1030 nm, while those
with 506 µmol of thiol groups/100 mg of PECs
were 1900 nm. Indeed increased aggregation of
the thiolated compared with nonthiolated PECs
appeared to be confirmed by optical microscopy
in a later study [66].
Importantly, the thiolated particles showed
little cytotoxicity when analyzed using an MTT
assay with even the highest thiol content PECs
(506 µmol of thiol groups/100 mg of PECs) at
the highest PEC concentration (2 mg/ml) show-
ing approximately 85% of Caco-2 cells remained
viable after 6 h incubation. This indicates that
any increase in permeability of cells caused by
these PECs should not be due to cellular damage,
or at least very low levels of damage, which may
be reversible.
These PECs were also shown to facilitate
transport of a macromolecular marker (fluores-
cein isothiocyanate-dextran 4000 g/mol [FD4])
across Caco-2 cell monolayers. This ability
was increased with the number of thiol groups
attached to the PECs and was shown to be depen-
dent upon the concentration of calcium in the
media used. As calcium concentration increased
the apparent permeability of FD4 decreased,
which would indicate that the transport of
FD4 across cell layers was due to calcium chela-
tion by the PECs, that is, it primarily involved
paracellular transport.
This was confirmed by the differences in
TEER found when Caco-2 cells were incubated
with thiolated PECs in the presence and absence
of calcium. In the absence of calcium TEER val-
ues dropped to 40% of their initial values within
30 min, compared with only 60% of initial val-
ues when calcium chloride concentration was
10 mM.
TEER reduction and concomitant transport
of FD4 was relatively similar for thiolated and
nonthiolated PECs at 0 mM calcium chloride,
as calcium concentrations increased, thiolated
PECs were able to reduce TEER and promote
FD4 transport to a greater extent than nonthio-
lated forms. This would suggest that thiol groups
are more effective chelators of metal cations than
carboxyl groups and so their use may be highly
advantageous in attempts to facilitate macromole-
cule absorption. However, when transport of FD4
across ex vivo rat intestinal tissue was determined
the presence of thiol groups appeared to make
little difference with apparent permeability of
FD4 being 7.4 × 10-7
and 7.9 × 10-7
 cm�����������/s for non-
thiolated and thiolated PECs, respectively. This
highlights the limitations of using Caco-2 cells
as an in vitro model for the GI tract epithelium.
Unlike the GI epithelium, Caco-2 cells do not
have a mucus layer and their TEER values are
markedly higher [114]. This is potentially a major
flaw with the use of Caco-2 cells as a model and
as such HT-29 cells may be more appropriate in
that their TEER values are in line with GI epi-
thelial cells and they produce a mucus layer and
M cells [26,78,114]. In particular the presence of a
mucus layer can have marked effects on PEC inti-
mate contact with mucosal membranes and tight
junctions resulting in a major reduction in the
ability of PECs to facilitate protein absorption.
The presence of thiol groups may result in disul-
fide formation between PECs and cysteine resi-
dues in mucus glycoproteins, which could mark-
edly inhibit their ability to reach and interact
with cell membranes [65]. Co-cultures of these cell
lines could provide an in vitro model, which more
closely mimics the intestinal epi­thelia, as together
they would produce monolayers which had both
absorptive and mucus secreting cells  [26,114].
However, it should be noted that Caco-2 has been
shown to be a useful model many times over and
has indeed generated in vitro data that correlates
well with in vivo findings [114].
In a later study by the same group, insulin was
loaded into the PECs rather than FD4 [66]. It
was notable that in vitro insulin degradation by a
mixture of trypsin and a-chymotrypsin (10 U/ml
each) was prevented to a greater extent with non-
thiolated PECs than thiolated versions. After
60 min, thiolated PECs had only approximately
10% of their initial insulin load nondegraded,
while nonthiolated PECs had approximately 25%
nondegraded. This is surprising given that one
expected advantage of thiolated PECs is their
increased ability to protect proteins from enzy-
matic breakdown [32,63,94–97,103]. Again this may
refer back to the propensity for surface-function-
alized PECs to form disulfides which in turn
Review |Thompson & Ibie
Therapeutic Delivery (2011) 2(12)1626 future science group
may limit their ability to chelate metal cations
which are responsible for enzyme activity [66]. It
should also be noted that regardless of PEC struc-
ture, total insulin degradation occurred within
90 min incubation, which suggests that neither
formulation may offer enough protection in vivo.
Conversely however, in vivo studies using dia-
betic rats demonstrated that blood glucose levels
were reduced by both thiolated and nonthiolated
PECs (with a dose of 50 IU/kg). Thiolated PECs
demonstrated a greater reduction of approxi-
mately 40% of initial glucose levels within 2 h
of administration compared with approximately
a 10% reduction for nonthiolated PECs. Blood
glucose levels in rats treated with thiolated
PECs gradually increased over the next 8 h, but
a reduction in glucose levels of approximately
10–20% was still maintained after 10 h. The
greatest reduction in glucose levels by nonthio-
lated PECs was seen after 6 h (20–30% reduc-
tion), which was quickly followed by a return to
pretest glucose levels within the next 2 h.
This highlights the difference in cellular
interaction of the PECs when thiol groups are
present. Even with the apparent crosslinking
between particles, enough ‘free’ thiols may be
present to rapidly chelate metal cations and inter-
act with tight-junction proteins to promote insu-
lin absorption. This could subsequently lead to
rapid insulin uptake and so a fairly rapid and sub-
stantial decline in blood glucose. The increased
ability of the thiolated PECs to adhere to the GI
tract mucosa may also explain their ability to
maintain this effect for a prolonged period. This
adherence to the GI tract wall could allow for an
insulin concentration gradient to be set up across
the epithelium into the circulation.
The prolonged reduction in glucose levels
found here is highly desirable in the treatment
of diabetes as it could maintain glucose control
without having to administer a large number of
doses per day, while potentially providing more
prolonged glycemic control than current sc.
dosing. However, as with the studies discussed
above, the reduction in glucose seen with PECs
is less than with a dose of insulin given sc. (70–
80% reduction within 90 min) and for PEC oral
formulations to be effective they must reduce
glucose levels to a great extent as well as main-
tain that effect over many hours (the accepted
gold standard for any insulin-dependent dia-
betes treatment is a reduction in glycosylated
hemoglobin levels to less than 6.5–7% [115]).
Therefore, although surface thiolation may
appear attractive in theory, in practice the very
high reactivity of thiol groups (after oxida-
tion) means that crosslinking between particles
becomes inevitable, thus limiting the ability of
these groups to aid enzymatic protection, muco-
adhesion and protein absorption. Further work in
this area should look at attempts to limit particle
crosslinking and reduce particle size to the nano-
range to improve both enzymatic protection as
well as increase absorption.
Future perspective
Oral protein delivery has still not been achieved
and although progress has been made, a great
deal of work remains to be done. The use of
PECs seems to allow for both protection against
degradation and promotion of absorption to a
limited extent. However, bioavailability of insu-
lin through the oral route is still much lower than
that of conventional parenteral delivery, requir-
ing the use of more than ten-times the sc. dose
of insulin to achieve and maintain hypoglycemia
in animal models.
There has been a general trend towards the use
of chemically modified chitosan either alone or in
combination with other polymers to form PECs.
As part of this, the use of thiomers is highly
advantageous in both promotion of absorption
and protection against enzymatic breakdown of
insulin. The use of PEG has also been a feature
of the research discussed in this review due to its
many potential benefits in terms of mucoadhe-
sion, maintenance of active protein conformation
and PEC stability. However, neither the use of
thiomers or the incorporation of PEG has so far
lead to a substantial breakthrough in oral pro-
tein delivery as they still do not provide for large
quantities of bioactive protein to pass across the
GI tract.
Interpolymer complexes seem to hold great
potential due to their ability to produce pH-
dependent release and mucoadhesion. However,
they too have their limitations in terms of effi-
cient insulin delivery to the circulation due
to continued loss of insulin during GI transit
through premature release and lack of complete
protection against degradation. This seems to be
primarily due to the limitation of utilizing PECs,
which rely solely on pH-dependent, coulombic
interactions for their formation (Table 2).
Therefore in future studies, it may be bene­
ficial to form interpolymer complexes via chemi-
cal crosslinks, which are not affected by the ionic
strength and changes in the pH of GI fluid.
However, this could lead to other issues in regards
to whether favorable protein conformations could
The oral delivery of proteins using interpolymer polyelectrolyte complexes | Review
www.future-science.com 1627future science group
be maintained and how proteins could be released
on passage through the GI epithelium.
In order to achieve high oral insulin bioavail-
ability, more work needs to be done to look at
developing polymers that have specific binding
sites for insulin (or other proteins), which pro-
tect their degradation target sites, maintain their
native conformation and also efficiently inter-
act with GI epithelial cell membranes and tight
junctions to rapidly and completely facilitate
absorption.
With this in mind, it is crucial to develop
better screening techniques for PECs to deter-
mine their uptake routes and exact mechanisms
of interaction with cell membranes and tight-
junction proteins in order that these structure
can be further fine tuned to aid protein trans-
port and translocation. This will require much
greater collaboration between polymer chemists,
human (gut) physiologists, biochemists and
pharmacologists to systematically develop PEC
systems from the ground up. Until this is done,
it is unlikely oral bioavailability of proteins can
be increased (reproducible, rapid and sustained
glycemic control can be achieved) to the point
where these delivery systems are commercially
viable alternatives to the status quo.
Financial & competing interests disclosure
The authors have no relevant affiliations or financial
involvement with any organization or entity with a finan-
cial interest in or financial conflict with the subject matter
or materials discussed in the manuscript. This includes
employment, consultancies, honoraria, stock ownership or
options, expert t­estimony, grants or patents received or
pending, or royalties.
No writing assistance was utilized in the production of
this manuscript.
Table 2. Summary of polymers used in interpolymer complexes.
Polymers Important findings Ref.
Chitosan and calcium-
crosslinked alginate
Lack of control over insulin release at intestinal pH
in vitro
Insulin stable in PECs for up to 1 month
Insulin bioavailability in diabetic rats of 7%
[22,30,83,87]
Carboxymethylated chitosan-
acrylate, calcium crosslinked
alginate and PEGylated chitosan
Coating of PECs with PEGylated chitosan and
removal of majority of primary amines from chitosan
structure offers control of protein release over pH
range of GI fluid
[99–102]
Chitosan, calcium-crosslinked
alginate, dextran sulfate,
PEG/poloxamer and albumin
Protection of insulin against pepsin degradation
in vitro
Insulin bioavailability in diabetic rats of 13%
Insulin appears to be absorbed transcellularly
[24–26,108]
Thiolated poly(aspartic acid)
and chitosan
Lack of control over release at intestinal pH in vitro
Slower release at pH 1.2 when thiolated form of
aspartic acid used
[84,90]
Poly(aspartic acid), chitosan
and PEG
Complexes more compact and stable on increases in
salt concentration when PEG incorporated into PEC
PMMA–PEG copolymer Lack of control over release at intestinal pH in vitro [13,21,65,66,88]
PMMA–PEG–chitosan copolymer
and Cb-CD insulin
Lack of control over release at intestinal pH in vitro
Cb-CD maintains insulin biological activity after
complexation with copolymer
Thiolated PMMA–Chitosan–PEG
copolymer
Presence of thiol groups increases transport of FD4
across Caco-2 monolayer
Thiol groups appear to increase chelation of calcium
in vitro thereby increasing reduction in Caco-2 TEER;
Thiol groups make no appreciable difference in FD4
transport across ex vivo mucosa
Carboxylic groups rather than thiols offer increased
protection against trypsin and a-chymotrypsin
in vitro
Thiolated PECs able to reduce glycemia by 40%
within 2 h of administration. Nonthiolated PECs
reduce glycemia by only 10% in same time.
Cb-CD: Carboxymethyl-b-cyclodextrin; FD4: Fluorescein isothiocyanate-dextran 4000 g/mol; PEC: Polyelectrolyte complex;
PEG: Polyetyhlene glycol; PMMA: Poly(methacrylic acid); TEER: Transepithelial electrical resistance.
Review |Thompson & Ibie
Therapeutic Delivery (2011) 2(12)1628 future science group
Executive summary
Current status of oral insulin delivery
„„ Proteins are becoming an increasingly popular option for treatment, due to high specificity and low toxicity.
„„ Oral delivery is preferential for patients and clinicians.
„„ A number of oral insulin formulations are in clinical trials at the moment.
Barriers to oral protein delivery
„„ Proteins are almost completely degraded in the GI tract due to enzymatic and chemical effects.
„„ Proteins cannot pass across epithelial layers unaided due to their large size, molecular weight and charged nature.
„„ There are numerous routes to pass across the gastrointestinal epithelium both para- and trans-cellularly. However, these are not
accessible to proteins when administered alone.
Technologies to facilitate oral protein delivery
„„ Numerous attempts have been made to promote absorption and prevent degradation of protein in the GI tract that have proved
unsuccessful due to problems in manufacturing of dosage forms and/or toxic side effects on administration.
„„ Polyelectrolyte complexes potentially offer a way to benignly encapsulate (via coulombic and hydrophobic interactions) and
deliver proteins.
Chitosan polyelectrolyte complexes
„„ Chitosan in its native state is ineffective in oral protein delivery due to its weakly basic nature.
„„ Modified forms of chitosan have been produced via trimethylation of primary amines or attachment of thiol-containing ligands. These
forms have shown improved ability to facilitate insulin absorption and improve protection in the intestines.
Interpolymer polyelectrolyte complexes
„„ Polyelectrolyte complexes formed from mixing (or copolymerization) of oppositely charged polymers has been achieved. Use of
polyelectrolyte mixtures can overcome some of the problems of using polycations/anions alone, namely premature release and polymer
precipitation.
„„ These have primarily been produced using mixtures of alginate and chitosan. However other polyanions have been used with chitosan
including poly(aspartic acid) and poly(methacrylic acid).
„„ Some of these systems have an ability to produce sustained hypoglycemic effects in rat models.
„„ Further work is required to maximize the efficiency of these systems, particularly in their ability to facilitate insulin absorption.
References
Papers of special note have been highlighted as:
n of interest
nn of considerable interest
1	 Hamman JH, Enslin GM, Kotzé AF. Oral
delivery of peptide drugs: barriers and
developments. BioDrugs 19(3), 165–177
(2005).
nn	 Detailed review of various barriers to oral
protein delivery and technologies to
overcome them.
2	 George M, Abraham TE. Polyionic
hydrocolloids for the intestinal delivery of
protein drugs: alginate and chitosan – a
review. J. Control. Release 114, 1–14 (2006).
3	 Morishita M, Peppas NA. Is the oral route
possible for peptide and protein drug delivery?
Drug Discov. Today 11(19–20), 905–910
(2006).
4	 Woitiski CB, Carvalho RA, Ribeiro AJ,
Neufeld RJ, Veiga F. Strategies toward the
improved oral delivery of insulin
nanoparticles via gastrointestinal uptake and
translocation. BioDrugs 22(4), 223–237
(2008).
5	 Peppas NA, Carr DA. Impact of absorption
and transport on intelligent therapeutics
and nanoscale delivery of protein
therapeutic agents. Chem. Eng. Sci. 64,
4553–4565 (2009).
6	 Park K, Kwan IC, Park K. Oral protein
delivery: current status and future prospect.
React. Funct. Polym. 71, 280–287 (2011)
nn	 Informative review of the status of oral
protein delivery and delivery systems
undergoing clinical trials.
7	 Singh R, Singh S, Lillard JW. Past, present,
and future technologies for oral delivery of
therapeutic proteins. J. Pharm. Sci. 97(7),
2497–2523 (2008).
n	 Review of a number of technologies with
potential to facilitate oral protein delivery.
8	 Woodley F. Enzymatic barriers for Gl
peptide and protein delivery. Crit. Rev. Ther.
Drug Carrier Syst. 11, 61–95 (1994).
nn	 Detailed review of enzymatic barriers to
oral protein delivery.
9	 Mahato RI, Narang AS, Thoma L, Miller
DD. Emerging trends in oral delivery of
peptide and protein drugs. Crit. Rev. Ther.
Drug Carrier Syst. 20(2–3), 153–214
(2003).
nn	 Highly detailed and comprehensive review
of all barriers to oral protein delivery and
perspective various technologies to
overcome those problems.
10	 Peppas NA, Kavimandan NJ. Nanoscale
ana­lysis of protein and peptide absorption:
Insulin absorption using complexation and
pH-sensitive hydrogels as delivery vehicles.
Eur. J. Pharm. Sci. 29, 183–197 (2006).
11	 Liu H, Wang Y, Li S. Advanced delivery of
ciclosporin A: Present state and perspective.
Expert Opin. Drug Deliv. 4, 349–358
(2007).
12	 van Kerrebroeck P, Rezapour M, Cortesse A,
Thüroff J, Riis A, Nørgaard JP.
Desmopressin in the treatment of nocturia: a
double-blind, placebo-controlled study. Eur.
Urol. 52(1), 221–229 (2007).
13	 Sajeesh S, Sharma CP. Novel polyelectrolyte
complexes based on poly(methacrylic
acid)-bis(2-aminopropyl)poly(ethylene
glycol) for oral protein delivery. J. Biomater.
Sci. Polym. Ed. 18, 1125–1139 (2007).
The oral delivery of proteins using interpolymer polyelectrolyte complexes | Review
www.future-science.com 1629future science group
14	 Lee Y-H, Sinko PJ. Oral delivery of salmon
calcitonin. Adv. Drug Deliv. Rev. 42(3),
225–238 (2000).
15	 Kidron M, Dinh S, Menachem Y et al. A
novel per-oral insulin formulation: proof of
concept study in non-diabetic subjects.
Diabet. Med. 21, 354–357 (2004).
16	 Morçöl T, Nagappan P, Nerenbaum L,
Mitchell A, Bell SJD. Calcium
phosphate–PEG–insulin–casein (CAPIC)
particles as oral delivery systems for insulin.
Int. J. Pharm. 277, 91–97 (2004).
17	 Clement S, Dandona R, Still JG, Kosutic G.
Oral modified insulin (HIM2) in patients
with type 1 diabetes mellitus: results from a
phase I/II clinical trial. Metabolism 53(1),
54–58 (2004).
18	 des Rieux A, Ragnarsson EG, Gullberg E,
Préat V, Schneider YJ, Artursson P.
Transport of nanoparticles across an in vitro
model of the human intestinal follicle
associated epithelium. Eur. J. Pharm. Sci.
25(4–5), 455–465 (2005).
19	 Skyler JS, Krischer JP, Wolfsdorf J et al.
Effects of oral insulin in relatives of patients
with type 1 diabetes: The Diabetes
Prevention Trial – Type 1. Diabetes Care
28(5), 1068–1076 (2005).
20	 Cilek A, Celebi N, Tirnaksiz F. Lecithin-
based microemulsion of a peptide for oral
administration: preparation,
characterization, and physical stability of
the formulation. Drug Deliv. 13, 19–24
(2006).
21	 Sajeesh S, Sharma CP. Cyclodextrin-insulin
complex encapsulated polymethacrylic acid
based nanoparticles for oral insulin delivery.
Int. J. Pharm. 325, 147–154 (2006).
22	 Sarmento B, Ferreira D, Jorgensen L, van de
Weert M. Probing insulin’s secondary
structure after entrapment into alginate–
chitosan nanoparticles. Eur. J. Pharm.
Biopharm. 65, 10–17 (2007).
23	 Bayat A, Larijani B, Ahmadian S, Junginger
HE, Rafiee-Tehrani M. Preparation and
characterization of insulin nanoparticles
using chitosan and its quaternized
derivatives. Nanomed. Nanotechnol. Biol.
Med. 4, 115–120 (2008).
24	 Woitiski CB, Veiga F, Ribeiro AJ, Neufeld
RJ. Design for optimization of nanoparticles
integrating biomaterials for orally dosed
insulin. Eur. J. Pharm. Biopharm. 73, 25–33
(2009).
25	 Woitiski CB, Neufeld RJ, Veiga F, Carvalho
RA, Figueiredo IV. Pharmacological effect
of orally delivered insulin facilitated by
multilayered stable nanoparticles. Eur.
J. Pharm. Sci. 41, 556–563 (2010).
26	 Woitiski CB, Sarmento B, Carvalho RA,
Neufeld RJ, Veiga F. Facilitated nanoscale
delivery of insulin across intestinal membrane
models. Int. J. Pharm. 412, 123–131 (2011).
27	 Florence AT. Issues in oral nanoparticle drug
carrier uptake and targeting. J. Drug Target.
12(2), 65–70 (2004).
28	 Löbenberg R, Amidon GL. Modern
bioavailability, bioequivalence and
biopharmaceutics classification system. New
scientific approaches to international
regulatory standards. Eur. J. Pharm.
Biopharm. 50, 3–12 (2000).
29	 Bastian SE, Walton PE, Ballard FJ, Belford
DA. Transport of IGF-I across epithelial cell
monolayers. J. Endocrinol. 162, 361–369
(1999).
30	 Sarmento B, Ribeiro A, Veiga F, Sampaio P,
Neufeld RJ, Ferreira D. Alginate–chitosan
nanoparticles are effective for oral insulin
delivery. Pharm. Res. 24(12), 2198–2206
(2007).
31	 Agarwal V, Khan MA. Current status of the
oral delivery of insulin. Pharm. Technol. 10,
76–90 (2001).
32	 Guggi D, Kast CE, Bernkop-Schnürch A.
In vivo evaluation of an oral salmon
calcitonin-delivery system based on a
thiolated chitosan carrier matrix. Pharm. Res.
20(12), 1989–1994 (2003).
33	 Desai MA, Mutlu M, Vadgama P. A study of
macromolecular diffusion through native
porcine mucus. Experientia 48, 22–26 (1992).
34	 Norris DA, Puri N, Sinko PJ. The effect of
physical barriers and properties on the oral
absorption of particulates. Adv. Drug Deliv.
Rev. 34(2–3), 135–154 (1998).
35	 Washington N, Washington C, Wilson CG.
Physiological Pharmaceutics. Barriers to Drug
Absorption (2nd Edition). Washington N,
Washington C, Wilson CG (Eds). Taylor and
Francis, London, UK (2002).
36	 Serra L, Domenech J, Peppas NA.
Engineering design and molecular dynamics
of mucoadhesive drug delivery systems as
targeting agents. Eur. J. Pharm. Biopharm.
71(3), 519–528 (2009).
37	 Bhalodia R, Basu B, Garala K, Joshi B,
Mehta K. Buccoadhesive drug delivery
systems: a review. Int. J. Pharm. Biosci. 1(2),
1–32 (2010).
38	 Strous GJ, Dekker J. Mucin-type
glycoproteins. Crit. Rev. Biochem. Mol. Biol.
27, 57–92 (1992).
39	 Liu Z, Wang S, Hu M. Chapter 11: Oral
absorption basics: pathways, physico-
chemical and biological factors affecting
absorption. In: Developing Solid Oral Dosage
Forms: Pharmaceutical Theory and Practice.
Qiu Y, Chen Y, Liu L, Zhang GGZ (Eds).
Academic Press, London, UK (2009).
40	 Salama NN, Eddington ND, Fasano A.
Tight junction modulation and its
relationship to drug delivery. Adv. Drug
Deliv. Rev. 58(1), 15–28 (2006).
41	 Fix AJ. Oral controlled release technology for
peptides: status and future prospects. Pharm.
Res. 13, 1760–1764 (1996).
42	 Wacher VJ, Salphati L, Benet LZ. Active
secretion and enterocytic drug metabolism
barriers to drug absorption. Adv. Drug Deliv.
Rev. 46, 89–102 (2001).
43	 Jung T, Kamm W, Breitenbach A, Kaiserling
E, Xiao JX, Kissel T. Biodegradable
nanoparticles for oral delivery of peptides: is
there a role for polymers to affect mucosal
uptake? Eur. J. Pharm. Biopharm. 50(1),
147–160 (2000).
44	 Yin L, Ding J, He C, Cui L, Tang C, Yin C.
Drug permeability and mucoadhesion
properties of thiolated trimethyl chitosan
nanoparticles in oral insulin delivery.
Biomaterials 30(29), 5691–5700 (2009).
45	 van der Lubben IM, Verhoef JC, van Aelst
AC, Borchard G, Junginger HE. Chitosan
microparticles for oral vaccination:
preparation, characterization and
preliminary in vivo uptake studies in murine
Peyer’s patches. Biomaterials 22(7), 687–694
(2001).
46	 Marks DL, Gores GJ, LaRusso NF. Hepatic
processing of peptides. In: Peptide-Based
Drug Design: Controlling Transport and
Metabolism. Taylor MD, Amidon GL. (Eds).
American Chemical Society, Washington,
DC, USA (1995).
47	 Swaan PW. Recent advances in intestinal
macromolecular drug delivery via
receptor-mediated transport pathways.
Pharm. Res. 15, 826–834 (1998).
48	 Schilling RJ, Mitra AK. Degradation of
insulin by trypsin and alpha-chymotrypsin.
Pharm. Res. 8(6), 721–727 (1991).
49	 Zhou XH. Overcoming enzymatic and
absorption barriers to non-parenterally
administered protein and peptide drugs.
J. Control. Release 29, 239–252 (1994).
50	 Calceti P, Salmaso S, Walker G,
Bernkop-Schnurch A. Development and
in vivo evaluation of an oral insulin–PEG
delivery system. Eur. J. Pharm. Sci. 22,
315–323 (2004).
51	 Aoki Y, Morishita M, Asai K, Akikusa B,
Hosoda S, Takayama K. Region-dependent
role of the mucous/glycocalyx layers in
insulin permeation across rat small intestinal
membrane. Pharm. Res. 22(11), 1854–1862
(2005).
Review |Thompson & Ibie
Therapeutic Delivery (2011) 2(12)1630 future science group
52	 Thompson CJ, Tetley L, Uchegbu IF, Cheng
WP. The complexation between novel comb
shaped amphiphilic polyallylamine and
insulin – towards oral insulin delivery. Int.
J. Pharm. 376, 46–55 (2009).
53	 Thompson CJ, Tetley L, Cheng WP. The
influence of polymer architecture on the
protective effect of novel comb shaped
amphiphilic poly(allyl amine) against
in vitro enzymatic degradation of insulin –
towards oral insulin delivery. Int. J. Pharm.
383, 216–227 (2010).
54	 Yamamoto A, Taniguchi T, Rikyuu K et al.
Effects of various protease inhibitors on the
intestinal absorption and degradation of
insulin in rats. Pharm. Res. 11(10),
1496–1500 (1994).
55	 Narayani R, Rao KP. Hypoglycemic effect of
gelatin microspheres with entrapped insulin
and protease inhibitor in normal and
diabetic rats. Drug Deliv. 2(1), 29–38
(1995).
56	 Mesiha M, Plakogiannis F, Vejosoth S.
Enhanced oral absorption of insulin from
desolvated fatty acid-sodium glycocholate
emulsions. Int. J. Pharm. 111(3), 213–216
(1994).
57	 Eaimtrakarn S, Rama Prasad YV, Ohno
T et al. Absorption enhancing effect of
labrasol on the intestinal absorption of
insulin in rats. J. Drug Target. 10(3),
255–260 (2002).
58	 Hodges G, Carr EA, Hazzard RA, Carr KE.
Uptake and translocation of microparticles
in small intestine. Digest. Dis. Sci. 40(5),
967–975 (1995).
59	 Spangler RS. Insulin administration via
liposomes. Diabetes Care 13(9), 911–922
(1990).
60	 des Rieux A, Fievez V, Garinot M, Schneider
YJ, Préat V. Nanoparticles as potential oral
delivery systems of proteins and vaccines: a
mechanistic approach. J. Control. Release
116(1), 1–27 (2006).
61	 Damgé C, Michel C, Aprahamian M,
Couvreur P, Devissaguet JP. Nanocapsules as
carriers for oral peptide delivery. J. Control.
Release 13(2–3), 233–239 (1990).
62	 Damgé C, Vranckx H, Balschmidt P,
Couvreur P. Poly(alkyl cyanoacrylate)
nanospheres for oral administration of
insulin. J. Pharm. Sci. 86(12), 1403–1409
(1997).
63	 Dünnhaupt S, Barthelmes J, Hombach J,
Sakloetsakun D, Arkhipova V, Bernkop-
Schnürch A. Distribution of thiolated
mucoadhesive nanoparticles on intestinal
mucosa. Int. J. Pharm. 408(1–2), 191–199
(2011).
64	 Hartig SM, Greene RR, DasGupta J et al.
Multifunctional nanoparticulate
polyelectrolyte complexes. Pharm. Res.
24(12), 2353–2369 (2007).
65	 Sajeesh S, Bouchemal K, Sharma CP,
Vauthier C. Surface-functionalized
polymethacrylic acid based hydrogel
microparticles for oral drug delivery. Eur.
J. Pharm. Biopharm. 74, 209–218 (2010).
66	 Sajeesh S, Bouchemal K, Sharma CP,
Vauthier C. Thiol-functionalized
polymethacrylic acid-based hydrogel
microparticles for oral insulin delivery. Acta
Biomaterialia 6, 3072–3080 (2010).
67	 Qiu Y, Park K. Environment-sensitive
hydrogels for drug delivery. Adv. Drug Deliv.
Rev. 53(3), 321–339 (2001).
68	 Ramkissoon-Ganorkar C, Liu F, Baudys M,
Kim SW. Modulating insulin-release profile
from pH/thermosensitive polymeric beads
through polymer molecular weight. J. Control.
Release 59(3), 287–298 (1999).
69	 Takeuchi H, Matsui Y, Yamamoto H,
Kawashima Y. Mucoadhesive properties of
carbopol or chitosan-coated liposomes and
their effectiveness in the oral administration
of calcitonin to rats. J. Control. Release
86(2–3), 235–242 (2003).
70	 Chalasani KB, Russell-Jones GJ, Jain AK,
Diwan PV, Jain SK. Effective oral delivery of
insulin in animal models using vitamin
B12
-coated dextran nanoparticles. J. Control.
Release 122(2), 141–150 (2007).
71	 Wood KM, Stone G, Peppas NA. Wheat
germ agglutinin functionalized complexation
hydrogels for oral insulin delivery.
Biomacromolecules 9, 1293–1298 (2008).
72	 Soppimath KS, Aminabhavi TM, Kulkarni
AR, Rudzinski WE. Biodegradable polymeric
nanoparticles as drug delivery devices.
J. Control. Release 70, 1–20 (2001).
73	 Cui F, Shi K, Zhang L, Tao A, Kawashima Y.
Biodegradable nanoparticles loaded with
insulin–phospholipid complex for oral
delivery: preparation, in vitro characterization
and in vivo evaluation. J. Control. Release 114,
242–250 (2006).
74	 Fan YF, Wang YN, Fan YG, Ma JB.
Preparation of insulin nanoparticles and their
encapsulation with biodegradable
polyelectrolytes via the layer-by-layer
adsorption. Int. J. Pharm. 324, 158–167
(2006).
75	 Simon M, Behrens I, Dailey LA, Wittmar M,
Kissel T. Nanosized insulin-complexes based
on biodegradable amine-modified graft
polyesters poly[(vinyl-3-(diethylamino)-
propylcarbamate-co-(vinyl acetate)-co-(vinyl
alcohol)]–graft–poly(l-lactic acid): protection
against enzymatic degradation, interaction
with Caco-2 cell monolayers, peptide
transport and cytotoxicity. Eur. J. Pharm.
Biopharm. 66(2), 165–172 (2007).
76	 Thompson CJ, Ding C, Qu X et al. The effect
of polymer architecture on the nano
self-assemblies based on novel comb-shaped
amphiphilic poly(allylamine). Colloid Polym.
Sci. 286, 1511–1526 (2008).
77	 Cheng WP, Thompson CJ, Ryan SM, Aguirre
T, Tetley L, Brayden DJ. In vitro and in vivo
characterisation of a novel peptide delivery
system: amphiphilic polyelectrolyte–salmon
calcitonin nanocomplexes. J. Control. Release
147, 289–297 (2010).
78	 Thompson CJ, Cheng WP. Chemically
modified polyelectrolytes for intestinal
protein and peptide delivery. In: Peptide and
Protein Delivery. Chris van der Walle (Ed.).
Academic Press/Elsevier, San Diego, CA,
USA (2011).
nn	 Comprehensive perspective on the use of
modified forms of chitosan and amphiphilic
polymers in oral protein delivery.
79	 Thompson CJ, Cheng WP, Gadad P et al.
Uptake and transport of novel amphiphilic
polyelectrolyte-insulin nanocomplexes by
Caco-2 cells – towards oral insulin. Pharm.
Res. 28, 886–896 (2011).
80	 Gowthamarajan K, Kulkarni GT. Oral
insulin – fact or fiction? Resonance 8(5),
38–46 (2003).
81	 van de Weert M, Hennink WE, Jiskoot W.
Protein instability in poly(lactic-co-glycolic
acid) microparticles. Pharm. Res. 17,
1159–1167 (2000).
82	 Mao S, Bakowsky U, Jintapattankit A, Kissel
T. Self-assembled polyelectrolyte
nanocomplexes between chitosan derivatives
and insulin. J. Pharm. Sci. 95, 1035–1048
(2006).
83	 Sarmento B, Ribeiro A, Veiga F, Ferreira D,
Neufeld RJ. Insulin-loaded nanoparticles are
prepared by alginate ionotropic pre-gelation
followed by chitosan polyelectrolyte
complexation. J. Nanosci. Nanotechnol. 7,
283–2841 (2007).
84	 Shu S, Zhang X, Teng D, Wang Z, Li C.
Polyelectrolyte nanoparticles based on
water-soluble chitosan-poly(l-aspartic
acid)-polyethylene glycol for controlled
protein release. Carbohydr. Res. 344,
1197–1204 (2009).
85	 Desai MP, Labhasetwar V, Walter E, Levy RJ,
Amidon GL. The mechanism of uptake of
biodegradable microparticles in Caco-2 cells
is size dependent. Pharm. Res. 14(11),
1568–1573 (1997).
86	 Bayat A, Dorkoosh FA, Dehpour AR et al.
Nanoparticles of quaternized chitosan
The oral delivery of proteins using interpolymer polyelectrolyte complexes | Review
www.future-science.com 1631future science group
derivatives as a carrier for colon delivery of
insulin: ex vivo and in vivo studies. Int.
J. Pharm. 356(1–2), 259–266 (2008).
87	 Sarmento B, Ferreira D, Veiga F, Ribeiro A.
Characterization of insulin-loaded alginate
nanoparticles produced by ionotropic
pre-gelation through DSC and FTIR studies.
Carbohydr. Polym. 66, 1–7 (2006).
88	 Sajeesh S, Sharma CP. Novel pH responsive
polymethacrylic acid-chitosan-polyethylene
glycol nanoparticles for oral peptide delivery.
J. Biomed. Mater. Res. B Appl. Biomater. 76B,
298–305 (2006).
89	 Sun W, Mao S, Mei D, Kissel T. Self-
assembled polyelectrolyte nanocomplexes
between chitosan derivatives and enoxaparin.
Eur. J. Pharm. Biopharm. 69, 471–425
(2008).
90	 Shu S, Wang X, Zhang X, Zhang X, Wang Z,
Li C. Disulfide cross-linked biodegradable
polyelectrolyte nanoparticles for the oral
delivery of protein drugs. New J. Chem. 33,
1882–1887 (2009).
91	 Mao S, Shuai X, Unger F, Wittmar M, Xie X,
Kissel T. Synthesis, characterization and
cytotoxicity of poly(ethylene glycol)-graft-
trimethyl chitosan block copolymers.
Biomaterials 26, 6343–6356 (2005).
92	 Mao S, Germershaus O, Fischer D, Linn T,
Schnepf R, Kissel T. Uptake and transport of
PEG-grsft-trimethyl-chitosan copolymer–
insulin nanocomplexes by epithelial cells.
Pharm. Res. 22, 2058–2068 (2005).
93	 Chen F, Zhang Z, Yuan F, Qin X, Wang M,
Huang Y. In vitro and in vivo study of
N-trimethyl chitosan nanoparticles for oral
protein delivery. Int. J. Pharm. 349, 226–233
(2008).
94	 Bernkop-Schnürch A. Thiomers: a new
generation of mucoadhesive polymers. Adv.
Drug Deliv. Rev. 57, 1569–1582 (2005).
95	 Albrecht K, Bernkop-Schnürch A. Thiomers:
forms, functions and applications to
nanomedicine. Nanomedicine 2(1), 41–50
(2007).
96	 Föger F, Schmitz T, Bernkop-Schnürch A.
In vivo evaluation of an oral delivery system
for P-gp substrates based on thiolated
chitosan. Biomaterials 27(23), 4250–4255
(2006).
97	 Sakloetsakun D, Iqbal J, Millotti G, Vetter A,
Bernkop-Schnürch A. Thiolated chitosans:
influence of various sulfhydryl ligands on
permeation-enhancing and P-gp inhibitory
properties. Drug Dev. Ind. Pharm. 37(6),
648–655 (2011).
98	 Hu Y, Jiang X, Ding Y, Ge H, Yuan Y, Yang
C. Synthesis and characterization of
chitosan-poly(acrylic acid) nanoparticles.
Biomaterials 23, 3193–3201 (2002).
99	 El-Sherbiny IM. Synthesis, characterization
and metal uptake capacity of a new
carboxymethyl chitosan derivative. Eur.
Polym. J. 45, 199–210 (2009).
100	 El-Sherbiny IM. Enhanced pH-responsive
carrier system based on alginate and
chemically modified carboxymethyl chitosan
for oral delivery of protein drugs: preparation
and in vitro assessment. Carbohydr. Polym. 80,
1125–1136 (2010).
101	 El-Sherbiny IM, Abdel-Bary EM, Harding
DRK. Preparation and in vitro evaluation of
new pH-sensitive hydrogel beads for oral
delivery of protein drugs. J. Appl. Polym. Sci.
115(5), 2828–2837 (2010).
102	 El-Sherbiny IM, Elmahdy MM. Preparation,
characterization, structure and dynamics of
carboxymethyl chitosan grafted with acrylic
acid sodium salt. J. Appl. Polym. Sci. 118(4),
2134–2145 (2010).
103	 Perera G, Greindl M, Palmberger T,
Bernkop-Schnürch A. Insulin-loaded
poly(acrylic acid)-cysteine nanoparticles:
stability studies towards digestive enzymes in
the intestine. Drug Deliv. 16, 254–260
(2009).
104	 Thurow H, Geisen K. Stabilization of
dissolved proteins against denaturation at
hydrophobie interfaces. Diabetologia 27(2),
212–218 (1984).
105	 Rahman NA, Mathiowitz E. Localization of
bovine serum albumin in double-walled
microspheres. J. Control. Release 94(1),
163–175 (2004).
106	 Lowman AM, Cowans BA, Peppas NA.
Investigation of interpolymer complexation in
swollen polyelectolyte networks using
solid-state NMR spectroscopy. J. Polym. Sci. B
Polym. Phys. 38(21), 2823–2831 (2000).
107	 Lankalapalli S, Kolapalli VRM.
Polyelectrolyte complexes: a review of their
applicability in drug delivery technology.
Indian J. Pharm. Sci. 71(5), 481–487 (2009).
108	 Woitiski CB, Neufeld RJ, Ribeiro AJ, Veiga F.
Colloidal carrier integrating biomaterials for
oral insulin delivery: influence of component
formulation on physicochemical and
biological parameters. Acta Biomater. 5,
2475–2484 (2009).
109	 Jintapattanakit A, Junyaprasert VB, Mao S,
Sitterberg J, Bakowsky U, Kissel T. Peroral
delivery of insulin using chitosan derivatives:
a comparative study of polyelectrolyte
nanocomplexes and nanoparticles. Int.
J. Pharm. 342, 240–249 (2007).
110	 Dupuy B, Arien A, Minnot AP. FT-IR of
membranes made with alginate/polylysine
complexes. Variations with the mannuronic or
guluronic content of the polysaccharides.
Artif. Cells Blood Substit. Immobil. Biotechnol.
22(1), 71–82 (1994).
111	 Foss AC, Peppas NA. Investigation of the
cytotoxicity and insulin transport of
acrylic-based copolymer protein delivery
systems in contact with Caco-2 cultures. Eur.
J. Pharm. Biopharm. 57, 447–455 (2004).
112	 López JE, Peppas NA. Effect of poly (ethylene
glycol) molecular weight and microparticle
size on oral insulin delivery from P(MAA-g-
EG) microparticles. Drug Dev. Ind. Pharm.
30(5), 497–504 (2004).
113	 Besheer A, Wood KM, Peppas NA, Mäder K.
Loading and mobility of spin-labeled insulin
in physiologically responsive complexation
hydrogels intended for oral administration.
J. Control. Release 111, 73–80 (2006).
114	 Wood KM, Stone G, Peppas NA. The effect
of complexation hydrogels on insulin
transport in intestinal epithelial cell models.
Acta Biomaterialia 6, 48–56 (2010).
115	 American Diabetes Association. Standards of
medical care in diabetes – 2009. Current
criteria for the diagnosis of diabetes. Diabetes
Care 32(Suppl. 1), S6–S12 (2009).
116	 Sugano K, Kansy M, Artursson P et al.
Coexistence of passive and carrier-mediated
processes in drug transport. Nat. Rev. Drug
Discov. 9, 597–614 (2010).
117	 Bae KH, Moon CW, Lee Y, Park TG.
Intracellular delivery of heparin complexed
with chitosan-g-poly(ethylene glycol) for
inducing apoptosis. Pharm. Res. 26(1),
93–100 (2009).

More Related Content

What's hot

Protein and peptide drug delivery system
Protein and peptide drug delivery systemProtein and peptide drug delivery system
Protein and peptide drug delivery systemPriyanka Goswami
 
Protein drug delivery and gene drug delivery
Protein drug delivery and gene drug deliveryProtein drug delivery and gene drug delivery
Protein drug delivery and gene drug deliveryDr Subodh Satheesh
 
Protein And Peptide Drug Delivery System
Protein And Peptide Drug Delivery SystemProtein And Peptide Drug Delivery System
Protein And Peptide Drug Delivery SystemKushal Saha
 
Protein and peptide drug delivery system (PPDDS)
Protein and peptide drug delivery system (PPDDS)Protein and peptide drug delivery system (PPDDS)
Protein and peptide drug delivery system (PPDDS)Sagar Savale
 
sustained and controlled Drug delivery system
sustained and controlled Drug delivery  systemsustained and controlled Drug delivery  system
sustained and controlled Drug delivery systemSripriyasekar1
 
PHARMACOKINETICS AND PHARMACODYNAMICS OF BIOTECHNOLOGY DRUGS : MONOCLONAL A...
PHARMACOKINETICS AND  PHARMACODYNAMICS OF  BIOTECHNOLOGY DRUGS : MONOCLONAL A...PHARMACOKINETICS AND  PHARMACODYNAMICS OF  BIOTECHNOLOGY DRUGS : MONOCLONAL A...
PHARMACOKINETICS AND PHARMACODYNAMICS OF BIOTECHNOLOGY DRUGS : MONOCLONAL A...Sai Adiseshu
 
Protein and peptide delivery
Protein and peptide deliveryProtein and peptide delivery
Protein and peptide deliveryShruti Tyagi
 
Protein and peptide drug delivery system
Protein and peptide drug delivery systemProtein and peptide drug delivery system
Protein and peptide drug delivery systemProdipta Chakraborty
 
TARGATED DRUG DELIVERY SYSTEM.
TARGATED DRUG DELIVERY SYSTEM.TARGATED DRUG DELIVERY SYSTEM.
TARGATED DRUG DELIVERY SYSTEM.MahefuzaaAraf
 
Protein and peptide drug delivery system
Protein and peptide drug delivery systemProtein and peptide drug delivery system
Protein and peptide drug delivery systemRajat Singh
 
Marketing of Proteins and Peptide Pharmaceuticals
Marketing of Proteins and Peptide PharmaceuticalsMarketing of Proteins and Peptide Pharmaceuticals
Marketing of Proteins and Peptide Pharmaceuticalsguest6c594976
 
A REVIEW ON COLON TARGATED DRUG DELIVERY SYSTEM
A REVIEW ON COLON TARGATED DRUG DELIVERY SYSTEMA REVIEW ON COLON TARGATED DRUG DELIVERY SYSTEM
A REVIEW ON COLON TARGATED DRUG DELIVERY SYSTEMMallikarjuna Mocharlla
 
Protein and peptide delivery (monika)
Protein and peptide delivery (monika)Protein and peptide delivery (monika)
Protein and peptide delivery (monika)Monika Targhotra
 
colon targetting
colon targettingcolon targetting
colon targettingROHIT
 
Barriers to Protein and peptide drug delivery system
Barriers to Protein and peptide drug delivery system   Barriers to Protein and peptide drug delivery system
Barriers to Protein and peptide drug delivery system JaskiranKaur72
 
protein and peptide drug delivery system
protein and peptide drug delivery system protein and peptide drug delivery system
protein and peptide drug delivery system Brajesh Kumar
 
Solid Lipid Nanoparticles: A Strategy to Improve Oral Delivery of the Biophar...
Solid Lipid Nanoparticles: A Strategy to Improve Oral Delivery of the Biophar...Solid Lipid Nanoparticles: A Strategy to Improve Oral Delivery of the Biophar...
Solid Lipid Nanoparticles: A Strategy to Improve Oral Delivery of the Biophar...BRNSS Publication Hub
 
final report fathima
final report fathimafinal report fathima
final report fathimaFathima P A
 
Polymer science in drug delivery system
Polymer science in drug delivery systemPolymer science in drug delivery system
Polymer science in drug delivery systemMd. Mominul Islam
 

What's hot (20)

Protein and peptide drug delivery system
Protein and peptide drug delivery systemProtein and peptide drug delivery system
Protein and peptide drug delivery system
 
Protein
ProteinProtein
Protein
 
Protein drug delivery and gene drug delivery
Protein drug delivery and gene drug deliveryProtein drug delivery and gene drug delivery
Protein drug delivery and gene drug delivery
 
Protein And Peptide Drug Delivery System
Protein And Peptide Drug Delivery SystemProtein And Peptide Drug Delivery System
Protein And Peptide Drug Delivery System
 
Protein and peptide drug delivery system (PPDDS)
Protein and peptide drug delivery system (PPDDS)Protein and peptide drug delivery system (PPDDS)
Protein and peptide drug delivery system (PPDDS)
 
sustained and controlled Drug delivery system
sustained and controlled Drug delivery  systemsustained and controlled Drug delivery  system
sustained and controlled Drug delivery system
 
PHARMACOKINETICS AND PHARMACODYNAMICS OF BIOTECHNOLOGY DRUGS : MONOCLONAL A...
PHARMACOKINETICS AND  PHARMACODYNAMICS OF  BIOTECHNOLOGY DRUGS : MONOCLONAL A...PHARMACOKINETICS AND  PHARMACODYNAMICS OF  BIOTECHNOLOGY DRUGS : MONOCLONAL A...
PHARMACOKINETICS AND PHARMACODYNAMICS OF BIOTECHNOLOGY DRUGS : MONOCLONAL A...
 
Protein and peptide delivery
Protein and peptide deliveryProtein and peptide delivery
Protein and peptide delivery
 
Protein and peptide drug delivery system
Protein and peptide drug delivery systemProtein and peptide drug delivery system
Protein and peptide drug delivery system
 
TARGATED DRUG DELIVERY SYSTEM.
TARGATED DRUG DELIVERY SYSTEM.TARGATED DRUG DELIVERY SYSTEM.
TARGATED DRUG DELIVERY SYSTEM.
 
Protein and peptide drug delivery system
Protein and peptide drug delivery systemProtein and peptide drug delivery system
Protein and peptide drug delivery system
 
Marketing of Proteins and Peptide Pharmaceuticals
Marketing of Proteins and Peptide PharmaceuticalsMarketing of Proteins and Peptide Pharmaceuticals
Marketing of Proteins and Peptide Pharmaceuticals
 
A REVIEW ON COLON TARGATED DRUG DELIVERY SYSTEM
A REVIEW ON COLON TARGATED DRUG DELIVERY SYSTEMA REVIEW ON COLON TARGATED DRUG DELIVERY SYSTEM
A REVIEW ON COLON TARGATED DRUG DELIVERY SYSTEM
 
Protein and peptide delivery (monika)
Protein and peptide delivery (monika)Protein and peptide delivery (monika)
Protein and peptide delivery (monika)
 
colon targetting
colon targettingcolon targetting
colon targetting
 
Barriers to Protein and peptide drug delivery system
Barriers to Protein and peptide drug delivery system   Barriers to Protein and peptide drug delivery system
Barriers to Protein and peptide drug delivery system
 
protein and peptide drug delivery system
protein and peptide drug delivery system protein and peptide drug delivery system
protein and peptide drug delivery system
 
Solid Lipid Nanoparticles: A Strategy to Improve Oral Delivery of the Biophar...
Solid Lipid Nanoparticles: A Strategy to Improve Oral Delivery of the Biophar...Solid Lipid Nanoparticles: A Strategy to Improve Oral Delivery of the Biophar...
Solid Lipid Nanoparticles: A Strategy to Improve Oral Delivery of the Biophar...
 
final report fathima
final report fathimafinal report fathima
final report fathima
 
Polymer science in drug delivery system
Polymer science in drug delivery systemPolymer science in drug delivery system
Polymer science in drug delivery system
 

Similar to The oral delivery of proteins using interpolymer polyelectrolyte complexes

Practical consideration of protien and peptides
Practical consideration of protien and peptidesPractical consideration of protien and peptides
Practical consideration of protien and peptidesSuchandra03
 
Novel Nanotechnology-based Drug Delivery Systems.pdf
Novel Nanotechnology-based Drug Delivery Systems.pdfNovel Nanotechnology-based Drug Delivery Systems.pdf
Novel Nanotechnology-based Drug Delivery Systems.pdfDoriaFang
 
Protein and Peptide Drug Delivery Oral Approaches.pptx
Protein and Peptide Drug Delivery Oral Approaches.pptxProtein and Peptide Drug Delivery Oral Approaches.pptx
Protein and Peptide Drug Delivery Oral Approaches.pptxjacksgamer
 
colon drug delivery systems
colon drug delivery systemscolon drug delivery systems
colon drug delivery systemsMaedeh Haedi
 
Nanotechnology based Oral Delivery of Insulin – A Retrospect
Nanotechnology based Oral Delivery of Insulin – A RetrospectNanotechnology based Oral Delivery of Insulin – A Retrospect
Nanotechnology based Oral Delivery of Insulin – A Retrospectpharmaindexing
 
Dm e obesidade manipulação da flora intestinal
Dm e obesidade manipulação da flora intestinal Dm e obesidade manipulação da flora intestinal
Dm e obesidade manipulação da flora intestinal Ruy Pantoja
 
Pharmacokinetics and pharmacodynamics of Biotechnological drugs-
 Pharmacokinetics and pharmacodynamics of Biotechnological drugs- Pharmacokinetics and pharmacodynamics of Biotechnological drugs-
Pharmacokinetics and pharmacodynamics of Biotechnological drugs-SnehalTidke
 
Protein and Peptides .ppt
Protein and Peptides .pptProtein and Peptides .ppt
Protein and Peptides .pptGirdhar Sahu
 
Colon drug delivery system ppt
Colon drug delivery system pptColon drug delivery system ppt
Colon drug delivery system pptMujeeb Rehman
 
j.ejpb.2009.06.001.pdf
j.ejpb.2009.06.001.pdfj.ejpb.2009.06.001.pdf
j.ejpb.2009.06.001.pdfbihonegn1
 
Pharmacokinetics and pharmacodynamics of biotechnological products
Pharmacokinetics  and  pharmacodynamics  of  biotechnological  productsPharmacokinetics  and  pharmacodynamics  of  biotechnological  products
Pharmacokinetics and pharmacodynamics of biotechnological productsPV. Viji
 
Colon Targeting by Novel Drug Delivery Drug System: Microsphere a Review Report
Colon Targeting by Novel Drug Delivery Drug System: Microsphere a Review ReportColon Targeting by Novel Drug Delivery Drug System: Microsphere a Review Report
Colon Targeting by Novel Drug Delivery Drug System: Microsphere a Review ReportBRNSSPublicationHubI
 
Treatment of familial mediterranean fever: colchicine and beyond
Treatment of familial mediterranean fever: colchicine and beyondTreatment of familial mediterranean fever: colchicine and beyond
Treatment of familial mediterranean fever: colchicine and beyondJosé Luis Moreno Garvayo
 
PHARMACOKINETICS AND PHARMACODYNAMICS OF BIOTECHNOLOGY DRUGS.pptx
PHARMACOKINETICS AND PHARMACODYNAMICS OF BIOTECHNOLOGY DRUGS.pptxPHARMACOKINETICS AND PHARMACODYNAMICS OF BIOTECHNOLOGY DRUGS.pptx
PHARMACOKINETICS AND PHARMACODYNAMICS OF BIOTECHNOLOGY DRUGS.pptxShubhamAhir1
 
Drug-MembraneInteractionsSignificanceforMedicinalChemistry.pdf
Drug-MembraneInteractionsSignificanceforMedicinalChemistry.pdfDrug-MembraneInteractionsSignificanceforMedicinalChemistry.pdf
Drug-MembraneInteractionsSignificanceforMedicinalChemistry.pdfVitalis Mbuya
 
Colon specific drug delivery system approaches and application
Colon specific drug delivery system approaches and applicationColon specific drug delivery system approaches and application
Colon specific drug delivery system approaches and applicationTukaram Patil
 

Similar to The oral delivery of proteins using interpolymer polyelectrolyte complexes (20)

Practical consideration of protien and peptides
Practical consideration of protien and peptidesPractical consideration of protien and peptides
Practical consideration of protien and peptides
 
Novel Nanotechnology-based Drug Delivery Systems.pdf
Novel Nanotechnology-based Drug Delivery Systems.pdfNovel Nanotechnology-based Drug Delivery Systems.pdf
Novel Nanotechnology-based Drug Delivery Systems.pdf
 
Protein and Peptide Drug Delivery Oral Approaches.pptx
Protein and Peptide Drug Delivery Oral Approaches.pptxProtein and Peptide Drug Delivery Oral Approaches.pptx
Protein and Peptide Drug Delivery Oral Approaches.pptx
 
proteinous drug delivery.pptx
proteinous drug delivery.pptxproteinous drug delivery.pptx
proteinous drug delivery.pptx
 
colon drug delivery systems
colon drug delivery systemscolon drug delivery systems
colon drug delivery systems
 
Nanotechnology based Oral Delivery of Insulin – A Retrospect
Nanotechnology based Oral Delivery of Insulin – A RetrospectNanotechnology based Oral Delivery of Insulin – A Retrospect
Nanotechnology based Oral Delivery of Insulin – A Retrospect
 
Dm e obesidade manipulação da flora intestinal
Dm e obesidade manipulação da flora intestinal Dm e obesidade manipulação da flora intestinal
Dm e obesidade manipulação da flora intestinal
 
Pharmacokinetics and pharmacodynamics of Biotechnological drugs-
 Pharmacokinetics and pharmacodynamics of Biotechnological drugs- Pharmacokinetics and pharmacodynamics of Biotechnological drugs-
Pharmacokinetics and pharmacodynamics of Biotechnological drugs-
 
Article 031
Article 031Article 031
Article 031
 
Protein and Peptides .ppt
Protein and Peptides .pptProtein and Peptides .ppt
Protein and Peptides .ppt
 
Colon drug delivery system ppt
Colon drug delivery system pptColon drug delivery system ppt
Colon drug delivery system ppt
 
j.ejpb.2009.06.001.pdf
j.ejpb.2009.06.001.pdfj.ejpb.2009.06.001.pdf
j.ejpb.2009.06.001.pdf
 
Pharmacokinetics and pharmacodynamics of biotechnological products
Pharmacokinetics  and  pharmacodynamics  of  biotechnological  productsPharmacokinetics  and  pharmacodynamics  of  biotechnological  products
Pharmacokinetics and pharmacodynamics of biotechnological products
 
Colon Targeting by Novel Drug Delivery Drug System: Microsphere a Review Report
Colon Targeting by Novel Drug Delivery Drug System: Microsphere a Review ReportColon Targeting by Novel Drug Delivery Drug System: Microsphere a Review Report
Colon Targeting by Novel Drug Delivery Drug System: Microsphere a Review Report
 
Treatment of familial mediterranean fever: colchicine and beyond
Treatment of familial mediterranean fever: colchicine and beyondTreatment of familial mediterranean fever: colchicine and beyond
Treatment of familial mediterranean fever: colchicine and beyond
 
PHARMACOKINETICS AND PHARMACODYNAMICS OF BIOTECHNOLOGY DRUGS.pptx
PHARMACOKINETICS AND PHARMACODYNAMICS OF BIOTECHNOLOGY DRUGS.pptxPHARMACOKINETICS AND PHARMACODYNAMICS OF BIOTECHNOLOGY DRUGS.pptx
PHARMACOKINETICS AND PHARMACODYNAMICS OF BIOTECHNOLOGY DRUGS.pptx
 
Drug-MembraneInteractionsSignificanceforMedicinalChemistry.pdf
Drug-MembraneInteractionsSignificanceforMedicinalChemistry.pdfDrug-MembraneInteractionsSignificanceforMedicinalChemistry.pdf
Drug-MembraneInteractionsSignificanceforMedicinalChemistry.pdf
 
Drug delivery to respiratory system
Drug delivery to respiratory systemDrug delivery to respiratory system
Drug delivery to respiratory system
 
Colon specific drug delivery system approaches and application
Colon specific drug delivery system approaches and applicationColon specific drug delivery system approaches and application
Colon specific drug delivery system approaches and application
 
Colon targeting
Colon targetingColon targeting
Colon targeting
 

The oral delivery of proteins using interpolymer polyelectrolyte complexes

  • 1. 1611ISSN 2041-5990Therapeutic Delivery (2011) 2(12), 1611–163110.4155/TDE.11.131 © 2011 Future Science Ltd Review Special Focus: Oral Delivery of Macromolecules Biopharmaceuticals, especially proteins, have become increasingly popular and relevant in the treatment of a number of disease states due to their high specificity and activity at low concentrations in comparison to traditional small-molecule therapies [1–6]. However, most proteins are still delivered parenterally rather than orally [1,2,5,7]. This is due to a number of factors including their structural complexity and sensitivity to degradation, as well as the numer- ous barriers to oral absorption [1–3,7–10]. When administered orally, the bioavailability of most proteins is virtually zero [5], exceptions being the hydrophobic immunosuppressive cyclospo- rin A [10,11] and the analogue of vasopressin, des- mopressin acetate [1,12]. It is thought that non- parenteral administration will increase patient compliance, which may be particularly pertinent in the case for patients who have a phobia of needles [1,5,10]. Increasing the oral absorption of proteins, therefore, has long been sought [1,3,5–7,9,10,13] and a number of oral protein technologies are already undergoing in vivo testing with some promis- ing results [5,6]. These include Eligen®  [14,15], BioOral®   [16], HIM2  [1,7,17], Oraldel®   [18], AI-401 [19], Macrulin®  [20] and Orasome®  [6]. The oral delivery of insulin is the main driver for the development of these technologies given its high (and increasing) level of clinical use in comparison to other proteins [4,5,10]. It is thought that oral insulin delivery will be advantageous not only in potential improve- ments in patient compliance, but also avoid- ing the peripheral hyperinsulinemic effects of parenteral delivery and replicating endogenous insulin hepatic delivery  [4,10,21–25] as well as potentially providing a more sustained control over blood glucose levels compared with current parenteral delivery [26]. Insulin has a plasma half- life of 6 min, and is cleared within 10–15 min, which means that any delivery system that pro- vides controlled release would be advantageous in terms of maintaining glycemic control for extended periods [4]. However, it is also possible that the delivery of proteins to non-endogenous sites may also have unwanted effects such as insulin-induced gastroparesis [27]. None of these delivery systems are yet widely available to the public [6]. This is due to a number of factors including: lack of reproducible biologi- cal effects and in vivo toxicity; toxic effects of cer- tain excipients on the protein and patient; produc- tion and formulation costs; deleterious effects of formulation on protein conformation/structure; and the two main barriers to oral protein delivery of facilitating protein absorption and protecting them from degradation in the gastrointestinal (GI) tract [1,6,7,27]. Barriers to oral protein delivery Due to the mainly hydrophilic nature of proteins and their large size and relatively high-molecular weights, oral absorption is difficult to achieve [6,7]. Most proteins are classified as Class III drugs (Biopharmaceutics Classification System), in that they are water soluble, but poorly soluble in lipid and so poorly absorbed (i.e., they have low log P values) [4,28]. (The exception being cyclosporin A, which is Biopharmaceutics Classification System Class IV.) For proteins to be absorbed orally they need to be taken up by specific receptors in the GI The oral delivery of proteins using interpolymer polyelectrolyte complexes In spite of the numerous barriers inherent in the oral delivery of therapeutically active proteins, research into the development of functional protein-delivery systems is still intense. The effectiveness of such oral protein-delivery systems depend on their ability to protect the incorporated protein from proteolytic degradation in the GI tract and enhance its intestinal absorption without significantly compromising the bioactivity of the protein. Among these delivery systems are polyelectrolyte complexes (PECs) which are composed of polyelectrolyte polymers complexed with a protein via coulombic and other interactions. This review will focus on the current status of PECs with a particular emphasis on the potential and limitations of multi- or inter-polymer PECs used to facilitate oral protein delivery. Colin Thompson* & Chidinma Ibie Robert Gordon University, School of Pharmacy and Life Sciences, Aberdeen, AB10 1FR, UK *Author for correspondence: Tel.: +44 1224 262 561 Fax: +44 1224 262 555 E-mail: c.thompson@rgu.ac.uk For reprint orders, please contact reprints@future-science.com
  • 2. Review |Thompson & Ibie Therapeutic Delivery (2011) 2(12)1612 future science group epithelium (e.g., IGF-I [29] and insulin [4,27,30]) or via endocytosis [4,31]. However, uptake via these routes does not lead to pharmacological effects in vivo due to simultaneous protein metabolism. Therefore attempts have been made to facilitate absorption and protect against metabolism by use of different delivery systems. Such delivery systems have a number of barriers to overcome in terms of the GI tract including: n Mucus layer on epithelial cells; n Unstirred water layer on epithelial cells; n Low permeability of cell membranes and GI tract to macromolecules; n The presence of efflux pumps in cell membranes; n Tight junctions between epithelial cells; n Chemical and enzymatic degradation/metab­ olism of proteins in GI fluid and on the brush borders and inside epithelial cells. The mucus layer on epithelial cells is continuously replenished every 12–24 h [32] and is a physical diffusion barrier to drug absorption, particularly macromolecules [1,5,33,34]. It can act to prevent absorption due to the adhesive nature of mucin (the major component of mucus) causing pro- teins to be trapped in the mucus layer and there- fore preventing their uptake/passage between cells  [1,35–37]. Amphiphilic structures such as proteins may also interact with mucin via non- specific hydrophobic interactions such as those seen with bacteria and surfactants [34], further reducing their ability to be absorbed. Cationic proteins may also electrostatically interact with the negatively charged groups on the side chains of mucin [34,38], limiting their uptake. The presence of an unstirred water layer on epithelial cells is less significant in absorption, but still can play a limiting role, particularly for hydrophobic drugs. However, in the case of most proteins they will readily pass across this layer due to their hydrophilic nature and their absorp- tion is mainly limited by their inability to diffuse through lipid cell membranes [1]. There are many potential mechanisms of drug absorption via either the paracellular or transcellular routes (Figure 1) [1,3,4,34,39]. The paracellular route is attractive for proteins due to the lack of enzymatic activity in the inter- cellular spaces [1,4,8,9]. However, access to intercel- lular spaces is limited by tight junctions between cells. The integrity of these tight junctions is maintained by a number of transmembrane (occludin and claudin) and cytoplasmic proteins (a family of zona occluding proteins, e.g., ZO-1) and when these junctions are closed only very small molecules, for example, water, may pass through them [4,10,40]. Also found within the tight junctions are cadherin glycoproteins, which rely on maintenance of physiological concentrations of calcium to preserve tight-junction integrity [4,10]. Therefore, paracellular transport is not normally an option for macromolecules. As stated above, the junctions found between cells only allow for the passage of very small hydrophilic molecules [4,5]. Molecules of 10–50 Å have been shown to pass between cells, although not usually more than 11.5 Å, due to the gap between cells being less than a nanometer [3,10,34]. There is also a limit in molecular weight of 200 g/mol via the paracellu- lar route, much smaller than most peptides [1,7,39]. Therefore, for macromolecules to have a chance to pass through tight junctions, they need to be opened to quite a significant extent [4,10]. Similarly, although a large number of small- drug molecules are absorbed via transcellular passive diffusion, this is not possible for pro- teins given their hydrophilicity, low log P values, high-molecular weights and charged nature, which mean they have very low absorption dif- fusion coefficients [1,3,4,7,35,41]. The presence of efflux pumps (e.g., P-glycoprotein [P-gp] and multidrug resistance protein) in GI epithelial cell membranes may also result in any protein that is absorbed into cells being removed before transport to the systemic circulation can be com- pleted [1,3,4,9,39,42]. However, this tends to be more of a problem with small molecules [39]. An excep- tion to this being cyclosporin A, which has been shown to be a substrate for P-gp [1,9,42]. Peyer’s patches are also be involved in trans­ cellular absorption. These are lymphoid nodules found primarily in the ileum of the small intes- tines and are covered by a different type of epi- thelial cells than the rest of the GI tract, called microfold or M cells [4]. The patches are predom­ inantly involved in mediating immune response to pathogens in the GI tract and transporting absorbed substances to the lymph nodes and spleen. However, their structure (underdeveloped villi and mucus layer) allows for nano-sized mate- rials to easily pass through, usually of 200 nm or less in diameter [4,43,44], while materials of 3000–5000 nm or more in diameter are trapped within the patches [4,45]. However, they const­ itute a limited part of the GI tract epithelium as a whole, which limits their role in absorption of therapeutic compounds [4,43,45].
  • 3. The oral delivery of proteins using interpolymer polyelectrolyte complexes | Review www.future-science.com 1613future science group Carrier-mediated uptake of some proteins is feasible due to a number of therapeutic proteins being of endogenous origin. These proteins can act as substrates for receptors found in the GI tract and, in doing so, are absorbed into epithe- lial cells [46,47]. However, this is only the case for a limited number of proteins and these pro- teins are subsequently enzymatically degraded within cells, prior to passage into the systemic circulation [3,4]. Further protein breakdown can occur in the GI fluid and this is primarily achieved by endo- peptidases: pepsin (stomach), trypsin, a-chymo- trypsin and elastase (intestines) [1,4,5,7,8,10,48–50]. Enzymatic degradation can take a number of forms with the hydrolysis of the peptide bonds within proteins being the most common, as well as oxidation, deamidation and phosphoryla- tion [1,9]. Each enzyme has its own target site(s) on an individual protein (Table  1)  [7,10,48,50], which makes simultaneous protection against all GI enzymes difficult to achieve. A further site for protein breakdown is in the brush-border membrane of the GI tract, where degradation can occur by exopeptidases such as the amino- and carboxy-peptidases [3,4,8,9,51]. Changes in pH can also play a role in protein degradation as well as stability and conforma- tion [1]. The low pH in the stomach (pH 1–3) can induce degradation of peptide bonds as it is crucial in the activation of pepsin from pepsino- gen [7,9]. In addition, the change in pH along the GI tract can have profound effects on pro- tein conformation, hence, their activity as well as susceptibility to enzymatic breakdown [1]. As proteins are amphoteric, their degree of ioniza- tion and so aqueous solubility will vary consider- ably with pH and their charge will alter between anionic and cationic around their isoelectric point (pI). Changes in charge on proteins will alter their ability to form ionic and hydrogen bonds, thus altering their secondary, tertiary and quaternary structures. This change in confor- mation can render them inactive and also more susceptible to degradation, as further enzymatic target sites may become exposed [1,3,52,53]. As is evident, oral administration presents numerous challenges as a route for protein Basolateral (B) side Paracellular permeation (passive) Passive transcellular transport Carrier-mediated efflux Carried-mediated uptake Tight junction Absorptive direction (A to B) Excretive direction (B to A) Concentration gradient Carrier-mediated uptake Carrier-mediated efflux Passive transcellular transport ABL permeation (passive) Apical (A) side Intestinal membrane Figure 1. GI tract epithelium physiology and transport routes across the GI tract. ABL: Aqueous boundary layer. Reproduced with permission from [116]. © Nature Publishing Group.
  • 4. Review |Thompson & Ibie Therapeutic Delivery (2011) 2(12)1614 future science group delivery. Therefore a number of different technol- ogies to improve oral bioavailability of proteins have been developed. Technologies to facilitate oral protein delivery Early steps in improving oral absorption involved co-administration/co-formulation of proteins with protease inhibitors, such as sodium glyco- cholate, aprotinin, bacitracin, soybean trypsin inhibitor, camostat mesilate [54,55] and absorp- tion enhancers such as bile salts, salicylates, sur- factants and chelating agents [56,57]. However, these were found to have deleterious effects on the GI tract wall, on normal digestion and the pancreas itself [1,3–7,9,10]. Subsequently, encapsulation of proteins in particulate structures of various sizes and mor- phologies has become one of the most widely used approaches to facilitate oral delivery [43,58– 62]. This concept is usually aimed at modifying the pharmacokinetic profile and physico­chemical characteristics of the resultant construct, as well as providing an outer polymeric membrane for the entrapped protein, which can serve as a protective coat from proteolytic enzymes [61,62]. Particulate carriers having different functional- ities can be tailor-made by careful selection of parent polymers that possess desirable proper- ties such as enzyme inhibition, mucoadhesion, absorption-enhancing or pH/environmentally sensitive properties [63–69]. These delivery sys- tems may also be functionalized by attachment of various receptor-recognizable/targeting ligands to preformed particles  [64,70,71]. Particulate carriers including liposomes [3,6,72], micropar- ticles [3,6,7,72], and polymeric colloids including polyelectrolyte nanoparticulates  [3,52,53,73–79], have been developed with varying degrees of success. For oral protein delivery, the use of liposomes (tiny, vesicular self-assemblies consisting of hydrophobic and hydrophilic domains formed when phospholipids are dispersed in aqueous media) is challenging due to the inability of lipid- based systems to entrap sufficient quantities of hydrophilic drugs resulting in problems such as low GI tract stability, low drug-loading capacity and leakage of entrapped drug [3,6,72,80]. Moreover, the fabrication of micropart­ icles often requires the use of organic solvents, high temperatures/pressures and hydropho- bic polymers, which can denature proteins. In addition, their use in oral protein delivery has led to highly variable levels of protein in the circulation [1,7,81–84]. In contrast, polymeric colloidal systems are thought to be advantageous, as they can be more quickly and completely absorbed due to their large surface areas and nano-range sizes, as well as protect proteins from degradation and improve their in vivo stability [3,6,52,75,85,86]. However, there can be a lack of control of protein release and levels of protein loading can be low [3,53]. Polyelectrolyte complexes (PECs) are nor- mally nanosized particulates that are composed of polyelectrolyte polymers and a protein. Their formation is driven by coulombic interactions, that is, electrostatic interactions and hydrogen bonding, as well as hydrophobic interactions in some cases (Figure 2) [52,76,78,82,86,87]. They can also involve prior copolymerization of oppo- sitely charged polymers, which can then inter- act electrostatically with each other and with a protein [65,66]. PECs are able to form spontaneously (self assemble) under benign conditions and can have high levels of protein loading, which is advanta- geous in comparison to some of the other delivery technologies mentioned above [78,82,84,86,88–90]. Key Terms Polyelectrolyte complex: Polyelectrolyte polymer–protein complex whose formation is primarily driven by coulombic forces. Chitosan: Cationic polysaccharide derived from crustaceans used in a large number of polyelectrolyte complexes. Thiomers: Polymers that contain moieties with thiol groups. Caco-2: Colorectal cancer cell line, one of the most common in vitro cell models for the GI tract. Quaternary ammonium moiety: Quaternary amine, commonly produced by the trimethylation of the primary amines of chitosan. Table 1. Various GI tract enzymes and their respective protein target sites. Enzyme Target sites Endopeptidases Nonterminal residues Trypsin After arginine, lysine unless followed by proline a-chymotrypsin After phenylalanine, tyrosine and tryptophan unless followed by proline; also after asparagine, leucine, methionine and histidine Pepsin Before leucine, phenylalanine, tyrosine and tryptophan unless preceded by proline; can also cleave at many other sites Elastase After glycine, alanine, serine and valine unless followed by proline Exopeptidases Terminal residues Carboxypeptidase A Aromatic/aliphatic hydrocarbon chain residues at the carboxy-terminal Carboxypeptidase B Cationic residues at carboxy-terminal Aminopeptidase Residues at amino-terminal
  • 5. The oral delivery of proteins using interpolymer polyelectrolyte complexes | Review www.future-science.com 1615future science group In addition, they can also fulfill the two main criteria for oral protein-delivery systems in that they have the potential to both facilitate absorp- tion and offer a degree of protection against degradation in the gut. „„ Chitosan PECs Chitosan is a weakly cationic polysaccha- ride derived from the shells of crustaceans. It has been used as a mucoadhesive, release- rate controlling excipient in a number of for- mulations due to its nontoxic, biocompatible nature [22,23,78,82–84,88–91]. It has also been shown to promote tight-junction opening, facilitating paracellular transport of active molecules [78,89,91]. It can do this by electrostatic interactions with tight-junction proteins to reversibly alter their morphology and increase the permeability of intercellular spaces [78,83,92,93]. Its highly positive charge can also facilitate cell uptake due to inter- actions with anionic cell membranes to a greater extent than anionic and non-ionic polymers [23]. However, its weakly basic nature means that it is fairly insoluble above pH 6.5 (pKa  = 6.5) [22,82– 84,88–91]. Therefore its ability to gel, to facili- tate junction opening and adhere to mucus membranes is limited in the intestines and colon  [2,23,78,89,92]. To overcome these prob- lems, derivatives of chitosan have been devel- oped, which have pH-independent charges; for example, trimethylated chitosan (TMC), as well as those that form stronger (of up to 140-fold), covalent mucoadhesive bonds (e.g., thiolated chi- tosan [thiomers]), in the intestines/colon than other forms of chitosan [2,78,89,91–95]. Chitosan and its derivatives have been shown to offer some protection against enzymatic break- down, particularly thiolated versions, which are vital for the function of some enzymes [78,94,95]. They are able to do this as thiol groups can chelate metal cations, such as zinc, which are responsible for enzyme function. These modified chitosans are also capable of facilitating protein absorption via both paracellu- larandtranscellularroutes.Thiolatedchitosanhas been shown to increase membrane permeability, possible due to an interaction with protein tyro- sine phosphatases (PTPs). PTPs de­phosphorylate the tight-junction protein occludin causing tight junctions to close. It is postulated that by forming disulfides with PTPs, thiolated chitosan is able to prevent junction closing for longer by prevent- ing dephosphorylation [65,95]. The mechanism by which thiomers first open tight junctions is as yet unclear. However, it is likely in part due to chelation of metal cations, such as calcium, which are responsible for maintenance of tight-junction integrity [65,66]. Nonetheless, the limited inter- cellular space available for paracellular transport means that protein transport via this route can be negligible for macromolecules regardless of the extent of junction opening [92]. Thiolated chitosan may even be able to inhibit P-gp efflux pumps [95] although, to date, this has only been shown with model molecules (e.g., rhodamine 123, which is 380 g/mol) and not a protein [96,97]. TMC modified by covalent attachment of poly(ethylene glycol)(PEG)-ylated has been shown to facilitate adsorptive endocytosis of insu- lin in a Caco-2 cell model [92]. Uptake was shown to linearly increase with polymer concentration andincubationtimeshowingthatthepolymerwas able to promote absorption for up to 4 h in vitro. Indeed, insulin uptake increased from approxi- mately 20 µg/mg protein (insulin only control) to 60 µg/mg protein when insulin was complexed with PEGylated TMC (100,000 g/mol) at physio­ logical pH. The promotion of cellular uptake is thought to be due to electrostatic interaction between the quaternary ammonium moieties on TMC and cell membranes, which facilitates endocytosis [92]. Another way to overcome the problems inher- ent with using chitosan alone in its native state in oral delivery systems has been to combine it with other polyelectrolytes to form interpolymer PECs. „„ Interpolymer PECs PECs of polyanions, such as alginate, and proteins, have also been produced. However, they suffer from similar problems to chitosan-­ based PECs, in particular pH-dependent ionization and so solubility and mucoad­ hesion  [2,4,22,24–26,30,66,83,84,88,90,98–103]. Therefore, modified forms, such as thiomers, Self-assembly <200 nm Polyelectrolyte complex micelle – – – – – – – – – – + + + + + + + + Figure 2. Formation of a polyelectrolyte complex. Reproduced with permission from [117]. © Springer.
  • 6. Review |Thompson & Ibie Therapeutic Delivery (2011) 2(12)1616 future science group interpolymer complexes of polyanions and poly- cations have been produced to try and mitigate the problems of using polyanions or polycations alone [2,84,90,103]. Electrostatically induced intermolecular com- plexation occurring between polycationic and polyanionic molecules present in aqueous solu- tion leads to the formation of discrete interpoly- mer complexes, which have been shown to be suitable delivery systems for oral proteins [104,105]. Interpolymer complexes are often prepared by mixing aqueous solutions containing oppositely charged polymers at a specific pH resulting in the formation of interpolymer PECs exhibit- ing varied morphologies and functionalities as dictated by their chemical composition [106,107]. However, tailoring the use of these interpoly- electrolyte structures to surmounting the bar- riers inherent in oral protein delivery involves careful selection of complementary polycations and polyanions, which interact optimally pro- viding a robust network suitable for the incor- poration and protection of proteins intended for oral administration. The concept of interpolymer complexation also conveniently creates a platform where unique properties of different polymers that have been shown to be capable of improving oral protein bioavailability can be imparted into a single delivery system, thereby potentiating the efficacy of the resultant construct. Drug release from the interpolyelectrolyte network is usually mediated by decomplexation of the PECs at pH conditions that reduce the electrostatic forces between the interacting polyelectrolytes leading to dissociation of the complex and release of the incorporated protein. Invariably, many research groups have exploited the versatility of chitosan by complex- ing it/its derivatives with various polyanions of which the majority are anionic polysaccharides of natural origin such as glucomannans and mucopolysaccharides as well as other natural and synthetic polymers [2,22,30,66,83,84,87,88,90]. As well as forming interpolymer PECs by mixing chitosan and polyanions, these can also be formed after prior copolymerization of poly- electrolytes. This can counteract some of the potential problems associated with PECs that rely solely on coulombic interactions. These include counterion effects and the high solubility of chi- tosan in the stomach and of polyanions in the lower GI tract, which can result in partial PEC collapse and subsequent protein dumping, limit- ing protection from degradation and promotion of absorption [2,24–26,65,66,78,82,89,99–102,108,109]. Alginate–chitosan PECs Alginate is an anionic polysaccharide extracted from bacteria or algae and is made of residues of mannuronicandguluronicacidofvaryingratios.It exists as salt forms due to cooperative binding with cations, for example, sodium alginate. Alginates form gels when mixed with divalent cations (at certain stoichiometric ratios), such as calcium, due to their exchange with sodium, which then results in inter-chain associations being formed (dimerization of the alginate) primarily between guluronic residues [2,22,24,30,83,100,108]. Alginates are already used in a number of foods as a thicken- ing agent/stabilizer and have been shown to be both biocompatible and biodegradable. They are approved for human use by the US FDA. In terms of their use as oral delivery vehicles, alginates have been shown to form strong mucoad- hesive bonds when compared with other polymers suchaschitosan [2,22,30,83].However,alginateshave a pH-solubility problem in that although the gels they form release very little protein in the stomach, once they reach the lower GI tract they can rapidly dissolve preventing controlled release of protein as well as providing very limited protection from degradation. In addition, protein electrostatic interaction with alginates can cause detrimental changes in protein conformation and loading of proteins can be low [2]. Therefore, attempts have been made to use alginate in concert with other polymers to overcome these problems. One such combination involves the use of algi- nate with chitosan, which forms electrostatic inter- actions between the carboxyl groups of alginate and amino groups of chitosan [2,22,24,30,83,100,108]. This predominantly occurs between carboxyls on mannuronic rather than guluronic residues of alginate and aminos on chitosan, probably due to the lower pKa of mannuronic residues (3.38 com- pared with 3.65 for guluronic) [110]. The potential advantage of this combination is that it allows the formation of a pH-sensitive gel network, which can complex/load a protein for delivery. This network does not entirely dissolve in the stom- ach or intestines due to the presence of alginate and chitosan, respectively. It can still swell in the intestines, as alginate does on its own, but due to the presence of chitosan does not completely dis- solve there and so may allow for mucoadhesion and controlled release to take place. These PECs are produced by first forming nuclei of alginate complexed with calcium and a protein followed by addition of chitosan to form PEC nanoparticles via crosslinking between chito- san and alginate, as well as between each polymer Key Term Interpolymer complex: Polyelectrolyte complex consisting of two or more oppositely charged polymers.
  • 7. The oral delivery of proteins using interpolymer polyelectrolyte complexes | Review www.future-science.com 1617future science group and the protein (Figure 3) [22,30,83]. Pregelation with calcium is thought to increase PEC stability by providing an additional level of crosslinking between alginate chains. Crucially these PECs have been shown, via derivatized Fourier trans- form IR and circular dichronism spectra, to have no significant effect on insulin secondary struc- ture and that insulin is stable in the PECs for up to one month, suggesting that the protein is still in its native conformation after complexation [22]. However, the limitation of these systems is that polymer–polymer and polymer–insulin interac- tions are also still pH dependent with insulin association efficiency falling from 81 to 60% of total insulin used in nanoparticle manufacture when PEC solution pH was increased from 4.7 to 5.2 due to a reduction in charge density of chitosan and insulin (pI = 5.3) [22]. This can lead to low protein loading and cause burst release at both low and high pH [2,7]. Indeed it was found that alginate–chitosan PECs showed burst release of insulin between 40–60% within 5 min in sim- ulated gastric (pH 1.2) and simulated intestinal fluid (pH 6.8) [22]. This was due to the variations in charge density on chitosan, alginate and insu- lin at these pH values resulting in weak interac- tions between the three components of the PEC and so uncontrolled release of insulin. In addition to this, some precipitation of alg­ inate and chitosan can still occur over the pH range in the GI tract resulting in large increases in particle size, which in turn could limit cellular interactions and uptake due to the reduction in PEC surface area. Sarmento and colleagues found that decreasing pH from 4.7 to 4.2 resulted in particle size increasing from 797 to 2858 nm, as the reduction in pH reduces the ionization of alginate (pKa between 3.3 and 3.7, depending on ratio of mannuronic to guluronic residues) [87]. Regardless of these problems, however, algi- nate–chitosan PECs have been shown to be effec- tive at reducing blood glucose levels in diabetic rats in vivo [30]. Oral doses of 50 or 100 IU/kg contained in alginate–chitosan PECs were shown to gradually reduce glycemia over a 14 h period to approximately 60% of initial values. This O HO HO O O O O O O HO HO HO HO HO HO HO O O O O O O O O O - - - - - - - - O O O O O HO HO OH OH O OO NH+ OHO OH HO OH HO HO n n HO HO HOHO Guluronic residue Calcium crosslink Deprotonated alginate Electrostatic interaction Mannuronic residue Protonated chitosan O O OH O OH OH OH OH OH O OO O O O O O O O O O O O Ca2+ OH OHOHOH + H3 N + H3 N + HN HO NH3 + NH3 + Figure 3. Polyelectrolyte complex of calcium-crosslinked alginate complexed with chitosan.
  • 8. Review |Thompson & Ibie Therapeutic Delivery (2011) 2(12)1618 future science group compared favorably with a 2.5 IU/kg dose of noncomplexed insulin given subcutaneously (sc.), which reduced glycemia by more than 50% within 2 h, but which returned to near initial blood glucose levels after 8 h. It was also differ- ent from an oral insulin solution and a physical mixture of insulin and empty nanoparticles (both 50 IU/kg), which resulted in a max­imum reduc- tion in glycemia of only 20% after 6 h, which then returned to near pretreatment levels rapidly [30]. This indicated that although some insulin would have been released in the stomach and overall control of release was probably poor (as well as possible precipitation/dissolution of the polymer complex), the PECs allowed insulin to pass through the GI epithelium intact and produce a hypoglycemic effect. This effect was delayed compared with sc. insulin, as it would be expected that the PECs adhere to the GI mucosa and produce a prolonged release of insulin. As well as this, the time taken for GI transit, trans- port of insulin through the various barriers of the GI tract epithelium, then the circulation and finally to the liver before it could exert its affect would take far longer than transit to the circulation after sc. administration [4]. These results are promising in the search for an oral insulin-delivery system in that it would appear that the alginate–chitosan PECs are able to protect insulin and facilitate its absorption to some extent. It would also suggest a prolonged control of blood glucose levels would be pos- sible, which would reduce patient dosing as well as potentially provide for a reduction in diabetic complications. However, unsurprisingly, the doses required to elicit this effect are high compared with sc. doses (a dose of 25 IU/kg given as an oral nanopar- ticle produced a much reduced effect [maxi- mum reduction of 25% glycemia after 8 h]). Bioavailability of the 50 and 100 IU/kg doses were 5.1 (equivalent to 7% of administered dose) and 9.5 ng/ml.h, respectively, which although much higher than oral insulin administered non- complexed (2.1 ng/ml.h), could still be improved in order to reduce dose sizes. It should also be borne in mind that these relatively high doses are required due to loss of insulin during GI transit and transport across the GI epithelium due to chemical and enzymatic breakdown. Therefore, further work on these systems’ ability to protect insulin from degradation could be investigated. Additionally, although several uptake mecha- nisms were postulated (absorptive endocytosis, paracellular transport, Peyer’s patches) the exact routes and mechanisms of uptake have not yet been determined. If the mechanisms by which these PECs facilitate insulin uptake could be fully elucidated, it could indicate any further changes to the PECs structure, which may aid uptake and also minimize insulin loss during delivery. Further alginate–chitosan PECs were devel- oped by El-Sherbiny and colleagues involving a carboxymethylchitosan(CMC)–sodium ­acrylate copolymer, which could be complexed with cal- cium-crosslinked alginate and a PEGylated form of chitosan [99–102]. By grafting sodium acrylate to the CMC, the number of carboxylic groups within the PEC structure were increased. This was done in order to limit the potential for swell- ing, and therefore protein release, at gastric pH. This is possible as the carboxylic groups can form hydrogen bonds with each other and the hydrox- yls of CMC as well as form strong ionic interac- tions with calcium ions thereby minimizing the ability of the PECs to swell at low pHs. The loss of a large proportion of the amine groups in the structure of chitosan due to carboxymethylation and grafting of sodium acrylate further limits ionization at low pH and therefore swelling due to repulsion between protonated amino groups. By coating sodium acrylate–CMC–alginate with PEGylated chitosan, this should offer ever greater protection against premature protein release in the stomach by forming strong electrostatic inter- actions with CMC–sodium acrylate and alginate (Figure 4) [100]. It was found that swelling was very low (less than 5% increase in weight of PECs after 2 h) in simulated gastric fluid (pH 2.1), but fairly exten- sive (up to 40% after a further 8 h) in simulated intestinal fluid (pH 7.4). Release of bovine serum albumin (BSA) from these PECs was monitored in the same buffers and was found to be lower with those PECs coated with PEGylated chitosan than those with no coating in simulated gastric fluid (~10 compared with ~25% release after 2 h) and fairly well controlled once transferred to sim- ulated intestinal fluid (70–85% after a further 8 h). PECs coated with PEGylated chitosan also had greater loading capacities than noncoated forms (63–83% encapsulation for coated PECs as compared with 31–47% for noncoated) [100]. From this, it can be seen that the additional hydrogel coating allowed for the formation of strong electrostatic interactions between it and the other polymers in the PEC thereby gaining greater control over swelling and release, but also allowed for greater interaction with a model protein to allow for higher rates of encapsulation.
  • 9. The oral delivery of proteins using interpolymer polyelectrolyte complexes | Review www.future-science.com 1619future science group This work appears to also be very promising in terms of controlled intestinal release. The for- mulation may also have beneficial mucoadhe- sive, absorption-enhancing and enzyme-inhib- iting properties due to the number of carboxylic groups present as well as the presence of PEG; however these have yet to be determined. Further multipolymer complexes for insulin delivery have been developed by Woitiski et al. again involving the use of calcium crosslinked alginate nuclei complexed with chitosan [108]. This PEC was further modified by addition of another polyanion, dextran sulfate, to try and aid protein encapsulation and control release by increasing interpolymer and polymer–insulin electrostatic interactions (as with the use of multi- ple polyelectrolytes by [100]). Attempts to improve steric stabilization of PECs were also made by hydrogen bonding of PEG or hydrophobic bonding of poloxamer 188 (Pluronic® ) (a tri- block copolymer of poly(ethylene oxide) and poly[propylene oxide]) to chitosan and alginate. PECs were coated with albumin in order for it to act as a competitive target for proteolysis to limit the degradation of insulin (Figure 5) [24,25,108], which is a fairly novel and interesting mechanism to reduce insulin degradation. Importantly, both an in vitro western blot bio- activity assay and in vivo sc. injection of insulin (collected after in vitro release from PECs) into diabetic rats demonstrated that insulin activ- ity was retained after formulation into these PECs [25,108]. However, a problem with using another protein in this formulation was that it also acted as a competitive binder to the polymers in the PEC, which reduced insulin encapsula- tion from 90% at 0.25% albumin to 75% at 1% albumin [108]. COO- C=O C=O C=O O- Cs-g-PEG (Thin layer coating) Alg–CMCs-g-AAs (Negative core of hydrogel microparticles) A B + NH3 HOOC COOH COOH COOH COOH COOH COOH HOOC COOH HOOC OH OH + NH3 NH3 + HOOC COOH HO - OOC COO- COO- COO- COO- COO- COO- COO- COO- - OOC - OOC - OOC O- O- + NH3 + NH3 + NH3 Ionic crosslinks (with Ca2+ ) Hydrogen bonds Repulsive interactions CMCs-g-AAs Na alginate Cs-g-PEG SIF (pH 7.4) SGF (pH 2.1) Figure 4. The different types of interactions in the developed alginate–CMCs-g-AAs hydrogel microspheres coated with Cs-g-PEG at both pH 2.1 (SGF) and pH 7.4 (SIF). CMC: Carboxymethylchitosan; PEG: Poly(ethylene glycol); SGF: Simulated gastric fluid; SIF: Simulated intestinal fluid. Reproduced with permission from [100]. © Elsevier.
  • 10. Review |Thompson & Ibie Therapeutic Delivery (2011) 2(12)1620 future science group Additionally, although some formulations limited release in simulated gastric fluid (to at best ~12% after 2 h), on transferring PECs to simulated intestinal fluid very rapid bursts (of up to another 40% of loaded insulin) of release were seen with all formulations. This would sug- gest a similar problem as found with the stan- dard alginate–chitosan PECs discussed above, which could also not control release on increases in pH due to polymer dissolution/precipitation, or swelling owing to electrostatic repulsion between ionized carboxyl groups [2,22,30,83]. It may be that the additional binding of dex- tran sulfate as well as PEG/poloxamer and albu- min to alginate–chitosan complexes reduces, rather than improves, their stability at intestinal pH values. These additional polymers and pro- teins may limit the number of binding sites avail- able for insulin in the PECs and result in weaker polymer–protein interactions as well as weaker interpolymer interactions limiting control over protein release. Further to this it was found that the PECs produced by Woitiski and colleagues were still able to protect insulin from pepsin degradation in vitro, compared with a noncomplexed insu- lin control (18 and 77% degraded, respectively) probably due to the lack of swelling of the PECs at gastric pH; the fact that the majority of the insulin is entrapped within PEC cores and not on the surface; and that albumin acted as a sacrificial target for the enzyme [25]. Furthermore, it was found that these PECs (using a dose of 50 IU/kg) were able to reduce blood glucose in diabetic rats gradually over 12 h to a peak reduction of 60% of initial values and were able to maintain levels to approximately 50% of initial values for a further 12 h and overall bioavailability was 13% (almost double that achieved by Sarmento et al. with the same dose [30]). This is a promising finding as it would indicate these PECs are able to deliver larger amounts of bioactive insulin through the GI epithelium and into the circulation over a long period of time and crucially maintain low glyce- mia for up to at least 24 h after administration. This (together with Sarmento and colleagues in vitro release and in vivo glycemia data dis- cussed above) would also suggest that the release of insulin in vivo is quite different than in vitro, given that hypoglycemia lasted for up to 24 h in vivo, but insulin release in vitro was complete within 180 min. This implied poor in vitro– in vivo correlation would indicate that the use of simulated gastric and intestinal fluid in these kinds of studies has limited use at best, particu- larly given the other substances, such as food, numerous enzymes and differences in buffer capacity and ionic strength of GI fluid compared with those used on the bench. Supplementary work with these PECs found that the presence of albumin also aided insu- lin transport across both in  vitro cell layers (Caco-2 alone and co-cultured with HT-29) and ex vivo rat intestinal mucosa. The perme- ability co­efficient for albumin-covered PECs was 28.4  ×  10-6 compared with 19.38 and 7.35 × 10-6  cm����������������������������������/s for albumin-free PECs and insu- lin alone, respectively, in ex vivo rat intestinal mucosa. This was possibly achieved by albumin acting as a stabilizer for insulin (i.e., it aids pas- sage across the mucus layer as well as limiting intracellular and extracellular enzymatic break- down). However the data for albumin-free PECs may be more instructive in terms of determin- ing permeability in vivo as the albumin layer should be degraded in the stomach and by gas- tric and intestinal enzymes prior to initiation of absorption of insulin in the intestines. Transepithelial electrical resistance (TEER) values of co-cultured Caco-2 and HT-29 cells also dropped on incubation with PECs. However, this was only by approximately 20% of initial values, which would suggest that transport was predominantly via transcellular routes [26]. The Alginate Dextran sulfate Poloxamer Albumin Chitosan Calcium Insulin Figure 5. Multilayered nanoparticles encapsulating insulin. Nanoparticles are formed by alginate and dextran sulfate nucleating around calcium ions and interacting with poloxamer, further stabilized with chitosan and subsequently coated with albumin. Reproduced with permission from [25] © Elsevier.
  • 11. The oral delivery of proteins using interpolymer polyelectrolyte complexes | Review www.future-science.com 1621future science group exact mechanisms and transporters involved in uptake and translocation through cells are also to be elucidated. Poly(aspartic acid)–chitosan PECs Poly(aspartic acid) is also a poly(carboxylic acid) like alginate. However, it is synthetic rather than derived from natural resources. It is pro- duced by polycondensation of aspartic acid to poly(succinimide) followed by hydrolytic ring- opening. As it is based on an amino acid, the polymer formed has a protein-like structure and is water soluble and biodegradable [84,90]. Shu and colleagues formed PECs between chitosan and thiolated polyaspartic acid with a view to produce pH-sensitive swellable nanoparticles for the oral delivery of insulin (Figure 6) [90]. The potential advantage of these PECs is that they may be more stable than other chitosan–polyanion-based PEC. Previously, the electrostatic interactions used to form PECs based on chitosan and other polyanions were weaker or stronger in different parts of the GI tract due to the variation in pH, depending upon their pKa , and were generally unstable in vivo due to salting-out effects [22,30,78,83,87,109]. However, the use of a thiolated polymer in this case has not only allowed for electrostatic interactions to take place, but for disulfide-bond crosslinking, which increased the strength of interactions and overall nanoparticle stability at low pH (pH 1.2–4) and in increasing salt concentrations in vitro. Additionally, in vitro release studies dem- onstrated that PECs, which relied solely upon electrostatic interactions, released insulin very quickly (~90% within 30  min) at pH  1.2, while chemically crosslinked particles showed a more controlled release at pH 1.2 (~40% after 120 min). Although after adjustment of the pH to 6.8, the remaining insulin from crosslinked nanoparticles was very rapidly released (~90% after a further 30 min). The rapid release in pH 6.8 is surprising, given that disulfide-bond formation should be increased at a pH above 5 due to the increasing oxidation of the thiol groups on the PECs [32,63,94– 97,103]. The opposite appears to have happened whereby the interpolymer interactions have become so weak that the nanoparticles can no longer retain insulin at increasing pHs. It may be that the use of chitosan has resulted in a degree of precipitation of the formulation at pH 6.8 given that it would be fairly insoluble at that pH. As the disulfides do not form between chitosan and aspartic acid, but only between aspartic acid chains this would not solve the problem of the lack of solubility of chitosan at higher pH. This together with the loss of electrostatic interaction between chitosan and aspartic acid could have caused the burst release. A variation on this formulation was also examined where a low-molecular weight (6000 g/mol), water-solu- ble chitosan was used instead and poly(aspartic acid) was used without thiolation. In this study, the polyelectrolytes were co-mixed with PEG in order to provide additional sites for inter- polymer interactions and, therefore, form more stable PECs [84]. PEG was also used in order to promote mucoadhesion by improving complex chain flexibility and mobility as well as maintain the secondary and tertiary conformation of BSA when encapsulated. The addition of PEG may have resulted in interpolymer interaction between its ether group and the carboxylic-acid group of the aspartic acid as well as between the primary amines on chitosan and the hydroxyl on PEG. As a result of this, the interpolymer network was strengthened and more compact nanoparticles were formed (from a radius of approximately 175 nm with no PEG to 125 nm at 30 mg/ml PEG). These nanoparticles were also stable along a range of salt concentrations with the particle radius not exceeding 300 nm up to 0.025 mol/l NaCl. This was in contrast to the previous study where particle radius increased up to 400 nm at the same salt concentration [90]. However it is dif- ficult to attribute this increased stability solely to the presence of PEG, given the number of other variables that changed between the two studies [84,90]. Regardless of the apparent increase in par- ticle stability (uncontrolled), pH-dependent release was again found. In vitro BSA release was approximately 20% at pH 2.5, 60% at pH 1.2 and 80% at pH 7.4 (after 240 min). Again, it can be seen that noncontrolled release occurred at a higher pH, which would not be beneficial in vivo. The reduced protonation of chitosan at pH 7.4 would again, regardless of its increased water solubility at low-molecular weight, result in very loose and diffuse PECs at this pH causing rapid and uncontrolled protein release. In order to improve upon this delivery sys- tem, quaternization and/or thiolation of chitosan may also be required in order that electrostatic forces between polymers are strengthened and are pH independent as well as potentially allow- ing the formation of disulfides between chitosan
  • 12. Review |Thompson & Ibie Therapeutic Delivery (2011) 2(12)1622 future science group and (thiolated) aspartic acid in order to further strengthen the interpolymer network. In addi- tion, conjugation of PEG to either polymer may also facilitate greater complex stability, solubil- ity, biocompatibility and mucoadhesion than the weaker hydrogen bonding seen above between PEG and chitosan [82,88,89,91,92]. Indeed, it has been postulated that the counterion effects of salt can be counteracted to a limited extent by steric hindrance of quaternary ammonium salt groups, grafted PEG and disulfide-bond formation [82,89]. However, it must also be borne in mind that the ability to control release of proteins is also dictated by their charge which will also change with pH. The ability of a protein to form elec- trostatic interactions with a polymer will alter drastically around their pI, as the charge on the polymer goes from cationic to anionic or vice versa [92]. Therefore any system that is to pro- duce controlled release in the GI tract and avoid release in the stomach, must not rely solely upon electrostatic interactions between the polymer and protein. Poly(methacrylic acid) copolymer PECs Poly(methacrylic acid) (PMMA) is another weak acid, which has been used to complex with pro- teins. However, as with other polyanions men- tioned above, when used alone it is unable to Electrostatic interaction HO HO HO HO HO HO HO HO + H3 N + H3 N + H3 N + H3 N + HN H2 H2 H2 H2 H2 CH2 CH2 CH2 CH2 CH2 CH2 H2 NH + OH OH OH OH OH OHOH O O O O O O CC C C C C = O NH NH C = O C O S S O O O O - - - - - - - - O O O O O O O O O O O O O O O O n x y n yx H N H N H N H N H N H N O O O O O OH Disulfide crosslink Ionised poly(aspartic acid) Protonated chitosanNH3 + NH3 + Figure 6. Polyelectrolyte complex of disulfide crosslinked thiolated poly(aspartic acid) complexed with chitosan.
  • 13. The oral delivery of proteins using interpolymer polyelectrolyte complexes | Review www.future-science.com 1623future science group eff­iciently complex with proteins at the pH range of the GI tract. Therefore attempts have been made to produce more stable complexes by copo- lymerizing with PEG. This increases complex stability as the PEG ether groups and carboxyl groups of PMMA can form hydrogen bonds to produce an interpolymer complex [13,111–113]. PMMA-based nanoparticles have been pro- duced by copolymerizing with an amine mod­ ified PEG, which facilitated interpolymer ionic crosslinking [13]. Individual PMMA–PEG par- ticles were approximately 800 nm in diameter. However, a high degree of aggregation occurred at gastric pH resulting in the form­ ation of micron-sized particulates. At intestinal/physi- ological pH, above the pKa of PMMA, increased interparticle repulsion occurred resulting in smaller, discrete particles. This is very impor- tant as smaller particles are more likely to inti- mately interact with the GI tract epithelium in terms of promoting both mucoadhesion and absorption [13,88]. PMMA–PEG particles were able to demon- strate ex vivo mucoadhesion for up to 3 h due to both the predominantly anionic charge on the particles at pH 7 and possibly the ability of PEG to entangle with mucus glycoproteins. However, longer lasting mucoadhesion would be desirable to set up and maintain a protein concentration gradient across the GI tract mucosa into the cir- culation and so further particle size reduction could be beneficial. BSA was loaded into these particles with encapsulation efficiencies of up to 80%. In vitro release from these particles showed a very slow and incomplete release at pH 1.2 (less than 10% after 3 h) and relatively rapid release at pH 7.4 (~90% after 3 h). This pH-dependent release could be attributed to the increase in swelling of the particles at elevated pH due to the increased ionization of carboxyl groups on PMMA caus- ing repulsion between PMMA chains above its pKa  [13]. Again, this is not ideal as although the protein would not be released in the stomach, the rapid release in the intestines would be unlikely to result in the maintenance of therapeutic levels and would most likely result in the requirement of repeated dosing. The same study also utilized the PMMA–PEG PECs to load insulin alone and precomplexed to carboxymethyl-b-cyclodextrin (Cb-CD). The Cb-CD was used to improve insulin stability in the delivery system, by pre- venting insulin hydrophobic intermolecular interactions which can lead to aggregation and a reduction in activity [21]. Encapsulation efficiency was high at approximately 78% (insulin alone) and 73% (insulin–Cb-CD) and insulin was shown to still be biologically active after encapsulation using an ELISA. However, insulin release from the particles was very rapid at pH 7.4 (in effect, dose dumping) with approximately 70% release within 60 min regardless whether insulin was encapsulated alone or complexed to Cb-CD. The more rapid release of insulin compared with BSA may in large part be due to its much smaller-molecular weight compared with BSA (5800 compared with 66,000 g/mol). The pore size of the swollen particles may be such that insulin can pass easily through them, while the larger BSA requires more time in order to be able to diffuse through the interpolymer complex. Therefore, more precise control over swelling (and hence pore size) is required for delivery of smaller proteins. One way to achieve this may be by increas- ing the molecular weight of the PEG used. Higher-molecular weight PEG could result in greater interpolymer interactions, thereby reduc- ing swelling and therefore pore size. However, this reduction in pore size could also limit insu- lin loading. Increasing PEG molecular weight in PECs from 400 to 1000 has been shown to reduce in vitro insulin release (from 41 to 39% after 2 h) and also reduced insulin encapsulation (from 38 to 32%) [112]. A reduction in insulin loading may be necessary to provide greater con- trol of insulin release. However, the reductions in release seen here are quite small and so other options need to be considered, perhaps including the use of considerably higher molecular weight PEGs, for example, ≥10000 g/mol. Other work with PMMA–PEG copolymers has involved the incorporation of chitosan into PECs. Sajeesh and Sharma copolymerized PMMA, chitosan and polyether (a copolymer of PEG and polypropylene glycol) [21]. In this case the electrostatic interaction was between the primary aminos on chitosan and the carboxyl groups on PMMA while polyether formed hydro- gen bonds with PMMA. The choice of polyether rather than PEG alone was due to the potential for polyethers to promote absorption across epi- thelial barriers via a surfactant-like effect on cell membranes (as with the Pluronics® ) [21]. This PEC was complexed with insulin either alone or after insulin was again complexed with a hydro- philic derivative of b-cyclodextrin (2-hydroxy- propoyl b-cyclodextrin [HPbCD]) to increase its stability in the formulation.
  • 14. Review |Thompson & Ibie Therapeutic Delivery (2011) 2(12)1624 future science group Encapsulation efficiency of insulin was approx- imately 85% regardless of whether insulin was used alone or as part of a CD complex indicating a lack of interaction between HPbCD and the interpolymer complex. This is borne out by the similarity in in vitro release from particles, which carried insulin alone or as part of a HPbCD com- plex; pH-dependent release was demonstrated with approximately 15% release within 2 h at pH 1.2, and 75–80% release at pH 7.4 over the same time regardless of the presence of HPbCD. The lack of control of release at pH 7.4 had still not been overcome by these further formula- tion changes, and would suggest that while the use of a CD ensures that insulin remains active when formulated it does nothing to aid control of insulin release as it cannot interact with the polymers present in the PEC. This study was also able to show that these PECs can show some concentration-depen- dent trypsin inhibitory activity when using N-a-benzoyl-l-arginine ethyl ester (BAEE) and casein as low- and high-molecular weight substrates, respectively. Greater protection was offered to casein than BAEE over a PEC concen- tration range of 0.25–1% (w/v), which may be due to differences in substrate molecular weight (24,000 g/mol for casein compared with approx­ imately 300 g/mol for BAEE). As the studies were conducted at pH 7.6, the PECs would have been swollen and so release of BAEE would have been even more rapid than insulin, while casein release could have been much slower. This difference in free and encapsulated substrate could have meant that as more casein was encapsulated for longer, then a physical barrier to trypsin binding, as well as any potential chelation of metal ions by the carboxyl groups of PMMA, would have provided greater protection to it than BAEE. However, the protective effects seen are modest and only reached at most to approximately 40% inhibition of trypsin at 1%(w/v) PEC concentration. The lack of a substantial inhibitory effect may be due to the lack of ‘free’ carboxyls on the PMMA after copolymerization. Electrostatic interactions and hydrogen bonding within the PEC between polymers and also the encapsu- lated protein maintain the PEC structure in solution. Therefore it may be that very few car- boxyls were not already interacting with other PEC components and so very few of them were available to chelate metal ions. This would seem to be confirmed by a simi- lar study, which showed that carbopol 974 had a greater inhibitory effect on trypsin activity against casein than PMMA–chitosan–PEG PECs; 80% inhibition compared with 50% inhibition at a concentration of 1%(w/v), respectively  [88]. Carbopols are weakly crosslinked hydrogels of acrylic acid and alkyl ethers and, as such, may have a fewer number of carboxyls involved in crosslinking than these PMMA-based PECs. However, it would be useful to compare PMMA alone with crosslinked versions to confirm this. In addition, attempts to reduce the degree of interaction between these polymers, perhaps by varying the molar ratios used for polymerization or molecular weights of the polymers, would provide greater insight into this. However, a reduction in interaction between polymers used in these systems would probably result in even faster release of encapsulated proteins due to fur- ther increases in pore size on swelling or perhaps collapse of PECs altogether on increases in pH. Further studies in this area have been con- ducted to try and improve control of protein release and enzyme inhibition of PMMA–chi- tosan–PEG nanoparticles by conjugation of cys- teine to the surface of these PECs [65,66]. As with other thiomers mentioned above, the presence of these thiol groups should confer additional properties to the PECs including covalent muco- adhesive bond formation and greater chelation of metal cations to facilitate both tight-junction opening and enzyme deactivation. These PMMA–chitosan–PEG thiomers are different in that the surface of the PECs is thio- lated and not the interior. To date, most thiomers have relied on disulfide crosslinking within PECs to form stable particles as the pH rises to five and above [32,63,94–97,103]. However, as mentioned above, this reduces the number of ‘free’ thiol groups available for mucoadhesion and enzyme inhibition. One way to improve these properties may be to attach thiol-­containing moieties to the surface of PECs instead, which could maximize the ‘free’ thiol content of the formulation. It has been found, however, that disulfides do still form between surface thiol groups. Sajeesh et al. found that thiol content of thiolated PMMA– chitosan–PEG PECs increased from 506 to 1214 µmol/100 mg of PECs after treatment with sodium borohydride to cleave disulfides [65]. The potential to form disulfides between par- ticles should also be considered, which would promote particle aggregation, particularly at ele- vated pHs, such as in the lower GI tract. If this were to occur, intimate contact between particles and the GI mucosa could be reduced as well as the ability to promote mucoadhesion, enzyme
  • 15. The oral delivery of proteins using interpolymer polyelectrolyte complexes | Review www.future-science.com 1625future science group inhibition and mucosa permeability, thus, limit- ing the ability to both protect a loaded protein and facilitate its absorption. Limiting disulfide formation may be achieved by the presence of other functional groups or ligands, which could increase steric hindrance between thiols, limit- ing their ability to conjugate, such as quaternary ammonium moieties or PEG [82,89]. Although no electron microscopy was per- formed on the thiolated PMMA–chitosan–PEG PECs in this study, particles would appear to be aggregated at pH 7, given that sizes increased with the number of thiol groups grafted to the particle surface. PECs with no thiol groups registered a hydrodynamic diameter of 1030 nm, while those with 506 µmol of thiol groups/100 mg of PECs were 1900 nm. Indeed increased aggregation of the thiolated compared with nonthiolated PECs appeared to be confirmed by optical microscopy in a later study [66]. Importantly, the thiolated particles showed little cytotoxicity when analyzed using an MTT assay with even the highest thiol content PECs (506 µmol of thiol groups/100 mg of PECs) at the highest PEC concentration (2 mg/ml) show- ing approximately 85% of Caco-2 cells remained viable after 6 h incubation. This indicates that any increase in permeability of cells caused by these PECs should not be due to cellular damage, or at least very low levels of damage, which may be reversible. These PECs were also shown to facilitate transport of a macromolecular marker (fluores- cein isothiocyanate-dextran 4000 g/mol [FD4]) across Caco-2 cell monolayers. This ability was increased with the number of thiol groups attached to the PECs and was shown to be depen- dent upon the concentration of calcium in the media used. As calcium concentration increased the apparent permeability of FD4 decreased, which would indicate that the transport of FD4 across cell layers was due to calcium chela- tion by the PECs, that is, it primarily involved paracellular transport. This was confirmed by the differences in TEER found when Caco-2 cells were incubated with thiolated PECs in the presence and absence of calcium. In the absence of calcium TEER val- ues dropped to 40% of their initial values within 30 min, compared with only 60% of initial val- ues when calcium chloride concentration was 10 mM. TEER reduction and concomitant transport of FD4 was relatively similar for thiolated and nonthiolated PECs at 0 mM calcium chloride, as calcium concentrations increased, thiolated PECs were able to reduce TEER and promote FD4 transport to a greater extent than nonthio- lated forms. This would suggest that thiol groups are more effective chelators of metal cations than carboxyl groups and so their use may be highly advantageous in attempts to facilitate macromole- cule absorption. However, when transport of FD4 across ex vivo rat intestinal tissue was determined the presence of thiol groups appeared to make little difference with apparent permeability of FD4 being 7.4 × 10-7 and 7.9 × 10-7  cm�����������/s for non- thiolated and thiolated PECs, respectively. This highlights the limitations of using Caco-2 cells as an in vitro model for the GI tract epithelium. Unlike the GI epithelium, Caco-2 cells do not have a mucus layer and their TEER values are markedly higher [114]. This is potentially a major flaw with the use of Caco-2 cells as a model and as such HT-29 cells may be more appropriate in that their TEER values are in line with GI epi- thelial cells and they produce a mucus layer and M cells [26,78,114]. In particular the presence of a mucus layer can have marked effects on PEC inti- mate contact with mucosal membranes and tight junctions resulting in a major reduction in the ability of PECs to facilitate protein absorption. The presence of thiol groups may result in disul- fide formation between PECs and cysteine resi- dues in mucus glycoproteins, which could mark- edly inhibit their ability to reach and interact with cell membranes [65]. Co-cultures of these cell lines could provide an in vitro model, which more closely mimics the intestinal epi­thelia, as together they would produce monolayers which had both absorptive and mucus secreting cells  [26,114]. However, it should be noted that Caco-2 has been shown to be a useful model many times over and has indeed generated in vitro data that correlates well with in vivo findings [114]. In a later study by the same group, insulin was loaded into the PECs rather than FD4 [66]. It was notable that in vitro insulin degradation by a mixture of trypsin and a-chymotrypsin (10 U/ml each) was prevented to a greater extent with non- thiolated PECs than thiolated versions. After 60 min, thiolated PECs had only approximately 10% of their initial insulin load nondegraded, while nonthiolated PECs had approximately 25% nondegraded. This is surprising given that one expected advantage of thiolated PECs is their increased ability to protect proteins from enzy- matic breakdown [32,63,94–97,103]. Again this may refer back to the propensity for surface-function- alized PECs to form disulfides which in turn
  • 16. Review |Thompson & Ibie Therapeutic Delivery (2011) 2(12)1626 future science group may limit their ability to chelate metal cations which are responsible for enzyme activity [66]. It should also be noted that regardless of PEC struc- ture, total insulin degradation occurred within 90 min incubation, which suggests that neither formulation may offer enough protection in vivo. Conversely however, in vivo studies using dia- betic rats demonstrated that blood glucose levels were reduced by both thiolated and nonthiolated PECs (with a dose of 50 IU/kg). Thiolated PECs demonstrated a greater reduction of approxi- mately 40% of initial glucose levels within 2 h of administration compared with approximately a 10% reduction for nonthiolated PECs. Blood glucose levels in rats treated with thiolated PECs gradually increased over the next 8 h, but a reduction in glucose levels of approximately 10–20% was still maintained after 10 h. The greatest reduction in glucose levels by nonthio- lated PECs was seen after 6 h (20–30% reduc- tion), which was quickly followed by a return to pretest glucose levels within the next 2 h. This highlights the difference in cellular interaction of the PECs when thiol groups are present. Even with the apparent crosslinking between particles, enough ‘free’ thiols may be present to rapidly chelate metal cations and inter- act with tight-junction proteins to promote insu- lin absorption. This could subsequently lead to rapid insulin uptake and so a fairly rapid and sub- stantial decline in blood glucose. The increased ability of the thiolated PECs to adhere to the GI tract mucosa may also explain their ability to maintain this effect for a prolonged period. This adherence to the GI tract wall could allow for an insulin concentration gradient to be set up across the epithelium into the circulation. The prolonged reduction in glucose levels found here is highly desirable in the treatment of diabetes as it could maintain glucose control without having to administer a large number of doses per day, while potentially providing more prolonged glycemic control than current sc. dosing. However, as with the studies discussed above, the reduction in glucose seen with PECs is less than with a dose of insulin given sc. (70– 80% reduction within 90 min) and for PEC oral formulations to be effective they must reduce glucose levels to a great extent as well as main- tain that effect over many hours (the accepted gold standard for any insulin-dependent dia- betes treatment is a reduction in glycosylated hemoglobin levels to less than 6.5–7% [115]). Therefore, although surface thiolation may appear attractive in theory, in practice the very high reactivity of thiol groups (after oxida- tion) means that crosslinking between particles becomes inevitable, thus limiting the ability of these groups to aid enzymatic protection, muco- adhesion and protein absorption. Further work in this area should look at attempts to limit particle crosslinking and reduce particle size to the nano- range to improve both enzymatic protection as well as increase absorption. Future perspective Oral protein delivery has still not been achieved and although progress has been made, a great deal of work remains to be done. The use of PECs seems to allow for both protection against degradation and promotion of absorption to a limited extent. However, bioavailability of insu- lin through the oral route is still much lower than that of conventional parenteral delivery, requir- ing the use of more than ten-times the sc. dose of insulin to achieve and maintain hypoglycemia in animal models. There has been a general trend towards the use of chemically modified chitosan either alone or in combination with other polymers to form PECs. As part of this, the use of thiomers is highly advantageous in both promotion of absorption and protection against enzymatic breakdown of insulin. The use of PEG has also been a feature of the research discussed in this review due to its many potential benefits in terms of mucoadhe- sion, maintenance of active protein conformation and PEC stability. However, neither the use of thiomers or the incorporation of PEG has so far lead to a substantial breakthrough in oral pro- tein delivery as they still do not provide for large quantities of bioactive protein to pass across the GI tract. Interpolymer complexes seem to hold great potential due to their ability to produce pH- dependent release and mucoadhesion. However, they too have their limitations in terms of effi- cient insulin delivery to the circulation due to continued loss of insulin during GI transit through premature release and lack of complete protection against degradation. This seems to be primarily due to the limitation of utilizing PECs, which rely solely on pH-dependent, coulombic interactions for their formation (Table 2). Therefore in future studies, it may be bene­ ficial to form interpolymer complexes via chemi- cal crosslinks, which are not affected by the ionic strength and changes in the pH of GI fluid. However, this could lead to other issues in regards to whether favorable protein conformations could
  • 17. The oral delivery of proteins using interpolymer polyelectrolyte complexes | Review www.future-science.com 1627future science group be maintained and how proteins could be released on passage through the GI epithelium. In order to achieve high oral insulin bioavail- ability, more work needs to be done to look at developing polymers that have specific binding sites for insulin (or other proteins), which pro- tect their degradation target sites, maintain their native conformation and also efficiently inter- act with GI epithelial cell membranes and tight junctions to rapidly and completely facilitate absorption. With this in mind, it is crucial to develop better screening techniques for PECs to deter- mine their uptake routes and exact mechanisms of interaction with cell membranes and tight- junction proteins in order that these structure can be further fine tuned to aid protein trans- port and translocation. This will require much greater collaboration between polymer chemists, human (gut) physiologists, biochemists and pharmacologists to systematically develop PEC systems from the ground up. Until this is done, it is unlikely oral bioavailability of proteins can be increased (reproducible, rapid and sustained glycemic control can be achieved) to the point where these delivery systems are commercially viable alternatives to the status quo. Financial & competing interests disclosure The authors have no relevant affiliations or financial involvement with any organization or entity with a finan- cial interest in or financial conflict with the subject matter or materials discussed in the manuscript. This includes employment, consultancies, honoraria, stock ownership or options, expert t­estimony, grants or patents received or pending, or royalties. No writing assistance was utilized in the production of this manuscript. Table 2. Summary of polymers used in interpolymer complexes. Polymers Important findings Ref. Chitosan and calcium- crosslinked alginate Lack of control over insulin release at intestinal pH in vitro Insulin stable in PECs for up to 1 month Insulin bioavailability in diabetic rats of 7% [22,30,83,87] Carboxymethylated chitosan- acrylate, calcium crosslinked alginate and PEGylated chitosan Coating of PECs with PEGylated chitosan and removal of majority of primary amines from chitosan structure offers control of protein release over pH range of GI fluid [99–102] Chitosan, calcium-crosslinked alginate, dextran sulfate, PEG/poloxamer and albumin Protection of insulin against pepsin degradation in vitro Insulin bioavailability in diabetic rats of 13% Insulin appears to be absorbed transcellularly [24–26,108] Thiolated poly(aspartic acid) and chitosan Lack of control over release at intestinal pH in vitro Slower release at pH 1.2 when thiolated form of aspartic acid used [84,90] Poly(aspartic acid), chitosan and PEG Complexes more compact and stable on increases in salt concentration when PEG incorporated into PEC PMMA–PEG copolymer Lack of control over release at intestinal pH in vitro [13,21,65,66,88] PMMA–PEG–chitosan copolymer and Cb-CD insulin Lack of control over release at intestinal pH in vitro Cb-CD maintains insulin biological activity after complexation with copolymer Thiolated PMMA–Chitosan–PEG copolymer Presence of thiol groups increases transport of FD4 across Caco-2 monolayer Thiol groups appear to increase chelation of calcium in vitro thereby increasing reduction in Caco-2 TEER; Thiol groups make no appreciable difference in FD4 transport across ex vivo mucosa Carboxylic groups rather than thiols offer increased protection against trypsin and a-chymotrypsin in vitro Thiolated PECs able to reduce glycemia by 40% within 2 h of administration. Nonthiolated PECs reduce glycemia by only 10% in same time. Cb-CD: Carboxymethyl-b-cyclodextrin; FD4: Fluorescein isothiocyanate-dextran 4000 g/mol; PEC: Polyelectrolyte complex; PEG: Polyetyhlene glycol; PMMA: Poly(methacrylic acid); TEER: Transepithelial electrical resistance.
  • 18. Review |Thompson & Ibie Therapeutic Delivery (2011) 2(12)1628 future science group Executive summary Current status of oral insulin delivery „„ Proteins are becoming an increasingly popular option for treatment, due to high specificity and low toxicity. „„ Oral delivery is preferential for patients and clinicians. „„ A number of oral insulin formulations are in clinical trials at the moment. Barriers to oral protein delivery „„ Proteins are almost completely degraded in the GI tract due to enzymatic and chemical effects. „„ Proteins cannot pass across epithelial layers unaided due to their large size, molecular weight and charged nature. „„ There are numerous routes to pass across the gastrointestinal epithelium both para- and trans-cellularly. However, these are not accessible to proteins when administered alone. Technologies to facilitate oral protein delivery „„ Numerous attempts have been made to promote absorption and prevent degradation of protein in the GI tract that have proved unsuccessful due to problems in manufacturing of dosage forms and/or toxic side effects on administration. „„ Polyelectrolyte complexes potentially offer a way to benignly encapsulate (via coulombic and hydrophobic interactions) and deliver proteins. Chitosan polyelectrolyte complexes „„ Chitosan in its native state is ineffective in oral protein delivery due to its weakly basic nature. „„ Modified forms of chitosan have been produced via trimethylation of primary amines or attachment of thiol-containing ligands. These forms have shown improved ability to facilitate insulin absorption and improve protection in the intestines. Interpolymer polyelectrolyte complexes „„ Polyelectrolyte complexes formed from mixing (or copolymerization) of oppositely charged polymers has been achieved. Use of polyelectrolyte mixtures can overcome some of the problems of using polycations/anions alone, namely premature release and polymer precipitation. „„ These have primarily been produced using mixtures of alginate and chitosan. However other polyanions have been used with chitosan including poly(aspartic acid) and poly(methacrylic acid). „„ Some of these systems have an ability to produce sustained hypoglycemic effects in rat models. „„ Further work is required to maximize the efficiency of these systems, particularly in their ability to facilitate insulin absorption. References Papers of special note have been highlighted as: n of interest nn of considerable interest 1 Hamman JH, Enslin GM, Kotzé AF. Oral delivery of peptide drugs: barriers and developments. BioDrugs 19(3), 165–177 (2005). nn Detailed review of various barriers to oral protein delivery and technologies to overcome them. 2 George M, Abraham TE. Polyionic hydrocolloids for the intestinal delivery of protein drugs: alginate and chitosan – a review. J. Control. Release 114, 1–14 (2006). 3 Morishita M, Peppas NA. Is the oral route possible for peptide and protein drug delivery? Drug Discov. Today 11(19–20), 905–910 (2006). 4 Woitiski CB, Carvalho RA, Ribeiro AJ, Neufeld RJ, Veiga F. Strategies toward the improved oral delivery of insulin nanoparticles via gastrointestinal uptake and translocation. BioDrugs 22(4), 223–237 (2008). 5 Peppas NA, Carr DA. Impact of absorption and transport on intelligent therapeutics and nanoscale delivery of protein therapeutic agents. Chem. Eng. Sci. 64, 4553–4565 (2009). 6 Park K, Kwan IC, Park K. Oral protein delivery: current status and future prospect. React. Funct. Polym. 71, 280–287 (2011) nn Informative review of the status of oral protein delivery and delivery systems undergoing clinical trials. 7 Singh R, Singh S, Lillard JW. Past, present, and future technologies for oral delivery of therapeutic proteins. J. Pharm. Sci. 97(7), 2497–2523 (2008). n Review of a number of technologies with potential to facilitate oral protein delivery. 8 Woodley F. Enzymatic barriers for Gl peptide and protein delivery. Crit. Rev. Ther. Drug Carrier Syst. 11, 61–95 (1994). nn Detailed review of enzymatic barriers to oral protein delivery. 9 Mahato RI, Narang AS, Thoma L, Miller DD. Emerging trends in oral delivery of peptide and protein drugs. Crit. Rev. Ther. Drug Carrier Syst. 20(2–3), 153–214 (2003). nn Highly detailed and comprehensive review of all barriers to oral protein delivery and perspective various technologies to overcome those problems. 10 Peppas NA, Kavimandan NJ. Nanoscale ana­lysis of protein and peptide absorption: Insulin absorption using complexation and pH-sensitive hydrogels as delivery vehicles. Eur. J. Pharm. Sci. 29, 183–197 (2006). 11 Liu H, Wang Y, Li S. Advanced delivery of ciclosporin A: Present state and perspective. Expert Opin. Drug Deliv. 4, 349–358 (2007). 12 van Kerrebroeck P, Rezapour M, Cortesse A, Thüroff J, Riis A, Nørgaard JP. Desmopressin in the treatment of nocturia: a double-blind, placebo-controlled study. Eur. Urol. 52(1), 221–229 (2007). 13 Sajeesh S, Sharma CP. Novel polyelectrolyte complexes based on poly(methacrylic acid)-bis(2-aminopropyl)poly(ethylene glycol) for oral protein delivery. J. Biomater. Sci. Polym. Ed. 18, 1125–1139 (2007).
  • 19. The oral delivery of proteins using interpolymer polyelectrolyte complexes | Review www.future-science.com 1629future science group 14 Lee Y-H, Sinko PJ. Oral delivery of salmon calcitonin. Adv. Drug Deliv. Rev. 42(3), 225–238 (2000). 15 Kidron M, Dinh S, Menachem Y et al. A novel per-oral insulin formulation: proof of concept study in non-diabetic subjects. Diabet. Med. 21, 354–357 (2004). 16 Morçöl T, Nagappan P, Nerenbaum L, Mitchell A, Bell SJD. Calcium phosphate–PEG–insulin–casein (CAPIC) particles as oral delivery systems for insulin. Int. J. Pharm. 277, 91–97 (2004). 17 Clement S, Dandona R, Still JG, Kosutic G. Oral modified insulin (HIM2) in patients with type 1 diabetes mellitus: results from a phase I/II clinical trial. Metabolism 53(1), 54–58 (2004). 18 des Rieux A, Ragnarsson EG, Gullberg E, Préat V, Schneider YJ, Artursson P. Transport of nanoparticles across an in vitro model of the human intestinal follicle associated epithelium. Eur. J. Pharm. Sci. 25(4–5), 455–465 (2005). 19 Skyler JS, Krischer JP, Wolfsdorf J et al. Effects of oral insulin in relatives of patients with type 1 diabetes: The Diabetes Prevention Trial – Type 1. Diabetes Care 28(5), 1068–1076 (2005). 20 Cilek A, Celebi N, Tirnaksiz F. Lecithin- based microemulsion of a peptide for oral administration: preparation, characterization, and physical stability of the formulation. Drug Deliv. 13, 19–24 (2006). 21 Sajeesh S, Sharma CP. Cyclodextrin-insulin complex encapsulated polymethacrylic acid based nanoparticles for oral insulin delivery. Int. J. Pharm. 325, 147–154 (2006). 22 Sarmento B, Ferreira D, Jorgensen L, van de Weert M. Probing insulin’s secondary structure after entrapment into alginate– chitosan nanoparticles. Eur. J. Pharm. Biopharm. 65, 10–17 (2007). 23 Bayat A, Larijani B, Ahmadian S, Junginger HE, Rafiee-Tehrani M. Preparation and characterization of insulin nanoparticles using chitosan and its quaternized derivatives. Nanomed. Nanotechnol. Biol. Med. 4, 115–120 (2008). 24 Woitiski CB, Veiga F, Ribeiro AJ, Neufeld RJ. Design for optimization of nanoparticles integrating biomaterials for orally dosed insulin. Eur. J. Pharm. Biopharm. 73, 25–33 (2009). 25 Woitiski CB, Neufeld RJ, Veiga F, Carvalho RA, Figueiredo IV. Pharmacological effect of orally delivered insulin facilitated by multilayered stable nanoparticles. Eur. J. Pharm. Sci. 41, 556–563 (2010). 26 Woitiski CB, Sarmento B, Carvalho RA, Neufeld RJ, Veiga F. Facilitated nanoscale delivery of insulin across intestinal membrane models. Int. J. Pharm. 412, 123–131 (2011). 27 Florence AT. Issues in oral nanoparticle drug carrier uptake and targeting. J. Drug Target. 12(2), 65–70 (2004). 28 Löbenberg R, Amidon GL. Modern bioavailability, bioequivalence and biopharmaceutics classification system. New scientific approaches to international regulatory standards. Eur. J. Pharm. Biopharm. 50, 3–12 (2000). 29 Bastian SE, Walton PE, Ballard FJ, Belford DA. Transport of IGF-I across epithelial cell monolayers. J. Endocrinol. 162, 361–369 (1999). 30 Sarmento B, Ribeiro A, Veiga F, Sampaio P, Neufeld RJ, Ferreira D. Alginate–chitosan nanoparticles are effective for oral insulin delivery. Pharm. Res. 24(12), 2198–2206 (2007). 31 Agarwal V, Khan MA. Current status of the oral delivery of insulin. Pharm. Technol. 10, 76–90 (2001). 32 Guggi D, Kast CE, Bernkop-Schnürch A. In vivo evaluation of an oral salmon calcitonin-delivery system based on a thiolated chitosan carrier matrix. Pharm. Res. 20(12), 1989–1994 (2003). 33 Desai MA, Mutlu M, Vadgama P. A study of macromolecular diffusion through native porcine mucus. Experientia 48, 22–26 (1992). 34 Norris DA, Puri N, Sinko PJ. The effect of physical barriers and properties on the oral absorption of particulates. Adv. Drug Deliv. Rev. 34(2–3), 135–154 (1998). 35 Washington N, Washington C, Wilson CG. Physiological Pharmaceutics. Barriers to Drug Absorption (2nd Edition). Washington N, Washington C, Wilson CG (Eds). Taylor and Francis, London, UK (2002). 36 Serra L, Domenech J, Peppas NA. Engineering design and molecular dynamics of mucoadhesive drug delivery systems as targeting agents. Eur. J. Pharm. Biopharm. 71(3), 519–528 (2009). 37 Bhalodia R, Basu B, Garala K, Joshi B, Mehta K. Buccoadhesive drug delivery systems: a review. Int. J. Pharm. Biosci. 1(2), 1–32 (2010). 38 Strous GJ, Dekker J. Mucin-type glycoproteins. Crit. Rev. Biochem. Mol. Biol. 27, 57–92 (1992). 39 Liu Z, Wang S, Hu M. Chapter 11: Oral absorption basics: pathways, physico- chemical and biological factors affecting absorption. In: Developing Solid Oral Dosage Forms: Pharmaceutical Theory and Practice. Qiu Y, Chen Y, Liu L, Zhang GGZ (Eds). Academic Press, London, UK (2009). 40 Salama NN, Eddington ND, Fasano A. Tight junction modulation and its relationship to drug delivery. Adv. Drug Deliv. Rev. 58(1), 15–28 (2006). 41 Fix AJ. Oral controlled release technology for peptides: status and future prospects. Pharm. Res. 13, 1760–1764 (1996). 42 Wacher VJ, Salphati L, Benet LZ. Active secretion and enterocytic drug metabolism barriers to drug absorption. Adv. Drug Deliv. Rev. 46, 89–102 (2001). 43 Jung T, Kamm W, Breitenbach A, Kaiserling E, Xiao JX, Kissel T. Biodegradable nanoparticles for oral delivery of peptides: is there a role for polymers to affect mucosal uptake? Eur. J. Pharm. Biopharm. 50(1), 147–160 (2000). 44 Yin L, Ding J, He C, Cui L, Tang C, Yin C. Drug permeability and mucoadhesion properties of thiolated trimethyl chitosan nanoparticles in oral insulin delivery. Biomaterials 30(29), 5691–5700 (2009). 45 van der Lubben IM, Verhoef JC, van Aelst AC, Borchard G, Junginger HE. Chitosan microparticles for oral vaccination: preparation, characterization and preliminary in vivo uptake studies in murine Peyer’s patches. Biomaterials 22(7), 687–694 (2001). 46 Marks DL, Gores GJ, LaRusso NF. Hepatic processing of peptides. In: Peptide-Based Drug Design: Controlling Transport and Metabolism. Taylor MD, Amidon GL. (Eds). American Chemical Society, Washington, DC, USA (1995). 47 Swaan PW. Recent advances in intestinal macromolecular drug delivery via receptor-mediated transport pathways. Pharm. Res. 15, 826–834 (1998). 48 Schilling RJ, Mitra AK. Degradation of insulin by trypsin and alpha-chymotrypsin. Pharm. Res. 8(6), 721–727 (1991). 49 Zhou XH. Overcoming enzymatic and absorption barriers to non-parenterally administered protein and peptide drugs. J. Control. Release 29, 239–252 (1994). 50 Calceti P, Salmaso S, Walker G, Bernkop-Schnurch A. Development and in vivo evaluation of an oral insulin–PEG delivery system. Eur. J. Pharm. Sci. 22, 315–323 (2004). 51 Aoki Y, Morishita M, Asai K, Akikusa B, Hosoda S, Takayama K. Region-dependent role of the mucous/glycocalyx layers in insulin permeation across rat small intestinal membrane. Pharm. Res. 22(11), 1854–1862 (2005).
  • 20. Review |Thompson & Ibie Therapeutic Delivery (2011) 2(12)1630 future science group 52 Thompson CJ, Tetley L, Uchegbu IF, Cheng WP. The complexation between novel comb shaped amphiphilic polyallylamine and insulin – towards oral insulin delivery. Int. J. Pharm. 376, 46–55 (2009). 53 Thompson CJ, Tetley L, Cheng WP. The influence of polymer architecture on the protective effect of novel comb shaped amphiphilic poly(allyl amine) against in vitro enzymatic degradation of insulin – towards oral insulin delivery. Int. J. Pharm. 383, 216–227 (2010). 54 Yamamoto A, Taniguchi T, Rikyuu K et al. Effects of various protease inhibitors on the intestinal absorption and degradation of insulin in rats. Pharm. Res. 11(10), 1496–1500 (1994). 55 Narayani R, Rao KP. Hypoglycemic effect of gelatin microspheres with entrapped insulin and protease inhibitor in normal and diabetic rats. Drug Deliv. 2(1), 29–38 (1995). 56 Mesiha M, Plakogiannis F, Vejosoth S. Enhanced oral absorption of insulin from desolvated fatty acid-sodium glycocholate emulsions. Int. J. Pharm. 111(3), 213–216 (1994). 57 Eaimtrakarn S, Rama Prasad YV, Ohno T et al. Absorption enhancing effect of labrasol on the intestinal absorption of insulin in rats. J. Drug Target. 10(3), 255–260 (2002). 58 Hodges G, Carr EA, Hazzard RA, Carr KE. Uptake and translocation of microparticles in small intestine. Digest. Dis. Sci. 40(5), 967–975 (1995). 59 Spangler RS. Insulin administration via liposomes. Diabetes Care 13(9), 911–922 (1990). 60 des Rieux A, Fievez V, Garinot M, Schneider YJ, Préat V. Nanoparticles as potential oral delivery systems of proteins and vaccines: a mechanistic approach. J. Control. Release 116(1), 1–27 (2006). 61 Damgé C, Michel C, Aprahamian M, Couvreur P, Devissaguet JP. Nanocapsules as carriers for oral peptide delivery. J. Control. Release 13(2–3), 233–239 (1990). 62 Damgé C, Vranckx H, Balschmidt P, Couvreur P. Poly(alkyl cyanoacrylate) nanospheres for oral administration of insulin. J. Pharm. Sci. 86(12), 1403–1409 (1997). 63 Dünnhaupt S, Barthelmes J, Hombach J, Sakloetsakun D, Arkhipova V, Bernkop- Schnürch A. Distribution of thiolated mucoadhesive nanoparticles on intestinal mucosa. Int. J. Pharm. 408(1–2), 191–199 (2011). 64 Hartig SM, Greene RR, DasGupta J et al. Multifunctional nanoparticulate polyelectrolyte complexes. Pharm. Res. 24(12), 2353–2369 (2007). 65 Sajeesh S, Bouchemal K, Sharma CP, Vauthier C. Surface-functionalized polymethacrylic acid based hydrogel microparticles for oral drug delivery. Eur. J. Pharm. Biopharm. 74, 209–218 (2010). 66 Sajeesh S, Bouchemal K, Sharma CP, Vauthier C. Thiol-functionalized polymethacrylic acid-based hydrogel microparticles for oral insulin delivery. Acta Biomaterialia 6, 3072–3080 (2010). 67 Qiu Y, Park K. Environment-sensitive hydrogels for drug delivery. Adv. Drug Deliv. Rev. 53(3), 321–339 (2001). 68 Ramkissoon-Ganorkar C, Liu F, Baudys M, Kim SW. Modulating insulin-release profile from pH/thermosensitive polymeric beads through polymer molecular weight. J. Control. Release 59(3), 287–298 (1999). 69 Takeuchi H, Matsui Y, Yamamoto H, Kawashima Y. Mucoadhesive properties of carbopol or chitosan-coated liposomes and their effectiveness in the oral administration of calcitonin to rats. J. Control. Release 86(2–3), 235–242 (2003). 70 Chalasani KB, Russell-Jones GJ, Jain AK, Diwan PV, Jain SK. Effective oral delivery of insulin in animal models using vitamin B12 -coated dextran nanoparticles. J. Control. Release 122(2), 141–150 (2007). 71 Wood KM, Stone G, Peppas NA. Wheat germ agglutinin functionalized complexation hydrogels for oral insulin delivery. Biomacromolecules 9, 1293–1298 (2008). 72 Soppimath KS, Aminabhavi TM, Kulkarni AR, Rudzinski WE. Biodegradable polymeric nanoparticles as drug delivery devices. J. Control. Release 70, 1–20 (2001). 73 Cui F, Shi K, Zhang L, Tao A, Kawashima Y. Biodegradable nanoparticles loaded with insulin–phospholipid complex for oral delivery: preparation, in vitro characterization and in vivo evaluation. J. Control. Release 114, 242–250 (2006). 74 Fan YF, Wang YN, Fan YG, Ma JB. Preparation of insulin nanoparticles and their encapsulation with biodegradable polyelectrolytes via the layer-by-layer adsorption. Int. J. Pharm. 324, 158–167 (2006). 75 Simon M, Behrens I, Dailey LA, Wittmar M, Kissel T. Nanosized insulin-complexes based on biodegradable amine-modified graft polyesters poly[(vinyl-3-(diethylamino)- propylcarbamate-co-(vinyl acetate)-co-(vinyl alcohol)]–graft–poly(l-lactic acid): protection against enzymatic degradation, interaction with Caco-2 cell monolayers, peptide transport and cytotoxicity. Eur. J. Pharm. Biopharm. 66(2), 165–172 (2007). 76 Thompson CJ, Ding C, Qu X et al. The effect of polymer architecture on the nano self-assemblies based on novel comb-shaped amphiphilic poly(allylamine). Colloid Polym. Sci. 286, 1511–1526 (2008). 77 Cheng WP, Thompson CJ, Ryan SM, Aguirre T, Tetley L, Brayden DJ. In vitro and in vivo characterisation of a novel peptide delivery system: amphiphilic polyelectrolyte–salmon calcitonin nanocomplexes. J. Control. Release 147, 289–297 (2010). 78 Thompson CJ, Cheng WP. Chemically modified polyelectrolytes for intestinal protein and peptide delivery. In: Peptide and Protein Delivery. Chris van der Walle (Ed.). Academic Press/Elsevier, San Diego, CA, USA (2011). nn Comprehensive perspective on the use of modified forms of chitosan and amphiphilic polymers in oral protein delivery. 79 Thompson CJ, Cheng WP, Gadad P et al. Uptake and transport of novel amphiphilic polyelectrolyte-insulin nanocomplexes by Caco-2 cells – towards oral insulin. Pharm. Res. 28, 886–896 (2011). 80 Gowthamarajan K, Kulkarni GT. Oral insulin – fact or fiction? Resonance 8(5), 38–46 (2003). 81 van de Weert M, Hennink WE, Jiskoot W. Protein instability in poly(lactic-co-glycolic acid) microparticles. Pharm. Res. 17, 1159–1167 (2000). 82 Mao S, Bakowsky U, Jintapattankit A, Kissel T. Self-assembled polyelectrolyte nanocomplexes between chitosan derivatives and insulin. J. Pharm. Sci. 95, 1035–1048 (2006). 83 Sarmento B, Ribeiro A, Veiga F, Ferreira D, Neufeld RJ. Insulin-loaded nanoparticles are prepared by alginate ionotropic pre-gelation followed by chitosan polyelectrolyte complexation. J. Nanosci. Nanotechnol. 7, 283–2841 (2007). 84 Shu S, Zhang X, Teng D, Wang Z, Li C. Polyelectrolyte nanoparticles based on water-soluble chitosan-poly(l-aspartic acid)-polyethylene glycol for controlled protein release. Carbohydr. Res. 344, 1197–1204 (2009). 85 Desai MP, Labhasetwar V, Walter E, Levy RJ, Amidon GL. The mechanism of uptake of biodegradable microparticles in Caco-2 cells is size dependent. Pharm. Res. 14(11), 1568–1573 (1997). 86 Bayat A, Dorkoosh FA, Dehpour AR et al. Nanoparticles of quaternized chitosan
  • 21. The oral delivery of proteins using interpolymer polyelectrolyte complexes | Review www.future-science.com 1631future science group derivatives as a carrier for colon delivery of insulin: ex vivo and in vivo studies. Int. J. Pharm. 356(1–2), 259–266 (2008). 87 Sarmento B, Ferreira D, Veiga F, Ribeiro A. Characterization of insulin-loaded alginate nanoparticles produced by ionotropic pre-gelation through DSC and FTIR studies. Carbohydr. Polym. 66, 1–7 (2006). 88 Sajeesh S, Sharma CP. Novel pH responsive polymethacrylic acid-chitosan-polyethylene glycol nanoparticles for oral peptide delivery. J. Biomed. Mater. Res. B Appl. Biomater. 76B, 298–305 (2006). 89 Sun W, Mao S, Mei D, Kissel T. Self- assembled polyelectrolyte nanocomplexes between chitosan derivatives and enoxaparin. Eur. J. Pharm. Biopharm. 69, 471–425 (2008). 90 Shu S, Wang X, Zhang X, Zhang X, Wang Z, Li C. Disulfide cross-linked biodegradable polyelectrolyte nanoparticles for the oral delivery of protein drugs. New J. Chem. 33, 1882–1887 (2009). 91 Mao S, Shuai X, Unger F, Wittmar M, Xie X, Kissel T. Synthesis, characterization and cytotoxicity of poly(ethylene glycol)-graft- trimethyl chitosan block copolymers. Biomaterials 26, 6343–6356 (2005). 92 Mao S, Germershaus O, Fischer D, Linn T, Schnepf R, Kissel T. Uptake and transport of PEG-grsft-trimethyl-chitosan copolymer– insulin nanocomplexes by epithelial cells. Pharm. Res. 22, 2058–2068 (2005). 93 Chen F, Zhang Z, Yuan F, Qin X, Wang M, Huang Y. In vitro and in vivo study of N-trimethyl chitosan nanoparticles for oral protein delivery. Int. J. Pharm. 349, 226–233 (2008). 94 Bernkop-Schnürch A. Thiomers: a new generation of mucoadhesive polymers. Adv. Drug Deliv. Rev. 57, 1569–1582 (2005). 95 Albrecht K, Bernkop-Schnürch A. Thiomers: forms, functions and applications to nanomedicine. Nanomedicine 2(1), 41–50 (2007). 96 Föger F, Schmitz T, Bernkop-Schnürch A. In vivo evaluation of an oral delivery system for P-gp substrates based on thiolated chitosan. Biomaterials 27(23), 4250–4255 (2006). 97 Sakloetsakun D, Iqbal J, Millotti G, Vetter A, Bernkop-Schnürch A. Thiolated chitosans: influence of various sulfhydryl ligands on permeation-enhancing and P-gp inhibitory properties. Drug Dev. Ind. Pharm. 37(6), 648–655 (2011). 98 Hu Y, Jiang X, Ding Y, Ge H, Yuan Y, Yang C. Synthesis and characterization of chitosan-poly(acrylic acid) nanoparticles. Biomaterials 23, 3193–3201 (2002). 99 El-Sherbiny IM. Synthesis, characterization and metal uptake capacity of a new carboxymethyl chitosan derivative. Eur. Polym. J. 45, 199–210 (2009). 100 El-Sherbiny IM. Enhanced pH-responsive carrier system based on alginate and chemically modified carboxymethyl chitosan for oral delivery of protein drugs: preparation and in vitro assessment. Carbohydr. Polym. 80, 1125–1136 (2010). 101 El-Sherbiny IM, Abdel-Bary EM, Harding DRK. Preparation and in vitro evaluation of new pH-sensitive hydrogel beads for oral delivery of protein drugs. J. Appl. Polym. Sci. 115(5), 2828–2837 (2010). 102 El-Sherbiny IM, Elmahdy MM. Preparation, characterization, structure and dynamics of carboxymethyl chitosan grafted with acrylic acid sodium salt. J. Appl. Polym. Sci. 118(4), 2134–2145 (2010). 103 Perera G, Greindl M, Palmberger T, Bernkop-Schnürch A. Insulin-loaded poly(acrylic acid)-cysteine nanoparticles: stability studies towards digestive enzymes in the intestine. Drug Deliv. 16, 254–260 (2009). 104 Thurow H, Geisen K. Stabilization of dissolved proteins against denaturation at hydrophobie interfaces. Diabetologia 27(2), 212–218 (1984). 105 Rahman NA, Mathiowitz E. Localization of bovine serum albumin in double-walled microspheres. J. Control. Release 94(1), 163–175 (2004). 106 Lowman AM, Cowans BA, Peppas NA. Investigation of interpolymer complexation in swollen polyelectolyte networks using solid-state NMR spectroscopy. J. Polym. Sci. B Polym. Phys. 38(21), 2823–2831 (2000). 107 Lankalapalli S, Kolapalli VRM. Polyelectrolyte complexes: a review of their applicability in drug delivery technology. Indian J. Pharm. Sci. 71(5), 481–487 (2009). 108 Woitiski CB, Neufeld RJ, Ribeiro AJ, Veiga F. Colloidal carrier integrating biomaterials for oral insulin delivery: influence of component formulation on physicochemical and biological parameters. Acta Biomater. 5, 2475–2484 (2009). 109 Jintapattanakit A, Junyaprasert VB, Mao S, Sitterberg J, Bakowsky U, Kissel T. Peroral delivery of insulin using chitosan derivatives: a comparative study of polyelectrolyte nanocomplexes and nanoparticles. Int. J. Pharm. 342, 240–249 (2007). 110 Dupuy B, Arien A, Minnot AP. FT-IR of membranes made with alginate/polylysine complexes. Variations with the mannuronic or guluronic content of the polysaccharides. Artif. Cells Blood Substit. Immobil. Biotechnol. 22(1), 71–82 (1994). 111 Foss AC, Peppas NA. Investigation of the cytotoxicity and insulin transport of acrylic-based copolymer protein delivery systems in contact with Caco-2 cultures. Eur. J. Pharm. Biopharm. 57, 447–455 (2004). 112 López JE, Peppas NA. Effect of poly (ethylene glycol) molecular weight and microparticle size on oral insulin delivery from P(MAA-g- EG) microparticles. Drug Dev. Ind. Pharm. 30(5), 497–504 (2004). 113 Besheer A, Wood KM, Peppas NA, Mäder K. Loading and mobility of spin-labeled insulin in physiologically responsive complexation hydrogels intended for oral administration. J. Control. Release 111, 73–80 (2006). 114 Wood KM, Stone G, Peppas NA. The effect of complexation hydrogels on insulin transport in intestinal epithelial cell models. Acta Biomaterialia 6, 48–56 (2010). 115 American Diabetes Association. Standards of medical care in diabetes – 2009. Current criteria for the diagnosis of diabetes. Diabetes Care 32(Suppl. 1), S6–S12 (2009). 116 Sugano K, Kansy M, Artursson P et al. Coexistence of passive and carrier-mediated processes in drug transport. Nat. Rev. Drug Discov. 9, 597–614 (2010). 117 Bae KH, Moon CW, Lee Y, Park TG. Intracellular delivery of heparin complexed with chitosan-g-poly(ethylene glycol) for inducing apoptosis. Pharm. Res. 26(1), 93–100 (2009).