23.10.2014 Views

Tailorable Trimethyl Chitosans as Adjuvant for ... - TI Pharma

Tailorable Trimethyl Chitosans as Adjuvant for ... - TI Pharma

Tailorable Trimethyl Chitosans as Adjuvant for ... - TI Pharma

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

TAILORABLE TRIMETHYL CHITOSANS<br />

AS ADJUVANT FOR INTRANASAL<br />

IMMUNIZA<strong>TI</strong>ON


The printing of this thesis w<strong>as</strong> financially supported by:<br />

J.E. Jurriaanse Stichting<br />

Utrecht Institute <strong>for</strong> <strong>Pharma</strong>ceutical Sciences<br />

ISBN: 978-90-39354292<br />

© 2010 RJ Verheul, Utrecht<br />

Cover: Drawings taken from ‘Poissons, écrevisses et crabs’ by Louis Renard. First<br />

published in 1719, the book describes exotic sea creatures (including a mermaid) from<br />

the Netherlands West-Indies, now Indonesia. Many thanks to Marijn van Hoorn,<br />

curator of the Teylers Museum, Haarlem<br />

The research presented in this thesis w<strong>as</strong> per<strong>for</strong>med under the framework of <strong>TI</strong><br />

<strong>Pharma</strong> project D5-106-1; Vaccine delivery: alternatives <strong>for</strong> conventional multiple<br />

injection vaccines<br />

Printed by: Vandenberg, Maarn


TAILORABLE TRIMETHYL CHITOSANS AS<br />

ADJUVANT FOR INTRANASAL IMMUNIZA<strong>TI</strong>ON<br />

Varieerbare <strong>Trimethyl</strong> Chitosanen als Adjuvans voor Intran<strong>as</strong>ale Vaccinatie<br />

(met een samenvatting in het Nederlands)<br />

Proefschrift<br />

ter verkrijging van de graad van doctor aan de Universiteit Utrecht op<br />

gezag van de rector magnificus, prof. dr. J.C. Stoof, ingevolge het besluit<br />

van het college voor promoties in het openbaar te verdedigen op<br />

maandag 8 november 2010 des middags te 2.30 uur<br />

door<br />

Rudolf Johannus Verheul<br />

geboren op 8 maart 1980 te Oss


Promotoren: Prof. dr. ir. W.E. Hennink<br />

Prof. dr. W. Jiskoot


Voor mijn ouders


Table of Contents<br />

Chapter 1<br />

General introduction<br />

Chapter 2<br />

Synthesis, characterization and in vitro biological properties of O-methyl free<br />

N,N,N,-trimethylated chitosan<br />

Chapter 3<br />

Influence of the degree of acetylation on the enzymatic degradation and in vitro<br />

biological properties of trimethylated chitosans<br />

Chapter 4<br />

Relationship between structure and adjuvanticity of trimethyl chitosan (TMC)<br />

structural variants in a n<strong>as</strong>al influenza vaccine<br />

Chapter 5A<br />

A step-by-step approach to study the influence of N-acetylation on the<br />

adjuvanticity of N,N,N-trimethyl chitosan (TMC) in an intran<strong>as</strong>al whole<br />

inactivated influenza virus vaccine<br />

Chapter 5B<br />

Maturation of human monocyte derived dendritic cells by trimethyl chitosan is<br />

correlated with its N-acetyl glucosamine (GlcNAc) content<br />

Chapter 6<br />

<strong>Tailorable</strong> thiolated trimethyl chitosans <strong>for</strong> covalently stabilized nanoparticles<br />

Chapter 7<br />

Covalently stabilized trimethyl chitosan-hyaluronic acid nanoparticles <strong>for</strong> n<strong>as</strong>al<br />

and intradermal vaccination<br />

Chapter 8<br />

Summary and future perspectives<br />

Appendices<br />

Affiliations of collaborating authors<br />

List of abbreviations<br />

Curriculum vitae<br />

List of publications<br />

Nederlandse samenvatting<br />

Dankwoord<br />

Page<br />

9<br />

25<br />

47<br />

69<br />

91<br />

113<br />

125<br />

149<br />

169<br />

183


CHAPTER 1<br />

GENERAL INTRODUC<strong>TI</strong>ON


General Introduction<br />

N<strong>as</strong>al Vaccination<br />

Ever since Edward Jenner’s successful inoculations with cowpox to prevent a potentially<br />

lethal smallpox infection in the end of the 18 th century, active vaccination h<strong>as</strong> proven to be the<br />

most (cost) effective tool in the fight against infectious dise<strong>as</strong>es. Active vaccination, or<br />

immunization, involves activation of the immune system by controlled exposure to a<br />

milder/inactivated <strong>for</strong>m of the pathogen causing the dise<strong>as</strong>e, or to components derived from it,<br />

thereby inducing (immunological) memory and <strong>for</strong>tifying the host’s response towards the real<br />

pathogen. Vaccines are usually made of live attenuated or inactivated pathogens (e.g. viruses<br />

or bacteria) or purified immunogenic protein(conjugate)s derived from these pathogens. After<br />

two hundred years, still, most vaccines are administered via parenteral injection due to the<br />

limited absorption and enhanced degradation of these large molecular structures when using<br />

alternative administration routes [1]. There is however a need <strong>for</strong> vaccines that can be<br />

administered via non-inv<strong>as</strong>ive other routes <strong>for</strong> re<strong>as</strong>ons pointed out below. Table I gives an<br />

overview of the (dis)advantages of the different immunization routes.<br />

Table I. Advantages and disadvantages of different immunization routes (adopted from Slütter et al.<br />

[2]).<br />

Immunization route Advantages Disadvantages<br />

Parenteral<br />

N<strong>as</strong>al<br />

Oral<br />

Pulmonary<br />

Dermal<br />

Powerful systemic immune<br />

response<br />

Accurate dosing<br />

Non-inv<strong>as</strong>ive<br />

Mucosal and systemic immune<br />

responses<br />

E<strong>as</strong>ily accessible<br />

Little antigen degradation<br />

(compared to oral)<br />

Non-inv<strong>as</strong>ive<br />

Mucosal and systemic immune<br />

responses<br />

Large surface area<br />

Non-inv<strong>as</strong>ive<br />

Mucosal and systemic immune<br />

responses<br />

Little degradation (compared to<br />

oral)<br />

Non or minimally inv<strong>as</strong>ive<br />

Large, e<strong>as</strong>ily accessible application<br />

area<br />

High density of immune cells in skin<br />

Mucosal and systemic immune<br />

responses<br />

Inv<strong>as</strong>ive<br />

Limited mucosal immune response<br />

Risk of contaminated needles<br />

Need <strong>for</strong> trained personnel<br />

Mucociliary clearance<br />

Inefficient uptake of antigen<br />

Application device needed<br />

Vaccine/antigen digestion in<br />

stomach and gut<br />

Inefficient uptake of antigen<br />

Mucosal tolerance<br />

Highly variable antigen delivery<br />

Dry powder inhaler or nebulizer<br />

needed<br />

Clearance from lungs<br />

May require inv<strong>as</strong>ive technology<br />

(e.g. tattooing, microneedles)<br />

Patch or application device needed<br />

Less established technology<br />

11


Chapter 1<br />

Of these alternatives, n<strong>as</strong>al vaccination in particular h<strong>as</strong> some interesting advantages. The<br />

nose is e<strong>as</strong>ily accessible, commonly used and accepted <strong>for</strong> drug administration (e.g. nosesprays).<br />

A relatively simple, painless, device can be used without the aid of trained personnel<br />

and the n<strong>as</strong>al environment is less harsh <strong>for</strong> vaccine components than the oral route [3]. Also,<br />

high numbers of antigen presenting cells (APCs) are present in the n<strong>as</strong>al mucosal linings<br />

mediating both mucosal and systemic immune responses against <strong>for</strong>eign pathogens that try to<br />

invade the human body through the respiratory tract. Importantly, this mucosal immunity,<br />

hardly induced by parenteral immunization, may highly contribute to overall protection<br />

against a <strong>for</strong>eign pathogen. The excretion of secretory immunoglobulins is not only limited to<br />

the area of antigen-exposure but can occur on mucosal surfaces across the body (e.g. n<strong>as</strong>al<br />

immunization against sexually transmitted human papilloma virus may thus be an interesting<br />

option) [2, 4, 5].<br />

All these potential benefits taken into consideration, it should be mentioned that currently<br />

only one n<strong>as</strong>al vaccine is on the market: a live attenuated influenza vaccine administered via a<br />

n<strong>as</strong>al spray, thereby mimicking the natural route of inv<strong>as</strong>ion of the pathogen. Although<br />

effective, immunization with live attenuated viruses is under debate because of their potential<br />

to mutate, thereby escaping the immune system and regaining their pathogenicity. As a<br />

consequence, elderly, small infants and immuno-compromised people (e.g. by AIDS) are<br />

excluded from such vaccines. Subunit vaccines consisting of better defined and characterized<br />

antigenic proteins cannot mutate into pathogenic <strong>for</strong>ms but are generally less immunogenic<br />

and need potent adjuvant(system)s to elicit an adequate immune response [3, 6]. As vaccine<br />

safety is now top priority, a number of hurdles needs to be tackled to make intran<strong>as</strong>al (i.n.)<br />

vaccination a success-story.<br />

Successful n<strong>as</strong>al vaccine delivery<br />

After administration of the vaccine to the nose with an appropriate device, several successive<br />

steps can be identified that should lead to an adequate immune response (Figure 1). In short, a<br />

successful i.n. vaccine <strong>for</strong>mulation h<strong>as</strong> to adhere to the mucosal surfaces of the n<strong>as</strong>al cavity and<br />

provide protection against the proteolytic degradation of the antigen in the n<strong>as</strong>al environment.<br />

Prolonging n<strong>as</strong>al residence time may be an important mode of action <strong>for</strong> adjuvants; normally<br />

rapid mucociliarly clearance of the antigen will limit contact of the antigen with the epithelial<br />

barrier. Next, sufficient uptake and/or transport of antigen through the epithelial barrier<br />

should be achieved. Macromolecules up to a certain size (some suggest 22 kDa [1]) may use<br />

12


General Introduction<br />

paracellular pathways via cellular tight junctions (that can be opened by penetration<br />

enhancers) to overcome the epithelial barrier [1].<br />

Figure 1. Schematic overview of the consecutive steps towards successful n<strong>as</strong>al vaccine delivery: 1)<br />

muco-adhesion; 2) antigen uptake, by M-cell transport; 3) delivery to and subsequent<br />

activation/maturation of DC; 4) induction of B- and T-cell responses. DC= dendritic cell, M-cell =<br />

microfold cell, Th cell = helper T cell (adopted from Slütter et al. [2]).<br />

Microfold (M) cells play an important role especially <strong>for</strong> particulate systems, since they are<br />

capable of transporting the antigen by transcytosis to n<strong>as</strong>al <strong>as</strong>sociated lymphoid tissue (NALT)<br />

[2-4]. NALT consists of agglomerates of cells involved in the initiation and execution of an<br />

immune response, like dendritic cells (DCs), T- and B-cells. Alternatively, DCs may establish<br />

contact with the antigen through close interaction with epithelial cells [3]. After transport<br />

through the epithelium, the antigen h<strong>as</strong> to be taken up by antigen presenting cells (APC), most<br />

likely DCs or macrophages, which subsequently should mature, migrate and interact with<br />

helper T (Th) cells. Here, DCs will present a peptide (epitope) of the (degraded) antigen via the<br />

13


Chapter 1<br />

major histocompatibility complex cl<strong>as</strong>s II (MHC II) to the Th cells. Upon recognition of the MHC<br />

II-peptide complex and co-stimulation of the APC, naïve Th cells differentiate into effector Th<br />

cells, which can be divided in two major subtypes: Th1 and Th2 cells. Th1 cells are mainly<br />

involved in activation and proliferation of the cellular immune system, i.e. stimulate the<br />

cytotoxic T cells (CTL). Th2 cells are involved in stimulation of B-cells to differentiate into<br />

pl<strong>as</strong>ma cells and incre<strong>as</strong>e the humoral immune responses i.e. the production of<br />

immunoglobulins (IgG) also named antibodies. Interestingly, mucosal B-cells may differentiate<br />

into mucosal pl<strong>as</strong>ma cells that secrete prote<strong>as</strong>e resistant secretory dimeric IgA (sIgA) into the<br />

lumen to prevent mucosal inv<strong>as</strong>ion [7].<br />

Ideally, adaptive immune responses elicited by vaccination should comprise both cellular<br />

and humoral immune responses and should result in long-lived specific T- and B-memory cells,<br />

<strong>as</strong> well <strong>as</strong> some readily available circulating antibodies [3]. Importantly, DC or APC signaling<br />

determines the fate of the naïve Th cell and this can be modulated by the use of delivery<br />

systems and/or adjuvants. Thus, a vaccine <strong>for</strong>mulation may not only enhance the antigen<br />

availability or incre<strong>as</strong>e the immune response (i.e. “the danger signal”) but can also influence<br />

the type of immune response (i.e. immunomodulation) [2, 3, 8].<br />

Polymeric carrier systems<br />

Polymeric carriers have the advantage over other delivery systems (e.g. liposomes, ISCOMs)<br />

because of their endless potential of chemical modifications thereby allowing fine-tuning of the<br />

physico-chemical properties of the carrier system [6]. Although many polymers have been<br />

studied <strong>for</strong> n<strong>as</strong>al immunization [4], most research h<strong>as</strong> been done with the synthetic polymer<br />

poly(lactide-co-glycolide) (PLGA) and chitosan (derivatives) which are obtained from naturally<br />

abundant chitin. Interestingly, chitosan and its derivatives are much more effective in eliciting<br />

immune responses in micro- or nanoparticulate <strong>for</strong>m than <strong>as</strong> plain polymer solution [9-12],<br />

most likely because of the particle’s resemblance to the original pathogen, their multimeric<br />

antigen presentation and improved protection of the antigen against degradation [7].<br />

Furthermore, particles are better taken up by APCs and can co-deliver antigen and adjuvant to<br />

the same cell (13). While PLGA is FDA approved <strong>for</strong> several therapeutic applications and is<br />

considered biodegradable and biocompatible in humans, chitosan derivatives are gaining<br />

interest because of their superior efficacy in mucosal antigen delivery [2, 6, 14].<br />

14


General Introduction<br />

Chitosan<br />

Chitin (poly β1→4 N-acetyl D-glucosamine) is the second most abundant natural biopolymer<br />

and is derived from exoskeletons of crustaceans, insects and cell walls of several fungi [15].<br />

Chitosan (Scheme 1) is a polysaccharide consisting of β1→4-D-glucosamine and β1→4 N-<br />

acetyl D-glucosamine units and is obtained from chitin by partial deacetylation, which makes it<br />

water-soluble in an acidic aqueous solution [1]. It can vary in size (molecular weight) and/or<br />

degree of acetylation and is used in many are<strong>as</strong> of drug delivery and tissue engineering [16]. It<br />

is approved <strong>for</strong> dietary applications in several countries and the FDA h<strong>as</strong> approved chitosan<br />

<strong>for</strong> use in wound dressings [16]. Chitosan’s muco-adhesiveness, penetration-enhancing<br />

abilities and its properties allowing the preparation of particles without the use of organic<br />

solvents h<strong>as</strong> led to many investigations on chitosan <strong>for</strong>mulations <strong>for</strong> mucosal vaccination [1, 5,<br />

17]. Additionally, the functional groups (amines and hydroxyls) allow many modifications<br />

depending on the <strong>for</strong>eseen application [15]. However, despite many studies on toxicity and<br />

biocompatibility, chitosan is not (yet) considered a GRAS (generally regarded <strong>as</strong> safe) material<br />

and no approval <strong>for</strong> use in drug delivery h<strong>as</strong> been given so far. One re<strong>as</strong>on might be that<br />

chitosan’s high chemical variability confuses regulatory authorities [16].<br />

Scheme 1. Chemical structure of chitosan. The degree of acetylation (x) is variable.<br />

Chitosan’s primary amine-groups have a pK a of around 6, resulting in low aqueous solubility<br />

and loss of penetration-enhancing abilities at neutral pH [1]. Carboxylation of the amines<br />

and/or hydroxyl moieties and N-PEGylation have proven to be useful strategies <strong>for</strong> improving<br />

the aqueous solubility of chitosan at neutral pH [18-20]. Other modifications encomp<strong>as</strong>s the<br />

introduction of quaternary amine groups thereby giving chitosan a permanent, pH<br />

independent positive charge and thus aqueous solubility also at neutral pH. While sometimes<br />

15


Chapter 1<br />

quaternary ammonium groups are introduced directly e.g. by coupling 2-diethylaminoethyl<br />

chloride to chitosan [21, 22], mostly quaternization is achieved by partial trimethylation of the<br />

amine resulting in N,N,N-trimethylated chitosan (TMC) [23].<br />

N,N,N,-<strong>Trimethyl</strong>ated chitosan (TMC)<br />

TMC (Scheme 2) h<strong>as</strong> been shown to have muco-adhesive properties [27] and is able to open<br />

tight junctions above a degree of quaternization (DQ) of 20% [25-29]. In addition, TMC h<strong>as</strong><br />

been used to complex and condense DNA to yield polyplexes <strong>for</strong> gene delivery purposes [30-<br />

32]. TMC is synthesized b<strong>as</strong>ed on the method first published by Domard and coworkers [33]<br />

and later modified by Sieval et al. [23]. They showed that alkylation of primary amines of<br />

chitosan occurred by reaction of this polymer in strong alkaline conditions with an excess of<br />

iodomethane at elevated temperature (60°C) using N-methyl-2-pyrrolidone (NMP) <strong>as</strong> solvent.<br />

These relatively vigorous reaction conditions also lead to polymer chain scission [34] <strong>as</strong> well<br />

<strong>as</strong> to partial and uncontrolled methylation of the C-3 and C-6 hydroxyl groups of chitosan [35,<br />

36].<br />

Scheme 2. Chemical structure of TMC. TMC can vary in degree of acetylation (x), quaternization (y),<br />

dimethylation (z) and O-methylation (z). The various substitutions are randomly distributed throughout<br />

the polymer; O-methylation (z) may also occur on the quaternized and acetylated units.<br />

Several studies have been per<strong>for</strong>med to determine the optimal DQ <strong>for</strong> either trans-epithelial<br />

delivery of low molecular weight drug molecules and/or proteins, or to incre<strong>as</strong>e the<br />

transfection potential of complexes of TMC with pl<strong>as</strong>mid DNA. A DQ of about 40-50% w<strong>as</strong><br />

found to be the optimum <strong>for</strong> transepithelial delivery of both low molecular weight compounds<br />

and proteins [28, 37-40]. Furthermore, TMC h<strong>as</strong> been used successfully <strong>for</strong> n<strong>as</strong>al immunization<br />

in mice [41-44]. Amidi et al. showed high levels of serum IgG and HI titers after i.n.<br />

administration of influenza A subunit encapsulated in TMC-tripolyphosphate (TPP)<br />

16


General Introduction<br />

nanoparticles <strong>as</strong> compared to plain antigen and antigen with TMC in solution [9]. Additionally,<br />

our results and those of others suggest that optimizing the DQ of TMC may lead to further<br />

improvement of the vaccine <strong>for</strong>mulation: Boonyo et al. proposed an optimal DQ of 40% b<strong>as</strong>ed<br />

upon an i.n. immunization with ovalbumin [41] and we found that TMC with a DQ of 37% may<br />

be superior to TMC with a DQ of 15% in an i.n. vaccination with whole inactivated influenza<br />

virus [44]. Importantly, in these studies the TMCs used also had a variable extent of O-<br />

methylation (DOM), degree of acetylation (DAc) and differences in polymer molecular weights.<br />

While incre<strong>as</strong>ing TMC polymer molecular weight leads to an incre<strong>as</strong>e in toxicity [45], the<br />

effects of the other polymer compositional variables are currently unknown. In particular the<br />

role of the DAc can be anticipated to be quite substantial: in chitosan the DAc is <strong>as</strong>sociated with<br />

its enzymatic degradability [16], penetration-enhancing capability [46] and stimulating effect<br />

on APCs [47-50]. Tailorability of these variables (DQ, DOM and DAc) will allow better<br />

understanding of the contributions of each of those side groups to the physico-chemical and<br />

biological properties of TMC. Additionally, the introduction of novel substitutions such <strong>as</strong> thiolmoieties<br />

may further broaden the potential pharmaceutical applications of TMC by enhancing<br />

its muco-adhesive potential and allowing further chemical derivatization reactions via<br />

reducible disulfide bridges [51-54].<br />

TMC nanoparticle preparation<br />

As mentioned above, nanoparticles composed of (subunit) antigen and TMC are more<br />

effective in i.n. vaccination than soluble antigen and polymer. In some c<strong>as</strong>es TMC is used <strong>as</strong><br />

coating <strong>for</strong> particulate systems (e.g. PLGA nanoparticles or whole inactivated influenza virus<br />

particle) but more frequently TMC nanoparticles are prepared by ionic crosslinking methods.<br />

The complexation between the positively charged TMC and oppositely charged<br />

(macro)molecules added drop-wise under stirring in low ionic strength buffer results in<br />

spontaneous <strong>for</strong>mation of nanoparticles. This ionic gelation method is simple and mild <strong>for</strong><br />

proteins <strong>as</strong> no chemical crosslinkers, organic solvents or elevated temperatures are required<br />

[1]. It h<strong>as</strong> been reported that is some c<strong>as</strong>es the antigen or therapeutic protein alone <strong>as</strong> such act<br />

<strong>as</strong> crosslinker to yield nanoparticles [55], but often an additional crosslinker is needed. Both<br />

small molecules such <strong>as</strong> tripolyphosphate (TPP) [9, 56-58] and larger anionic macromolecules<br />

like polyglutamic acid [59] and hyaluronic acid [60] have been successfully used <strong>for</strong><br />

nanoparticle preparation. However, the physico-chemical stability of these complexes is<br />

dependent on the characteristics of the crosslinker used and they often have a limited stability<br />

17


Chapter 1<br />

in physiological saline [60, 61] or when the pH drops [59, 62]. So, although TMC nanoparticles<br />

are successfully used in n<strong>as</strong>al vaccination, many variables such <strong>as</strong> the type of crosslinker used<br />

and particle stability leave room <strong>for</strong> improvement.<br />

TMC mode of action<br />

Several studies have shown the superiority of positively charged particles above negatively<br />

charged or neutral particles in n<strong>as</strong>al immunization. E.g. chitosan coated PLGA particles<br />

improved immunogenicity of ‘naked’ PLGA particles [63], TMC particles were superior in i.n.<br />

vaccination with tetanus toxoid <strong>as</strong> compared to negatively charged particles made from<br />

carboxylated chitosan [19] and TMC coated whole inactivated influenza virus resulted in<br />

protection against a live virus challenge, in contr<strong>as</strong>t with the uncoated virus which w<strong>as</strong> not<br />

effective [44]. Often, muco-adhesion due to charge interactions between polymer and mucus is<br />

suggested <strong>as</strong> potential mechanism of action <strong>for</strong> these penetration-enhancing polymers. It h<strong>as</strong><br />

further been postulated that muco-adhesive polymers incre<strong>as</strong>e n<strong>as</strong>al residence time and<br />

improve interaction of the antigen with the epithelial cell barrier leading to an incre<strong>as</strong>ed<br />

antigen uptake [1, 24, 43]. Alternatively, the opening of cellular tight junctions (<strong>as</strong> determined<br />

in a transepithelial electrical resistance (TEER) <strong>as</strong>say) may result in enhanced antigen/protein<br />

uptake [59].<br />

Chitosan and TMC are known to induce a rearrangement of cytoskeletal F-actin and tight<br />

junction protein ZO-1, leading to enhanced permeability of the epithelium [59, 64]. However,<br />

fully quaternized diethyl aminoethyl dextran w<strong>as</strong> ineffective <strong>as</strong> absorption enhancer indicating<br />

that solely cationic charge is not sufficient <strong>for</strong> a polymer to have adjuvant activity [65].<br />

Additionally, nanoparticles are considered too big to p<strong>as</strong>s through these junctions [1] and also<br />

non-TEER reducing TMCs can improve immunogenicity of i.n. vaccine <strong>for</strong>mulations [9, 44].<br />

Also, the direct adjuvant effect of TMC(-particles) on APCs or DCs h<strong>as</strong> hardly been studied, but<br />

<strong>as</strong> mentioned be<strong>for</strong>e, the N-acetyl glucosamine units may interact with C-type lectin receptors<br />

present on these cells [47-50]. Taken together, clearly, the exact mode of action of TMC is not<br />

yet fully understood.<br />

18


General Introduction<br />

Aim and thesis outline<br />

As pointed out in the previous paragraphs, TMCs are promising polymers <strong>for</strong> use in n<strong>as</strong>al<br />

immunization. However, many opportunities <strong>for</strong> optimizing TMC structure and TMC- b<strong>as</strong>ed<br />

nanoparticles still remain and mechanistic insights in the mode of action of TMC is lacking. The<br />

aim of this thesis is to develop synthetic routes to make TMC structural variants in a<br />

controllable and tailorable manner and introduce substitutions such <strong>as</strong> thiol-moieties that may<br />

further improve TMC’s properties. In this way, structure-activity relationships can be<br />

investigated in in vitro <strong>as</strong>says and in in vivo n<strong>as</strong>al vaccination studies. Also, novel covalently<br />

stabilized TMC-b<strong>as</strong>ed particles are prepared and investigated in in vivo immunization studies.<br />

Chapter 2 challenges the current synthetic method that introduces several side-reactions<br />

such <strong>as</strong> O-methylation and chain scission. A novel synthesis route is proposed that allows<br />

tailorability of the DQ without introducing other modifications. These novel TMC polymers are<br />

studied <strong>for</strong> physico-chemical properties, evaluated <strong>for</strong> cytotoxicity and the ability to open tight<br />

junctions, and compared with TMCs synthesized via the traditional method.<br />

Chapter 3 describes the synthesis of TMC with a high degree of acetylation (DAc) and<br />

investigates the lysozyme-catalyzed degradability of TMCs with different DAcs and with or<br />

without O-methylated groups, and evaluates these polymers <strong>for</strong> their physico-chemical<br />

properties, cytotoxicity and ability to open tight junctions.<br />

In Chapter 4 the adjuvant effect of several TMC types is investigated <strong>for</strong> i.n. administered<br />

whole inactivate influenza virus (WIV). In particular, O-methyl free TMCs with varying DQs,<br />

reacetylated O-methyl free TMC (TMC-RA), and conventional O-methylated TMCs with similar<br />

DQ are compared. The TMC-WIV vaccines are physicochemically characterized and the<br />

immunogenicity and protectivity of the vaccines are <strong>as</strong>sessed in a murine challenge model.<br />

Additionally, the influence of TMC:WIV ratio on the type and extent of humoral immune<br />

responses is investigated.<br />

To understand the differences in adjuvanticity of the different TMCs in Chapter 5A TMC-<br />

WIV, TMC-RA-WIV and WIV <strong>for</strong>mulations are compared at six potentially critical steps in the<br />

induction of an immune response after i.n. administration. In particular, (1) the degradation of<br />

TMC and TMC-RA in n<strong>as</strong>al w<strong>as</strong>hes, (2) n<strong>as</strong>al residence time, (3) n<strong>as</strong>al distribution patterns, (4)<br />

cellular uptake and (5) transport through an epithelial (Calu-3) cell line of WIV <strong>for</strong>mulated<br />

19


Chapter 1<br />

without or with TMC(-RA), and (6) the effect of the different <strong>for</strong>mulations on maturation of<br />

murine bone marrow derived dendritic cells (DCs) are studied. In Chapter 5B the effects of<br />

TMC and TMC-RA in solution and in combination with WIV on human monocyte-derived<br />

human DCs are investigated.<br />

Chapter 6 introduces a novel synthetic method to yield thiolated TMCs with a high DQ. The<br />

different thiolated TMCs are physico-chemically characterized, evaluated in in vitro<br />

cytotoxicity <strong>as</strong>says and used <strong>for</strong> preparation of covalently stabilized nanoparticles with<br />

thiolated hyaluronic acid.<br />

Physico-chemically characterized covalently stabilized and/or PEGylated TMC-hyaluronic<br />

acid nanoparticles are used in an intran<strong>as</strong>al and intradermal vaccination study in mice in<br />

Chapter 7. For comparison, non-stabilized particles are used <strong>as</strong> control. Both the type and the<br />

extent of the immune responses are investigated.<br />

Chapter 8 summarizes the findings and conclusions of this thesis and addresses the future<br />

opportunities <strong>for</strong> these novel TMCs and TMC-b<strong>as</strong>ed delivery systems.<br />

20


General Introduction<br />

References<br />

1. M. Amidi, E. M<strong>as</strong>trobattista, W. Jiskoot, and W. E. Hennink. Chitosan-b<strong>as</strong>ed delivery systems<br />

<strong>for</strong> protein therapeutics and antigens. Adv Drug Deliv Rev 62: 59-82 (2010).<br />

2. B. Slütter, N. Hagenaars, and W. Jiskoot. Rational design of n<strong>as</strong>al vaccines. J Drug Target 16: 1-17<br />

(2008).<br />

3. N. Csaba, M. Garcia-Fuentes, and M. J. Alonso. Nanoparticles <strong>for</strong> n<strong>as</strong>al vaccination. Adv Drug<br />

Deliv Rev 61: 140-157 (2009).<br />

4. S. Chadwick, C. Kriegel, and M. Amiji. Nanotechnology solutions <strong>for</strong> mucosal immunization.<br />

Adv Drug Deliv Rev 62: 394-407 (2009).<br />

5. L. Illum, I. Jabbal-Gill, M. Hinchcliffe, A. N. Fisher, and S. S. Davis. Chitosan <strong>as</strong> a novel n<strong>as</strong>al<br />

delivery system <strong>for</strong> vaccines. Adv Drug Deliv Rev 51: 81-96 (2001).<br />

6. N. Mishra, A. K. Goyal, S. Tiwari, R. Paliwal, S. R. Paliwal, B. Vaidya, S. Mangal, M. Gupta, D.<br />

Dube, A. Mehta, and S. P. Vy<strong>as</strong>. Recent advances in mucosal delivery of vaccines: Role of<br />

mucoadhesive/biodegradable polymeric carriers. Exp Opin Ther Patents 20: 661-679 (2010).<br />

7. M. R. Neutra and P. A. Kozlowski. Mucosal vaccines: The promise and the challenge. Nat Rev<br />

Immunol 6: 148-158 (2006).<br />

8. V. E. Schijns. Immunological concepts of vaccine adjuvant activity. Curr Opin Immunol 12: 456-<br />

463 (2000).<br />

9. M. Amidi, S. G. Romeijn, J. C. Verhoef, H. E. Junginger, L. Bungener, A. Huckriede, D. J. A.<br />

Crommelin, and W. Jiskoot. N-<strong>Trimethyl</strong> chitosan (TMC) nanoparticles loaded with influenza<br />

subunit antigen <strong>for</strong> intran<strong>as</strong>al vaccination: Biological properties and immunogenicity in a mouse<br />

model. Vaccine 25: 144-153 (2007).<br />

10. L. Feng, X. J. Zhou, and X. R. Qi. Preparation, rele<strong>as</strong>e and immunogenicity evaluation of<br />

HBsAg-PLGA microspheres. Beijing da xue xue bao. Yi xue ban = Journal of Peking University. Health<br />

sciences 37: 527-531 (2005).<br />

11. D. T. O'Hagan, M. Singh, and J. B. Ulmer. Microparticle-b<strong>as</strong>ed technologies <strong>for</strong> vaccines.<br />

Methods 40: 10-19 (2006).<br />

12. M. J. Shephard, D. Todd, B. M. Adair, A. Li Wan Po, D. P. Mackie, and E. M. Scott.<br />

Immunogenicity of bovine parainfluenza type 3 virus proteins encapsulated in nanoparticle<br />

vaccines, following intran<strong>as</strong>al administration to mice. Res Vet Sci 74: 187-190 (2003).<br />

13. A. C. Rice-Ficht, A. M. Aren<strong>as</strong>-Gamboa, M. M. Kahl-McDonagh, and T. A. Ficht. Polymeric<br />

particles in vaccine delivery. Curr Opin Microbiology 13: 106-112 (2010).<br />

14. B. Slütter, L. Plapied, V. Fievez, M. Alonso Sande, A. des Rieux, Y. J. Schneider, E. Van Riet,<br />

W. Jiskoot, and V. Préat. Mechanistic study of the adjuvant effect of biodegradable<br />

nanoparticles in mucosal vaccination. J Control Rele<strong>as</strong>e 138: 113-121 (2009).<br />

15. V. K. Mourya and N. N. Inamdar. Chitosan-modifications and applications: Opportunities<br />

galore. React Funct Polym 68: 1013-1051 (2008).<br />

16. T. Kean and M. Thanou. Biodegradation, biodistribution and toxicity of chitosan. Adv Drug<br />

Deliv Rev 62: 3-11 (2010).<br />

17. H. C. Arca, M. Günbeyaz, and S. Şenel. Chitosan-b<strong>as</strong>ed systems <strong>for</strong> the delivery of vaccine<br />

antigens. Exp Rev Vaccines 8: 937-953 (2009).<br />

18. Y. Ohya, R. Cai, H. Nishizawa, K. Hara, and T. Ouchi. Preparation of PEG-grafted chitosan<br />

nanoparticles <strong>as</strong> peptide drug carriers. S.T.P. <strong>Pharma</strong> Sciences 10: 77-82 (2000).<br />

19. B. Sayin, S. Somavarapu, X. W. Li, M. Thanou, D. Sesardic, H. O. Alpar, and S. Şenel. Mono-<br />

N-carboxymethyl chitosan (MCC) and N-trimethyl chitosan (TMC) nanoparticles <strong>for</strong> noninv<strong>as</strong>ive<br />

vaccine delivery. Int J Pharm 363: 139-148 (2008).<br />

20. M. Thanou, M. T. Nihot, M. Jansen, J. C. Verhoef, and H. E. Junginger. Mono-Ncarboxymethyl<br />

chitosan (MCC), a polyampholytic chitosan derivative, enhances the intestinal<br />

absorption of low molecular weight heparin across intestinal epithelia in vitro and in vivo. J<br />

Pharm Sci 90: 38-46 (2001).<br />

21


Chapter 1<br />

21. Y. Xu, Y. Du, R. Huang, and L. Gao. Preparation and modification of N-(2-hydroxyl) propyl-3-<br />

trimethyl ammonium chitosan chloride nanoparticle <strong>as</strong> a protein carrier. Biomaterials 24: 5015-<br />

5022 (2003).<br />

22. Y. Zambito, C. Zaino, G. Uccello-Barretta, F. Balzano, and G. Di Colo. Improved synthesis of<br />

quaternary ammonium-chitosan conjugates (N+-Ch) <strong>for</strong> enhanced intestinal drug permeation.<br />

Eur J Pharm Sci 33: 343-350 (2008).<br />

23. A. B. Sieval, M. Thanou, A. F. Kotze, J. C. Verhoef, J. Brussee, and H. E. Junginger.<br />

Preparation and NMR characterization of highly substituted N-trimethyl chitosan chloride.<br />

Carbohydr Polym 36: 157-165 (1998).<br />

24. V. Grabovac, D. Guggi, and A. Bernkop-Schnürch. Comparison of the mucoadhesive<br />

properties of various polymers. Adv Drug Deliv Rev 57: 1713-23 (2005).<br />

25. A. F. Kotzé, H. L. Lueßen, B. J. De Leeuw, B. G. De Boer, J. C. Verhoef, and H. E. Junginger.<br />

N-<strong>Trimethyl</strong> chitosan chloride <strong>as</strong> a potential absorption enhancer across mucosal surfaces: In<br />

vitro evaluation in intestinal epithelial cells (Caco-2). Pharm Res 14: 1197-1202 (1997).<br />

26. G. Sandri, M. C. Bonferoni, S. Rossi, F. Ferrari, C. Boselli, and C. Caramella. Insulin-loaded<br />

nanoparticles b<strong>as</strong>ed on N-trimethyl chitosan: In vitro (caco-2 model) and ex vivo (excised rat<br />

jejunum, duodenum, and ileum) evaluation of penetration enhancement properties. AAPS<br />

PharmSciTech 11: 362-371 (2010).<br />

27. D. Snyman, J. H. Hamman, and A. F. Kotze. Evaluation of the mucoadhesive properties of N-<br />

trimethyl chitosan chloride. Drug Develop Ind Pharm 29: 61-69 (2003).<br />

28. M. M. Thanou, A. F. Kotze, T. Scharringhausen, H. L. Lueßen, A. G. De Boer, J. C. Verhoef,<br />

and H. E. Junginger. Effect of degree of quaternization of N-trimethyl chitosan chloride <strong>for</strong><br />

enhanced transport of hydrophilic compounds across intestinal Caco-2 cell monolayers. J<br />

Control Rele<strong>as</strong>e 64: 15-25 (2000).<br />

29. I. M. Van der Lubben, J. C. Verhoef, M. M. Fretz, O. Van, I. Mesu, G. Kersten, and H. E.<br />

Junginger. <strong>Trimethyl</strong> chitosan chloride (TMC) <strong>as</strong> a novel excipient <strong>for</strong> oral and n<strong>as</strong>al<br />

immunisation against diphtheria. S.T.P. <strong>Pharma</strong> Sciences 12: 235-242 (2002).<br />

30. T. Kean, S. Roth, and M. Thanou. <strong>Trimethyl</strong>ated chitosans <strong>as</strong> non-viral gene delivery vectors:<br />

Cytotoxicity and transfection efficiency. J Control Rele<strong>as</strong>e 103: 643-653 (2005).<br />

31. Z. Mao, M. Lie, Y. Jiang, M. Yan, C. Gao, and J. Shen. N,N,N-trimethylchitosan chloride <strong>as</strong> a<br />

gene vector: Synthesis and application. Macromol Biosci 7: 855-863 (2007).<br />

32. M. Thanou, B. I. Florea, M. Geldof, H. E. Junginger, and G. Borchard. Quaternized chitosan<br />

oligomers <strong>as</strong> novel gene delivery vectors in epithelial cell lines. Biomaterials 23: 153-159 (2002).<br />

33. A. Domard, M. Rinaudo, and C. Terr<strong>as</strong>sin. New method <strong>for</strong> the quaternization of chitosan. Int J<br />

Bio Macromol 8: 105-107 (1986).<br />

34. D. Snyman, J. H. Hamman, J. S. Kotze, J. E. Rollings, and A. F. Kotze. The relationship<br />

between the absolute molecular weight and the degree of quaternisation of N-trimethyl chitosan<br />

chloride. Carbohydr Polym 50: 145-150 (2002).<br />

35. E. Curti, D. De Britto, and S. P. Campana-Filho. Methylation of chitosan with iodomethane:<br />

Effect of reaction conditions on chemoselectivity and degree of substitution. Macromol Biosci 3:<br />

571-576 (2003).<br />

36. A. Polnok, G. Borchard, J. C. Verhoef, N. Sarisuta, and H. E. Junginger. Influence of<br />

methylation process on the degree of quaternization of N-trimethyl chitosan chloride. Eur J<br />

Pharm Biopharm 57: 77-83 (2004).<br />

37. J. H. Hamman, C. M. Schultz, and A. F. Kotzé. N-trimethyl chitosan chloride: Optimum degree<br />

of quaternization <strong>for</strong> drug absorption enhancement across epithelial cells. Drug Develop Ind<br />

Pharm 29: 161-172 (2003).<br />

38. A. F. Kotze, M. M. Thanou, H. L. Luessen, A. B. G. De Boer, J. C. Verhoef, and H. E.<br />

Junginger. Effect of the degree of quaternization of N-trimethyl chitosan chloride on the<br />

permeability of intestinal epithelial cells (Caco-2). Eur J Pharm Biopharm 47: 269-274 (1999).<br />

22


General Introduction<br />

39. G. Sandri, S. Rossi, M. C. Bonferoni, F. Ferrari, Y. Zambito, G. Di Colo, and C. Caramella.<br />

Buccal penetration enhancement properties of N-trimethyl chitosan: Influence of quaternization<br />

degree on absorption of a high molecular weight molecule. Int J Pharm 297: 146-155 (2005).<br />

40. S. M. Van Der Merwe, J. C. Verhoef, J. H. M. Verheijden, A. F. Kotzé, and H. E. Junginger.<br />

<strong>Trimethyl</strong>ated chitosan <strong>as</strong> polymeric absorption enhancer <strong>for</strong> improved peroral delivery of<br />

peptide drugs. Eur J Pharm Biopharm 58: 225-235 (2004).<br />

41. W. Boonyo, H. E. Junginger, N. Waranuch, A. Polnok, and T. Pitaksuteepong. Chitosan and<br />

trimethyl chitosan chloride (TMC) <strong>as</strong> adjuvants <strong>for</strong> inducing immune responses to ovalbumin in<br />

mice following n<strong>as</strong>al administration. J Control Rele<strong>as</strong>e 121: 168-175 (2007).<br />

42. M. Amidi, S. G. Romeijn, G. Borchard, H. E. Junginger, W. E. Hennink, and W. Jiskoot.<br />

Preparation and characterization of protein-loaded N-trimethyl chitosan nanoparticles <strong>as</strong> n<strong>as</strong>al<br />

delivery system. J Control Rele<strong>as</strong>e 111: 107-116 (2006).<br />

43. B. C. Baudner, J. C. Verhoef, M. M. Giuliani, S. Peppoloni, R. Rappuoli, G. Del Giudice, and H.<br />

E. Junginger. Protective immune responses to meningococcal C conjugate vaccine after<br />

intran<strong>as</strong>al immunization of mice with the LTK63 mutant plus chitosan or trimethyl chitosan<br />

chloride <strong>as</strong> novel delivery plat<strong>for</strong>m. J Drug Target 13: 489-498 (2005).<br />

44. N. Hagenaars, E. M<strong>as</strong>trobattista, R. J. Verheul, I. Mooren, H. L. Glansbeek, J. G. M. Heldens,<br />

H. Van Den Bosch, and W. Jiskoot. Physicochemical and immunological characterization of<br />

N,N,N-trimethyl chitosan-coated whole inactivated influenza virus vaccine <strong>for</strong> intran<strong>as</strong>al<br />

administration. Pharm Res 26: 1353-1364 (2009).<br />

45. S. Mao, X. Shuai, F. Unger, M. Wittmar, X. Xie, and T. Kissel. Synthesis, characterization and<br />

cytotoxicity of poly(ethylene glycol)-graft-trimethyl chitosan block copolymers. Biomaterials 26:<br />

6343-6356 (2005).<br />

46. N. G. M. Schipper, K. M. Vårum, and P. Artursson. <strong>Chitosans</strong> <strong>as</strong> absorption enhancers <strong>for</strong><br />

poorly absorbable drugs. 1: Influence of molecular weight and degree of acetylation on drug<br />

transport across human intestinal epithelial (Caco-2) cells. Pharm Res 13: 1686-1692 (1996).<br />

47. K. Nishimura, S. Nishimura, and H. Seo. Macrophage activation with multi-porous beads<br />

prepared from partially deacetylated chitin. J Biomed Mat Res 20: 1359-1372 (1986).<br />

48. M. J. Robinson, D. Sancho, E. C. Slack, S. LeibundGut-Landmann, and C. R. Sousa. Myeloid C-<br />

type lectins in innate immunity. Nat Immunology 7: 1258-1265 (2006).<br />

49. J. Nadesalingam, A. W. Dodds, K. B. M. Reid, and N. Palaniyar. Mannose-binding lectin<br />

recognizes peptidoglycan via the N-acetyl glucosamine moiety, and inhibits ligand-induced<br />

proinflammatory effect and promotes chemokine production by macrophages. J Immunol 175:<br />

1785-1794 (2005).<br />

50. P. Zhang, S. Snyder, P. Feng, P. Azadi, S. Zhang, S. Bulgheresi, K. E. Sanderson, J. He, J.<br />

Klena, and T. Chen. Role of N-acetylglucosamine within core lipopolysaccharide of several<br />

species of Gram-negative bacteria in targeting the DC-SIGN (CD209). J Immunol 177: 4002-<br />

4011 (2006).<br />

51. K. Albrecht and A. Bernkop-Schnürch. Thiomers: Forms, functions and applications to<br />

nanomedicine. Nanomedicine 2: 41-50 (2007).<br />

52. A. Bernkop-Schnürch and A. Greimel. Thiomers: The next generation of mucoadhesive<br />

polymers. Amer J Drug Deliv 3: 141-154 (2005).<br />

53. T. M<strong>as</strong>uko, A. Minami, N. Iw<strong>as</strong>aki, T. Majima, S. I. Nishimura, and Y. C. Lee. Thiolation of<br />

chitosan. Attachment of proteins via thioether <strong>for</strong>mation. Biomacromol 6: 880-884 (2005).<br />

54. L. Yin, J. Ding, C. He, L. Cui, C. Tang, and C. Yin. Drug permeability and mucoadhesion<br />

properties of thiolated trimethyl chitosan nanoparticles in oral insulin delivery. Biomaterials 30:<br />

5691-5700 (2009).<br />

55. S. Mao, U. Bakowsky, A. Jintapattanakit, and T. Kissel. Self-<strong>as</strong>sembled polyelectrolyte<br />

nanocomplexes between chitosan derivatives and insulin. J Pharm Sci 95: 1035-1048 (2006).<br />

56. F. Chen, Z. R. Zhang, and Y. Huang. Evaluation and modification of N-trimethyl chitosan<br />

chloride nanoparticles <strong>as</strong> protein carriers. Int J Pharm 336: 166-173 (2007).<br />

23


Chapter 1<br />

57. F. Chen, Z. R. Zhang, F. Yuan, X. Qin, M. Wang, and Y. Huang. In vitro and in vivo study of<br />

N-trimethyl chitosan nanoparticles <strong>for</strong> oral protein delivery. Int J Pharm 349: 226-233 (2008).<br />

58. G. Sandri, M. C. Bonferoni, S. Rossi, F. Ferrari, S. Gibin, Y. Zambito, G. Di Colo, and C.<br />

Caramella. Nanoparticles b<strong>as</strong>ed on N-trimethylchitosan: Evaluation of absorption properties<br />

using in vitro (Caco-2 cells) and ex vivo (excised rat jejunum) models. Eur JPharm Biopharm 65:<br />

68-77 (2007).<br />

59. F. L. Mi, Y. Y. Wu, Y. H. Lin, K. Sonaje, Y. C. Ho, C. T. Chen, J. H. Juang, and H. W. Sung.<br />

Oral delivery of peptide drugs using nanoparticles self-<strong>as</strong>sembled by poly(γ-glutamic acid) and a<br />

chitosan derivative functionalized by trimethylation. Bioconj Chem 19: 1248-1255 (2008).<br />

60. S. Boddohi, N. Moore, P. A. Johnson, and M. J. Kipper. Polysaccharide-b<strong>as</strong>ed polyelectrolyte<br />

complex nanoparticles from chitosan, heparin, and hyaluronan. Biomacromol 10: 1402-1409<br />

(2009).<br />

61. D. V. Pergushov, H. M. Buchhammer, and K. Lunkwitz. Effect of a low-molecular-weight salt<br />

on colloidal dispersions of interpolyelectrolyte complexes. Colloid Polym Sci 277: 101-107 (1999).<br />

62. A. Bernkop-Schnürch, A. Weithaler, K. Albrecht, and A. Greimel. Thiomers: Preparation and in<br />

vitro evaluation of a mucoadhesive nanoparticulate drug delivery system. Int J Pharm 317: 76-81<br />

(2006).<br />

63. K. S. Jaganathan and S. P. Vy<strong>as</strong>. Strong systemic and mucosal immune responses to surfacemodified<br />

PLGA microspheres containing recombinant Hepatitis B antigen administered<br />

intran<strong>as</strong>ally. Vaccine 24: 4201-4211 (2006).<br />

64. N. G. M. Schipper, S. Olsson, J. A. Hoogstraate, A. G. DeBoer, K. M. Vårum, and P.<br />

Artursson. <strong>Chitosans</strong> <strong>as</strong> absorption enhancers <strong>for</strong> poorly absorbable drugs 2: Mechanism of<br />

absorption enhancement. Pharm Res 14: 923-929 (1997).<br />

65. G. Di Colo, S. Burgal<strong>as</strong>si, Y. Zambito, D. Monti, and P. Chetoni. Effects of different N-<br />

trimethyl chitosans on in vitro/in vivo ofloxacin transcorneal permeation. J Pharm Sci 93: 2851-<br />

2862 (2004).<br />

24


CHAPTER 2<br />

SYNTHESIS, CHARACTERIZA<strong>TI</strong>ON AND IN<br />

VITRO BIOLOGICAL PROPER<strong>TI</strong>ES OF O-<br />

METHYL FREE N,N,N-TRIMETHYLATED<br />

CHITOSAN<br />

Rolf J. Verheul, Maryam Amidi, Steffen van der Wal,<br />

Elly van Riet, Wim Jiskoot, Wim E. Hennink.<br />

Biomaterials 2008, 29, 3642-3649


Chapter 2<br />

Abstract<br />

N,N,N-trimethylated chitosan (TMC) with varying degrees of quaternization (DQs) is<br />

currently being investigated in mucosal drug, vaccine and in gene delivery. However, besides<br />

N-methylation, also O-methylation and chain scission occur during the synthesis of this<br />

polymer. Since both side reactions may affect the polymer characteristics, there is a need <strong>for</strong><br />

TMCs without O-methylation and disparities in chain lengths while varying the DQ. In this<br />

study, O-methyl free TMC with varying DQs w<strong>as</strong> successfully synthesized by using a two-step<br />

method. First, chitosan w<strong>as</strong> quantitatively dimethylated using <strong>for</strong>mic acid and <strong>for</strong>maldehyde.<br />

Then, in presence of an excess amount of iodomethane, TMC w<strong>as</strong> obtained with different DQs<br />

by varying reaction time. TMC obtained by this two-step method showed no detectable O-<br />

methylation ( 1 H-NMR) and a slight incre<strong>as</strong>e in molecular weight with incre<strong>as</strong>ing DQ (GPC),<br />

implying that no chain scission occurred during synthesis. The solubility in aqueous solutions<br />

at pH 7 of O-methyl free TMC with DQ < 22% w<strong>as</strong> less <strong>as</strong> compared to O-methylated TMC with<br />

the same DQ. On the other hand, O-methyl free TMC with DQ > 30% had a good aqueous<br />

solubility. On Caco-2 cells O-methyl free TMCs demonstrated a larger decre<strong>as</strong>e in transepithelial<br />

electrical resistance (TEER) than O-methylated TMCs. Also, with incre<strong>as</strong>ing DQ, an<br />

incre<strong>as</strong>e in cytotoxicity (MTT) and membrane permeability (LDH) w<strong>as</strong> observed.<br />

26


Synthesis and Characterization of O-methyl Free TMC<br />

Introduction<br />

Chitosan is a polysaccharide consisting of β1→4-D-glucosamine and β1→4 N-acetyl D-<br />

glucosamine units and is obtained by partial deacetylation of the natural polymer chitin. In<br />

recent years, chitosan h<strong>as</strong> been under investigation <strong>for</strong> various biomedical and pharmaceutical<br />

applications [1-4]. However, its poor aqueous solubility and loss of penetration-enhancing<br />

activity above pH 6 is a major drawback <strong>for</strong> its use at physiological conditions [5]. Partial<br />

quaternization of chitosan’s primary amine groups h<strong>as</strong> been used to obtain a chitosan<br />

derivative that is soluble at physiological conditions. N,N,N-trimethylated chitosan (TMC) h<strong>as</strong><br />

been shown to have muco-adhesive properties and is able to open tight junctions above a<br />

certain degree of quaternization (DQ) [6-10]. In addition, TMC h<strong>as</strong> been used to complex and<br />

condense DNA to yield polyplexes <strong>for</strong> gene delivery purposes [11, 12].<br />

Without exceptions, TMC is synthesized b<strong>as</strong>ed on the method first published by Domard and<br />

coworkers [13] and later modified by Sieval et al [7]. They showed the alkylation of primary<br />

amines of chitosan by reaction of this polymer in strong alkaline conditions with an excess of<br />

iodomethane using N-methyl-2-pyrrolidone (NMP) <strong>as</strong> solvent. The relatively vigorous reaction<br />

conditions lead to polymer chain scission [14] and, importantly, partial and uncontrolled<br />

methylation of the C-3 and C-6 hydroxyl groups of chitosan [15, 16]. Furthermore, the DQ<br />

proved difficult to control and often multiple reaction steps are required to obtain the desired<br />

TMC [17].<br />

Several studies have been per<strong>for</strong>med to determine the optimal DQ <strong>for</strong> either transepithelial<br />

delivery of low molecular weight drug molecules and/or proteins, or to incre<strong>as</strong>e the<br />

transfection potential of complexes of TMC with pl<strong>as</strong>mid DNA. It h<strong>as</strong> been reported that a DQ<br />

of about 40-50% is the optimum <strong>for</strong> transepithelial delivery of both low molecular weight<br />

compounds [17, 18] and proteins [19]. In these studies the TMCs used also had a variable<br />

extent of O-methylation and disparities in polymer chain lengths. Since O-methylation and<br />

variations in polymer chain length may affect the physicochemical properties and, likely, also<br />

the biological properties of TMC, there is a need <strong>for</strong> a synthetic method that yields TMC<br />

without O-methylated groups and prevents chain scission.<br />

Muzzarelli and Tanfani reported on the synthesis of TMC using iodomethane and N-dimethyl<br />

chitosan (DMC) obtained by reaction of chitosan with <strong>for</strong>maldehyde and sodium borohydride<br />

[20]. They showed that up to 60% of the amine groups could be trimethylated by this method,<br />

but no investigations were done on the tailorability of the DQ. Interestingly, this two-step<br />

method likely prevents chain scission and deacetylation of remaining N-acetyl groups, and<br />

might result in TMC without O-methylation. Nevertheless, this w<strong>as</strong> not investigated by these<br />

27


Chapter 2<br />

researchers. Recently, several adjustments to this method were presented by Jia et al. [21] and<br />

Guo et al. [22] introducing sodium hydroxide in the second step to incre<strong>as</strong>e the degree of<br />

substitution. However, the use of a strong b<strong>as</strong>e when trimethylating DMC will very likely result<br />

in O-methylation. So far, TMC synthesized by these methods h<strong>as</strong> only been studied <strong>for</strong> its<br />

antibacterial activity [20-22].<br />

The aim of this study w<strong>as</strong> to investigate a two-step method to synthesize TMC with tailorable<br />

DQ avoiding O-methylation and chain scission <strong>as</strong> side reactions. The synthesized TMC with<br />

different DQs were studied <strong>for</strong> physico-chemical properties, evaluated <strong>for</strong> cytotoxicity and the<br />

ability to open tight junctions, and compared with TMC synthesized via the traditional method.<br />

Materials and Methods<br />

Materials. Chitosan with a residual degree of acetylation of 17% (determined by NMR) and<br />

M n of 25 kDa, M w of 42 kDa (determined by Viscotek triple detection system <strong>as</strong> described<br />

below) w<strong>as</strong> purch<strong>as</strong>ed from Primex (Siglufjordur, Iceland). N-methyl-2-pyrrolidone (NMP),<br />

<strong>for</strong>maldehyde (37% stabilized with methanol), <strong>for</strong>mic acid, thiazolyl blue tetrazolium bromide<br />

(MTT), sodium acetate, acetic acid (anhydrous), sodium hydroxide and hydrochloric acid were<br />

obtained from Sigma-Aldrich Chemical Co. Dulbecco’s Modified Eagle’s Medium (DMEM),<br />

Hank’s balanced salt solution (HBSS), Fetal calf serum (FCS) were obtained from Invitrogen<br />

(Breda, The Netherlands). Sodium dodecyl sulfate (SDS) w<strong>as</strong> ordered from Merck (Darmstadt,<br />

Germany). Iodomethane 99% stabilized with copper w<strong>as</strong> obtained from Acros Organics (Geel,<br />

Belgium). All other chemicals used were of analytical grade.<br />

Synthesis of dimethylated chitosan (DMC). The two-step reaction pathway to synthesize<br />

TMC avoiding O-methylation is depicted in scheme 1 and is b<strong>as</strong>ed on the method published by<br />

Muzzarelli and Tanfani [20], with some modifications. In detail, a <strong>for</strong>mic acid-<strong>for</strong>maldehyde<br />

methylation (Eschweiler-Clarke) w<strong>as</strong> used to synthesize N,N-dimethylated chitosan [23].<br />

Instead of sodium borohydride <strong>as</strong> the reducing agent [20-22], we used <strong>for</strong>mic acid that allows<br />

chitosan to dissolve in the aqueous solution without the use of an acetate buffer. Ten grams of<br />

chitosan w<strong>as</strong> transferred into a 500 ml roundbottom fl<strong>as</strong>k. Next, 30 ml of <strong>for</strong>mic acid w<strong>as</strong><br />

added followed by 40 ml of 37% <strong>for</strong>maldehyde and 180 ml of distilled water yielding a total<br />

volume of about 250 ml. A reflux condenser w<strong>as</strong> attached and the solution w<strong>as</strong> heated to 70 °C<br />

and stirred using a magnetic stirrer <strong>for</strong> 118 hours. Then, the slightly yellow, viscous solution<br />

w<strong>as</strong> evaporated under reduced pressure and 1 M NaOH solution w<strong>as</strong> used to incre<strong>as</strong>e the pH<br />

28


Synthesis and Characterization of O-methyl Free TMC<br />

to 12 at which gel <strong>for</strong>mation occurred. This gel w<strong>as</strong> w<strong>as</strong>hed with deionized water over a gl<strong>as</strong>s<br />

filter to remove impurities. Then, the DMC w<strong>as</strong> dissolved in deionized water at pH 4 (adjusted<br />

with 1 M HCl), filtered over a gl<strong>as</strong>s filter and dialyzed against deionized water <strong>for</strong> three days<br />

(changing buffer twice-daily). Finally, the product w<strong>as</strong> filtered through a 0.8 µm filter and<br />

freeze dried.<br />

Synthesis of trimethylated chitosan from DMC. DMC w<strong>as</strong> reacted with iodomethane to<br />

yield TMC following the method described by Muzzarelli, with some modifications (see Scheme<br />

1) [20]. To prevent O-methylation, the reaction of DMC with iodomethane w<strong>as</strong> done in NMP,<br />

without the addition of a b<strong>as</strong>e catalyst [21, 22]. In detail, 250 mg of DMC w<strong>as</strong> dissolved in 40<br />

ml deionized water and the pH w<strong>as</strong> adjusted to 11 with NaOH by which gel <strong>for</strong>mation<br />

occurred. This step is per<strong>for</strong>med to ensure deprotonation of the tertiary amino groups of the<br />

DMC. Then, the gel w<strong>as</strong> w<strong>as</strong>hed with water and finally three times with acetone. Next, DMC<br />

w<strong>as</strong> suspended in 50 ml NMP and 2 ml iodomethane w<strong>as</strong> added. The dispersion w<strong>as</strong> stirred at<br />

40°C <strong>for</strong> the desired time and subsequently dropped in 150 ml of an ethanol/diethyl ether<br />

mixture (50/50). The precipitate (TMC) w<strong>as</strong> isolated by centrifugation and w<strong>as</strong>hed extensively<br />

with diethyl ether. After drying overnight, the TMC w<strong>as</strong> dissolved in 100 ml of an aqueous 10%<br />

NaCl solution and put on a shaker <strong>for</strong> a minimum of 18 hours <strong>for</strong> ion-exchange. Finally, the<br />

TMC w<strong>as</strong> dialyzed against deionized water <strong>for</strong> 3 days changing buffer twice daily, filtered<br />

through a 0.8 µm filter and freeze dried. Analysis on NMR 500 MHz w<strong>as</strong> per<strong>for</strong>med to<br />

determine the degree of quaternization (DQ) and to confirm the absence of O-methylation.<br />

TMCs with different DQs were synthesized starting with one gram of DMC following the same<br />

procedure but using 3 ml of iodomethane, 100 ml of NMP and a 250 ml round bottom fl<strong>as</strong>k and<br />

varying reaction time.<br />

29


Chapter 2<br />

Scheme 1. Two-step synthetic pathway <strong>for</strong> the preparation of TMC avoiding O-methylation.<br />

Synthesis of O-methylated TMC from chitosan. TMC w<strong>as</strong> synthesized by methylation of<br />

chitosan with iodomethane in presence of an aqueous solution of NaOH essentially <strong>as</strong><br />

described previously [7]. In detail, 1 g of chitosan and 2.4 g of sodium iodide were added to a<br />

mixture of 40 ml of NMP and 6 ml of 15% w/v aqueous NaOH solution. Subsequently, the<br />

mixture w<strong>as</strong> heated to 60°C and after stirring <strong>for</strong> 20 min, 6 ml of methyl iodide w<strong>as</strong> added and<br />

30


Synthesis and Characterization of O-methyl Free TMC<br />

the reaction mixture w<strong>as</strong> refluxed <strong>for</strong> 60 min. The reaction w<strong>as</strong> stopped by dropping the<br />

mixture in a 200 ml mixture of diethyl ether and ethanol (50/50). The obtained precipitate<br />

w<strong>as</strong> w<strong>as</strong>hed extensively with diethyl ether. In this way, TMC with a DQ of about 20% and<br />

comparable O-methylation w<strong>as</strong> obtained (step 1). To synthesize TMC with a DQ of around 40-<br />

50%, be<strong>for</strong>e precipitation, 3 ml of 15% NaOH solution and 3 ml of iodomethane were added<br />

and the solution w<strong>as</strong> stirred <strong>for</strong> another 60 minutes be<strong>for</strong>e stopping the reaction <strong>as</strong> described<br />

above (step 1.5). TMCs with higher DQs (60% to 90%) were synthesized by dissolving the<br />

dried TMC (DQ ~ 20) together with 2.4 g of sodium iodide in 40 ml of NMP at 60°C.<br />

Subsequently, 5.5 ml of an aqueous 15% w/v NaOH solution w<strong>as</strong> added and, after stirring <strong>for</strong><br />

20 min 3.5 ml CH 3I w<strong>as</strong> added and the reaction w<strong>as</strong> done <strong>for</strong> 45 minutes under refluxing to<br />

yield TMC with a DQ of about 60 to 70% (step 2). To obtain TMC with a DQ of about 80 to 90%,<br />

after 45 minutes, 0.6 g NaOH and 1 ml iodomethane were added and the reaction w<strong>as</strong><br />

continued <strong>for</strong> 60 min at 60°C (step 2.5). The reaction w<strong>as</strong> terminated by precipitation the<br />

reaction mixture in 200 ml of a mixture of diethyl ether and ethanol (50:50) and the<br />

precipitate w<strong>as</strong> w<strong>as</strong>hed extensively with diethyl ether.<br />

Finally, the products were dissolved in 50 ml aqueous 10% w/v NaCl solution and put on a<br />

shaker <strong>for</strong> a minimum of 18 hours <strong>for</strong> ion-exchange. The obtained solution w<strong>as</strong> dialyzed at<br />

room temperature against deionized water <strong>for</strong> 3 days changing buffer twice daily, filtered<br />

through a 0.8 µm filter and freeze dried. Analysis on NMR 500MHz w<strong>as</strong> per<strong>for</strong>med to<br />

determine the DQ and the degree of O-methylation.<br />

31


Chapter 2<br />

Scheme 2. Synthetic pathway <strong>for</strong> the preparation of TMC according to the method of Sieval et al [7].<br />

Determination of the degrees of dimethylation, quaternization and O-methylation. The<br />

1H-NMR spectra of the DMC and the various TMCs were recorded with a Varian INOVA<br />

500MHz NMR spectrometer (Varian Inc., Palo Alto, Ca, USA) at 80°C in D 2O. The degree of<br />

dimethylation of the DMC w<strong>as</strong> calculated <strong>as</strong> follows:<br />

DDM = [(CH 3) 2]/[H2-H6] x 100<br />

Here, [(CH 3) 2] is the integral of the hydrogens of the dimethyl amino groups at 2.9 ppm and<br />

[H2-H6] is the integral corresponding the H-2 to H-6 protons between 4.0 and 3.2 ppm.<br />

32


Synthesis and Characterization of O-methyl Free TMC<br />

The DQ, degree of dimethylation (DM) and degree of 3- and 6-O-methylation (DOM-3 and<br />

DOM-6, respectively) of the TMCs were calculated according to previous described methods[7,<br />

15, 24].<br />

DQ = [[(CH 3) 3]/[H] × 1/9] × 100<br />

DM = [[(CH 3) 2]/[H] × 1/6] × 100<br />

DOM = [[CH 3]/[H] × 1/3] × 100<br />

Here, [(CH 3) 3], [(CH 3) 2] and [CH 3] are the integrals of the hydrogens of the trimethylated amino<br />

groups at 3.3 ppm, the dimethylated amino groups at 2.9 ppm and the methylated hydroxyl<br />

groups at either 3.4 (DOM-6) or 3.5 (DOM-3) ppm, respectively. [H] is the integral of the H-1<br />

peaks between 4.7 and 5.7 ppm; the signal related to the hydrogen atoms bound to the C-1’s of<br />

the TMC molecule. For the DMC and TMC synthesized with the mild method and a DQ below<br />

25% addition of 0.05 ml of DCl w<strong>as</strong> needed to dissolve the polymers in D 2O.<br />

Determination of M n and M w of chitosan and TMC. M n and M w of chitosan and the various<br />

TMCs were determined by gel permeation chromatography (GPC) on a Viscotek-triple<br />

detection system using a Shodex OHPak SB-806 column (30 cm) and 0.3 M sodium acetate pH<br />

4.4 (adjusted with acetic acid) <strong>as</strong> running buffer [25]. To remove residual water, chitosan and<br />

the TMC samples were dried in a vacuum oven at 40°C overnight. Then, the samples were<br />

dissolved overnight in the running buffer at a concentration of 5 mg/ml, filtered through a 0.2<br />

µm filter and injected (50 µl); the flow rate w<strong>as</strong> 0.7 ml/min. Data from the l<strong>as</strong>er photometer (λ<br />

= 670 nm) (right (90 0 ) and low (7 0 ) angle light scattering), refracting index detector and<br />

viscosity detector were integrated using the provided Viscotek-software to calculate the M n,<br />

M w and dn/dc of the different samples. Pullulan (M n = 102 kDa, M w = 106 kDa) obtained from<br />

Viscotek Benelux (Oss, the Netherlands) w<strong>as</strong> used <strong>for</strong> calibration.<br />

Water solubility of chitosan and various TMCs. Aqueous solubility of the different<br />

polymers w<strong>as</strong> determined at pH 7 at room temperature. First the polymers were dissolved<br />

overnight in a 0.5% acetic acid solution at 2.5 mg/ml. Then the pH w<strong>as</strong> adjusted to 7 using 1 M<br />

NaOH and the transmittance of the solutions at 500 nm w<strong>as</strong> me<strong>as</strong>ured on an UV/VIS<br />

spectrophotometer (UV-2450, Shimadzu, Japan). The polymers were considered insoluble<br />

when the transmittance w<strong>as</strong> less than 90% compared to the transmittance of 0.5% acetic acid<br />

solution [26].<br />

33


Chapter 2<br />

Transepithelial electrical resistance (TEER) me<strong>as</strong>urements. Caco-2 cells were seeded at<br />

a density of 2x10 5 cells per well on 12-transwell plates with a microporous membrane. The<br />

cells were cultured in Dulbecco’s Modified Eagle’s Medium (DMEM) <strong>for</strong> 10 days until a<br />

confluent cell layer w<strong>as</strong> <strong>for</strong>med. The medium w<strong>as</strong> replaced by Hank’s Balanced Salt Solution<br />

(HBSS) at the b<strong>as</strong>olateral side 10 minutes be<strong>for</strong>e the start of the experiments. Then, 0.5 ml<br />

solution of TMC (with various DQs, with or without O-methylation, dissolved (2 mg/ml) in<br />

HBSS, pH adjusted to 7 with 0.1 M NaOH) w<strong>as</strong> applied at the apical site of the cell monolayers.<br />

SDS (10 mg/ml) w<strong>as</strong> used <strong>as</strong> positive control and HBSS <strong>as</strong> reference. The resistance me<strong>as</strong>ured<br />

of the membrane without cells w<strong>as</strong> used <strong>as</strong> blank. The TEER of the Caco-2 cells w<strong>as</strong> me<strong>as</strong>ured<br />

with a Millicell-ERS (Millipore, Billerica, USA) me<strong>as</strong>uring device at certain time points (0, 15,<br />

30, 45, 60 and 90 min.) after addition of the stimuli. After 90 minutes the cells were w<strong>as</strong>hed<br />

with HBSS and incubation of the cells w<strong>as</strong> continued in DMEM <strong>for</strong> 24 hours at 37°C, CO 2 5% to<br />

determine the recovery of the TEER [24, 27].<br />

MTT cell toxicity <strong>as</strong>say. Caco-2 cells were seeded in a 96-well plate at a density of 4x10 4<br />

cells per well and incubated <strong>for</strong> 2 days at 37°C, CO 2 5% in culture medium (DMEM, high<br />

glucose, 10% FCS, L-glutamine, pyruvate, non essential amino acids). The medium w<strong>as</strong><br />

removed and the cells were incubated <strong>for</strong> 2.5 hours with 100 µl TMC solutions in HBSS (TMC<br />

concentrations were 0.1, 1 and 10 mg/ml, pH set at 7 with 0.1 M NaOH). SDS (10 mg/ml) w<strong>as</strong><br />

used <strong>as</strong> positive control and HBSS <strong>as</strong> reference <strong>for</strong> 100% cell viability. Thereafter, the HBSS<br />

w<strong>as</strong> removed and the cells were w<strong>as</strong>hed with phosphate buffered saline. One hundred µl of a<br />

freshly prepared solution of 0.5 mg/ml MTT in DMEM, without any additions, w<strong>as</strong> added and<br />

the cells were incubated <strong>for</strong> 3 hours at 37°C and 5% CO 2. Subsequently, the wells were<br />

emptied, 100 µl of DMSO w<strong>as</strong> used to dissolve the <strong>for</strong>med <strong>for</strong>mazan crystals and the<br />

absorbance w<strong>as</strong> read at 595 nm [28].<br />

LDH <strong>as</strong>say. Caco-2 cells were seeded in a 96-well plate at a density of 4x10 4 cells per well<br />

and incubated <strong>for</strong> 2 days at 37°C, CO 2 5% in culture medium (see section on MTT <strong>as</strong>say <strong>for</strong><br />

composition). The medium w<strong>as</strong> removed and the cells were w<strong>as</strong>hed with HBSS and incubated<br />

<strong>for</strong> 2.5 hours with 100 µl TMC solutions in HBSS (TMC concentrations were 0.1, 1 and 10<br />

mg/ml, pH set at 7 with 0.1 M NaOH). After incubation, the concentration of LDH present in the<br />

supernatant of the samples w<strong>as</strong> determined with the Cytotoxicity Detection Kit-Plus (Roche<br />

Diagnostics, Mannheim, Germany) by me<strong>as</strong>uring absorbance at 490 nm with 650 nm <strong>as</strong> a<br />

reference wavelength. A calibration curve w<strong>as</strong> made with the lysis buffer provided by the<br />

34


Synthesis and Characterization of O-methyl Free TMC<br />

manufacturer, setting the LDH concentration me<strong>as</strong>ured with the undiluted lysis buffer at 100%<br />

LDH rele<strong>as</strong>e. HBSS w<strong>as</strong> used <strong>as</strong> a negative control.<br />

Results and discussion<br />

Synthesis and characterization. To synthesize O-methyl free TMC, chitosan w<strong>as</strong> first<br />

converted into DMC using the Eschweiler-Clarke reaction with <strong>for</strong>maldehyde and <strong>for</strong>mic acid<br />

(Figure 1). 1 H-NMR analysis showed that the obtained polymer had a degree of dimethylation<br />

of 83%. Since the chitosan used in this study had a degree of acetylation of around 17 % it can<br />

be concluded that the free amines were quantitatively dimethylated. In the next step, DMC<br />

dissolved in NMP w<strong>as</strong> converted into TMC using an excess amount of iodomethane. Figure 1<br />

shows that the degree of quaternization can be accurately tailored by varying the reaction time<br />

while keeping reaction temperature and DMC/CH 3I ratio constant. In other published<br />

procedures to synthesize TMC, sodium iodide w<strong>as</strong> added to the reaction mixture of chitosan<br />

and iodomethane [7, 21, 22]. However, the presence of sodium iodide (0.011 M) during the<br />

synthesis of TMC from DMC using CH 3I did not affect the obtained DQ in our studies (data not<br />

shown).<br />

degree of quaternization<br />

(%)<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

0 10 20 30 40 50 60 70<br />

reaction time (h)<br />

Figure 1. Effect of reaction time on the DQ of TMC. The reaction temperature w<strong>as</strong> 40°C. Error bars<br />

represent standard deviation of three independent syntheses.<br />

35


Chapter 2<br />

Figure 2. 1 H-NMR-spectra of TMC with a DQ of about 60% synthesized with a) one-step method under<br />

alkaline conditions and b) the two-step method presented in this article. Red oval indicates signals<br />

correlated to O-methylation.<br />

Typical 1 H-NMR spectra of TMC synthesized with the new procedure described in this paper<br />

and with the procedure described by Sieval et al. are shown in Figure 2. From the spectra it can<br />

be concluded that both TMCs have a DQ of about 60%. The spectrum of TMC synthesized<br />

36


Synthesis and Characterization of O-methyl Free TMC<br />

according to the procedure of Sieval et al. (Figure 2a) clearly shows O-methylation of the<br />

hydroxyl groups at the C-3 and C-6 of the glucosamine units (peaks observed at 3.5 and 3.4<br />

ppm, respectively). It w<strong>as</strong> calculated that DOM-6 and DOM-3 are 56 and 44%, respectively. The<br />

higher degree of methylation observed on the C-6 hydroxyl groups is likely because this<br />

hydroxyl group is less sterically hindered than the hydroxyl group on the C-3. An overview of<br />

the 1 H-NMR results of the O-methylated TMCs is presented in Table 1. As also reported by<br />

others [15, 16], both the DQ and the DOM-3 and DOM-6 incre<strong>as</strong>ed with the number of reaction<br />

steps, but the DQ and DOM proved hard to control. In contr<strong>as</strong>t, 1 H-NMR analysis of TMC with a<br />

DQ of 56% synthesized with the new method described in this paper shows no peaks at 3.4<br />

and 3.5 ppm (Figure 2b), demonstrating the absence of O-methylation.<br />

Table 1. Degree of quaternization (DQ), dimethylation (DM), O-methylation on C-6 (DOM-6) and on<br />

C-3 (DOM-3) of TMCs synthesized according to Sieval et al. [7] <strong>as</strong> determined by 1 H-NMR analysis.<br />

DQ DM DOM-6 DOM-3<br />

Step 1 22% 61% 18% 12%<br />

Step 1.5 50% 35% 45% 40%<br />

Step 2 61% 25% 56% 44%<br />

Step 2.5 86% 1% 76% 72%<br />

In line with the results of Snyman et al. [14], GPC analysis showed a slightly decre<strong>as</strong>ed<br />

molecular weight of TMC synthesized in presence of aqueous NaOH with incre<strong>as</strong>ing DQ (Table<br />

2). Under the alkaline reaction conditions hydrolysis of some glycosidic bonds linking the<br />

glucosamine units likely occurs, thus with incre<strong>as</strong>ing reaction time, the molecular weight and<br />

the polymer chain length of the O-methylated TMCs decre<strong>as</strong>es. Although acid hydrolysis of<br />

chitosan h<strong>as</strong> been reported [29], we show that the molecular weight incre<strong>as</strong>es after converting<br />

chitosan into DMC (Table 2) which demonstrates that no or only limited chain scission occurs<br />

during reaction. The slight incre<strong>as</strong>e in molecular weight can be <strong>as</strong>cribed to the addition of<br />

methyl groups to the molecule. Further, GPC analysis also shows that TMC synthesized from<br />

DMC shows a slight incre<strong>as</strong>e in molecular weight with incre<strong>as</strong>ing DQ (Table 2). There<strong>for</strong>e, <strong>as</strong><br />

expected, the trimethylation of DMC with the synthetic method described in this thesis is not<br />

<strong>as</strong>sociated with chain scission, and all O-methyl free TMCs will have the same polymer chain<br />

length. This is in contr<strong>as</strong>t with the ‘standard’ method to synthesize TMC <strong>as</strong> described by Sieval<br />

et al [7] where TMCs with varying DQ will also have discrepancies in polymer chain lengths.<br />

37


Chapter 2<br />

Table 2. Molecular weights and water solubility of various derivatives of chitosan.<br />

M n M w dn/dc Solubility in water<br />

at pH 7<br />

Chitosan 25 kDa 42 kDa 0.18 -<br />

TMC-OM 22% 34 kDa 56 kDa 0.15 +<br />

TMC-OM 50% 32 kDa 49 kDa 0.15 +<br />

TMC-OM 61% 31 kDa 49 kDa 0.14 +<br />

TMC-OM 86% 29 kDa 44 kDa 0.14 +<br />

DMC 28 kDa 57 kDa 0.16 -<br />

TMC 22% 31 kDa 60 kDa 0.16 -<br />

TMC 30% 33 kDa 59 kDa 0.15 +<br />

TMC 43% 36 kDa 75 kDa 0.15 +<br />

TMC 56% 37 kDa 78 kDa 0.15 +<br />

TMC 68% 39 kDa 84 kDa 0.15 +<br />

All TMCs tested (with or without O-methylation) and a DQ >22% were readily soluble in<br />

aqueous solutions at pH 7. As expected, chitosan and DMC became insoluble when the pH w<strong>as</strong><br />

incre<strong>as</strong>ed to 7. Remarkably, O-methyl free TMC with a DQ 22% w<strong>as</strong> insoluble at pH 7 at a<br />

concentration of 2.5 mg/ml while TMC with around the same molecular weight and DQ (22%)<br />

and a DOM of 18 and 12% at C-6 and C-3, respectively, w<strong>as</strong> readily dissolved in the same<br />

solvent. This is rather unexpected since O-methylated glucosamine units can only act <strong>as</strong><br />

hydrogen-acceptor, where<strong>as</strong> non O-methylated units can both accept and donate hydrogen<br />

bridges in interaction with aqueous solvents. Possibly, TMC with low DQs will still partially<br />

resemble the behavior of chitosan, where intra- and intermolecular interactions decre<strong>as</strong>e the<br />

aqueous solubility at pH 7 [1]. Partial O-methylation might reduce these interactions and<br />

there<strong>for</strong>e result in better aqueous solubility of O-methylated TMCs with a DQ


Synthesis and Characterization of O-methyl Free TMC<br />

50% <strong>for</strong> penetration enhancing effects [10, 18, 19]. However, the possible contribution of O-<br />

methylation on these effects is unknown. The availability of TMCs with varying DQ but without<br />

O-methylated groups or discrepancies in polymer chain lengths allows a better evaluation of<br />

the effect of DQ on the biological properties of TMC. Figures 3a and 3b show the effect of TMCs<br />

with different DQ with and without (partial) O-methylation on the TEER. The results of the O-<br />

methylated TMCs are in line with those reported by other researchers [6, 10, 18, 24], namely<br />

TMC with a DQ of 22% and DOM-3 and DOM-6 of 12 and 18%, respectively, showed no effect<br />

on the TEER, indicating that the polymer is unable to open the tight junctions. The TMC with a<br />

DQ of 61% and similar O-methylation demonstrates the largest decre<strong>as</strong>e in the TEER and<br />

showed marginal toxicity using the MTT <strong>as</strong>say (Figure 4a) implying that this polymer induces<br />

opening of tight junctions without exerting major acute toxic effects. However, since the LDH<br />

levels (Figure 5a) are elevated, this TMC apparently induces some membrane damage to cells.<br />

Interestingly, <strong>as</strong> compared to O-methylated TMC with a DQ of 61%, TMC with the highest DQ<br />

(86%) had similar effect on the TEER while it showed substantial cellular toxicity (MTT <strong>as</strong>say)<br />

at a concentration of 10 mg/ml. The highly elevated LDH rele<strong>as</strong>e upon exposure of the Caco-2<br />

cells to this polymer supports the data obtained with the MTT <strong>as</strong>say. Other groups [10, 18]<br />

reported the highest tight-junction opening activity of O-methylated TMC with a DQ of about<br />

40-50%. However, the molecular weights of the TMCs used in those studies were higher which<br />

may account <strong>for</strong> the somewhat different optimal DQ found in these studies.<br />

39


Chapter 2<br />

TEER (% HBSS)<br />

A<br />

SDS 1%<br />

100 TMC-OM 22%<br />

TMC-OM 50%<br />

80<br />

TMC-OM 61%<br />

60<br />

TMC-OM 86%<br />

40<br />

20<br />

0<br />

0 15 30 45 60 75 90<br />

Time (min.)<br />

1140<br />

TEER (% HBSS)<br />

B<br />

100 SDS 1%<br />

80<br />

60<br />

40<br />

20<br />

TMC 30%<br />

TMC 43%<br />

TMC 56%<br />

TMC 68%<br />

0<br />

0 15 30 45 60 75 90<br />

Time (min.)<br />

1140<br />

Figure 3. Effect of (a) O-methylated TMC and (b) O-methyl free TMC with various DQs on the TEER of<br />

Caco-2 cells at a TMC concentration of 2 mg/ml. SDS 10 mg/ml w<strong>as</strong> used <strong>as</strong> a positive control. Error bars<br />

represent the standard deviation of six me<strong>as</strong>urements.<br />

Figure 3 shows that O-methyl free TMC with a DQ of 30% had a similar TEER effect <strong>as</strong> the O-<br />

methylated TMC with a DQ of 56%, where<strong>as</strong> the MTT and LDH <strong>as</strong>says showed some toxicity<br />

above a concentration of 1 mg/ml (Figures 4b and 5b). Figure 3b also shows that the effect on<br />

the TEER slightly incre<strong>as</strong>ed (reduction from 35% to 20% remaining resistance compared with<br />

HBSS) with incre<strong>as</strong>ing DQ (from 30 to 68%). This might indicate that with incre<strong>as</strong>ing DQ the<br />

polymers have greater capacity to open the tight junctions, but the decre<strong>as</strong>e in TEER can also<br />

be the result of acute toxic effects on the Caco-2 cells (see Figures 4b and 5b). The MTT <strong>as</strong>say<br />

clearly displays a DQ depending cytotoxicity <strong>for</strong> O-methyl free TMCs and these polymers show<br />

40


Synthesis and Characterization of O-methyl Free TMC<br />

a larger decre<strong>as</strong>e in cell viability than O-methylated TMC. The LDH <strong>as</strong>say shows a DQdependent<br />

LDH rele<strong>as</strong>e <strong>for</strong> both O-methylated and O-methyl free TMCs. Interestingly, although<br />

the O-methyl free TMCs demonstrate a higher LDH rele<strong>as</strong>e than the O-methylated TMCs, the<br />

difference is much less prominent <strong>as</strong> with the MTT <strong>as</strong>say; O-methylated TMC with a DQ of 56%<br />

induces a similar LDH rele<strong>as</strong>e <strong>as</strong> O-methyl free TMC with a DQ of 30% and the LDH rele<strong>as</strong>e of<br />

O-methylated TMC with a DQ of 86% is comparable to the amount of LDH rele<strong>as</strong>ed by<br />

application of O-methyl free TMC with a DQ of 68%. Partial recovery of the TEER w<strong>as</strong> observed<br />

(24 hours after removal of the polymer) <strong>for</strong> all TMCs (with or without O-methylation), which<br />

indicates that cells were still viable after the removal of the stimulus. Taken together, O-methyl<br />

free TMCs lead to a larger decre<strong>as</strong>e of the TEER than O-methylated TMCs but the penetrationenhancing<br />

effects of the TMCs without O-methylation remain to be established.<br />

Cell viability (%)<br />

A<br />

100<br />

80<br />

60<br />

40<br />

20<br />

TMC-OM 22%<br />

TMC-OM 50%<br />

TMC-OM 61%<br />

TMC-OM 86%<br />

0<br />

0.1 1 10<br />

Conc. (mg/ml)<br />

Cell viability (%)<br />

B<br />

100 TMC 30%<br />

TMC 43%<br />

80<br />

TMC 56%<br />

60<br />

TMC 68%<br />

40<br />

20<br />

0<br />

0.1 1 10<br />

Conc. (mg/ml)<br />

Figure 4. Effect of (a) O-methylated TMC and (b) O-methyl free TMC with various DQs on the viability of<br />

Caco-2 cells (MTT <strong>as</strong>say) at different concentrations. Error bars represent the standard deviation of six<br />

me<strong>as</strong>urements.<br />

41


Chapter 2<br />

The synthesis method described in this paper provides a better defined TMC compared to<br />

the traditionally synthesized TMC with unavoidable O-methylation [15]. Moreover, the lack of<br />

chain scission during the quaternization of TMC allows studies of TMCs with varying DQ<br />

without discrepancies in polymer chain lengths.<br />

Our studies demonstrate that O-methyl free TMCs have a stronger effect on the Caco-2 cells<br />

in the TEER, MTT and LDH <strong>as</strong>says than the O-methylated TMCs. Since the molecular weights of<br />

the polymers used in the present study are comparable (Table 2), the results imply a reduction<br />

in toxicity of TMC by the (partial) O-methylation of the polymer. Additionally, TMC without O-<br />

methylation resembles to a larger extent the original chitosan polymer. This may have<br />

beneficial effects on the enzymatic degradability of the polymer.<br />

LDH rele<strong>as</strong>e (%)<br />

A<br />

30<br />

20<br />

10<br />

HBSS<br />

TMC-OM 22%<br />

TMC-OM 50%<br />

TMC-OM 61%<br />

TMC-OM 86%<br />

0<br />

0.1 1.0 10.0<br />

Conc. (mg/ml)<br />

LDH rele<strong>as</strong>e (%)<br />

B<br />

30<br />

20<br />

10<br />

HBSS<br />

TMC 30%<br />

TMC 43%<br />

TMC 56%<br />

TMC 68%<br />

0<br />

0.1 1.0 10.0<br />

Conc. (mg/ml)<br />

Figure 5. Effect of (a) O-methylated TMC and (b) O-methyl free TMC with various DQs on the viability of<br />

Caco-2 cells (LDH rele<strong>as</strong>e) at different concentrations. Error bars represent the standard deviation of<br />

three me<strong>as</strong>urements.<br />

42


Synthesis and Characterization of O-methyl Free TMC<br />

TMCs without O-methylation may improve the penetration-enhancing effects of the<br />

‘traditional’ TMCs, demonstrating a strong (reversible) effect of the opening of tight junctions.<br />

Furthermore, it shows limited toxicity at concentrations used <strong>for</strong> cell transfection experiments<br />

[11, 12]. Since incorporation into nanoparticles reduces the toxicity of TMC [24], this non O-<br />

methylated TMC may be very suitable <strong>for</strong> incorporation into nanoparticles <strong>for</strong> mucosal<br />

vaccination, which is currently under investigation.<br />

Conclusion<br />

This paper shows a straight<strong>for</strong>ward method to synthesize TMC avoiding O-methylation while<br />

preventing chain scission and to tailor the DQ of the TMC by varying the reaction time. O-<br />

methyl free TMCs with a DQ > 30% are readily soluble in aqueous solutions at pH 7. In the<br />

TEER, MTT and LDH <strong>as</strong>says O-methyl free TMCs have a stronger effect on the Caco-2 cells than<br />

the O-methylated TMCs. In conclusion, these new well-characterized O-methyl free TMCs have<br />

potentially favorable biological characteristics over the traditional TMC and allow to effectively<br />

study the influence of the DQ of TMC in various delivery systems.<br />

Acknowledgement. This research w<strong>as</strong> per<strong>for</strong>med under the framework of <strong>TI</strong> <strong>Pharma</strong><br />

project number D5-106-1; Vaccine delivery: alternatives <strong>for</strong> conventional multiple injection<br />

vaccines.<br />

43


Chapter 2<br />

References<br />

1. Kumar MNVR, Muzzarelli RAA, Muzzarelli C, S<strong>as</strong>hiwa H, Domb AJ. Chitosan chemistry<br />

and pharmaceutical perspectives. Chem Rev 104:6017-84 (2004)<br />

2. Agnihotri SA, Mallikarjuna NN, Aminabhavi TM. Recent advances on chitosan-b<strong>as</strong>ed<br />

micro- and nanoparticles in drug delivery. J Control Rele<strong>as</strong>e 100:5-28 (2004).<br />

3. Aspden TJ, Illum L, Skaugrud Ø. Chitosan <strong>as</strong> a n<strong>as</strong>al delivery system: evaluation of insulin<br />

absorption enhancement and effect on n<strong>as</strong>al membrane integrity using rat models. Eur J<br />

Pharm Sci 4:23-31 (1996).<br />

4. Van Der Lubben IM, Kersten G, Fretz MM, Beuvery C, Verhoef JC, Junginger HE.<br />

Chitosan microparticles <strong>for</strong> mucosal vaccination against diphtheria: oral and n<strong>as</strong>al efficacy<br />

studies in mice. Vaccine 21:1400-8 (2003).<br />

5. Kotze AF, Luessen HL, de Boer AG, Verhoef JC, Junginger HE. Chitosan <strong>for</strong> enhanced<br />

intestinal permeability: prospects <strong>for</strong> derivatives soluble in neutral and b<strong>as</strong>ic environments.<br />

Eur J Pharm Sci 7:145-51 (1999).<br />

6. Kotze AF, Luessen HL, de Leeuw BJ, de Boer AG, Verhoef JC, Junginger HE. N-<strong>Trimethyl</strong><br />

chitosan chloride <strong>as</strong> a potential absorption enhancer across mucosal surfaces: in vitro evalutation<br />

in intestinal epithelial cells (Caco-2). Pharm Res 14:1197-202 (1997).<br />

7. Sieval AB, Thanou M, Kotze AF, Verhoef JC, Brussee J, Junginger HE. Preparation and<br />

NMR characterization of highly substituted N-trimethyl chitosan chloride. Carbohydrate<br />

Polym 36:157-65 (1998).<br />

8. Van der Lubben IM, Verhoef JC, Fretz MM, Van O, Mesu I, Kersten G, et al. <strong>Trimethyl</strong><br />

chitosan chloride (TMC) <strong>as</strong> a novel excipient <strong>for</strong> oral and n<strong>as</strong>al immunisation against<br />

diphtheria. STP Pharm Sci 12:235-42 (2002).<br />

9. Snyman D, Hamman JH, Kotze AF. Evaluation of the mucoadhesive properties of N-<br />

trimethyl chitosan chloride. Drug Develop Ind Pharm 29:61-9 (2003).<br />

10. Thanou MM, Kotze AF, Scharringhausen T, Lueßen HL, De Boer AG, Verhoef JC, et al.<br />

Effect of degree of quaternization of N-trimethyl chitosan chloride <strong>for</strong> enhanced transport of<br />

hydrophilic compounds across intestinal Caco-2 cell monolayers. J Control Rele<strong>as</strong>e 64:15-25<br />

(2000).<br />

11. Thanou M, Florea BI, Geldof M, Junginger HE, Borchard G. Quaternized chitosan oligomers<br />

<strong>as</strong> novel gene delivery vectors in epithelial cell lines. Biomaterials 23:153-9 (2002).<br />

12. Mao Z, Ma L, Jiang Y, Yan M, Gao C, Shen J. N,N,N-<strong>Trimethyl</strong>chitosan chloride <strong>as</strong> a<br />

gene vector: synthesis and application. Macromol Biosci 7:855-63 (2007).<br />

13. Domard A, Rinaudo M, Terr<strong>as</strong>sin C. New method <strong>for</strong> the quaternization of chitosan. Int J Biol<br />

Macromol 8:105-7 (1986).<br />

14. Snyman D, Hamman JH, Kotze JS, Rollings JE, Kotze AF. The relationship between the<br />

absolute molecular weight and the degree of quaternisation of N-trimethyl chitosan chloride.<br />

Carbohydrate Polym 50:145-50 (2002).<br />

15. Polnok A, Borchard G, Verhoef JC, Sarisuta N, Junginger HE. Influence of methylation<br />

process on the degree of quaternization of N-trimethyl chitosan chloride. Eur J Pharm<br />

Biopharm 57:77-83 (2004).<br />

16. Curti E, De Britto D, Campana-Filho SP. Methylation of chitosan with iodomethane: effect of<br />

reaction conditions on chemoselectivity and degree of substitution. Macromol Biosci 3:571-6<br />

(2003).<br />

17. Di Colo G, Burgal<strong>as</strong>si S, Zambito Y, Monti D, Chetoni P. Effects of different N-trimethyl<br />

chitosans on in vitro/in vivo ofloxacin transcorneal permeation. J Pharm Sci 93:2851-62 (2004).<br />

18. Hamman JH, Stander M, Kotze AF. Effect of the degree of quaternisation of N-trimethyl<br />

chitosan chloride on absorption enhancement: in vivo evaluation in rat n<strong>as</strong>al epithelia. Int J<br />

Pharm 232:235-42 (2002).<br />

44


Synthesis and Characterization of O-methyl Free TMC<br />

19. Boonyo W, Junginger HE, Waranuch N, Polnok A, Pitaksuteepong T. Chitosan and<br />

<strong>Trimethyl</strong> chitosan chloride (TMC) <strong>as</strong> adjuvants <strong>for</strong> inducing immune responses to ovalbumin<br />

in mice following n<strong>as</strong>al administration. J Control Rele<strong>as</strong>e 121:168-75 (2007).<br />

20. Muzzarelli RAA, Tanfani F. The N-permethylation of chitosan and the preparation of N-<br />

trimethyl chitosan iodide. Carbohydr Polym 5:297-307 (1985).<br />

21. Jia Z, shen D, Xu W. Synthesis and antibacterial activities of quaternary ammonium salt of<br />

chitosan. Carbohydr Res 333:1-6 (2001).<br />

22. Guo Z, Xing R, Liu S, Zhong Z, Ji X, Wang L, Li P. Antifungal properties of Schiff b<strong>as</strong>es of<br />

chitosan, N-substituted chitosan and quaternized chitosan. Carbohydr Res 342:1329-32 (2007).<br />

23. Pine SH, Sanchez BL. Formic acid-<strong>for</strong>maldehyde methylation of amines. J Org Chem 36:829-32<br />

(1971).<br />

24. Amidi M, Romeijn SG, Borchard G, Junginger HE, Hennink WE, Jiskoot W. Preparation<br />

and characterization of protein-loaded N-trimethyl chitosan nanoparticles <strong>as</strong> n<strong>as</strong>al delivery<br />

system. J Control Rele<strong>as</strong>e 111:107-16 (2006).<br />

25. Jiang X, van der Horst A, van Steenbergen M, Akeroyd N, van Nostrum C, Schoenmakers P,<br />

Hennink WE. Molar-m<strong>as</strong>s characterization of cationic polymers <strong>for</strong> gene delivery by<br />

aqueous size-exclusion chromatography. Pharm Res 23:595-603 (2006).<br />

26. Mao S, Shuai X, Unger F, Wittmar M, Xie X, Kissel T. Synthesis, characterization and<br />

cytotoxicity of poly(ethylene glycol)-graft-trimethyl chitosan block copolymers. Biomaterials<br />

26:6343-56 (2005).<br />

27. Nerurkar MM, Burton PS, Borchardt RT. The use of surfactants to enhance the permeability<br />

of peptides through Caco-2 cells by inhibition of an apically polarized efflux system. Pharm Res<br />

13:528-34 (1996).<br />

28. Mosmann TJ. Rapid colorimetric <strong>as</strong>say <strong>for</strong> cellular growth and survival: application to<br />

proliferation and cytotoxicity <strong>as</strong>says. J Immunol Methods 65:55-65 (1983).<br />

29. Varum KM, Ottoy MH, Smidsrod O. Acid hydrolysis of chitosans. Carbohydrate Polym 46:89-98<br />

(2001).<br />

45


CHAPTER 3<br />

INFLUENCE OF THE DEGREE OF<br />

ACETYLA<strong>TI</strong>ON ON THE ENZYMA<strong>TI</strong>C<br />

DEGRADA<strong>TI</strong>ON AND IN VITRO BIOLOGICAL<br />

PROPER<strong>TI</strong>ES OF TRIMETHYLATED<br />

CHITOSANS<br />

Rolf J. Verheul, Maryam Amidi, Mies van Steenbergen,<br />

Elly van Riet, Wim Jiskoot, Wim E. Hennink.<br />

Biomaterials 2009, 30, 3129-3135


Chapter 3<br />

Abstract<br />

Chitosan derivatives such <strong>as</strong> N,N,N-trimethylated chitosan (TMC) are currently being<br />

investigated <strong>for</strong> the delivery of drugs, vaccines and genes. However, the influence of the extent<br />

of N-acetylation of these polymers on their enzymatic degradability and biological properties<br />

is unknown. In this study, TMCs with a degree of acetylation (DAc) ranging from 11 to 55%<br />

were synthesized by using a three-step method. First, chitosan w<strong>as</strong> partially re-acetylated<br />

using acetic anhydride followed by quantitative dimethylation using <strong>for</strong>maldehyde and sodium<br />

borohydride. Then, in presence of an excess amount of iodomethane, TMC w<strong>as</strong> synthesized.<br />

The TMCs obtained by this method showed neither detectable O-methylation nor loss in acetyl<br />

groups ( 1 H-NMR) and a slight incre<strong>as</strong>e in molecular weight (GPC) with incre<strong>as</strong>ing degree of<br />

substitution, implying that no chain scission occurred during synthesis. The extent of<br />

lysozyme-catalyzed degradation of TMC, and that of its precursors chitosan and dimethyl<br />

chitosan, w<strong>as</strong> highly dependent on the DAc and polymers with the highest DAc showed the<br />

largest decre<strong>as</strong>e in molecular weight. On Caco-2 cells, TMCs with a high DAc (~50%), a DQ of<br />

around 44% and with or without O-methylated groups, were not able to open tight junctions in<br />

the trans-epithelial electrical resistance (TEER) <strong>as</strong>say, in contr<strong>as</strong>t with TMCs (both O-<br />

methylated and O-methyl free; concentration 2.5 mg/ml) with a similar DQ but a lower DAc<br />

which were able to reduce the TEER with 30 and 70%, respectively. Additionally, TMCs with a<br />

high DAc (~50%) demonstrated no cell toxicity (MTT, LDH rele<strong>as</strong>e) up to a concentration of 10<br />

mg/ml.<br />

48


Influence of Degree of Acetylation on Degradation and Properties of TMC<br />

Introduction<br />

Chitosan is a polysaccharide consisting of β1→4-D-glucosamine and β1→4 N-acetyl D-<br />

glucosamine units and is obtained from the natural polymer chitin (poly β1→4 N-acetyl D-<br />

glucosamine) by partial deacetylation. Chitosan h<strong>as</strong> been under investigation <strong>for</strong> various<br />

biomedical and pharmaceutical applications due to its biocompatibility, low toxicity and its<br />

muco-adhesive properties [1-4]. It h<strong>as</strong> been reported that chitosan can be degraded in human<br />

tissues by several enzymes (e.g. lysozyme, chitin<strong>as</strong>e), by acid hydrolysis and by oxidativereductive<br />

depolymerization (ORD) reactions [5, 6]. Nordtveit et al. demonstrated that the ORD<br />

of chitosan driven by hydrogen peroxide is independent of the amount of residual N-acetylated<br />

units [7]. However, in vitro studies on the degradation of chitosan by acid hydrolysis [8] and by<br />

enzymes [6, 7, 9-14] revealed a major dependency on the degree of acetylation (DAc).<br />

Completely de-acetylated chitosan shows very limited degradation by (enzyme catalyzed)<br />

hydrolysis while the extent of degradation and degradation kinetics incre<strong>as</strong>e with the extent of<br />

acetylation. Lysozyme is widely present in human body fluids (e.g. serum, saliva, tears) and is<br />

actively secreted by macrophages and neutrophils [15], and consequently the enzymatic<br />

degradation of chitosan by lysozyme h<strong>as</strong> been studied in depth [7, 9-13, 16, 17]. Temperature,<br />

pH and ionic strength have been found to influence degradation kinetics [10]. Additionally, in<br />

vivo experiments on the degradability of chitosan have shown that the DAc also plays a key<br />

role in the depolymerization of chitosan in living animals [13, 17]. Furthermore, uptake-,<br />

toxicity- and cell transfection studies with chitosan suggest that the degree of acetylation h<strong>as</strong><br />

an important influence on biological polymer characteristics <strong>as</strong> well [18-20]. Finally, partially<br />

acetylated chitosan h<strong>as</strong> been found to have better adjuvant properties <strong>for</strong> macrophage<br />

stimulation compared to chitosan with a low amount of residual acetylated groups or chitin<br />

[21-23], and recently, it w<strong>as</strong> reported that N-acetylated glucosamines can bind to C-type lectin<br />

receptors on denditric cells thereby working <strong>as</strong> an adjuvant [24].<br />

In contr<strong>as</strong>t to chitosan, N,N,N,-trimethylated chitosan (TMC), a partially quaternized<br />

derivative of chitosan, is water-soluble at neutral pH. TMC h<strong>as</strong> been widely studied in the<br />

biomedical field <strong>as</strong> drug, vaccine and gene delivery vehicle [25-31]. It h<strong>as</strong> been shown in<br />

several in vitro and in vivo models that TMC h<strong>as</strong> limited toxicity, possesses muco-adhesive<br />

properties and can incre<strong>as</strong>e the uptake of small drug molecules <strong>as</strong> well <strong>as</strong> proteins via various<br />

mucosal routes [25-27, 29, 30, 32-35]. Several investigators have studied the optimal degree of<br />

quaternization (DQ) of TMC <strong>for</strong> mucosal transport and gene delivery [26, 33, 36, 37]. However,<br />

molecular weight [38, 39] and, <strong>as</strong> demonstrated in chapter 2, O-methylation [40] also have a<br />

major impact on the physical and biological characteristics of TMC. Besides its solubility at pH<br />

49


Chapter 3<br />

7, another often suggested potentially beneficial characteristic of TMC is its biodegradability.<br />

These biodegradability claims, however, are b<strong>as</strong>ed on degradation studies per<strong>for</strong>med on<br />

chitosan, and no investigations on the enzymatic degradability of TMC have been carried out so<br />

far. The aim of this paper w<strong>as</strong> to investigate the lysozyme-catalyzed degradability of TMCs<br />

with different DAcs and with or without O-methylated groups, and to evaluate these polymers<br />

<strong>for</strong> their physico-chemical properties, cytotoxicity and ability to open tight junctions.<br />

Materials and Methods<br />

Materials. Chitosan with a residual degree of acetylation of 17% (determined with 1 H-NMR<br />

<strong>as</strong> described in below) and a number average molecular weight (M n) and weight average<br />

molecular weight (M w) of 28 and 43 kDa, (determined with GPC-TD <strong>as</strong> described below),<br />

respectively, w<strong>as</strong> purch<strong>as</strong>ed from Primex (Siglufjodur, Iceland). Hen-egg white lysozyme<br />

(41800 units per mg solid), acetic anhydride, sodium borohydride, <strong>for</strong>mic acid, <strong>for</strong>maldehyde<br />

37% (stabilized with methanol), DCl 35% w/w in D 2O, sodium azide, deuterium oxide, sodium<br />

acetate, acetic acid (anhydrous), sodium hydroxide and hydrochloric acid were obtained from<br />

Sigma-Aldrich Chemical Co. Dulbecco’s Modified Eagle’s Medium (DMEM), Hank’s balanced salt<br />

solution (HBSS) and fetal calf serum (FCS) were obtained from Invitrogen (Breda, The<br />

Netherlands). Sodium dodecyl sulfate (SDS) and Sicapent were ordered from Merck<br />

(Darmstadt, Germany). Iodomethane 99% stabilized with copper w<strong>as</strong> obtained from Acros<br />

Organics (Geel, Belgium). All other chemicals used were of analytical grade.<br />

Acetylation of chitosan. Chitosan (degree of acetylation 17%) w<strong>as</strong> used <strong>as</strong> obtained and<br />

acetylation w<strong>as</strong> carried out according to a previously described method [19]. Briefly, chitosan<br />

(10 g) w<strong>as</strong> dissolved in 1% acetic acid (500 ml). Then, 500 ml methanol and 2.7 ml acetic<br />

anhydride were added. The resulting mixture w<strong>as</strong> stirred overnight at room temperature in a<br />

roundbottom fl<strong>as</strong>k. Next, the reacetylated chitosan w<strong>as</strong> precipitated by dropping the reaction<br />

mixture into an aqueous solution of 1 M NaOH, filtrated using a gl<strong>as</strong>s filter and w<strong>as</strong>hed<br />

extensively with methanol. Then, the precipitate w<strong>as</strong> dissolved in an aqueous acidic solution<br />

(pH 4, adjusted with 1 M HCl) and dialyzed against deionized water <strong>for</strong> 3 days changing buffer<br />

twice-daily. Finally, the solution w<strong>as</strong> filtered through a 0.8 µm filter and the polymer w<strong>as</strong><br />

collected after freeze-drying.<br />

50


Influence of Degree of Acetylation on Degradation and Properties of TMC<br />

Synthesis of dimethylated chitosan. Dimethylated chitosan (DMC) with a degree of<br />

acetylation of 17% w<strong>as</strong> synthesized <strong>as</strong> described previously in chapter 2 [40]. In short,<br />

chitosan (5 g) w<strong>as</strong> dissolved in a mixture of 90 ml deionized water and 15 ml <strong>for</strong>mic acid.<br />

Subsequently, 20 ml of 37% <strong>for</strong>maldehyde w<strong>as</strong> added, and the solution w<strong>as</strong> heated to 70 °C<br />

and stirred under refluxing <strong>for</strong> 2 days. The slightly yellow, viscous solution w<strong>as</strong> evaporated<br />

under reduced pressure and the pH w<strong>as</strong> adjusted to 8 using 1 M NaOH resulting in the<br />

<strong>for</strong>mation of a gel, which w<strong>as</strong> extensively w<strong>as</strong>hed with deionized water to remove impurities.<br />

Then, the DMC w<strong>as</strong> dissolved in deionized water at pH 4 (adjusted with 1 M HCl), filtered over<br />

a gl<strong>as</strong>s filter and dialyzed against deionized water <strong>for</strong> three days (changing buffer twice-daily).<br />

Finally, the product w<strong>as</strong> filtered through a 0.8 µm filter and collected after freeze-drying.<br />

GPC analysis revealed that under these conditions, some chain scission occurred with<br />

chitosan with a DAc of 55% and there<strong>for</strong>e less vigorous reaction conditions were applied to<br />

obtain dimethylated chitosan with a high degree of acetylation. Here, sodium borohydride<br />

instead of <strong>for</strong>mic acid w<strong>as</strong> used <strong>as</strong> a reductor to allow the reaction to take place at room<br />

temperature [41, 42]. In detail, chitosan (1 g) with a degree of acetylation of 55% w<strong>as</strong><br />

dissolved in 50 ml of 2% acetic acid (v/v). Next, 10 ml of 37% <strong>for</strong>maldehyde solution w<strong>as</strong><br />

added, the pH w<strong>as</strong> adjusted to 5 with 1M NaOH and the mixture w<strong>as</strong> stirred <strong>for</strong> 2 hours at<br />

room temperature. Subsequently, sodium borohydride (2.5 g) w<strong>as</strong> added in portions of 250 mg<br />

over a period of 24 hours, adjusting the pH to 5 with 1M HCl after each addition. After<br />

precipitation with acetone, the product w<strong>as</strong> w<strong>as</strong>hed extensively with acetone and dried<br />

overnight.<br />

Synthesis of O-methyl free trimethylated chitosan. DMC with different degrees of<br />

acetylation synthesized from chitosan with a DAc of 55% or 17% <strong>as</strong> described above, were<br />

reacted with iodomethane to yield trimethylated chitosan (TMC). To prevent O-methylation,<br />

the reaction of DMC with iodomethane w<strong>as</strong> done in NMP, without the addition of a b<strong>as</strong>e<br />

catalyst. TMC with a degree of acetylation of 17% and a degree of quaternization of around<br />

45% w<strong>as</strong> obtained by the method described previously [40]. TMC with a degree of acetylation<br />

of 55% and a DQ of 45% w<strong>as</strong> obtained with a slightly modified method. In detail, DMC (500<br />

mg) with a degree of acetylation of 55% w<strong>as</strong> dissolved in 80 ml deionized water and the pH<br />

w<strong>as</strong> adjusted to 11 with a 1 M solution of NaOH, resulting in gel <strong>for</strong>mation. Then, the gel w<strong>as</strong><br />

w<strong>as</strong>hed with water followed by acetone. To remove residual solvents, the DMC w<strong>as</strong> dried<br />

under vacuum <strong>for</strong> 4 hours. Next, DMC w<strong>as</strong> suspended in 125 ml NMP followed by the addition<br />

of 8 ml iodomethane. The dispersion w<strong>as</strong> stirred at 40°C <strong>for</strong> 50 hours and subsequently<br />

51


Chapter 3<br />

dropped into 400 ml of an ethanol/diethyl ether mixture (50/50) to precipitate the <strong>for</strong>med<br />

TMC, which w<strong>as</strong> collected by centrifugation and subsequently extensively w<strong>as</strong>hed with diethyl<br />

ether. After drying overnight at room temperature, the obtained TMC w<strong>as</strong> dissolved in 100 ml<br />

of an aqueous 10% NaCl solution <strong>for</strong> ion-exchange. Finally, the TMC w<strong>as</strong> dialyzed against<br />

deionized water <strong>for</strong> 3 days changing buffer twice daily, filtered through a 0.8 µm filter and<br />

collected after freeze-drying.<br />

Synthesis of O-methylated TMC from chitosan. O-methylated TMC with a DQ of about 45%<br />

w<strong>as</strong> synthesized by methylation of chitosan with iodomethane in a mixture of an aqueous<br />

solution of NaOH and NMP essentially <strong>as</strong> described previously [43]. In detail, chitosan (500<br />

mg) with a degree of acetylation of 17 or 55% and sodium iodide (1.2 g) were dispersed in a<br />

mixture of 40 ml of NMP and 5 ml of 15% w/v aqueous NaOH solution. Subsequently, the<br />

mixture w<strong>as</strong> heated to 60°C and after stirring <strong>for</strong> 20 min, 4 ml of methyl iodide w<strong>as</strong> added and<br />

the reaction mixture w<strong>as</strong> refluxed <strong>for</strong> 60 min. Then, 2 ml of 15% NaOH solution and 1.5 ml of<br />

iodomethane were added and the solution w<strong>as</strong> stirred <strong>for</strong> 60 minutes. Next, the reaction<br />

mixture w<strong>as</strong> dropped into 200 ml of a mixture of diethyl ether and ethanol (50/50) to<br />

precipitate the O-methylated TMC, which w<strong>as</strong> subsequently w<strong>as</strong>hed extensively with diethyl<br />

ether. Finally, the product w<strong>as</strong> dissolved in 50 ml aqueous 10% w/v NaCl solution, put on a<br />

shaker <strong>for</strong> 18 hours <strong>for</strong> ion-exchange and the obtained solution w<strong>as</strong> dialyzed at room<br />

temperature against deionized water <strong>for</strong> 3 days changing buffer twice daily, filtered through a<br />

0.8 µm filter and freeze dried.<br />

Determination of the degrees of acetylation, dimethylation, and quaternization. The<br />

1H-NMR spectra of the various chitosans, DMC and TMCs were recorded with a Varian INOVA<br />

500MHz NMR spectrometer (Varian Inc., Palo Alto, Ca, USA) at 80°C in D 2O. For the DMC and<br />

the chitosans addition of DCl (35% w/w in D 2O) w<strong>as</strong> needed to dissolve the polymers.<br />

The degree of acetylation of the chitosans, DMC and TMCs w<strong>as</strong> calculated <strong>as</strong> described<br />

previously [44]:<br />

DAc = [[CH 3]/[H2-H6] x 1/2] x 100<br />

Here, [CH 3] is the integral of the three hydrogens of the acetyl groups at 2.0 ppm and [H2-H6]<br />

is the integral corresponding the six H-2 to H-6 protons between 3.9 and 3.0 ppm.<br />

52


Influence of Degree of Acetylation on Degradation and Properties of TMC<br />

The degree of dimethylation of the DMC w<strong>as</strong> calculated <strong>as</strong> follows:<br />

DDM = [(CH 3) 2]/[H2-H6] x 100<br />

Here, [(CH 3) 2] is the integral of the hydrogens of the dimethyl amino groups at 2.9 ppm and<br />

[H2-H6] is the integral corresponding the H-2 to H-6 protons between 3.9 and 3.0 ppm.<br />

The DQ and degree of dimethylation (DM) of the TMCs were calculated <strong>as</strong> previously described<br />

[25, 43, 45].<br />

DQ = [[(CH 3) 3]/[H] × 1/9] × 100<br />

DM = [[(CH 3) 2]/[H] × 1/6] × 100<br />

Here, [(CH 3) 3] and [(CH 3) 2] are the integrals of the hydrogens of the trimethylated amino<br />

groups at 3.3 ppm and the dimethylated amino groups at 2.9 ppm, respectively. [H] is the<br />

integral of the H-1 peaks between 4.7 and 5.7 ppm; the signal related to the hydrogen atoms<br />

bound to the C-1’s of TMC.<br />

Enzymatic degradation of various TMCs. Table 1 gives an overview of the different<br />

chitosans and TMCs used <strong>for</strong> evaluation of the lysozyme-catalyzed degradation. The different<br />

polymers were placed overnight in a vacuum oven at 40 o C in presence of Sicapent to remove<br />

residual water. Next, the samples were dissolved in 2% acetic acid solution <strong>for</strong> at le<strong>as</strong>t 18<br />

hours. Then, lysozyme dissolved in H 2O, and 1 M NaCl and 1 M NaOH were added to the<br />

polymer solutions to obtain a lysozyme concentration of 38 µg/ml, polymer concentrations of<br />

5 mg/ml, 150 mM NaCl, 1% acetic acid and a pH of 4.5.<br />

The degradation of the different TMC polymers w<strong>as</strong> also studied at physiological pH in a<br />

phosphate-buffered salt (PBS) buffer (8.2 g/l NaCl, 3.1 g/l Na 2HPO 4 12 H 2O, 0.3 g/l NaH 2PO 4 2<br />

H 2O, pH 7.4). In detail, the different polymers were placed overnight in a vacuum oven at 40 o C<br />

in presence of Sicapent to remove residual water. Next, the samples were dissolved in PBS with<br />

0.02% sodium azide <strong>for</strong> 18 hours. Then, lysozyme dissolved in PBS (also containing 0.02%<br />

sodium azide) w<strong>as</strong> added to the polymer solutions to obtain a lysozyme concentration of 38<br />

µg/ml, polymer concentrations of 5 mg/ml and a pH of 7.4. Additionally, the influence of<br />

lysozyme concentration on polymer degradation w<strong>as</strong> studied. To this end, TMC-RA w<strong>as</strong><br />

dissolved in PBS with 0.02% sodium azide <strong>as</strong> described above and incubated with varying<br />

concentrations of lysozyme (9 µg/ml to 38 µg/ml) in PBS with 0.02% sodium azide.<br />

53


Chapter 3<br />

Polymer solutions were incubated at 37 °C and samples were taken at various time points<br />

during 6 days. Polymer molecular weights were determined with a GPC-triple detection<br />

described below. Polymers not exposed to lysozyme were used <strong>as</strong> controls.<br />

Determination of M n and M w of the different polymers. The number average weight (M n)<br />

and weight average weight (M w) of chitosan and the various TMCs were determined by gel<br />

permeation chromatography (GPC) on a Viscotek-triple detection system using a Shodex<br />

OHPak SB-806 column (15 cm) and 0.3 M sodium acetate pH 4.4 (adjusted with acetic acid) <strong>as</strong><br />

running buffer [46]. Data from the l<strong>as</strong>er photometer (λ = 670 nm) (right (90 0 ) and low (7 0 )<br />

angle light scattering), refractive index detector and viscosity detector were integrated using<br />

the provided Omnisec-software to calculate the M n, M w, dn/dc and the intrinsic viscosity ([η])<br />

of the different samples. Pullulan (M n = 102 kDa, M w = 106 kDa) obtained from Viscotek<br />

Benelux (Oss, the Netherlands) w<strong>as</strong> used <strong>for</strong> calibration.<br />

Transepithelial electrical resistance (TEER) me<strong>as</strong>urements. Caco-2 cells were seeded at<br />

a density of 2x10 5 cells per well on 12-transwell plates with a microporous membrane. The<br />

cells were cultured in Dulbecco’s Modified Eagle’s Medium (DMEM) <strong>for</strong> 10 days until a<br />

confluent cell layer w<strong>as</strong> <strong>for</strong>med. The medium w<strong>as</strong> replaced by Hank’s Balanced Salt Solution<br />

(HBSS) at the b<strong>as</strong>olateral side 10 minutes be<strong>for</strong>e the start of the experiments. Then, 0.5 ml<br />

solution of TMC (with various DQs, DAcs and with or without O-methylation, dissolved (2.5<br />

mg/ml) in HBSS, pH adjusted to 7 with 0.1 M NaOH) w<strong>as</strong> applied at the apical site of the cell<br />

monolayers. SDS (10 mg/ml) w<strong>as</strong> used <strong>as</strong> positive control and HBSS <strong>as</strong> reference. The<br />

resistance me<strong>as</strong>ured of the membrane without cells w<strong>as</strong> used <strong>as</strong> blank. The TEER of the Caco-2<br />

cells at certain time points (0, 15, 30, 45, 60 and 90 min.) after addition of the stimuli w<strong>as</strong><br />

me<strong>as</strong>ured with a Millicell-ERS (Millipore, Billerica, USA) me<strong>as</strong>uring device. After 90 minutes<br />

the cells were w<strong>as</strong>hed with HBSS and incubation of the cells w<strong>as</strong> continued in DMEM <strong>for</strong> 24<br />

hours at 37°C, CO 2 5% to determine the recovery of the TEER [25, 47].<br />

MTT cell toxicity <strong>as</strong>say. Caco-2 cells were seeded in a 96-well plate at a density of 4x10 4<br />

cells per well and incubated <strong>for</strong> 2 days at 37°C, CO 2 5% in culture medium (DMEM, high<br />

glucose, 10% FCS, L-glutamine, pyruvate, non essential amino acids). The medium w<strong>as</strong><br />

removed and the cells were incubated <strong>for</strong> 2.5 hours with 100 µl TMC solutions in HBSS (TMC<br />

concentrations were 0.1, 1 and 10 mg/ml, pH set at 7 with 0.1 M NaOH). SDS (10 mg/ml) w<strong>as</strong><br />

used <strong>as</strong> positive control and HBSS <strong>as</strong> reference <strong>for</strong> 100% cell viability. Thereafter, the HBSS<br />

54


Influence of Degree of Acetylation on Degradation and Properties of TMC<br />

w<strong>as</strong> removed and the cells were w<strong>as</strong>hed with phosphate buffered saline. One hundred µl of a<br />

freshly prepared solution of 0.5 mg/ml MTT in DMEM (without any additions), w<strong>as</strong> added and<br />

the cells were incubated <strong>for</strong> 3 hours at 37°C and 5% CO 2. Subsequently, the wells were<br />

emptied, 100 µl of DMSO w<strong>as</strong> used to dissolve the <strong>for</strong>med <strong>for</strong>mazan crystals and the<br />

absorbance w<strong>as</strong> read at 595 nm [48].<br />

LDH <strong>as</strong>say. Caco-2 cells were seeded in a 96-well plate at a density of 4x10 4 cells per well<br />

and incubated <strong>for</strong> 2 days at 37°C, CO 2 5% in culture medium (see section above <strong>for</strong><br />

composition). The medium w<strong>as</strong> removed and the cells were w<strong>as</strong>hed with HBSS and incubated<br />

<strong>for</strong> 2.5 hours with 100 µl TMC solutions in HBSS (TMC concentrations were 0.1, 1 and 10<br />

mg/ml, pH set at 7 with 0.1 M NaOH). After incubation, the concentration of LDH present in the<br />

supernatant of the samples w<strong>as</strong> determined with the Cytotoxicity Detection Kit-Plus (Roche<br />

Diagnostics, Mannheim, Germany) by me<strong>as</strong>uring absorbance at 490 nm with 650 nm <strong>as</strong> a<br />

reference wavelength. A calibration curve w<strong>as</strong> made with the lysis buffer provided by the<br />

manufacturer, setting the LDH concentration me<strong>as</strong>ured with the undiluted lysis buffer at 100%<br />

LDH rele<strong>as</strong>e. HBSS w<strong>as</strong> used <strong>as</strong> a negative control.<br />

Results and discussion<br />

Synthesis and characterization of different polymers. To investigate the influence of the<br />

degree of acetylation on the biological properties and enzymatic degradability of trimethylated<br />

chitosan (TMC), chitosan (CS, DAc 17%) w<strong>as</strong> partially re-acetylated using acetic anhydride to a<br />

DA of 55% (CS-RA, DAc 55%) <strong>as</strong> determined by 1 H-NMR analysis. Both CS and CS-RA were<br />

subsequently trimethylated to a degree of about 45% according to the method described by<br />

Sieval [43]. 1 H-NMR analysis (Table 1) showed that this synthesis method also leads to<br />

considerable O-methylation and loss of N-acetylated units, likely due to the alkaline reaction<br />

conditions.<br />

Besides the frequently used method of Sieval et al, CS and CS-RA were also trimethylated to a<br />

degree of 45% via the two-step synthesis route described recently by Verheul et al. [40] (with<br />

some modifications <strong>for</strong> the synthesis of TMC-RA) which avoids O-methylation and loss of N-<br />

acetylated units (Table 1). Chitosan with a DAc of 17% w<strong>as</strong> quantitatively dimethylated with<br />

<strong>for</strong>maldehyde using <strong>for</strong>mic acid <strong>as</strong> reducing agent. However, using these reaction conditions to<br />

dimethylate chitosan with a DAc of 55%, which is more susceptible to acidic hydrolysis than<br />

55


Chapter 3<br />

chitosan with a DAc of 17% [8], resulted in the reduction of molecular weights (data not<br />

shown). When sodium borohydride w<strong>as</strong> used <strong>as</strong> reducing agent, relatively mild reaction<br />

conditions (pH 4, room temperature) could be applied to quantitively dimethylate chitosan.<br />

Table 1. Characteristics of the chitosans and TMCs. Degree of acetylation (DAc), quaternization (DQ),<br />

dimethylation (DM), O-methylation on C-6 (DOM-6) and on C-3 (DOM-3) of TMCs <strong>as</strong> determined by 1 H-<br />

NMR analysis. M n, M w and dn/dc were determined by GPC-triple detection.<br />

DAc DQ DM DOM-6 DOM-3 Mn Mw dn/dc<br />

(kDa) (kDa)<br />

CS 17% - - - - 28 43 0.18<br />

TMC 17% 43% 40% - - 39 64 0.16<br />

TMC-OM 11% 45% 44% 25% 16% 35 48 0.16<br />

CS-RA 55% - - - - 35 65 0.16<br />

DMC-RA 55% - 45% - - 42 71 0.16<br />

TMC-RA 55% 44% 1% - - 43 79 0.16<br />

TMC-RA-OM 49% 46% 5% 77% 57% 40 63 0.15<br />

Table 1 shows that particularly in c<strong>as</strong>e of the re-acetylated chitosan, these milder reaction<br />

conditions prevented polymer chain scission; the molecular weights of CS, CS-RA, DMC-RA and<br />

TMC-RA slightly incre<strong>as</strong>ed with incre<strong>as</strong>ing extent of substitutions (TMC-RA>DMC-RA>CS-<br />

RA>CS). As reported previously, TMC synthesized according to the method of Sieval et al.<br />

resulted in polymers with a lower molecular weight than the polymers obtained via the<br />

method without the use of a strong b<strong>as</strong>e [40]. Likely, the strong b<strong>as</strong>ic reaction conditions<br />

caused some chain scission during synthesis [40, 49]. Importantly, the mild two-step synthetic<br />

method described in this paper yields TMC with varying DQs while O-methylation, polymer<br />

chain scission and the hydrolysis of N-acetylated groups do not occur. Previous studies showed<br />

that changes in TMC molecular weight [38, 39] and the presence of O-methyl groups [40]<br />

influence polymer characteristics, such <strong>as</strong> cytotoxicity and the ability to open tight junctions.<br />

GPC-TD analysis showed that re-acetylated chitosan and TMCs contained some high molecular<br />

weight material (about 3% of the injected amount) which can be likely <strong>as</strong>cribed to aggregation<br />

[50-52]. However, this high-molecular-weight material had no impact on the degradation<br />

studies due to its relatively small contribution and it w<strong>as</strong> e<strong>as</strong>ily separated on the GPC column<br />

from the free polymer.<br />

Enzymatic degradation of various chitosans and TMCs. Figure 1 shows the decre<strong>as</strong>e of<br />

the M n of TMC-RA in presence of various concentrations of hen egg-white lysozyme at pH 7.4 at<br />

37 °C. Clearly, the degradation rate incre<strong>as</strong>es with lysozyme concentration while without<br />

lysozyme almost no decre<strong>as</strong>e in molecular weight w<strong>as</strong> observed. The enzymatic degradation<br />

essentially occurred in the first 24 hours and after this time period hardly any decre<strong>as</strong>e of the<br />

56


Influence of Degree of Acetylation on Degradation and Properties of TMC<br />

M n of TMC-RA w<strong>as</strong> detected. Addition of fresh lysozyme after 70 hours did not result in a<br />

further drop of molecular weight (results not shown), implying that arrest of the polymer<br />

degradation is not due to inactivation of lysozyme but that lysozyme is not able to catalyze the<br />

hydrolysis of the remaining glyosidic bonds. As expected, the molecular weight obtained after<br />

prolonged exposure to lysozyme w<strong>as</strong> independent of the concentration of the enzyme. Overall,<br />

in the presence of lysozyme, the M n of TMC-RA dropped from 43 kDa to about 11 kDa<br />

corresponding on average to approximately three cuts per molecule. Figure 2 shows the<br />

changes in M n, M w and intrinsic viscosity ([η]) of TMC-RA in presence of lysozyme (38 µg/ml)<br />

at pH 7.4 at 37 °C <strong>as</strong> obtained with the GPC-triple detection method. As expected, both the M w<br />

and the [η] followed a similar pattern <strong>as</strong> the M n of TMC-RA. This w<strong>as</strong> observed <strong>for</strong> all analyzed<br />

polymers (data not shown).<br />

4.0×10 4 TMC-RA 38 μg/ml lyso<br />

Mn (Da)<br />

3.0×10 4<br />

2.0×10 4<br />

TMC-RA 0μg/ml lyso<br />

TMC-RA 9 μg/ml lyso<br />

TMC-RA 19 μg/ml lyso<br />

1.0×10 4<br />

0<br />

0 5 10<br />

100 200<br />

time (h)<br />

Figure 1. Degradation of TMC-RA by lysozyme (various concentrations) at pH 7.4 and 37°C. Error bars<br />

represent the standard deviation of three me<strong>as</strong>urements.<br />

The lysozyme catalyzed enzymatic degradation of different chitosans and TMCs w<strong>as</strong><br />

compared <strong>as</strong> shown in Figure 3. Since CS, CS-RA and DMC-RA are not soluble at physiological<br />

pH, degradation w<strong>as</strong> studied at pH 4.5 and at 37 °C in presence of 38 µg/ml lysozyme. This<br />

figure clearly demonstrates that the extent of lysozyme-catalyzed degradation of the<br />

investigated polymers is highly dependent on the degree of acetylation. Polymers with a DAc ≤<br />

17% (CS, TMC and TMC-OM) are less susceptible to lysozyme-catalyzed degradation than the<br />

re-acetylated polymers with a DAc ≥49%. Although the changes in molecular weights are less<br />

57


Chapter 3<br />

prominent, it is obvious that TMC-OM, with the lowest DAc (DAc 11%), w<strong>as</strong> also to the lowest<br />

extent degraded by lysozyme after 6 days (decre<strong>as</strong>e in M n from 35 kDa to 30 kDa). In contr<strong>as</strong>t,<br />

the M n of TMC-RA with a DAc of 55% dropped from 43 kDa to 10 kDa under the same<br />

conditions. Re-acetylated CS-RA (DAc 55%), in agreement with others [6, 7, 10, 14, 17, 53], w<strong>as</strong><br />

much more susceptible to lysozyme-catalyzed degradation than CS (DAc 17%), and their M n’s<br />

decre<strong>as</strong>ed from 35 kDa to 7 kDa and from 28 kDa to 20 kDa, respectively.<br />

0.4<br />

Mn or Mw (Da)<br />

8.0×10 4 100 200<br />

6.0×10 4<br />

4.0×10 4<br />

2.0×10 4<br />

[η]<br />

Mn<br />

Mw<br />

0.3<br />

0.2<br />

0.1<br />

[η] (dl/g)<br />

0<br />

0 5 10<br />

time (h)<br />

0.0<br />

Figure 2. Changes in the M n, M w and intrinsic viscosity of TMC-RA by lysozyme (38 µg/ml) at pH 7.4 and<br />

37°C. Error bars represent the standard deviation of three me<strong>as</strong>urements.<br />

Nordtveit et al. demonstrated that at le<strong>as</strong>t 3 to 4 acetylated units are required in the<br />

hexameric binding pocket <strong>for</strong> lysozyme to allow enzymatic cleavage [7]. Since these domains<br />

containing the required number of acetylated units are destroyed through cleavage of the<br />

glycosidic bond [7, 16], degradation by lysozyme occurs till a minimum molecular weight<br />

depending on the DAc of the chitosan [11]. As DMC-RA w<strong>as</strong> degraded to practically the same<br />

molecular weight <strong>as</strong> CS-RA (Figure 3), it can be concluded that methylation of the free NH 2<br />

groups of chitosan did not affect its extent of lysozyme-catalyzed degradation. TMC-RA w<strong>as</strong><br />

cleaved up to a slightly higher M n than CS-RA (10 kDa vs. 7 kDa, respectively), however, this<br />

can be explained by its higher degree of substitution due to the quaternization of the free<br />

amines thereby incre<strong>as</strong>ing the weight per glucosamine-residue. There<strong>for</strong>e, it can be stated that<br />

quaternized chitosan synthesized according to the two-step method presented here is<br />

58


Influence of Degree of Acetylation on Degradation and Properties of TMC<br />

degraded by lysozyme to the same extent <strong>as</strong> the original chitosan. In contr<strong>as</strong>t, chitosan<br />

quaternized using iodomethane and a strong b<strong>as</strong>e (TMC-RA-OM) w<strong>as</strong> cleaved less than the<br />

original chitosan (CS-RA) or the O-methyl free TMC-RA, likely due to loss of N-acetylated units<br />

during synthesis. To explain, the M n of TMC-RA-OM with a DAc of 49% decre<strong>as</strong>ed from 40 kDa<br />

to 18 kDa. Although TMC-RA-OM h<strong>as</strong> a modestly higher weight per glucosamine unit than CS-<br />

RA, the difference in final M n of 7 kDa of CS-RA and 18 kDa of TMC-RA-OM h<strong>as</strong> to be attributed<br />

to less chain scission by lysozyme <strong>for</strong> TMC-RA-OM due to its lower DAc <strong>as</strong> compared to the<br />

other re-acetylated polymers. Interestingly, TMC-RA and, especially DMC-RA, were f<strong>as</strong>ter<br />

cleaved by lysozyme than CS-RA implying that enzyme-affinity and/or maximum substrate<br />

conversion rates may be enhanced by the methylation of the free amines.<br />

Mn (Da)<br />

4.0×10 4 CS<br />

CS-RA<br />

DMC-RA<br />

3.0×10 4<br />

TMC-RA<br />

TMC-RA-OM<br />

TMC<br />

2.0×10 4<br />

TMC-OM<br />

1.0×10 4<br />

0<br />

0 5 10<br />

100 200<br />

time (h)<br />

Figure 3. Degradation of various chitosans and TMCs by lysozyme (38 µg/ml) at pH 4.5 and 37°C. Error<br />

bars represent the standard deviation of three me<strong>as</strong>urements.<br />

To investigate the lysozyme-catalyzed degradation of quaternized chitosans under<br />

physiological conditions, changes in molecular weight of TMCs in the presence of lysozyme<br />

were studied in PBS at pH 7.4. The extent of lysozyme-catalyzed degradation of quaternized<br />

chitosans w<strong>as</strong> comparable to the degradation observed at pH 4.5 (Figure 4). In detail, at pH 7.4<br />

59


Chapter 3<br />

the polymer with the lowest DAc of 11% (TMC-OM) showed the smallest decre<strong>as</strong>e in molecular<br />

weight (M n from 34 to 29 kDa) while the TMC with a DAc of 55% (TMC-RA) w<strong>as</strong> readily<br />

degraded (M n from 43 to 11 kDa). At pH 4.5 the lysozyme-catalyzed degradation of the reacetylated<br />

TMC-RA and TMC-RA-OM w<strong>as</strong> slightly f<strong>as</strong>ter than at pH 7.4 likely due to higher<br />

enzyme activity at lower pH [9]. Overall, our data show that the degradation of quaternized<br />

chitosans by lysozyme both in terms of kinetics and in final molecular weight of the<br />

degradation products is hardly affected by the pH in the range studied (pH 4.5-7.4). This w<strong>as</strong><br />

not observed <strong>for</strong> the degradation of chitosan, where the degradation rate at pH 7.4 is<br />

substantially slower than that in acidic environment [14]. At a pH above 6.5 chitosans are<br />

insoluble and degradation by enzymes decre<strong>as</strong>es due to limited accessibility of the binding<br />

sites <strong>for</strong> lysozyme. Since TMCs are soluble at physiological pH, solubility does not limit the<br />

lysozyme-catalyzed degradation.<br />

4.0×10 4 TMC<br />

Mn (Da)<br />

3.0×10 4<br />

2.0×10 4<br />

TMC-RA<br />

TMC-OM<br />

TMC-RA-OM<br />

1.0×10 4<br />

0<br />

0 5 10<br />

100 200<br />

time (h)<br />

Figure 4. Degradation of various TMCs by lysozyme (38 µg/ml) at pH 7.4 and 37°C. Error bars represent<br />

the standard deviation of three me<strong>as</strong>urements.<br />

The degradability of TMCs will have important consequences <strong>for</strong> its pharmaceutical<br />

applications. Onishi and co-workers demonstrated in mice that, after intraperitoneally<br />

injection, chitosan with a DAc of 50% is readily degraded to a molecular weight of less than 10<br />

60


Influence of Degree of Acetylation on Degradation and Properties of TMC<br />

kDa and after 24 hours the full dose w<strong>as</strong> excreted by the kidneys [54]. Since glomerular<br />

filtration by kidneys is size-dependent, our results imply that the degradation products of the<br />

re-acetylated polymers, due to their lower final molecular weight, will be more e<strong>as</strong>ily removed<br />

from the body than the chitosans with a low DAc. Also, TMC-carrier systems may be tailored<br />

<strong>for</strong> its degradation characteristics; ideally, the carrier molecule will degrade after delivery at<br />

the target site thereby rele<strong>as</strong>ing its pharmaceutically active compounds.<br />

In conclusion, our data demonstrate that the lysozyme-catalyzed degradation of TMCs is, like<br />

that of chitosan, facilitated <strong>for</strong> polymers with a high DAc. However, the degradation of TMC is<br />

similar at pH 4.5 and 7.4.<br />

Evaluation of various TMCs on TEER, MTT and LDH <strong>as</strong>says. TMC is capable to open tight<br />

junctions of epithelial cells and the highest activity w<strong>as</strong> observed with DQs of around 40-50%<br />

[29, 35, 55]. Recently, it w<strong>as</strong> shown that O-methylation substantially reduces the activity of<br />

TMC on transepithelial electrical resistance (TEER) [40]. In this study, TMCs were tested with<br />

only minor variations in DQ (see Table 1), so the charge ratio and the cationic character of<br />

these polymers were considered similar. However, the polymers differed in DAc and in extent<br />

of O-methylation.<br />

TEER (% HBSS)<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

SDS 1%<br />

TMC<br />

TMC-OM<br />

TMC-RA<br />

TMC-RA-OM<br />

0<br />

0 15 30 45 60 75 90<br />

time (min.)<br />

1140<br />

Figure 5. Effect of TMCs with various DAcs and with or without O-methylated units on the TEER of<br />

Caco-2 cells at a TMC concentration of 2.5 mg/ml. SDS 10 mg/ml w<strong>as</strong> used <strong>as</strong> a positive control. Error<br />

bars represent the standard deviation of six me<strong>as</strong>urements.<br />

Figure 5 shows the effect of various TMCs on the TEER of a Caco-2 monolayer. In line with<br />

previous data [40], O-methyl free TMC with a DQ of 43% and a DAc of 17% demonstrated the<br />

61


Chapter 3<br />

largest decre<strong>as</strong>e in TEER (about 70% reduction), while the partially O-methylated TMC with a<br />

DQ of 42% and a DAc of 11% reduced the TEER with about 30% compared to the TEER<br />

observed with HBSS. Importantly, the Caco-2 monolayer partially recovered after removal of<br />

these polymers implying that no permanent toxicity w<strong>as</strong> induced on the cells. F<strong>as</strong>cinatingly,<br />

TMC polymers with a similar DQ but with higher DAcs had no effect on the TEER, indicating<br />

that these polymers, although positively charged, were unable to open tight junctions. Also, the<br />

results of the MTT and LDH <strong>as</strong>says (Figures 6 and 7) obtained with TMC and TMC-OM were <strong>as</strong><br />

expected [40], showing a higher toxicity <strong>for</strong> O-methyl free TMC. Additionally, even at a<br />

concentration of 10 mg/ml both TMC-RA and TMC-RA-OM clearly show neither a decre<strong>as</strong>e in<br />

cell viability nor LDH rele<strong>as</strong>e. Studies with chitosan have shown that an incre<strong>as</strong>e in DAc lead to<br />

lower penetration through a Caco-2 monolayer [20], a decre<strong>as</strong>e in cytotoxicity [18, 20] and a<br />

reduction in DNA transfection [19]. However, these experiments were carried out in slightly<br />

acidic environment to solubilize the chitosan polymer. In acidic environment the primary<br />

amines will be (partially) protonated and thus become positively charged. A higher DAc will<br />

reduce the number of amines available <strong>for</strong> protonation thereby decre<strong>as</strong>ing the charge density<br />

of chitosan. There<strong>for</strong>e, with chitosan, it w<strong>as</strong> not possible to study the effect of DAc on polymer<br />

characteristics without altering the charge density of the polymer.<br />

cell viability (%)<br />

125<br />

100<br />

75<br />

50<br />

25<br />

TMC<br />

TMC-OM<br />

TMC-RA<br />

TMC-RA-OM<br />

0<br />

0.1 1 10<br />

conc. (mg/ml)<br />

Figure 6. Effect of TMCs with various DAcs and with or without O-methylated units on the viability of<br />

Caco-2 cells (MTT <strong>as</strong>say) at different concentrations. Error bars represent the standard deviation of six<br />

me<strong>as</strong>urements.<br />

62


Influence of Degree of Acetylation on Degradation and Properties of TMC<br />

LDH rele<strong>as</strong>e (%)<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

HBSS<br />

TMC-RA-OM<br />

TMC-RA<br />

TMC-OM<br />

TMC<br />

0<br />

0.1 1 10<br />

conc. (mg/ml)<br />

Figure 7. Effect of TMCs with various DAcs and with or without O-methylated units on the viability of<br />

Caco-2 cells (LDH rele<strong>as</strong>e) at different concentrations. Error bars represent the standard deviation of six<br />

me<strong>as</strong>urements.<br />

Our studies were done at physiological pH and with TMCs with a similar DQ (and thus charge<br />

density) allowing evaluation of the influence of the DAc without changing other polymer<br />

characteristics. Consequently the remarkable decre<strong>as</strong>e in toxicity and ability to open tight<br />

junctions when incre<strong>as</strong>ing the DAc h<strong>as</strong> to be contributed to intra- and/or intermolecular<br />

changes induced by the N-acetylated units. Possibly, the N-acetylated glucosamine parts of the<br />

TMC can <strong>for</strong>m hydrophobic domains, thereby shielding the positively charged quaternized<br />

parts of the macromolecule. However, further studies are needed to provide more insight into<br />

the macromolecular changes that occur when incre<strong>as</strong>ing the DAc.<br />

Conclusion<br />

This paper presents a method to synthesize TMC with a high DAc without introducing other<br />

alterations of the polymer such <strong>as</strong> O-methylation, chain scission or loss of N-acetylation. The<br />

enzymatic degradation of TMC by lysozyme w<strong>as</strong> essentially identical to the degradation of<br />

chitosan (thus highly dependent on the degree of acetylation) however, in contr<strong>as</strong>t to chitosan,<br />

this degradation w<strong>as</strong> pH independent. TMCs with a DAc of ~50%, a DQ of around 44% and<br />

with or without O-methylation, were not able to open tight junctions of Caco-2 cells, in contr<strong>as</strong>t<br />

to TMCs (O-methylated or O-methyl free) with a similar DQ but a lower DAc. Additionally,<br />

TMCs with a high DAc (~50%) showed no cell toxicity up to a concentration of 10 mg/ml. In<br />

conclusion, the degree of N-acetylation h<strong>as</strong> great impact on the biological characteristics of<br />

63


Chapter 3<br />

TMC and future research will be done to determine the optimal polymer structure <strong>for</strong> the<br />

various potential pharmaceutical applications of TMC.<br />

Acknowledgement. This research w<strong>as</strong> per<strong>for</strong>med under the framework of <strong>TI</strong> <strong>Pharma</strong><br />

project number D5-106-1; Vaccine delivery: alternatives <strong>for</strong> conventional multiple injection<br />

vaccines. The authors thank Sara Riera Riv<strong>as</strong> <strong>for</strong> her contributions to the project.<br />

64


Influence of Degree of Acetylation on Degradation and Properties of TMC<br />

References<br />

1. Mourya VK, Inamdar NN. Chitosan-modifications and applications: Opportunities galore. React<br />

Funct Polym 68:1013-51 (2008).<br />

2. Agnihotri SA, Mallikarjuna NN, Aminabhavi TM. Recent advances on chitosan-b<strong>as</strong>ed microand<br />

nanoparticles in drug delivery. J Control Rele<strong>as</strong>e 100:5-28 (2004).<br />

3. Kumar MNVR, Muzzarelli RAA, Muzzareilli C, S<strong>as</strong>hiwa H, Domb AJ. Chitosan chemistry and<br />

pharmaceutical perspectives. Chem Rev 104:6017-84 (2004).<br />

4. Van Der Lubben IM, Kersten G, Fretz MM, Beuvery C, Verhoef JC, Junginger HE. Chitosan<br />

microparticles <strong>for</strong> mucosal vaccination against diphtheria: Oral and n<strong>as</strong>al efficacy studies in<br />

mice. Vaccine 21:1400-8 (2003).<br />

5. Muzzarelli RAA. Human enzymatic activities related to the therapeutic administration of chitin<br />

derivatives. Cel Mol Life Sci 53:131-40 (1997).<br />

6. Vårum KM, Myhr MM, Hjerde RJN, Smidsrød O. In vitro degradation rates of partially N-<br />

acetylated chitosans in human serum. Carbohydr Res 299:99-101 (1997).<br />

7. Nordtveit RJ, Varum KM, Smidsrod O. Degradation of fully water-soluble, partially N-<br />

acetylated chitosans with lysozyme. Carbohydr Polym 23:253-60 (1994).<br />

8. Varum KM, Ottoy MH, Smidsrod O. Acid hydrolysis of chitosans. Carbohydr Polym 46:89-98<br />

(2001).<br />

9. Freier T, Koh HS, Kazazian K, Shoichet MS. Controlling cell adhesion and degradation of<br />

chitosan films by N-acetylation. Biomaterials 26:5872-78 (2005).<br />

10. Nordtveit RJ, Varum KM, Smidsrod O. Degradation of partially N-acetylated chitosans with<br />

hen egg white and human lysozyme. Carbohydr Polym 29:163-7 (1996).<br />

11. Ren D, Yi H, Wang W, Ma X. The enzymatic degradation and swelling properties of chitosan<br />

matrices with different degrees of N-acetylation. Carbohydr Res 340:2403-10 (2005).<br />

12. S<strong>as</strong>hiwa H, Saimoto H, Shigem<strong>as</strong>a Y, Ogawa R, Tokura S. Lysozyme susceptibility of partially<br />

deacetylated chitin. Int J Biol Macromol 12:295-6 (1990).<br />

13. Tomihata K, Ikada Y. In vitro and in vivo degradation of films of chitin and its deacetylated<br />

derivatives. Biomaterials 18:567-75 (1997).<br />

14. Shigem<strong>as</strong>a Y, Saito K, S<strong>as</strong>hiwa H, Saimoto H. Enzymatic degradation of chitins and partially<br />

deacetylated chitins. Int J Biol Macromol 16:43-9 (1994).<br />

15. Lewis CE, McCarthy SP, Lorenzen J, McGee JOD. Differential effects of LPS, IFN-γ and<br />

TNFα on the secretion of lysozyme by individual human mononuclear phagocytes: relationship<br />

to cell maturity. Immunology 9:402-8 (1990).<br />

16. Pangburn SH, Trescony PV, Heller J. Lysozyme degradation of partially deacetylated chitin, its<br />

films and hydrogels. Biomaterials 3:105-8 (1982).<br />

17. Yang Y, Hu W, Wang X, Gu X. The controlling biodegradation of chitosan fibers by N-<br />

acetylation in vitro and in vivo. J Mater Sci: Mater Med 18:2117-21 (2007).<br />

18. Huang M, Khor E, Lim LY. Uptake and cytotoxicity of chitosan molecules and nanoparticles:<br />

effects of molecular weight and degree of deacetylation. Pharm Res 21:344-53 (2004).<br />

19. Kiang T, Wen J, Lim HW, Leong KWKW. The effect of the degree of chitosan deacetylation<br />

on the efficiency of gene transfection. Biomaterials 25:5293-301 (2004).<br />

20. Schipper NGM, Vårum KM, Artursson P. <strong>Chitosans</strong> <strong>as</strong> absorption enhancers <strong>for</strong> poorly<br />

absorbable drugs. 1: Influence of molecular weight and degree of acetylation on drug transport<br />

across human intestinal epithelial (Caco-2) cells. Pharm Res 13:1686-92 (1996).<br />

21. Nishimura K, Ishihara C, Ukei S, Tokura S, Azuma I. Stimulation of cytokine production in<br />

mice using deacetylated chitin. Vaccine 4:151-56 (1986).<br />

22. Nishimura K, Nishimura S, Nishi N. Immunological activity of chitin and its derivatives. Vaccine<br />

2:93-98 (1984).<br />

23. Nishimura K, Nishimura S-i, Nishi N, Numata F, Tone Y, Tokura S et al. <strong>Adjuvant</strong> activity of<br />

chitin derivatives in mice and guinea-pigs. Vaccine 3:379-84 (1985).<br />

65


Chapter 3<br />

24. Robinson MJ, Sancho D, Slack EC, LeibundGut-Landmann S, Reis e Sousa C. Myeloid C-type<br />

lectins in innate immunity. Nat Immunol 7:1258-65 (2006).<br />

25. Amidi M, Romeijn SG, Borchard G, Junginger HE, Hennink WE, Jiskoot W. Preparation and<br />

characterization of protein-loaded N-trimethyl chitosan nanoparticles <strong>as</strong> n<strong>as</strong>al delivery system. J<br />

Control Rele<strong>as</strong>e 111:107-16 (2006).<br />

26. Boonyo W, Junginger HE, Waranuch N, Polnok A, Pitaksuteepong T. Chitosan and trimethyl<br />

chitosan chloride (TMC) <strong>as</strong> adjuvants <strong>for</strong> inducing immune responses to ovalbumin in mice<br />

following n<strong>as</strong>al administration. J Control Rele<strong>as</strong>e 121:168-75 (2007).<br />

27. Chen F, Zhang Z-R, Yuan F, Qin X, Wang M, Huang Y. In vitro and in vivo study of N-<br />

trimethyl chitosan nanoparticles <strong>for</strong> oral protein delivery. Int J Pharm 349:226-33 (2008).<br />

28. Kean T, Roth S, Thanou M. <strong>Trimethyl</strong>ated chitosans <strong>as</strong> non-viral gene delivery vectors:<br />

Cytotoxicity and transfection efficiency. J Control Rele<strong>as</strong>e 103:643-53 (2005).<br />

29. Van der Lubben IM, Verhoef JC, Fretz MM, Van O, Mesu I, Kersten G, Junginger HE.<br />

<strong>Trimethyl</strong> chitosan chloride (TMC) <strong>as</strong> a novel excipient <strong>for</strong> oral and n<strong>as</strong>al immunisation against<br />

diphtheria. STP Pharm Sci 12:235-42 (2002).<br />

30. Sayin B, Somavarapu S, Li XW, Thanou M, Sesardic D, Alpar HO et al. Mono-Ncarboxymethyl<br />

chitosan (MCC) and N-trimethyl chitosan (TMC) nanoparticles <strong>for</strong> non-inv<strong>as</strong>ive<br />

vaccine delivery. Int J Pharm 363:139-48 (2008).<br />

31. Thanou M, Florea BI, Geldof M, Junginger HE, Borchard G. Quaternized chitosan oligomers<br />

<strong>as</strong> novel gene delivery vectors in epithelial cell lines. Biomaterials 23:153-9 (2002).<br />

32. Chen F, Zhang ZR, Huang Y. Evaluation and modification of N-trimethyl chitosan chloride<br />

nanoparticles <strong>as</strong> protein carriers. Int J Pharm 336:166-73 (2007).<br />

33. Di Colo G, Burgal<strong>as</strong>si S, Zambito Y, Monti D, Chetoni P. Effects of different N-trimethyl<br />

chitosans on in vitro/in vivo ofloxacin transcorneal permeation. J Pharm Sci 93:2851-62 (2004).<br />

34. Snyman D, Hamman JH, Kotze AF. Evaluation of the mucoadhesive properties of N-trimethyl<br />

chitosan chloride. Drug Develop Ind Pharm 29:61-9 (2003).<br />

35. Thanou MM, Kotze AF, Scharringhausen T, Lueßen HL, De Boer AG, Verhoef JC et al. Effect<br />

of degree of quaternization of N-trimethyl chitosan chloride <strong>for</strong> enhanced transport of<br />

hydrophilic compounds across intestinal Caco-2 cell monolayers. J Control Rele<strong>as</strong>e 64:15-25<br />

(2000).<br />

36. Hamman JH, Stander M, Kotze AF. Effect of the degree of quaternisation of N-trimethyl<br />

chitosan chloride on absorption enhancement: In vivo evaluation in rat n<strong>as</strong>al epithelia. Int J<br />

Pharm 232:235-42 (2002).<br />

37. Sandri G, Bonferoni MC, Rossi S, Ferrari F, Gibin S, Zambito Y et al. Nanoparticles b<strong>as</strong>ed on<br />

N-trimethylchitosan: Evaluation of absorption properties using in vitro (Caco-2 cells) and ex<br />

vivo (excised rat jejunum) models. Eur J Pharm and Biopharm 65:68-77 (2007).<br />

38. Mao S, Shuai X, Unger F, Wittmar M, Xie X, Kissel T. Synthesis, characterization and<br />

cytotoxicity of poly(ethylene glycol)-graft-trimethyl chitosan block copolymers. Biomaterials<br />

26:6343-56 (2005).<br />

39. Guo Z, Xing R, Liu S, Zhong Z, Ji X, Wang L et al. The influence of molecular weight of<br />

quaternized chitosan on antifungal activity. Carbohydr Polym 71:694-7 (2008).<br />

40. Verheul RJ, Amidi M, van der Wal S, van Riet E, Jiskoot W, Hennink WE. Synthesis,<br />

characterization and in vitro biological properties of O-methyl free N,N,N-trimethylated<br />

chitosan. Biomaterials 29:3642-9 (2008).<br />

41. Guo Z, Xing R, Liu S, Zhong Z, Ji X, Wang L et al. Antifungal properties of Schiff b<strong>as</strong>es of<br />

chitosan, N-substituted chitosan and quaternized chitosan. Carbohydr Res 2007;342:1329-32.<br />

42. Jia Z, shen D, Xu W. Synthesis and antibacterial activities of quaternary ammonium salt of<br />

chitosan. Carbohydr Res 333:1-6 (2001).<br />

43. Sieval AB, Thanou M, Kotze AF, Verhoef JC, Brussee J, Junginger HE. Preparation and NMR<br />

characterization of highly substituted N-trimethyl chitosan chloride. Carbohydr Polym 36:157-65<br />

(1998).<br />

66


Influence of Degree of Acetylation on Degradation and Properties of TMC<br />

44. Lavertu M, Xia Z, Serreqi AN, Berrada M, Rodrigues A, Wang D et al. A validated 1H NMR<br />

method <strong>for</strong> the determination of the degree of deacetylation of chitosan. J Pharm Biomed Anal<br />

32:1149-58 (2003).<br />

45. Polnok A, Borchard G, Verhoef JC, Sarisuta N, Junginger HE. Influence of methylation<br />

process on the degree of quaternization of N-trimethyl chitosan chloride. Eur J Pharm Biopharm<br />

57:77-83 (2004).<br />

46. Jiang X, van der Horst A, van Steenbergen M, Akeroyd N, van Nostrum C, Schoenmakers P et<br />

al. Molar-m<strong>as</strong>s characterization of cationic polymers <strong>for</strong> gene delivery by aqueous size-exclusion<br />

chromatography. Pharm Res 23:595-603 (2006).<br />

47. Nerurkar MM, Burton PS, Borchardt RT. The use of surfactants to enhance the permeability of<br />

peptides through Caco-2 cells by inhibition of an apically polarized efflux system. Pharm Res<br />

13:528-34 (1996).<br />

48. Mosmann T. Rapid colorimetric <strong>as</strong>say <strong>for</strong> cellular growth and survival: Application to<br />

proliferation and cytotoxicity <strong>as</strong>says. J Immunol Methods 65:55-65 (1983).<br />

49. Snyman D, Hamman JH, Kotze JS, Rollings JE, Kotze AF. The relationship between the<br />

absolute molecular weight and the degree of quaternisation of N-trimethyl chitosan chloride.<br />

Carbohydr Polym 50:145-50 (2002).<br />

50. Hu Y, Du Y, Yang J, Tang Y, Li J, Wang X. Self-aggregation and antibacterial activity of N-<br />

acylated chitosan. Polymer 48:3098-106 (2007).<br />

51. Schatz C, Viton C, Delair T, Pichot C, Domard A. Typical physicochemical behaviors of<br />

chitosan in aqueous solution. Biomacromolecules 4:641-8 (2003).<br />

52. Yanagisawa M, Kato Y, Yoshida Y, Isogai A. SEC-MALS study on aggregates of chitosan<br />

molecules in aqueous solvents: Influence of residual N-acetyl groups. Carbohydr Polym 66:192-98<br />

(2006).<br />

53. Zhang H, Neau SH. In vitro degradation of chitosan by bacterial enzymes from rat cecal and<br />

colonic contents. Biomaterials 23:2761-66 (2002).<br />

54. Onishi H, Machida Y. Biodegradation and distribution of water-soluble chitosan in mice.<br />

Biomaterials 20:175-182 (1999).<br />

55. Kotze AF, Luessen HL, de Leeuw BJ, de Boer AG, Verhoef JC, Junginger HE. N-<strong>Trimethyl</strong><br />

chitosan chloride <strong>as</strong> a potential absorption enhancer across mucosal surfaces: in vitro evalutation<br />

in intestinal epithelial cells (Caco-2). Pharm Res 14:1197-202 (1997).<br />

67


CHAPTER 4<br />

RELA<strong>TI</strong>ONSHIP BETWEEN STRUCTURE AND<br />

ADJUVAN<strong>TI</strong>CITY OF N,N,N-TRIMETHYL<br />

CHITOSAN (TMC) STRUCTURAL VARIANTS<br />

IN A NASAL INFLUENZA VACCINE<br />

Rolf J. Verheul*, Niels Hagenaars*, Imke Mooren, P<strong>as</strong>cal H.J.L.F. de Jong,<br />

Enrico M<strong>as</strong>trobattista, Harrie L. Glansbeek, Jacco G.M. Heldens<br />

Han van den Bosch, Wim E. Hennink, Wim Jiskoot.<br />

*authors contributed equally<br />

Journal of Controlled Rele<strong>as</strong>e 2009, 140, 126-133


Chapter 4<br />

Abstract<br />

The aim of this study w<strong>as</strong> to <strong>as</strong>sess the influence of structural properties of N,N,N-trimethyl<br />

chitosan (TMC) on its adjuvanticity. There<strong>for</strong>e, TMCs with varying degrees of quaternization<br />

(DQ, 22-86%), O-methylation (DOM, 0-76%) and acetylation (DAc 9-54%) were <strong>for</strong>mulated<br />

with whole inactivated influenza virus (WIV). The <strong>for</strong>mulations were characterized<br />

physicochemically and evaluated <strong>for</strong> their immunogenicity in an intran<strong>as</strong>al (i.n.)<br />

vaccination/challenge study in mice.<br />

Simple mixing of the TMCs with WIV at a 1:1 (w/w) ratio resulted in comparable positively<br />

charged nanoparticles, indicating coating of WIV with TMC. The amount of free TMC in solution<br />

w<strong>as</strong> comparable <strong>for</strong> all TMC-WIV <strong>for</strong>mulations. After i.n. immunization of mice with WIV and<br />

TMC-WIV on day 0 and 21, all TMC-WIV <strong>for</strong>mulations induced stronger total IgG, IgG1 and<br />

IgG2a/c responses than WIV alone, except WIV <strong>for</strong>mulated with reacetylated TMC with a DAc<br />

of 54% and a DQ of 44% (TMC-RA44). No significant differences in antibody titers were<br />

observed <strong>for</strong> TMCs that varied in DQ or DOM, indicating that these structural characteristics<br />

play a minor role in their adjuvant properties. TMC with a DQ of 56% (TMC56) <strong>for</strong>mulated<br />

with WIV at a ratio of 5:1 (w/w) resulted in significantly lower IgG2a/c:IgG1 ratio’s compared<br />

to TMC56 mixed in ratios of 0.2:1 and 1:1, implying a shift towards a Th2 type immune<br />

response. Challenge of vaccinated mice with aerosolized virus demonstrated protection <strong>for</strong> all<br />

TMC-WIV <strong>for</strong>mulations with the exception of TMC-RA44-WIV.<br />

In conclusion, <strong>for</strong>mulating WIV with TMCs strongly enhances the immunogenicity and<br />

induced protection after i.n. vaccination with WIV. The adjuvant properties of TMCs <strong>as</strong> i.n.<br />

adjuvant are strongly decre<strong>as</strong>ed by reacetylation of TMC, where<strong>as</strong> the DQ and DOM hardly<br />

affect the adjuvanticity of TMC.<br />

70


Relationship between Structure and <strong>Adjuvant</strong>icity of TMC<br />

Introduction<br />

Intran<strong>as</strong>al (i.n.) vaccination offers several advantages over the intramuscular (i.m.) route,<br />

like simple, needle-free administration without the need <strong>for</strong> trained personnel, potentially less<br />

adverse effects and the induction of local mucosal immune responses [1]. On the other hand,<br />

vaccines administered via the i.n. route generally induce low systemic immune responses<br />

when compared to i.m. administration likely due to mucociliary clearance and low antigen<br />

uptake. Mucoadhesive polymers have been used to incre<strong>as</strong>e the immunogenicity of i.n.<br />

vaccines by incre<strong>as</strong>ing n<strong>as</strong>al residence time and enhancing antigen presentation [1].<br />

Chitosan, a polysaccharide that is obtained by deacetylation of the natural polymer chitin,<br />

h<strong>as</strong> mucoadhesive properties and showed promising results <strong>as</strong> an adjuvant in n<strong>as</strong>al vaccines<br />

[2-5]. However, the unfavorable pH-dependent solubility and charge density led to the<br />

synthesis of its quaternized derivative N,N,N-trimethyl chitosan (TMC) (Scheme 1), which is<br />

well soluble in aqueous solution at neutral pH.<br />

Scheme 1. General structure of TMC. Depending on the synthesis route TMCs can be varied in degree of<br />

acetylation (see block ‘x’), quaternization (see block ‘y’), and O-methylation (see block ‘z’). The various<br />

substitutions are randomly distributed throughout the polymer; O-methylation (block ‘z’) may also<br />

occur on the quaternized and acetylated units (blocks ‘x’ and ‘y’).<br />

TMC is traditionally synthesized by reaction of chitosan with excess iodomethane in strong<br />

alkaline conditions with N-methyl-2-pyrrolidone (NMP) <strong>as</strong> solvent and the degree of<br />

quaternization (DQ) can be varied by varying the number of reaction steps [6]. Besides N-<br />

methylation this synthesis method also introduces substantial O-methylation on the hydroxyl<br />

groups located at the C-3 and C-6 of the glucosamine unit. The degree of O-methylation (DOM)<br />

incre<strong>as</strong>es with incre<strong>as</strong>ing DQ (up to 80-90%) [7, 8]. Recently, O-methyl free TMC w<strong>as</strong><br />

synthesized using a novel two-step synthesis procedure, allowing good control of the DQ<br />

without altering other structural properties. Both DQ and DOM were found to influence<br />

toxicity and transepithelial electrical resistance (TEER), an indicator <strong>for</strong> opening of tight<br />

71


Chapter 4<br />

junctions, in a Caco-2 cell model (Chapter 2). A higher DQ leads to more toxicity and a stronger<br />

TEER effect, with a maximum effect on TEER at a DQ above 60%. Furthermore, O-methyl free<br />

TMC h<strong>as</strong> a much stronger effect on TEER than O-methylated TMC (TMC-OM) and shows more<br />

in vitro cell toxicity [8]. Another characteristic of TMCs is the degree of N-acetylation (DAc).<br />

Partial reacetylation of TMC (from 17 to 54%) decre<strong>as</strong>ed the in vitro cell toxicity and effect on<br />

TEER but incre<strong>as</strong>ed the enzymatic degradability of TMC by lysozyme (Chapter 3) [9].<br />

Little is known about the relationship between the structural characteristics and adjuvant<br />

properties of TMCs in vivo. For TMC-OM solutions in i.n. vaccination with ovalbumin an<br />

optimal DQ of 40% w<strong>as</strong> reported although differences were small [10]. Previously, whole<br />

inactivated influenza virus (WIV) vaccine w<strong>as</strong> <strong>for</strong>mulated with TMC-OM with a DQ of 15% or<br />

37%. This resulted in positively charged nanoparticles with partially bound TMC-OM. These<br />

particles had an intact viral ultr<strong>as</strong>tructure. Strong, protective immune responses were induced<br />

after i.n. vaccination [11]. No significant differences were observed between the two different<br />

TMC-OMs. Most likely, TMC exerts its adjuvant effect by an improved antigen delivery, through<br />

an incre<strong>as</strong>ed n<strong>as</strong>al residence time and/or enhanced uptake through the epithelium and by<br />

antigen presenting cells. Besides differences in DQ, the TMC-OMs used in these studies also<br />

differed in DOM and, likely, polymer molecular weight. So, the individual contributions of DQ,<br />

DAc and DOM on the adjuvant effect of TMC are unknown.<br />

In the present study we investigated <strong>for</strong> i.n. administered WIV the adjuvant properties of O-<br />

methyl free TMCs with varying DQs and reacetylated O-methyl free TMC in comparison to<br />

conventional TMC-OMs with similar DQ. The TMC-WIV vaccines were physicochemically<br />

characterized and the immunogenicity and protectivity of the vaccines were <strong>as</strong>sessed in a<br />

murine challenge model. Additionally, the influence of TMC:WIV ratio on the quality and<br />

quantity of humoral immune responses w<strong>as</strong> investigated.<br />

Materials and Methods<br />

Materials. Chitosan with a DAc of 17% (determined with 1 H-NMR <strong>as</strong> described in [9]) and a<br />

number average molecular weight (M n) and weight average molecular weight (M w) of 28 and<br />

43 kDa, <strong>as</strong> determined by gel permeation chromatography (GPC) <strong>as</strong> described in [8],<br />

respectively, w<strong>as</strong> purch<strong>as</strong>ed from Primex (Siglufjordur, Iceland). Acetic anhydride, sodium<br />

borohydrate, <strong>for</strong>mic acid, <strong>for</strong>maldehyde 37% (stabilized with methanol), deuterium oxide,<br />

sodium acetate, acetic acid (anhydrous), sodium hydroxide and hydrochloric acid were<br />

obtained from Sigma-Aldrich Chemical Co. Iodomethane 99% stabilized with copper w<strong>as</strong><br />

72


Relationship between Structure and <strong>Adjuvant</strong>icity of TMC<br />

obtained from Acros Organics (Geel, Belgium). Live, egg-grown, mouse adapted influenza<br />

A/Puerto Rico/8/34 virus (A/PR/8/34) and purified, cell culture-grown (Madin-Darby Canine<br />

Kidney (MDCK) cells), β-propiolacton (BPL)-inactivated A/PR/8/34, <strong>as</strong> well <strong>as</strong> polyclonal<br />

rabbit anti-A/PR/8/34 serum were from Nobilon International BV, Boxmeer, The Netherlands.<br />

PO-labeled goat anti mouse -IgG (H+L), -IgG1, -IgG2a/c and -IgA(Fc) were purch<strong>as</strong>ed from<br />

Nordic Immunological Laboratories (Tilburg, The Netherlands). All other chemicals used were<br />

of analytical grade.<br />

Synthesis and characterization of methylated chitosans. N,N,N-<strong>Trimethyl</strong>ated chitosans<br />

with varying DQ and DAc, and DOM were synthesized from chitosan <strong>as</strong> described previously<br />

[8, 9]. Briefly, O-methyl free TMCs were made with a two-step method: first quantitative<br />

dimethylation of the free amino-groups of chitosan with <strong>for</strong>maldehyde and <strong>for</strong>mic acid w<strong>as</strong><br />

carried out, followed by reaction of the dimethylated chitosan with an excess of iodomethane.<br />

By varying the reaction time the DQ of the TMCs could be tailored [8]. TMC-OMs, with<br />

substantial O-methylation of the hydroxyl groups on the C-3 and C-6 of the glucosamine units,<br />

were synthesized according to the method of Sieval et al. [6, 8]. Here, chitosan w<strong>as</strong><br />

trimethylated to various extents by reacting with iodomethane in the presence of a strong b<strong>as</strong>e<br />

(NaOH) <strong>for</strong> several times depending on the desired DQ. Finally, to obtain TMC with a high<br />

degree of acetylation, chitosan w<strong>as</strong> first re-acetylated using acetic anhydride [9, 12]. Then, this<br />

re-acetylated chitosan w<strong>as</strong> quantitatively dimethylated with <strong>for</strong>maldehyde and sodium<br />

borohydrate followed by complete trimethylation of the dimethylated amino groups with<br />

iodomethane [9]. All synthesized TMCs were dissolved in an aqueous 10% w/v NaCl solution,<br />

put on a shaker overnight <strong>for</strong> ion-exchange and the obtained solution w<strong>as</strong> dialyzed at room<br />

temperature against deionized water <strong>for</strong> 3 days changing water twice daily, filtered through a<br />

0.8 µm filter and freeze dried.<br />

The DQ, DAc and DOM of the hydroxyl groups on C-3 and C-6 (DOM-3 and DOM-6,<br />

respectively) of the TMCs were determined with 1 H-NMR on a Varian INOVA 500MHz NMR<br />

spectrometer (Varian Inc., Palo Alto, Ca, USA) at 80 °C in D 2O [9]. Furthermore, M n and M w of<br />

the various TMCs were determined, <strong>as</strong> described previously [9, 13], by GPC on a Viscotek<br />

system detecting refractive index, viscosity and light scattering. A Shodex OHPak SB-806<br />

column (30 cm) w<strong>as</strong> used with 0.3 M sodium acetate pH 4.4 (adjusted with acetic acid) <strong>as</strong><br />

running buffer. The structural characteristics of the synthesized TMCs are summarized in<br />

Table 1.<br />

73


Chapter 4<br />

Preparation of TMC-WIV <strong>for</strong>mulations. Purified, cell culture-derived, BPL-inactivated<br />

A/PR/8/34 suspended in a 10 mM phosphate buffered saline solution (150 mM NaCl, pH 7.4)<br />

(PBS) w<strong>as</strong> concentrated by centrifugation at 22,000 x g <strong>for</strong> 30 min at 4 °C and resuspended in 5<br />

mM HEPES buffer (pH 7.4). The WIV concentration is expressed <strong>as</strong> mg total protein/ml <strong>as</strong><br />

determined by DC protein <strong>as</strong>say (Bio-Rad, Hercules, CA, USA). The amount of hemagglutinin<br />

(HA) w<strong>as</strong> approximately 35 % of the total protein content, <strong>as</strong> determined previously [14]. The<br />

TMC-WIV vaccines were prepared by adding equal volumes of TMC solution (in 5 mM HEPES,<br />

pH 7.4) to a WIV dispersion (in 5 mM HEPES pH 7.4) at a 1:1 w/w ratio using a Gilson pipette<br />

while gently mixing <strong>for</strong> 5 seconds. TMC-WIV w<strong>as</strong> <strong>for</strong>mulated at a final WIV concentration of<br />

1.25 mg/ml, except <strong>for</strong> the samples used <strong>for</strong> dynamic light scattering (DLS) and zeta-potential<br />

me<strong>as</strong>urements, which were per<strong>for</strong>med at lower concentrations (62.5 µg/ml) <strong>for</strong> optimal testconditions.<br />

To study the immunogenicity of TMC-WIV at other ratios, TMC56 w<strong>as</strong> also<br />

<strong>for</strong>mulated with WIV at ratios of 0.2:1 (TMC56-WIV(0.2:1))and 5:1(TMC56-WIV(5:1)), by<br />

varying the TMC concentration added to WIV.<br />

Dynamic light scattering and zeta-potential me<strong>as</strong>urements. WIV and TMC-WIV<br />

<strong>for</strong>mulations were prepared at a final WIV concentration of 62.5 µg/ml in 5 mM HEPES buffer<br />

pH 7.4. Particle size w<strong>as</strong> me<strong>as</strong>ured by dynamic light scattering (DLS) using a Malvern ALV CGS-<br />

3 (Malvern Instruments, Malvern, UK). DLS results are given <strong>as</strong> a z-average particle size<br />

diameter and a polydispersity index (PDI). The PDI can range from 0 (indicating monodisperse<br />

particles) to 1 (a completely heterodisperse system). Zeta-potential w<strong>as</strong> me<strong>as</strong>ured using a<br />

Zet<strong>as</strong>izer Nano (Malvern Instruments, Malvern, UK).<br />

Quantification of unbound TMCs by GPC. The fraction of the various TMCs that w<strong>as</strong> not<br />

bound to WIV w<strong>as</strong> quantified in the supernatant of centrifuged TMC-WIV <strong>for</strong>mulations, using<br />

the GPC method described earlier [11]. TMC-WIV <strong>for</strong>mulations were centrifuged <strong>for</strong> 40 min at<br />

22,000 x g at 4 °C and the supernatant w<strong>as</strong> collected. Prior to injection, 20 µl GPC running<br />

buffer w<strong>as</strong> added to 100 µl supernatant to adjust the pH of the sample to pH 4.4. The sample<br />

concentration w<strong>as</strong> determined using refractive index detection.<br />

Immunization protocol. Animal experiments were conducted according to the guidelines<br />

provided by the Dutch Animal Protection Act and were approved by a Committee <strong>for</strong> Animal<br />

Experimentation. For all experiments 6-8 weeks old female C57-BL/6 mice (Charles River)<br />

were used. Mice were housed in groups of 7-11 mice and food and water were provided ad<br />

74


Relationship between Structure and <strong>Adjuvant</strong>icity of TMC<br />

libitum. Prime and boost immunizations at day 0 and 21, respectively, were per<strong>for</strong>med without<br />

anesthesia. Groups of 11 mice were vaccinated i.n. with the various TMC-WIV <strong>for</strong>mulations at a<br />

dose of 12.5 µg WIV (corresponding to approximately 4.3 µg HA). All TMC-WIV vaccines were<br />

freshly prepared by mixing WIV dispersion with solutions of the various TMCs, <strong>as</strong> described<br />

above. Additionally, a group of 11 mice were vaccinated with WIV i.n. without TMC. As<br />

negative control groups, one group of 11 mice w<strong>as</strong> treated i.n. with 5 mM HEPES and another<br />

with TMC56 solution (1.25 mg/ml in 5 mM HEPES pH 7.4) without WIV. For i.n. immunization,<br />

mice were held in supine position without anesthesia and the <strong>for</strong>mulations were administered<br />

to the left and right nostril in a total volume of 10 µl. As a reference, one group of mice w<strong>as</strong><br />

vaccinated i.m. with WIV at a dose of 12.5 µg protein in a volume of 100 μl in the left and right<br />

quadriceps <strong>for</strong> prime and boost vaccination, respectively.<br />

Blood sampling and n<strong>as</strong>al w<strong>as</strong>hes. Blood samples were collected by orbital puncture in<br />

MINICOLLECT® serum separator tubes coated with SiO 2 (Greiner Bio-One, Alphen a/d Rijn,<br />

the Netherlands) 20 days after prime vaccination and 17 days after boost vaccination.<br />

Coagulated blood samples were centrifuged at 6,500 x g <strong>for</strong> 8 min at room temperature to<br />

obtain serum samples. Individual serum samples were stored at -20 °C until further analysis.<br />

Seventeen days after boost vaccination, 4 mice from each group were sacrificed by a lethal<br />

intraperitoneal injection of 100 µl sodium pentobarbital (200 mg/ml). The trachea w<strong>as</strong> then<br />

cannulated towards the n<strong>as</strong>opharyngeal duct with a PVC tube (inner/outer diameter 0.5/1.0<br />

mm). PBS (500 μl) containing complete Mini, EDTA free prote<strong>as</strong>e inhibitor (Roche Diagnostics,<br />

Indianapolis, IN, USA) at a concentration of 1 tablet / 7 ml PBS w<strong>as</strong> flushed through the n<strong>as</strong>al<br />

cavity and collected from the nostrils 3 times. The combined n<strong>as</strong>al w<strong>as</strong>hes were stored at -70<br />

°C until further analysis.<br />

Challenge. Twenty one days after the boost vaccination, mice were challenged with 50 ml (2<br />

x 10 8 x the 50% egg infectious dose (EID50)/ml) aerosolized, egg-grown A/PR/8/34 using a<br />

DeVilbiss Ultra-Neb 2000 ultr<strong>as</strong>onic nebulizer (Direct Medical Ltd, Lecarrow, Ireland) <strong>for</strong> 25 min.<br />

After challenge, mice were put back in their cages and any observed signs of illness like<br />

lethargy, standing fur and curved back were recorded. Additionally, their body weight w<strong>as</strong><br />

monitored daily <strong>for</strong> 15 days. For comparison of loss in body weight, the average area under the<br />

curve (AUC) w<strong>as</strong> calculated <strong>for</strong> each group from relative body weight curves of individual mice.<br />

All i.n. groups were compared to the negative control group (PBS i.n.) and the positive control<br />

75


Chapter 4<br />

group (WIV i.m.) by the average (AUC) of individual mice using a one-way ANOVA and<br />

Bonferroni’s correction <strong>for</strong> multiple comparisons.<br />

Hemagglutination inhibition test. First, 25 μl serum w<strong>as</strong> incubated <strong>for</strong> 18 h at 37 °C with<br />

75 μl Receptor Destroying Enzyme (RDE) solution (Denka Seiken UK Ltd, Coventry, UK) to<br />

suppress nonspecific hemagglutination inhibition. RDE w<strong>as</strong> then inactivated by incubating the<br />

mixture <strong>for</strong> 30 min at 56 °C. Next, 150 μl PBS w<strong>as</strong> added to obtain a final 10-fold serum<br />

dilution. Fifty μl diluted serum w<strong>as</strong> transferred in duplicate to V-bottom 96-wells plates<br />

(Greiner, Alphen a/d Rijn, The Netherlands) and serially diluted twofold in PBS. Next, 4<br />

hemagglutination units (HAU) of A/PR/8/34 (in 25 μl PBS) were added and the mixture w<strong>as</strong><br />

incubated <strong>for</strong> 40 min at room temperature. Finally, 50 μl 0.5% (v/v) chicken erythrocytes in<br />

PBS w<strong>as</strong> added. Plates were incubated <strong>for</strong> 1 h at room temperature. The HI titer is expressed <strong>as</strong><br />

the reciprocal value of the highest serum dilution capable of completely inhibiting the virusinduced<br />

agglutination of chicken erythrocytes. If no complete inhibition could be detected in<br />

the first lane, serum w<strong>as</strong> arbitrarily scored 5. Comparison between different experimental<br />

groups from the same dose w<strong>as</strong> made by a one-way ANOVA test and the Tukey post test on the<br />

log trans<strong>for</strong>med HI titers.<br />

Antibody <strong>as</strong>says. Antigen specific serum antibody responses were determined by a<br />

sandwich ELISA. Maxisorp ELISA plates (Nunc, Roskilde, Denmark) were coated overnight<br />

with polyclonal rabbit anti-A/PR/8/34 serum (dilution 1:1620). Plates were w<strong>as</strong>hed in<br />

between all prescribed steps with w<strong>as</strong>h buffer (0.64 M NaCl, 3 mM KCl, 0.15% polysorbate 20<br />

in 10 mM phosphate buffer pH 7.2) using a Skanw<strong>as</strong>her 300 (Molecular Devices, Sunnyvale,<br />

CA, USA). Next, plates were incubated <strong>for</strong> 1 h at 37 °C with blocking buffer (0.2% (w/w) c<strong>as</strong>ein,<br />

4% (w/w) sucrose, 0.05% (w/w) Triton X-100 and 0.01% (w/w) sodium azide in 30 mM TRIS<br />

pH 7.4), followed by incubation <strong>for</strong> 1 h at 37 °C with egg-grown, BPL-inactivated A/PR/8/34<br />

(25.6 HAU/ml). Plates were then incubated with twofold serially diluted sera (100 μl/well) <strong>for</strong><br />

1 h at 37 °C. Next, plates were incubated with 100 μl of a 1:2500 dilution of horseradish<br />

peroxid<strong>as</strong>e linked goat anti mouse -IgG (H+L), -IgG1, -IgG2a/c or -IgA(Fc) (Nordic<br />

Laboratories, Tilburg, the Netherlands) <strong>for</strong> 30 min, and w<strong>as</strong>hed twice. Finally, 100 μl 3,3’,5,5’-<br />

Tetramethylbenzidine (TMB) substrate solution w<strong>as</strong> added and plates were incubated <strong>for</strong> 15<br />

min at room temperature be<strong>for</strong>e enzymatic conversion w<strong>as</strong> stopped by adding 50 μl 2 M<br />

sulfuric acid. Optical density (OD) w<strong>as</strong> then me<strong>as</strong>ured at 450 nm using a Tecan Sunrise plate<br />

reader (Tecan Trading AG, Zurich Switzerland). Titers are given <strong>as</strong> the reciprocal sample<br />

76


Relationship between Structure and <strong>Adjuvant</strong>icity of TMC<br />

dilution corresponding to 20% of the maximal ELISA signal above background. Seronegative<br />

sera were arbitrarily scored with a titer of 15. Comparison between different experimental<br />

groups w<strong>as</strong> made using the log trans<strong>for</strong>med titers by a one-way ANOVA test and the Tukey<br />

post test. Boost IgG2a/c:IgG1 ratios were calculated and compared using the log trans<strong>for</strong>med<br />

data by a one-way ANOVA test and Bonferroni’s correction <strong>for</strong> multiple comparison.<br />

Results and discussion<br />

Structural properties of TMCs. The structural properties of the TMCs used in this study are<br />

summarized in Table 1. The DQ of the O-methyl free TMCs ranged between 30 and 68%,<br />

allowing us to selectively study the influence of trimethylation on adjuvant properties of TMC.<br />

With the TMC-OM group (DQ varying from 22 to 86%, O-methylation from 12 to 76% along<br />

with incre<strong>as</strong>ing DQ) the combined effect of charge density and O-methylation on adjuvanticity<br />

can be studied. The effect of O-methylation can be evaluated by comparing the O-methylated<br />

TMCs with O-methyl free TMCs. Finally, the re-acetylated TMC with a DAc of 54% and a DQ of<br />

44% can be used to <strong>as</strong>sess the role of N-acetylated units in the adjuvant properties of TMCs.<br />

Especially the comparison of TMC43, TMC-OM45 and TMC-RA44 will provide insight into the<br />

optimal structural properties of TMC <strong>for</strong> its use in n<strong>as</strong>al vaccine delivery. Importantly, it is<br />

unlikely that the minor differences in molecular weights between the various TMCs (Table 1)<br />

caused by variation in the degrees of substitutions and/or synthesis methods will affect the<br />

biological properties of the polymers [15].<br />

Table 1. Structural properties of synthesized TMCs. 1)<br />

M n<br />

M w<br />

Abbreviation<br />

(kDa) (kDa)<br />

DQ<br />

(%)<br />

DAc<br />

(%)<br />

DOM-6<br />

(%)<br />

DOM-3<br />

(%)<br />

TMC30 33 59 30 17 - -<br />

TMC43 36 75 43 17 - -<br />

TMC56 37 78 56 17 - -<br />

TMC68 39 84 68 17 - -<br />

TMC-OM22 34 56 22 12 18 12<br />

TMC-OM45 32 49 45 11 25 16<br />

TMC-OM61 31 49 61 10 56 44<br />

TMC-OM86 29 44 86 9 76 72<br />

TMC-RA44 43 83 44 54 - -<br />

1) Degree of acetylation (DAc), quaternization (DQ), O-methylation on C-6 (DOM-6) and on<br />

C-3 (DOM-3) of TMCs were determined by 1 H-NMR analysis. M n, M w were determined by GPC.<br />

77


Chapter 4<br />

Characterization of TMC-WIV <strong>for</strong>mulations. Various TMC-WIV <strong>for</strong>mulations were<br />

prepared by mixing the two components at a 1:1 (w/w) ratio. For comparison, TMC56-WIV<br />

<strong>for</strong>mulations were also prepared at 0.2:1 and 5:1 (w/w) ratios. Particle size and size<br />

distribution were determined and compared with those of plain WIV. Figure 1 shows that plain<br />

WIV had a diameter of approximately 170 nm with a polydispersity index (PDI) of 0.2,<br />

indicating a fairly homogeneous particle size distribution. Formulating the WIV particles with<br />

the various TMCs led to a small incre<strong>as</strong>e in particle size (200-220 nm) and comparable size<br />

distributions, independent of the type of TMC used or the TMC/WIV ratio.<br />

Figure 1. Diameter and size distribution (PDI) of TMC-WIV <strong>for</strong>mulations <strong>as</strong> determined by dynamic<br />

light scattering. Error bars represent standard deviations of three me<strong>as</strong>urements.<br />

Furthermore, the zeta-potential of the TMC-WIV particles w<strong>as</strong> analyzed in 5 mM HEPES pH<br />

7.4. The addition of TMC to negatively charged WIV (-13 mV) in a 1:1 (w/w) ratio resulted in<br />

positively charged particles (+12 to +20 mV) <strong>as</strong> seen in Figure 2, suggesting that all TMCs<br />

adsorb onto WIV in a similar f<strong>as</strong>hion. The zeta-potential incre<strong>as</strong>ed slightly with incre<strong>as</strong>ing DQ<br />

and the lowest surface charge w<strong>as</strong> observed with TMC-OM22-WIV. As expected, compared to<br />

the 1:1 ratio a lower TMC56:WIV ratio (0.2:1 (w/w)) resulted in a lower surface charge (+12<br />

mV vs. +18 mV), but a higher ratio (5:1 (w/w)) did not result in a higher zeta-potential.<br />

78


Relationship between Structure and <strong>Adjuvant</strong>icity of TMC<br />

Figure 2. Zeta-potential of TMC-WIV <strong>for</strong>mulations determined in 5 mM HEPES pH 7.4 Error bars<br />

indicate standard deviations of three samples.<br />

The amount of free TMC present in the TMC-WIV <strong>for</strong>mulations w<strong>as</strong> quantified using GPC. The<br />

concentration and relative amount of free TMC are depicted in Table 2. TMC-WIV <strong>for</strong>mulations<br />

at a ratio of 1:1 (w/w) had an average free TMC content between 0.95 and 1.15 mg/ml<br />

(corresponding to 76 and 91% of total TMC in the <strong>for</strong>mulation), independent of the DQ of the<br />

polymers. TMC-RA44 had the highest amount of TMC bound to the WIV particles. In the<br />

TMC56-WIV(0.2:1) mixture about 50% of the TMC56 remained free in solution, where<strong>as</strong> in the<br />

TMC56-WIV(5:1) almost 98% of the total TMC56 in the <strong>for</strong>mulation remained unbound. These<br />

results, together with the results of the zeta-potential me<strong>as</strong>urements, indicate that WIV is<br />

likely saturated with TMCs at TMC:WIV ratios 1:1 and 5:1 (w/w).<br />

79


Chapter 4<br />

Table 2. Specifications of TMC in various TMC-WIV <strong>for</strong>mulations.<br />

Total TMC in<br />

WIV<br />

TMC:WIV<br />

Formulation<br />

<strong>for</strong>mulation<br />

(µg)<br />

(w/w) ratio<br />

(µg)<br />

Unbound TMC<br />

(mg/ml) ab<br />

Unbound TMC<br />

(% of total) b<br />

TMC30-WIV 12.5 12.5 1:1 1.13± 0.04 90.0±3.3<br />

TMC43-WIV 12.5 12.5 1:1 1.14± 0.03 91.2±2.6<br />

TMC56-WIV 12.5 12.5 1:1 1.09± 0.01 87.1±0.8<br />

TMC68-WIV 12.5 12.5 1:1 1.08± 0.08 86.1±6.2<br />

TMC-OM22-WIV 12.5 12.5 1:1 1.02± 0.04 81.8±3.4<br />

TMC-OM45-WIV 12.5 12.5 1:1 1.05± 0.06 83.9±4.4<br />

TMC-OM61-WIV 12.5 12.5 1:1 1.01± 0.04 80.7±3.2<br />

TMC-OM86-WIV 12.5 12.5 1:1 1.06± 0.01 85.1±0.9<br />

TMC-RA44-WIV 12.5 12.5 1:1 0.95± 0.00 76.0±0.4<br />

TMC56-WIV (0.2:1) 12.5 2.5 0.2:1 0.10± 0.05 49.3±1.5<br />

TMC56-WIV (5:1) 12.5 62.5 5:1 6.11± 0.07 97.7±1.1<br />

WIV 12.5 - - - -<br />

a Amount of free TMC in the TMC-WIV <strong>for</strong>mulations determined by GPC.<br />

b Values are presented <strong>as</strong> the average of three samples ± standard deviation.<br />

Summarizing, mixing WIV with various TMCs in a ratio of 1:1 resulted in particles with<br />

similar size, size distribution, surface charge and amount of free TMC, thus allowing a fair<br />

evaluation of the influence of the structural characteristics of the TMCs on their adjuvant<br />

properties. Altering the TMC:WIV ratio mostly changed the amount of free TMC present in the<br />

<strong>for</strong>mulation and here the contribution of free TMC can be <strong>as</strong>sessed.<br />

Serum antigen specific total IgG after prime and boost vaccination. The TMC-WIV<br />

<strong>for</strong>mulations were compared to WIV alone in an intran<strong>as</strong>al vaccination study. Serum samples<br />

were analyzed <strong>for</strong> antigen-specific total IgG responses three weeks after prime and boost<br />

vaccination (Figure 3). IgG responses were induced by all <strong>for</strong>mulations containing WIV after<br />

prime vaccination and had incre<strong>as</strong>ed after boost vaccination. TMCs with a DAc ≤ 17% had a<br />

strong adjuvant effect that is not critically affected by DOM and DQ. All TMC-WIV <strong>for</strong>mulations,<br />

except TMC-RA44-WIV, showed a higher number of responding mice and elicited significantly<br />

higher IgG titers than WIV alone. Interestingly, TMC-RA44-WIV induced significantly lower<br />

total IgG responses than all other TMC-WIV <strong>for</strong>mulations with a TMC:WIV ratio of 1:1.<br />

Additionally, no significant differences were observed between the various TMC56-WIV ratio<br />

<strong>for</strong>mulations, indicating that free TMC does not strongly influence total IgG titers and that the<br />

lowest dose of TMC that w<strong>as</strong> tested already significantly incre<strong>as</strong>ed immune responses after i.n.<br />

vaccination with WIV.<br />

80


Relationship between Structure and <strong>Adjuvant</strong>icity of TMC<br />

Figure 3. Geometric mean antigen specific total IgG titers three weeks after prime and boost<br />

vaccination. Error bars indicate 95% confidence intervals (n=11). Non-responding mice were arbitrarily<br />

given a total IgG serum titer of 15. Below the x-axis the number of seropositive mice per group after<br />

boost vaccination are depicted. All titers after boost vaccination were compared using a one-way ANOVA<br />

test and Tukey’s post test. *** p


Chapter 4<br />

the IgG2a/c:IgG1 ratio per individual mouse illustrates that the Th1/Th2 balance of humoral<br />

immune responses shifts towards Th2 with incre<strong>as</strong>ing TMC:WIV ratio (Figure 4B).<br />

A<br />

B<br />

Figure 4. A) IgG1 and IgG2a/c titers elicited by WIV <strong>for</strong>mulations with varying TMC56:WIV ratios<br />

after prime and boost vaccination. Error bars indicate 95% confidence intervals (n=11). Indicated<br />

above the bars is the number of mice that developed detectable IgG1 or IgG2a/c titers (e.g. 1/11<br />

indicates one out of eleven mice). ** p


Relationship between Structure and <strong>Adjuvant</strong>icity of TMC<br />

HI titers. After boost vaccination, HI titers were hardly detectable in any of the i.n.<br />

vaccinated groups (data not shown). Only WIV i.m. induced substantial HI titers (average HI<br />

titer of 160). This is in line with previous results [11, 14]<br />

Antigen-specific secretory IgA (sIgA) in n<strong>as</strong>al w<strong>as</strong>hes. N<strong>as</strong>al w<strong>as</strong>hes were per<strong>for</strong>med 17<br />

days after boost vaccination to determine whether antigen-specific sIgA were elicited by any of<br />

the various <strong>for</strong>mulations. As shown in Figure 5 only in the TMC-WIV <strong>for</strong>mulations some mice<br />

developed detectable sIgA responses in the n<strong>as</strong>al cavity. Interestingly, although WIV i.m.<br />

immunization showed the highest serum IgG titers, no sIgA w<strong>as</strong> found in the n<strong>as</strong>al w<strong>as</strong>hes of<br />

any of these mice. Altogether, the determination of sIgA in the n<strong>as</strong>al mucus showed a relatively<br />

high variation within the <strong>for</strong>mulation-groups (<strong>as</strong> observed by others [10]) likely due to the<br />

collection and detection methods.<br />

OD 450<br />

1.4<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

TMC30-WIV<br />

TMC43-WIV<br />

TM56-WIV<br />

TMC68-WIV<br />

TMC-OM20-WIV<br />

TMC-OM45-WIV<br />

TMC-OM61-WIV<br />

TMC-OM86-WIV<br />

TMC-RA44-WIV<br />

TMC56-WIV (0.2:1)<br />

TMC56-WIV (5:1)<br />

WIV<br />

TMC56 sol<br />

WIV i.m.<br />

HEPES<br />

Figure 5. Antigen-specific secretory IgA levels in n<strong>as</strong>al w<strong>as</strong>hes of four individual mice per <strong>for</strong>mulation<br />

17 days after boost vaccination. Only in groups represented by black dots sIgA positive w<strong>as</strong>hes were<br />

found.<br />

Challenge with live, aerosolized virus. To <strong>as</strong>sess the protective effect of the immunization,<br />

seven mice per group were challenged with potentially lethal, homologous, egg-grown<br />

influenza virus and loss of body weight <strong>as</strong> a me<strong>as</strong>ure of illness w<strong>as</strong> monitored. All mice<br />

vaccinated i.n. with O-methyl free or O-methylated TMC-WIV <strong>for</strong>mulations were protected<br />

against the live virus. As representative examples the average body weight over 15 days after<br />

83


Chapter 4<br />

the challenge of the mice immunized with TMC43-WIV and TMC-OM45-WIV <strong>for</strong>mulations are<br />

shown in Figure 6A and B.<br />

A<br />

B<br />

C<br />

D<br />

Figure 6. Average relative body weight curves after challenge of mice (n=7) vaccinated with a) TMC43-<br />

WIV; b) TMC-OM45-WIV; c) TMC-RA44-WIV and d) WIV i.n.. For comparison each panel contains the<br />

curves of WIV i.m. and HEPES i.n.. Error bars indicate the 95% confidence intervals. *** p< 0.001 and **<br />

p< 0.01 indicate that average AUCs are significantly higher than the HEPES i.n. group. ### p< 0.001 and #<br />

p


Relationship between Structure and <strong>Adjuvant</strong>icity of TMC<br />

Additionally, there were no signs of illness observed in these groups and the AUC of the<br />

average body weight w<strong>as</strong> significantly higher (p


Chapter 4<br />

Furthermore, O-methylated TMCs had similar effects on both antibody titers and protection <strong>as</strong><br />

O-methyl free TMCs, indicating that the DOM does not affect the adjuvant properties of TMC.<br />

Interestingly, these findings do not correlate with the large influence of DQ and O-methylation<br />

in vitro on cell toxicity and TEER. This teaches us that DQ and DOM have a much more<br />

pronounced effect on in vitro cell toxicity and TEER <strong>as</strong>says than on mucosal adjuvant<br />

properties. It should be noted that the DOM and DQ may have an influence on the activity of<br />

TMC <strong>as</strong> a penetration enhancer in mucosal drug delivery, <strong>as</strong> suggested by reports on the<br />

influence of DQ [18-20]. As HA likely is too big to p<strong>as</strong>s even fully opened tight junctions, it is<br />

unlikely that opening of tight junctions will directly improve antigen uptake from the n<strong>as</strong>al<br />

cavity.<br />

The reacetylation of TMC on the other hand induced a strong decre<strong>as</strong>e in adjuvant effect<br />

illustrated by lower antibody titers and poor protection against challenge with live influenza<br />

virus. Since the mucoadhesion of particulate systems can be attributed to positive charge and<br />

hydrophobic effect [3], it is likely that TMC-RA44-WIV h<strong>as</strong> similar or even better<br />

mucoadhesive properties <strong>as</strong> the other TMC-WIV vaccines. Also, the zeta-potential of the<br />

particles w<strong>as</strong> similar to the other TMC-coated particles, implying that the introduction of a<br />

positive surface charge alone is not sufficient to improve adjuvanticity. There<strong>for</strong>e other factors<br />

must be responsible <strong>for</strong> the decre<strong>as</strong>ed adjuvant effect of reacetylated TMC.<br />

The first explanation <strong>for</strong> the decre<strong>as</strong>ed adjuvant effect of TMC-RA44 is a difference in<br />

interaction with cells, illustrated by a much lower in vitro toxicity and TEER effect than most<br />

other TMCs on Caco-2 cells [9]. TMC-OM22, however, hardly induced a TEER effect or cell<br />

toxicity either but showed to be a good mucosal adjuvant. This indicates that TEER and toxicity<br />

studies, <strong>as</strong> carried out <strong>for</strong> these TMCs [8, 9] cannot fully explain the loss of adjuvant effect by<br />

reacetylation.<br />

A second explanation <strong>for</strong> the poor adjuvant properties of TMC-RA44 is its enhanced<br />

enzymatic degradation by lysozyme compared to other TMCs. Previous research showed that<br />

the extent of lysozyme-catalyzed degradation of TMC is highly dependent on the DAc; TMC-<br />

RA44 showed a large decre<strong>as</strong>e in molecular weight while TMCs with DAc of ≤17% were only<br />

slightly degraded in presence of lysozyme [9]. Lysozyme is a strong antibacterial cationic<br />

protein that is excreted in high concentrations in the n<strong>as</strong>al cavity [21]. This may result in rapid<br />

degradation of the TMC-RA44 after i.n. administration, thereby strongly limiting its adjuvant<br />

effect on the WIV vaccination. Finally, chitin, an insoluble polysaccharide of N-<br />

acetylglucosamine units, which are also present in TMC-RA44, can be recognized by specific<br />

receptors of the innate immune system [22]. It h<strong>as</strong> recently been suggested that chitin induces<br />

86


Relationship between Structure and <strong>Adjuvant</strong>icity of TMC<br />

pro-inflammatory but also anti-inflammatory signals depending on its size [23]. Further<br />

studies should be done to investigate whether TMC-RA44 also induces anti-inflammatory<br />

signals through interaction with the innate immune system.<br />

In addition to changes in the type of TMC, the influence of free TMC w<strong>as</strong> studied by varying<br />

the TMC concentration in the <strong>for</strong>mulation. An excess of free TMC induced a shift towards<br />

higher IgG1 titers after boost vaccination. Several studies investigated the influence of TMC on<br />

the quality of immune responses [24-26] while other studies determined the effect of adjuvant<br />

dose on the type of responses elicited, using different adjuvants [27, 28]. However, since the<br />

type of adjuvant, animal model used, route of administration and animal model can affect the<br />

quality of immune responses, it is difficult to compare these studies to our data. This is the first<br />

time that the effect of adjuvant dose w<strong>as</strong> studied <strong>for</strong> TMC and additional studies should be<br />

per<strong>for</strong>med to provide mechanistical insight.<br />

Conclusion<br />

All TMC-WIV <strong>for</strong>mulations had comparable physicochemical properties, and there<strong>for</strong>e<br />

observed differences in immunogenicity are related to the various chemical structures of the<br />

TMCs. Formulating WIV with TMCs strongly enhances the immunogenicity and protection of<br />

i.n. vaccination with WIV. The adjuvant properties of TMCs <strong>as</strong> i.n. adjuvant are strongly<br />

decre<strong>as</strong>ed by reacetylation of TMC, where<strong>as</strong> the DQ and DOM did not significantly affect the<br />

adjuvant effect of TMC.<br />

Acknowledgement. This research w<strong>as</strong> partially per<strong>for</strong>med under the framework of <strong>TI</strong><br />

<strong>Pharma</strong> project number D5-106-1; Vaccine delivery: alternatives <strong>for</strong> conventional multiple<br />

injection vaccines. The authors acknowledge Frouke Kuijer <strong>for</strong> making the graphical abstract.<br />

87


Chapter 4<br />

References<br />

1. B. Slütter, N. Hagenaars, W. Jiskoot. Rational design of n<strong>as</strong>al vaccines. J Drug Target 16:1-17<br />

2008).<br />

2. V. Grabovac, D. Guggi, A. Bernkop-Schnürch. Comparison of the mucoadhesive properties of<br />

various polymers. Adv Drug Deliv Rev 57:1713-23 (2005).<br />

3. S.E. Harding. Mucoadhesive interactions. Biochem Soc Trans 31:1036-41 (2003).<br />

4. R.C. Read, S.C. Naylor, C.W. Potter, J. Bond, I. Jabbal-Gill, A. Fisher, L. Illum, R. Jennings.<br />

Effective n<strong>as</strong>al influenza vaccine delivery using chitosan. Vaccine 23:4367-74 (2005).<br />

5. A. Vila, A. Sanchez, K. Janes, I. Behrens, T. Kissel, J.L. Vila Jato, M.J. Alonso. Low molecular<br />

weight chitosan nanoparticles <strong>as</strong> new carriers <strong>for</strong> n<strong>as</strong>al vaccine delivery in mice. Eur J Pharm<br />

Biopharm 57:123-31 (2004).<br />

6. A.B. Sieval, M. Thanou, A.F. Kotze, J.E. Verhoef, J. Brussee, H.E. Junginger. Preparation and<br />

NMR characterization of highly substituted N-trimethyl chitosan chloride. Carbohydr<br />

Polym 36:157-65 (1998).<br />

7. A. Polnok, G. Borchard, J.C. Verhoef, N. Sarisuta, H.E. Junginger. Influence of methylation<br />

process on the degree of quaternization of N-trimethyl chitosan chloride. Eur J Pharm<br />

Biopharm 57:77-83 (2004).<br />

8. R.J. Verheul, M. Amidi, S. van der Wal, E. van Riet, W. Jiskoot, W.E. Hennink. Synthesis,<br />

characterization and in vitro biological properties of O-methyl free N,N,N-trimethylated<br />

chitosan. Biomaterials 29:3642-9 (2008).<br />

9. R.J. Verheul, M. Amidi, M.J. van Steenbergen, E. van Riet, W. Jiskoot, W.E. Hennink. Influence<br />

of the degree of acetylation on the enzymatic degradation and in vitro biological properties of<br />

trimethylated chitosans. Biomaterials 30:3129-35 (2009).<br />

10. W. Boonyo, H.E. Junginger, N. Waranuch, A. Polnok, T. Pitaksuteepong. Chitosan and<br />

trimethyl chitosan chloride (TMC) <strong>as</strong> adjuvants <strong>for</strong> inducing immune responses to ovalbumin in<br />

mice following n<strong>as</strong>al administration. J Control Rele<strong>as</strong>e 121:168-75 (2007).<br />

11. N. Hagenaars, E. M<strong>as</strong>trobattista, R.J. Verheul, I. Mooren, H.L. Glansbeek, J.G. Heldens, H. van<br />

den Bosch, W. Jiskoot. Physicochemical and Immunological Characterization of N,N,N-<br />

<strong>Trimethyl</strong> Chitosan-Coated Whole Inactivated Influenza Virus Vaccine <strong>for</strong> Intran<strong>as</strong>al<br />

Administration. Pharm Res 26:1353-64 (2009)<br />

12. T. Kiang, J. Wen, H.W. Lim, K.W. Leong. The effect of the degree of chitosan deacetylation on<br />

the efficiency of gene transfection. Biomaterials 25:5293-301 (2004).<br />

13. X. Jiang, A. van der Horst, M.J. van Steenbergen, N. Akeroyd, C.F. van Nostrum, P.J.<br />

Schoenmakers, W.E. Hennink. Molar-m<strong>as</strong>s characterization of cationic polymers <strong>for</strong> gene<br />

delivery by aqueous size-exclusion chromatography. Pharm Res 23:595-603 (2006).<br />

14. N. Hagenaars, E. M<strong>as</strong>trobattista, H. Glansbeek, J. Heldens, H. van den Bosch, V. Schijns, D.<br />

Betbeder, H. Vromans, W. Jiskoot. Head-to-head comparison of four nonadjuvanted<br />

inactivated cell culture-derived influenza vaccines: Effect of composition, spatial organization<br />

and immunization route on the immunogenicity in a murine challenge model. Vaccine 26<br />

:6555-63 (2008).<br />

15. S. Mao, X. Shuai, F. Unger, M. Wittmar, X. Xie, T. Kissel. Synthesis, characterization and<br />

cytotoxicity of poly(ethylene glycol)-graft-trimethyl chitosan block copolymers. Biomaterials 26<br />

:6343-56 (2005).<br />

16. D. Mei, S. Mao, W. Sun, Y. Wang, T. Kissel. Effect of chitosan structure properties and<br />

molecular weight on the intran<strong>as</strong>al absorption of tetramethylpyrazine phosphate in rats. Eur J<br />

Pharm Biopharm 70:874-881 (2008).<br />

17. M. Huang, E. Khor, L.Y. Lim. Uptake and cytotoxicity of chitosan molecules and nanoparticles:<br />

effects of molecular weight and degree of deacetylation. Pharm Res 21:344-53 (2004).<br />

18. M.M. Thanou, A.F. Kotze, T. Scharringhausen, H.L. Luessen, A.G. de Boer, J.C. Verhoef, H.E.<br />

Junginger. Effect of degree of quaternization of N-trimethyl chitosan chloride <strong>for</strong> enhanced<br />

88


Relationship between Structure and <strong>Adjuvant</strong>icity of TMC<br />

transport of hydrophilic compounds across intestinal caco-2 cell monolayers. J Control Rele<strong>as</strong>e<br />

64:15-25 (2000).<br />

19. J.H. Hamman, M. Stander, A.F. Kotze. Effect of the degree of quaternisation of N-trimethyl<br />

chitosan chloride on absorption enhancement: in vivo evaluation in rat n<strong>as</strong>al epithelia. Int J<br />

Pharm 232:235-42 (2002).<br />

20. G. Sandri, S. Rossi, M.C. Bonferoni, F. Ferrari, Y. Zambito, G. Di Colo, C. Caramella. Buccal<br />

penetration enhancement properties of N-trimethyl chitosan: Influence of quaternization degree<br />

on absorption of a high molecular weight molecule. Int J Pharm 297:146-55 (2005).<br />

21. A.M. Cole, H.I. Liao, O. Stuchlik, J. Tilan, J. Pohl, T. Ganz. Cationic polypeptides are required<br />

<strong>for</strong> antibacterial activity of human airway fluid. J Immunol 169:6985-91 (2002).<br />

22. C.G. Lee, C.A. Da Silva, J.Y. Lee, D. Hartl, J.A. Eli<strong>as</strong>. Chitin regulation of immune responses:<br />

an old molecule with new roles. Curr Opin Immunol 20:684-9 (2008).<br />

23. C.A. Da Silva, C. Chalouni, A. Williams, D. Hartl, C.G. Lee, J.A. Eli<strong>as</strong>. Chitin is a sizedependent<br />

regulator of macrophage TNF and IL-10 production. J Immunol 182:3573- 82 (2009).<br />

24. M. Amidi, H.C. Pellikaan, H. Hirschberg, A.H. de Boer, D.J. Crommelin, W.E. Hennink, G.<br />

Kersten, W. Jiskoot. Diphtheria toxoid-containing microparticulate powder <strong>for</strong>mulations <strong>for</strong><br />

pulmonary vaccination: preparation, characterization and evaluation in guinea pigs. Vaccine 25<br />

:6818-29 (2007).<br />

25. M. Amidi, S.G. Romeijn, J.C. Verhoef, H.E. Junginger, L. Bungener, A. Huckriede, D.J.<br />

Crommelin, W. Jiskoot. N-trimethyl chitosan (TMC) nanoparticles loaded with influenza<br />

subunit antigen <strong>for</strong> intran<strong>as</strong>al vaccination: biological properties and immunogenicity in a mouse<br />

model. Vaccine 25:144-53 (2007).<br />

26. B. Sayin, S. Somavarapu, X.W. Li, M. Thanou, D. Sesardic, H.O. Alpar, S. Senel. Mono-Ncarboxymethyl<br />

chitosan (MCC) and N-trimethyl chitosan (TMC) nanoparticles <strong>for</strong> non-inv<strong>as</strong>ive<br />

vaccine delivery. Int J Pharm 363:139-48 (2008).<br />

27. M.J. McCluskie, R.D. Weeratna, J.D. Clements, H.L. Davis. Mucosal immunization of mice<br />

using CpG DNA and/or mutants of the heat-labile enterotoxin of Escherichia coli <strong>as</strong> adjuvants.<br />

Vaccine 19:3759-68 (2001).<br />

28. K. Riedl, R. Riedl, A. von Gabain, E. Nagy, K. Lingnau. The novel adjuvant IC31 strongly<br />

improves influenza vaccine-specific cellular and humoral immune responses in young adult and<br />

aged mice. Vaccine 26:3461-8 (2008).<br />

89


CHAPTER 5A<br />

A STEP-BY-STEP APPROACH<br />

TO STUDY THE INFLUENCE OF N-<br />

ACETYLA<strong>TI</strong>ON ON THE ADJUVAN<strong>TI</strong>CITY OF<br />

N,N,N-TRIMETHYL CHITOSAN (TMC) IN AN<br />

INTRANASAL WHOLE INAC<strong>TI</strong>VATED<br />

INFLUENZA VIRUS VACCINE<br />

Rolf J. Verheul*, Niels Hagenaars*, Thom<strong>as</strong> van Es, Ethlinn van Gaal,<br />

P<strong>as</strong>cal H.J.L.F. de Jong, Sven Bruins, Ivo Que, Bram Slütter,<br />

Enrico M<strong>as</strong>trobattista, Harrie L. Glansbeek, Jacco G.M. Heldens,<br />

Han van den Bosch, Wim E. Hennink, Wim Jiskoot.<br />

*authors contributed equally<br />

Manuscript submitted


Chapter 5A<br />

Abstract<br />

In a previous study, we observed that reacetylation of N,N,N-trimethyl chitosan (TMC)<br />

reduced the adjuvant effect of TMC in mice after intran<strong>as</strong>al (i.n.) administration of whole<br />

inactivated influenza virus (WIV) vaccine. The aim of the present study w<strong>as</strong> to elucidate the<br />

re<strong>as</strong>on <strong>for</strong> the lack of adjuvanticity of reacetylated TMC (TMC-RA) by comparing TMC-RA<br />

(degree of acetylation 54%) with TMC (degree of acetylation 17%) at six potentially critical<br />

steps in the induction of an immune response after i.n. administration in mice: chemical<br />

stability of the polymer in a n<strong>as</strong>al w<strong>as</strong>h, local i.n. distribution of WIV, n<strong>as</strong>al residence time of<br />

WIV, cellular uptake of WIV by epithelial cells, transport of WIV by epithelial cells, and capacity<br />

of the <strong>for</strong>mulation to induce maturation of murine bone marrow derived dendritic cells (DCs).<br />

GPC analysis showed that TMC-RA w<strong>as</strong> degraded in a n<strong>as</strong>al w<strong>as</strong>h to a slightly larger extent<br />

than TMC. The local i.n. distribution and n<strong>as</strong>al clearance were similar <strong>for</strong> both TMC types.<br />

Fluorescently labeled WIV w<strong>as</strong> taken up more efficiently by Calu-3 cells when <strong>for</strong>mulated with<br />

TMC-RA compared to TMC and both TMCs significantly reduced transport of WIV over a Calu-3<br />

monolayer. Murine bone-marrow derived dendritic cell activation w<strong>as</strong> similar <strong>for</strong> plain WIV,<br />

and WIV <strong>for</strong>mulated with TMC-RA or TMC.<br />

In conclusion, the inferior adjuvant effect of TMC-RA over that of TMC might be caused by a<br />

slightly lower stability of TMC-RA-WIV in the n<strong>as</strong>al cavity, rather than by any of the other<br />

factors studied in this paper.<br />

92


Influence of N-Acetylation on the <strong>Adjuvant</strong>icity of TMC<br />

Introduction<br />

Intran<strong>as</strong>al (i.n.) immunization offers several advantages over traditional parenteral routes,<br />

such <strong>as</strong> painless, needle-free administration without the need <strong>for</strong> trained personnel and the<br />

potential induction of both systemic and local immune responses [1-3]. Nonetheless, its<br />

success h<strong>as</strong> been limited due to delivery issues <strong>as</strong>sociated with the physiology of the n<strong>as</strong>al<br />

epithelium and the tolerogenic nature of immune cells in the n<strong>as</strong>al <strong>as</strong>sociated lymphoid tissue.<br />

This is illustrated by the strong immune responses that the same vaccines elicit after<br />

intramuscular administration compared to n<strong>as</strong>al administration [4]. A successful i.n. vaccine<br />

<strong>for</strong>mulation should there<strong>for</strong>e adhere to the mucosal surfaces of the n<strong>as</strong>al cavity and provide<br />

protection against the degradation of the antigen in the n<strong>as</strong>al environment. Next, the<br />

<strong>for</strong>mulation should induce the uptake and/or transport of antigen through the epithelium, to<br />

ensure that a sufficient amount of antigen will reach the antigen presenting cells in the<br />

subepithelial space. Finally, these APCs (most noticeably dendritic cells (DCs)), have to be<br />

activated (matured) either by the antigen or by addition of an adjuvant, in order <strong>for</strong> them to<br />

migrate to the nearby lymph nodes and elicit the desired type of immune response [5-8].<br />

N,N,N-trimethyl chitosan (TMC), a quaternized, water-soluble derivative of chitosan, h<strong>as</strong><br />

previously been shown to possess mucoadhesive [9] and absorption enhancing [10-12]<br />

characteristics and when <strong>for</strong>mulated with whole inactivated influenza virus vaccine (WIV)<br />

greatly enhanced the efficacy of the i.n. administered vaccine in mice [13]. To further improve<br />

its adjuvant effect, the structural properties of TMC, like the degree of quaternization (DQ), the<br />

degree of O-methylation (DOM), and the degree of N-acetylation (DAc), can be varied during<br />

synthesis [14, 15]. In a previous study on the influence of these structural properties on the<br />

adjuvant effect of TMC in i.n. WIV vaccines in mice [16], we showed that both the DQ and the<br />

DOM of TMC did not have a significant effect on the immunogenicity of the i.n. WIV vaccines,<br />

where<strong>as</strong> the DAc did. A striking loss in adjuvanticity of TMC w<strong>as</strong> observed when WIV w<strong>as</strong><br />

<strong>for</strong>mulated with re-acetylated TMC (TMC-RA) with a DAc of 54% and a DQ of 44% (compared<br />

to a DAc of 17% <strong>for</strong> other tested TMCs). Importantly, the physical characteristics of WIV<br />

<strong>for</strong>mulated with TMC-RA or with TMC (both polymers having a DQ of about 45%) were similar<br />

<strong>for</strong> size, zeta-potential and amount of unbound TMC, indicating that these physical properties<br />

can not be held accountable <strong>for</strong> the difference in adjuvant effect. In contr<strong>as</strong>t to TMC, TMC-RA is<br />

a substrate <strong>for</strong> lysozyme, h<strong>as</strong> a lower toxicity profile than TMC and h<strong>as</strong> no effect on the<br />

transepithelial electrical resistance (TEER) of Caco-2 cells [15].<br />

The aim of this study, w<strong>as</strong> to understand why TMC-RA, in contr<strong>as</strong>t to TMC, does not act <strong>as</strong> an<br />

adjuvant <strong>for</strong> an i.n. WIV vaccine in mice. This will provide more insight into the mechanism of<br />

93


Chapter 5A<br />

adjuvanticity of TMC and may aid in the rational, structural design of TMC <strong>as</strong> an adjuvant. TMC-<br />

WIV, TMC-RA-WIV and WIV <strong>for</strong>mulations were compared at six potentially critical steps in the<br />

induction of an immune response after i.n. administration. In particular, we studied (i) the<br />

stability of TMC and TMC-RA in n<strong>as</strong>al w<strong>as</strong>hes, (ii) the effect of both polymers on the n<strong>as</strong>al<br />

residence time of WIV, (iii) n<strong>as</strong>al distribution patterns, (iv) cellular uptake and (v) transport of<br />

WIV through an epithelial (Calu-3) cell line and (vi) the effect of the different <strong>for</strong>mulations on<br />

maturation of murine bone-marrow derived dendritic cells (DCs).<br />

Materials and Methods<br />

Materials. Chitosan w<strong>as</strong> purch<strong>as</strong>ed from Primex (Siglufjordur, Iceland) and had a DAc of 17%<br />

and a number average molecular weight (M n) and weight average molecular weight (M w) of 28<br />

and 43 kDa, respectively. Acetic anhydride, sodium borohydrate, <strong>for</strong>mic acid, <strong>for</strong>maldehyde<br />

37% (stabilized with methanol), deuterium oxide, sodium acetate, acetic acid (anhydrous),<br />

sodium hydroxide and hydrochloric acid were obtained from Sigma-Aldrich Chemical Co.<br />

Iodomethane 99% stabilized with copper w<strong>as</strong> obtained from Acros Organics (Geel, Belgium).<br />

Purified, cell culture-grown (Madin-Darby Canine Kidney (MDCK) cells), β-propiolacton (BPL)-<br />

inactivated A/PR/8/34, <strong>as</strong> well <strong>as</strong> polyclonal rabbit anti-A/PR/8/34 serum were from Nobilon<br />

International BV (Boxmeer, The Netherlands). Alexa Fluor® 488-labeled goat anti-rabbit IgG,<br />

Alexa Fluor® 488 labeling kit and 4,6-diamidino-2-phenylindole dihydrochloride (DAPI) were<br />

purch<strong>as</strong>ed from Invitrogen (Breda, the Netherlands). Osteosoft decalcifier (10% EDTA) w<strong>as</strong><br />

from Merck Serono (Schiphol, the Netherlands). Citric acid, Triton-X, trypsin from bovine<br />

pancre<strong>as</strong>, normal goat serum and normal rabbit serum were obtained from Sigma<br />

(Zwijndrecht, the Netherlands). Phosphate buffered saline (10 mM phosphate buffer, pH 7.4,<br />

150 mM NaCl; PBS) w<strong>as</strong> purch<strong>as</strong>ed from Braun (Melsungen, Germany). IRDye800CW®-labeled<br />

epidermal growth factor (EGF) and the IRDye800CW® protein labeling kit were from LI-COR<br />

Biosciences (Lincoln, NE, USA). Fixative (4% <strong>for</strong>malin in phosphate buffer pH 7.0) w<strong>as</strong><br />

obtained from Klinipath (Duiven, the Netherlands). Calu-3 cells were obtained from ATCC<br />

(Man<strong>as</strong>s<strong>as</strong>, VA, USA). Virally trans<strong>for</strong>med human bronchial epithelial cell line 16HBE14o- w<strong>as</strong><br />

a kind gift from Dr. D. Gruenert (University of Cali<strong>for</strong>nia at San Francisco). Trypsin/EDTA 10x,<br />

Plain DMEM (Dulbecco’s modified Eagle’s medium, with 3.7 g/l sodium bicarbonate, 1 g/l L-<br />

glucose, L-glutamine) and antibiotics/antimycotics (penicillin, streptomycin sulfate,<br />

amphotericin B) were from PAA Laboratories GmbH (P<strong>as</strong>ching, Austria) and propidium iodide<br />

(PI) and MEM with 0.292g/L L-glutamine, 1 g/L Glucose, 2.2 g/L NaHCO3, w<strong>as</strong> from Invitrogen<br />

94


Influence of N-Acetylation on the <strong>Adjuvant</strong>icity of TMC<br />

(Breda, The Netherlands). Matrigel ® w<strong>as</strong> acquired from Becton Dickinson (Franklin Lakes, NJ<br />

USA). All other chemicals used were of analytical grade. Animal experiments were conducted<br />

according to the guidelines provided by the Dutch Animal Protection Act and were approved<br />

by a Committee <strong>for</strong> Animal Experimentation.<br />

Synthesis and characterization of trimethylated chitosans. O-methyl free TMCs with a<br />

DQ of 45% and varying DAc were synthesized from chitosan <strong>as</strong> described previously [15]. The<br />

DQ and DAc of the TMCs were determined with 1H-NMR on a Varian INOVA 500 MHz NMR<br />

spectrometer (Varian Inc., Palo Alto, Ca, USA) at 80 °C in D 2O. Furthermore, number average<br />

weight (M n) and weight average weight (M w) of the TMCs were determined <strong>as</strong> described be<strong>for</strong>e<br />

[15] by GPC on a Viscotek system detecting refractive index, viscosity and light scattering. A<br />

Shodex OHPak SB-806 column (30 cm) w<strong>as</strong> used with 0.3 M sodium acetate pH 4.4 (adjusted<br />

with acetic acid) <strong>as</strong> running buffer. The structural characteristics of the synthesized TMCs are<br />

summarized in Table 1.<br />

Preparation and characterization of TMC-WIV <strong>for</strong>mulations. TMC-WIV <strong>for</strong>mulations<br />

were prepared <strong>as</strong> described previously (16). Briefly, purified, cell culture-grown (Madin-Darby<br />

Canine Kidney (MDCK) cells), β-propiolacton (BPL)-inactivated mouse adapted influenza<br />

A/Puerto Rico/8/34 virus (A/PR/8/34) suspended in a 10 mM phosphate buffered saline (150<br />

mM NaCl, pH 7.4) w<strong>as</strong> concentrated by centrifugation at 22,000xg <strong>for</strong> 30 min at 4 °C and<br />

resuspended in 5 mM HEPES buffer (pH 7.4). WIV suspension w<strong>as</strong> mixed with equal volumes<br />

of TMC or TMC-RA in 1:1 or 5:1 weight/weight ratio. For dynamic light scattering (DLS)<br />

(Malvern ALV CGS-3, Malvern Instruments, Malvern, UK) and zeta-potential me<strong>as</strong>urements<br />

(Zet<strong>as</strong>izer Nano, Malvern Instruments, Malvern, UK) a final WIV concentration of 62.5 μg/ml<br />

(expressed <strong>as</strong> total protein concentration) in 5 mM HEPES pH 7.4 w<strong>as</strong> used. DLS results are<br />

given <strong>as</strong> z-average particle size diameter and a polydispersity index (PDI). The PDI can vary<br />

from 0 (indicating monodisperse particles) to 1 (a completely heterodisperse particle size<br />

distribution)).<br />

Degradation of TMCs in n<strong>as</strong>al w<strong>as</strong>h. Six female 6-8 weeks old Balb/c nu/nu mice (Charles<br />

River, L’Arbresle, France) were sacrificed by a lethal intraperitoneal injection of 100 μl sodium<br />

pentobarbital (200 mg/ml). The trachea of each mouse w<strong>as</strong> then cannulated towards the<br />

n<strong>as</strong>opharyngeal duct with a PVC tube (inner/outer diameter 0.5/1.0 mm). PBS (600 μl) w<strong>as</strong><br />

flushed through the n<strong>as</strong>al cavity and collected from the nostrils 3 times; the collected samples<br />

95


Chapter 5A<br />

were pooled. TMC and TMC-RA were dissolved in 5 mM HEPES at 2.5 mg/ml, mixed 1:1 (v/v)<br />

with n<strong>as</strong>al w<strong>as</strong>h and the mixtures were incubated at 37 °C. Samples were taken after 4 and 24<br />

hours and 5 and 9 days. M n and M w were determined by GPC on a Viscotek system detecting<br />

refractive index, viscosity and light scattering. A Shodex OHPak SB806 column (15 cm) w<strong>as</strong><br />

used with 0.3 M sodium acetate, pH 4.4 (adjusted with acetic acid), <strong>as</strong> running buffer [17].<br />

Pullulan (M n 102 kDa, M n 106 kDa) obtained from Viscotek Benelux (Oss, The Netherlands)<br />

w<strong>as</strong> used <strong>for</strong> calibration and polymer solutions mixed with PBS 1:1 (v/v) were used <strong>as</strong><br />

controls. Prior to injection, 30 μl GPC running buffer w<strong>as</strong> added to 120 μl sample to adjust the<br />

pH of the sample to pH 4.4.<br />

In vivo imaging of n<strong>as</strong>al residence time of TMC-WIV <strong>for</strong>mulations. In vivo imaging of<br />

TMC-WIV <strong>for</strong>mulations w<strong>as</strong> done according to a previous study [18]. Female nude (Balb/c<br />

nu/nu) mice were obtained from Charles River (L’Arbresle, France) and received plain<br />

IRDye800CW®-labeled WIV (n=9) or <strong>for</strong>mulated with TMC (n=9) or TMC-RA (n=3) in a ratio<br />

of 1:1 (w/w) under light anesthesia with isoflurane inhalation. Next, mice were scanned with<br />

an IVIS Spectrum imaging system from Caliper Life Sciences (Hopkinton, MA, USA). Scans were<br />

per<strong>for</strong>med regularly over at le<strong>as</strong>t 2 h. Scanned images were analyzed using Living Image 3.1<br />

software from Caliper Life Sciences (Hopkinton, MA, USA). The threshold w<strong>as</strong> set using the<br />

background scan made from each mouse be<strong>for</strong>e administration of the IRDye800CW®-labeled<br />

<strong>for</strong>mulations. The excitation wavelength w<strong>as</strong> set at 710 nm and emitted light w<strong>as</strong> me<strong>as</strong>ured at<br />

760; 780; 800; 820 and 840 nm. Spectral unmixing w<strong>as</strong> per<strong>for</strong>med to decompose the emitted<br />

light into auto fluorescence and label-specific fluorescence. To compare the fluorescence in<br />

different mice and different groups, the absolute fluorescence w<strong>as</strong> converted to relative<br />

fluorescence (% of the maximal fluorescence in the n<strong>as</strong>al cavity). The are<strong>as</strong> under the curve<br />

(AUC) of the relative fluorescence of individual mice were used to compare the fluorescence<br />

over time <strong>for</strong> the different groups.<br />

Immunostaining of TMC-WIV <strong>for</strong>mulations in n<strong>as</strong>al cross-sections. Immunostaining of<br />

TMC-WIV <strong>for</strong>mulations w<strong>as</strong> carried out essentially <strong>as</strong> described be<strong>for</strong>e [18]. Six to eight weeks<br />

old female C57-BL/6 mice from Charles River (L’Arbresle, France) were i.n. vaccinated with<br />

TMC-RA-WIV or TMC-WIV in a (w/w) ratio of 1:1 or plain WIV <strong>for</strong>mulations. As negative<br />

controls, mice received PBS or solutions of TMCs and 4 mice were left untreated. After 20<br />

minutes or 1 hour, animals were sacrificed by cervical dislocation and the n<strong>as</strong>al cavity w<strong>as</strong><br />

isolated by removing the brains, lower jaw, skin and muscle tissue. The n<strong>as</strong>al cavities were<br />

96


Influence of N-Acetylation on the <strong>Adjuvant</strong>icity of TMC<br />

then fixed in 10% <strong>for</strong>malin, decalcified, embedded in paraffin and 3.5 µm thick cross-sections<br />

were made at different depths of the n<strong>as</strong>al cavity using a Microm HM 355 S microtome from<br />

Thermo Fisher Scientific (Walldorf, Germany). Cross-sections <strong>for</strong> immunohistochemistry were<br />

mounted on superfrost plus gl<strong>as</strong>s slides from Menzel-Gläser (Braunschweig, Germany). After<br />

deparaffinization, hydration and a heat-induced epitope retrieval (HIER) step with citrate<br />

buffer, cross-sections were w<strong>as</strong>hed 3 x 5 min with 0.2% Triton-X in PBS. Subsequently, crosssections<br />

were incubated first with normal goat serum <strong>for</strong> 15 minutes and then with a 1:500<br />

dilution of polyclonal WIV-specific rabbit antiserum overnight at 4°C. As a negative control,<br />

one of the two paraffin sections w<strong>as</strong> incubated with nonspecific rabbit serum <strong>as</strong> primary<br />

antibody. After w<strong>as</strong>hing the slides 3x <strong>for</strong> 5 min with 0.2% Triton-X in PBS, a 1:200 dilution of<br />

Alexa Fluor® 488-labeled goat anti-rabbit IgG w<strong>as</strong> applied. The samples were incubated <strong>for</strong> 45<br />

minutes at room temperature. Next, the slides were w<strong>as</strong>hed 3x with PBS and stained with a<br />

1:25000 dilution of DAPI <strong>for</strong> 6 min at room temperature and w<strong>as</strong>hed 3x with PBS again.<br />

Finally, the slides were mounted with Fluorosave from Calbiochem (San Diego, CA, USA). All<br />

slides were examined using a fluorescence microscope (Nikon Eclipse TE-2000 Nikon,<br />

Amstelveen, the Netherlands) equipped with a Digital Sight DS-2Mv camera. Slides were<br />

blindly scored <strong>for</strong> the presence, location, pattern and intensity of antigen staining and pictures<br />

were taken with fixed camera settings <strong>for</strong> a fair comparison.<br />

Uptake by Calu-3 epithelial cells of fluorescently labeled TMC-WIV <strong>for</strong>mulations. Calu-<br />

3 were grown in MEM with 0.292g/L L Glutamine, 1g /L Glucose, 2.2g/L NaHCO3 (Invitrogen,<br />

Breda, The Netherlands) supplemented with antibiotics/antimycotics and 10% heatinactivated<br />

fetal calf serum (FCS; Integro, Zaandam, The Netherlands). Cells were cultured in<br />

fibronectin-coated fl<strong>as</strong>ks and maintained at 37°C in a 5% CO2 humidified air atmosphere and<br />

split once a week.<br />

Cells were transferred into a flat bottom 24-well plate (500,000 cells/well) and after<br />

maturation <strong>for</strong> 2 days cells were incubated with 500 μl of plain AlexaFluor-488®-labeled WIV<br />

or <strong>for</strong>mulated with TMC or TMC-RA in PBS at a TMC:WIV (w/w) ratio of 5:1 and diluted <strong>as</strong><br />

indicated (50µg WIV/ml diluted 10-160x) <strong>for</strong> 1 h at 37 ºC or 4 ºC. After incubation, cells were<br />

w<strong>as</strong>hed with PBS (400 μl) and detached with 100 μl trypsin/EDTA at 37 ºC. Then, 150 μl MEM<br />

+ 10% FCS w<strong>as</strong> added and the cells were transferred into a 96 well plate. Cells were<br />

centrifuged (250 x g <strong>for</strong> 5 minutes at 4 ºC) and w<strong>as</strong>hed with three times with 200 μl phosphate<br />

buffered albumin (PBA, 1 g albumin per 100 ml PBS) and finally resuspended in 200 μl PBA.<br />

Immediately prior to me<strong>as</strong>urement, 20 μl propidium iodide solution (PI; 1 μg/ml in water) w<strong>as</strong><br />

97


Chapter 5A<br />

added <strong>for</strong> live/dead cell discrimination. Flow cytometric analysis w<strong>as</strong> per<strong>for</strong>med on a FACS<br />

Canto (Becton and Dickinson, Mountain View, CA, USA) using a 15 mW 488 nm, air-cooled<br />

argon-ion l<strong>as</strong>er and data were analyzed using FACS Diva software (Becton and Dickinson,<br />

Mountain View, CA, USA). 10,000 cells were recorded per sample to determine bead-uptake<br />

(FL1-channel) and PI-staining (FL3-channel).<br />

Transport of fluorescently labeled TMC-WIV <strong>for</strong>mulations over Calu-3 monolayer.<br />

Calu-3 cells were seeded at a density of 5x10 5 cells per well on 12-transwell plates with<br />

Matrigel ® coated 2 µm microporous membranes. The cells were cultured in Dulbecco’s<br />

Modified Eagle’s Medium (DMEM) <strong>for</strong> 14 days until a confluent cell layer w<strong>as</strong> <strong>for</strong>med. The<br />

medium w<strong>as</strong> replaced by Hank’s Balanced Salt Solution (HBSS) at the apical (300 μl) and<br />

b<strong>as</strong>olateral (1.2 ml) sides and after 30 minutes equilibration the transepithelial electrical<br />

resistance (TEER) w<strong>as</strong> me<strong>as</strong>ured using a home made dipstick electrode. Then, HBSS w<strong>as</strong><br />

replaced by 300 μl of plain AlexaFluor-488®-labeled WIV or <strong>for</strong>mulated with TMC or TMC-RA<br />

at a TMC:WIV (w/w) ratio of 5:1 (25µg WIV/ml in HBSS). After incubation <strong>for</strong> 60 minutes at 37<br />

ºC the TEER w<strong>as</strong> me<strong>as</strong>ured again and the amount of fluorescently labeled WIV particles in the<br />

b<strong>as</strong>olateral acceptor compartment w<strong>as</strong> determined by flow cytometry.<br />

Maturation of murine bone marrow derived dendritic cells. The femur and tibia of C57-<br />

BL/6 mice were removed, both ends were cut and bone marrow w<strong>as</strong> flushed with Iscove’s<br />

Modified Dulbecco’s Medium (IMDM; Gibco, CA, USA) using a syringe with a 0.45 mm diameter<br />

needle. The bone marrow suspension w<strong>as</strong> vigorously resuspended and p<strong>as</strong>sed over a 100 μm<br />

gauge to obtain a single cell suspension. After w<strong>as</strong>hing, cells were seeded 2x10 6 cells per 100<br />

mm petridish (Greiner Bio-One, Alphen aan den Rijn, The Netherlands) in 10 ml IMDM,<br />

supplemented with 2 mM L-glutamine, 50 U/ml penicillin, 50 ug/ml streptomycin and 50 μM<br />

β-mercaptoethanol (Merck, Darmstadt, Germany) and 30 ng/ml recombinant murine GM-CSF<br />

(rmGM-CSF). At day 2, 10 ml medium containing 30 ng/ml rmGM-CSF w<strong>as</strong> added. At day 5<br />

another 30 ng/ml rmGM-CSF w<strong>as</strong> added to each plate. From day 6 onwards, the non-adherent<br />

DCs were harvested and used <strong>for</strong> subsequent experiments.<br />

These immature DCs were seeded (50,000 cells/well) in a 96-well plate in a total volume of<br />

100 μl RPMI 10% FCS in the presence of TMC (0.001-25 μg/ml), TMC-WIV <strong>for</strong>mulations at a<br />

1:1 or 5:1 (w/w) ratios or lipopolysaccharide (LPS) (0.01-100 ng/ml) <strong>as</strong> a positive control. DCs<br />

were incubated with the samples <strong>for</strong> 16 h at 37 °C. DC maturation w<strong>as</strong> determined by<br />

98


Influence of N-Acetylation on the <strong>Adjuvant</strong>icity of TMC<br />

analyzing cell-surface expression of co-stimulatory molecules (CD40) on CD11 and MHC II<br />

positive cells by flow cytometry [19].<br />

Results<br />

Properties of TMC-WIV <strong>for</strong>mulations. The structural characteristics of the synthesized<br />

TMCs are summarized in Table 1.<br />

Table 1. Structural characteristics of TMC and TMC-RA used in this study.<br />

Polymer M n (kDa) M w (kDa) DQ (%) DAc (%)<br />

TMC 36 75 43 17<br />

TMC-RA 43 83 44 54<br />

As in a previous study [16], WIV w<strong>as</strong> mixed in a 1:1 or 5:1 w/w ratio with TMC or TMC-RA<br />

and analyzed <strong>for</strong> size, PDI and zeta potential. Figure 1 shows that mixing of WIV with TMC led<br />

to a minor incre<strong>as</strong>e in size and PDI, and the zeta potential reversed from -17 mV (plain WIV) to<br />

about +18 mV, independent of type or amount of TMC used <strong>for</strong> coating. It can there<strong>for</strong>e be<br />

stated that all <strong>for</strong>mulations were physically similar, except <strong>for</strong> the amount of free polymer;<br />

<strong>for</strong>mulations with a 5:1 w/w polymer to WIV ratio will very likely contain more free polymer<br />

[16].<br />

Diameter (nm)<br />

A<br />

400<br />

300<br />

200<br />

100<br />

PDI<br />

Diameter<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

PDI<br />

Zeta potential (mV)<br />

B<br />

25<br />

15<br />

5<br />

-5<br />

-15<br />

0<br />

WIV<br />

TMC-WIV 1:1<br />

TMC-WIV 5:1<br />

TMC-RA-WIV 1:1<br />

TMC-RA-WIV 5:1<br />

0.0<br />

-25<br />

WIV<br />

TMC-WIV 1:1<br />

TMC-WIV 5:1<br />

TMC-RA-WIV 1:1<br />

TMC-RA-WIV 5:1<br />

Figure 1. Z-average diameter and PDI (A) and zeta potential (B) of the WIV <strong>for</strong>mulations. Error bars<br />

represent the standard deviation (n=5).<br />

99


Chapter 5A<br />

Biodegradation. An important difference between TMC and TMC-RA is their enzymatic<br />

biodegradability by lysozyme [15], an enzyme which is present in the n<strong>as</strong>al cavity [20].<br />

There<strong>for</strong>e, after i.n. immunization TMC-RA-WIV may be more rapidly degraded in the n<strong>as</strong>al<br />

cavity. To study whether TMC and TMC-RA are also degraded upon contact with enzymes<br />

present in the n<strong>as</strong>al cavity, solutions of these TMCs were incubated at 37 °C with murine n<strong>as</strong>al<br />

w<strong>as</strong>hes. At different time points, the solutions were analyzed by GPC to determine the<br />

molecular weight (M n and M w) of the TMCs. As expected, TMCs in PBS that were taken <strong>as</strong><br />

controls, did not show decre<strong>as</strong>e in M n/M w. As shown in Figure 2, TMC-RA degraded to a greater<br />

extent than TMC in a n<strong>as</strong>al w<strong>as</strong>h. However, the degradation of both TMC-RA and TMC w<strong>as</strong><br />

relatively slow in the n<strong>as</strong>al w<strong>as</strong>h (compared to the in vitro findings) and up to 24 h of<br />

incubation, no significant differences were observed between the two TMCs.<br />

Figure 2. Biodegradation of TMC (squares) and TMC-RA (dots) in pooled n<strong>as</strong>al w<strong>as</strong>h of mice. Error bars<br />

indicate the standard deviation (n=3).<br />

In vivo fluorescence imaging. In vivo fluorescence imaging w<strong>as</strong> used to compare the effect<br />

of TMC and TMC-RA on the n<strong>as</strong>al clearance of WIV. Fluorescently labeled WIV w<strong>as</strong> <strong>for</strong>mulated<br />

with TMC and TMC-RA and the <strong>for</strong>mulations were administered i.n. to mice. Imaging of the<br />

fluorescence in the n<strong>as</strong>al cavity <strong>as</strong> a function of time did not reveal a significant difference<br />

between the <strong>for</strong>mulations (p>0.05), <strong>as</strong> calculated from the AUC (see Figure 3). After an initial<br />

incre<strong>as</strong>e in fluorescence, likely due to fluorescence dequenching [18], all <strong>for</strong>mulations showed<br />

a comparable decre<strong>as</strong>e in fluorescence over time. This suggests that the n<strong>as</strong>al clearance of<br />

plain WIV, TMC-WIV and TMC-RA-WIV is comparable.<br />

100


Influence of N-Acetylation on the <strong>Adjuvant</strong>icity of TMC<br />

Relative fluorescence<br />

(%)<br />

110<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

0 25 50 75 100 125<br />

Time (min)<br />

WIV<br />

TMC-WIV<br />

TMC-RA-WIV<br />

Figure 3. Average relative fluorescence in the n<strong>as</strong>al cavity over time after intran<strong>as</strong>al administration of<br />

fluorescently labeled WIV <strong>for</strong>mulations. Error bars indicate standard deviation (n=3 mice).<br />

Local distribution in the n<strong>as</strong>al cavity of intran<strong>as</strong>ally applied vaccine <strong>for</strong>mulations. In a<br />

previous study it w<strong>as</strong> shown that in the murine n<strong>as</strong>al cavity, TMC-WIV <strong>for</strong>mulations exhibit a<br />

completely different distribution pattern <strong>as</strong> plain WIV: TMC coated WIV w<strong>as</strong> located tightly<br />

along the epithelial cells while plain WIV w<strong>as</strong> mainly observed in mucosal blobs [18]. To<br />

investigate whether the distribution pattern is different <strong>for</strong> TMC-RA-WIV, mice were<br />

administered TMC-RA-WIV and TMC-WIV and n<strong>as</strong>al sections were stained to visualize the<br />

location of WIV. After 20 min, WIV-specific staining w<strong>as</strong> found on the epithelial surfaces of the<br />

n<strong>as</strong>o- and maxilloturbinates in all mice, <strong>for</strong> both TMC-WIV and TMC-RA-WIV (see Figure 4 <strong>for</strong><br />

representative examples, n=5). This indicates that the degree of acetylation is unlikely to affect<br />

the contact between WIV with the mucosal surfaces. After 1 hour, WIV w<strong>as</strong> still present on the<br />

n<strong>as</strong>o- and maxilloturbinates but the staining w<strong>as</strong> less intense (results not shown). TMC<br />

solutions showed no fluorescence upon excitation nor did staining with non-specific rabbit<br />

serum [18].<br />

101


Chapter 5A<br />

A<br />

B<br />

Figure 4. Representative picture of fluorescently labeled WIV (depicted in green) on the mucosal<br />

surfaces of the n<strong>as</strong>oturbinates when <strong>for</strong>mulated with TMC-RA (A) and TMC (B) 20 minutes after<br />

administration.<br />

102


Influence of N-Acetylation on the <strong>Adjuvant</strong>icity of TMC<br />

A<br />

Fluorescence Geo Mean<br />

(arb. units)<br />

2000<br />

1500<br />

1000<br />

500<br />

0<br />

5.00<br />

*<br />

2.50<br />

*<br />

1.25<br />

*<br />

0.62<br />

WIV<br />

TMC-WIV<br />

TMC-RA-WIV<br />

PBS<br />

** **<br />

0.31<br />

B<br />

Fluorescence Geo Mean<br />

(arb. units)<br />

1250<br />

1000<br />

750<br />

500<br />

250<br />

0<br />

**<br />

**<br />

5.00<br />

** ***<br />

2.50<br />

** ** ***<br />

***<br />

1.25<br />

WIV<br />

TMC-WIV<br />

TMC-RA-WIV<br />

PBS<br />

0.62<br />

0.31<br />

Conc. WIV (μg/ml)<br />

Conc. WIV (μg/ml)<br />

Figure 5. Uptake/<strong>as</strong>sociation of fluorescently labeled WIV, uncoated or coated with TMC or TMC-RA by<br />

Calu-3 cells when incubated with solution of different concentrations <strong>for</strong> 1 hour at 37°C (A) or 4°C (B). *<br />

p


Chapter 5A<br />

Change in TEER:<br />

% of WIV transported<br />

1.0×10 -2 -1% (±9) -26% (±7) -7% (±4)<br />

7.5×10 -3<br />

5.0×10 -3<br />

2.5×10 -3<br />

0<br />

oo<br />

***<br />

WIV<br />

TMC-WIV<br />

***<br />

TMC-RA-WIV<br />

Figure 6. Transport of fluorescently labeled WIV, uncoated or coated with TMC or TMC-RA, by Calu-3<br />

human bronchial epithelial cells after incubation at 37°C <strong>for</strong> 1 h. Changes in TEER values (± standard<br />

deviation) after 1 h incubation are depicted above the bars. Error bars indicate standard deviations<br />

(n=3). *** value significantly (p


Influence of N-Acetylation on the <strong>Adjuvant</strong>icity of TMC<br />

CD40 exp. (MFI)<br />

500<br />

400<br />

300<br />

200<br />

100<br />

medium<br />

0<br />

WIV<br />

TMC-WIV 5:1<br />

TMC-RA-WIV 5:1<br />

LPS 100 ng/ml<br />

TMC 25 μg/ml<br />

TMC-RA 25 μg/ml<br />

0.63μg/ml WIV<br />

0.31<br />

0.16<br />

0.08<br />

0.04<br />

Controls<br />

Figure 7. A representative example of DC maturation (n=2) (monitored by CD40 expression) by plain<br />

WIV, TMC-WIV and TMC-RA-WIV <strong>for</strong>mulations at 5:1 TMC:WIV (w/w) ratio in different concentrations.<br />

Discussion<br />

As pointed out be<strong>for</strong>e, adjuvants <strong>for</strong> n<strong>as</strong>al immunization can act by enhancing the<br />

immunoavailability of the antigen, or <strong>as</strong> immune potentiator by providing or incre<strong>as</strong>ing a<br />

“danger” signal [1, 2, 5, 6]. In general, TMC is <strong>as</strong>sociated with improving antigen delivery<br />

through its muco-adhesiveness [6, 13, 21, 22], thereby incre<strong>as</strong>ing n<strong>as</strong>al residence time and/or<br />

improving antigen-epithelial barrier interactions [18]. Although of importance <strong>for</strong> mucosal<br />

peptide and protein delivery [6, 12], TMC’s capability to open tight junctions seems not to be of<br />

major relevance <strong>for</strong> n<strong>as</strong>al vaccination <strong>as</strong> several studies showed successful n<strong>as</strong>al immunization<br />

with TMCs with low DQ (


Chapter 5A<br />

In the present study, we found that TMC-RA is degraded to a larger extent in n<strong>as</strong>al w<strong>as</strong>h than<br />

TMC, although this degradation appeared to be rather slow. However, note that the<br />

biodegradability of TMC in a n<strong>as</strong>al w<strong>as</strong>h may not accurately reflect the degradation in the n<strong>as</strong>al<br />

cavity after i.n. administration. It is likely that the concentration of lysozyme and other<br />

prote<strong>as</strong>es in a n<strong>as</strong>al w<strong>as</strong>h is much lower than the concentration in the n<strong>as</strong>al cavity, firstly<br />

because of a strong dilution effect in the w<strong>as</strong>h and secondly because most of the lysozyme is<br />

stored and excreted from submucosal glands [20] and may not be efficiently extracted by a<br />

n<strong>as</strong>al w<strong>as</strong>h. Thus, although in vitro degradation occurred quite slowly, the intran<strong>as</strong>al<br />

degradation of TMC-RA during its n<strong>as</strong>al residence may be significant in vivo. Our results<br />

consequently suggest that TMC-RA is enzymatically degraded more extensively than TMC.<br />

However, both the extent of degradation and the degradation rate in the n<strong>as</strong>al cavity are<br />

difficult to establish quantitatively.<br />

This reduction of polymer molecular weight may lead to a decre<strong>as</strong>e of cytotoxicity [24],<br />

however, TMC-RA h<strong>as</strong> already a very low toxicity profile showing no reduction of cell viability<br />

or incre<strong>as</strong>ed lactate dehydrogen<strong>as</strong>e (LDH) rele<strong>as</strong>e in Caco-2 cells even up to a concentration of<br />

10 mg/ml [15]. Furthermore, both in vivo WIV n<strong>as</strong>al clearance and local n<strong>as</strong>al distribution<br />

patterns are similar <strong>for</strong> TMC and TMC-RA coated WIV, indicating that possible n<strong>as</strong>al<br />

degradation of TMC-RA does not affect antigen exposure to the epithelial barrier. There<strong>for</strong>e,<br />

although n<strong>as</strong>al degradation of TMC-RA in vivo may occur, its effects on the exposure of the<br />

antigen to the epithelial barrier are likely rather small.<br />

Importantly, our data indicate that enhancing the interaction between the antigen and the<br />

epithelial barrier alone is not sufficient <strong>for</strong> an adequate immune response and consequently<br />

other factors have to play a role <strong>as</strong> well. One possibility is that a co-stimulatory ‘danger’ signal<br />

due to toxicity of the TMC is crucial in provoking an adequate immune response [25]. As TMC-<br />

RA showed the lowest in vitro cyto-toxicity of all TMC polymers [15] this polymer may not be<br />

capable of inducing such a danger signal. Interestingly, this difference in toxicity between TMC<br />

and TMC-RA may become even more pronounced due to the higher extent of n<strong>as</strong>al degradation<br />

of TMC-RA.<br />

The uptake and binding of TMC coated WIV by Calu-3 cells decre<strong>as</strong>ed compared to TMC-RA<br />

coated WIV and uncoated WIV. In contr<strong>as</strong>t, in other cell-types (e.g. 16-HBE14o and HeLa cells),<br />

TMC coated WIV w<strong>as</strong> taken up to a higher extent than plain WIV and TMC-RA coated WIV<br />

(results not shown). This implies that the extent of uptake of TMC coated WIV is highly<br />

dependent on the model cell line used. Calu-3 cells are, although derived from human<br />

bronchial epithelium, used extensively <strong>for</strong> ‘n<strong>as</strong>al’ transepithelial electrical resistance (TEER)<br />

106


Influence of N-Acetylation on the <strong>Adjuvant</strong>icity of TMC<br />

studies [26, 27] <strong>as</strong> they <strong>for</strong>m tight monolayers and also secrete mucus [28]. Several<br />

investigation have shown incre<strong>as</strong>ed transport of proteins or peptides through Caco-2 and Calu-<br />

3 monolayers by chitosan and TMC [29, 30], however the transport and uptake by Calu-3 cells<br />

of virus particles h<strong>as</strong> not been studied. Strong interactions with mucus (which is w<strong>as</strong>hed away<br />

after incubation) may be a re<strong>as</strong>on <strong>for</strong> the observed reduced uptake and binding of TMC-WIV by<br />

these cells <strong>as</strong> compared to WIV [31]. Interestingly, this w<strong>as</strong> not observed with the TMC-RA<br />

coated WIV, despite its similar zeta potential compared to TMC-WIV. Perhaps due to steric<br />

hindrance, the electrostatic interactions between TMC-RA and WIV are weaker than those<br />

between TMC and WIV, resulting in better <strong>as</strong>sociation properties <strong>for</strong> WIV in these mucusproducing<br />

cells due to loss of the TMC-RA coating. This suggests that the intran<strong>as</strong>al stability in<br />

the presence of mucus of the TMC-RA-WIV <strong>for</strong>mulation may be inferior compared to the TMC-<br />

WIV particle. This inferior stability <strong>for</strong> TMC-RA-WIV particles w<strong>as</strong> also observed in saline<br />

conditions [32] and may result in decre<strong>as</strong>ed protection against proteolytic enzymes.<br />

Additionally, the results obtained by studying the transport of WIV over the Calu-3<br />

monolayer (WIV>>TMC-RA-WIV>TMC-WIV) support the <strong>as</strong>sumption that WIV particles are<br />

too large to be transported through epithelial tight junctions, since the only <strong>for</strong>mulation that<br />

significantly decre<strong>as</strong>ed the TEER values (TMC-WIV) showed the lowest transport rate (Figure<br />

6). As the negatively charged uncoated WIV w<strong>as</strong> transported to the largest extent, interactions<br />

between the coated particles and mucus may play a key role: the positively charged TMC- and<br />

TMC-RA-WIV particles likely adhere to the negatively charged mucus and are there<strong>for</strong>e unable<br />

to p<strong>as</strong>s through the cell monolayer. Importantly, <strong>as</strong> in vivo i.n. immunization h<strong>as</strong> shown that<br />

TMC greatly enhances antigen immunogenicity, most likely by improved delivery [6, 22], our<br />

results imply that the uptake by and transport through n<strong>as</strong>al epithelial cells is not the main<br />

route <strong>for</strong> improved antigen delivery by TMC. This suggests that the uptake and transport of<br />

TMC b<strong>as</strong>ed particulate systems is mediated through M-cells [1, 3, 6] or direct DC sampling [2].<br />

Interestingly, influenza viruses and influenza derivated virosomes are known to specifically<br />

interact with M-cells [33] most likely due to recognition of sialic acid and galactose residues by<br />

lectin receptors [34]. Additionally, N-acetyl glucosamine moieties, which are more abundant in<br />

TMC-RA, are known to interact with several lectin receptors present on DCs and/or<br />

macrophages [35-38]. However, the effect of coating of WIV with TMC(-RA) on these<br />

interactions remains to be clarified. More advanced in vitro or ex vivo models [39] should be<br />

developed to further study the transport and uptake of WIV particulate systems over mucosal<br />

epithelial surfaces [40]. Possibly co-culturing Calu-3 cells with microfold (M)-cells may<br />

107


Chapter 5A<br />

improve the predictability of these models [21] <strong>for</strong> in vivo applications, <strong>as</strong> w<strong>as</strong> observed <strong>for</strong> a<br />

Caco-2/M-cell co-culture model [41].<br />

Recently it w<strong>as</strong> shown that TMC, both in particulate <strong>for</strong>m and in solution, can exert<br />

significant immunostimulatory effects on human monocyte derived DCs [41-43]. The<br />

mechanism, however, remains unclear although suggestions were made about residual N-<br />

acetylated glucosamine units capable of interacting with C-type lectin receptors present on<br />

DCs [36, 43]. However, since the differences in adjuvanticity between TMC and TMC-RA coated<br />

WIV were observed in mice, it is more appropriate to investigate the effect of TMC and TMC-RA<br />

coated WIV on the maturation of murine DCs. Since TMC-RA contains a higher number of N-<br />

acetyl glucosamine units, an altered interaction with DCs could be expected [36-38, 44]. Our<br />

data, however, suggest that WIV is the main inducer of DC maturation and that there is no<br />

additive effect of TMC or TMC-RA, or any effect of plain TMC or TMC-RA (Figure 7).<br />

Importantly, murine DCs express different C-type lectin receptors on their surface than human<br />

DCs and there<strong>for</strong>e adjuvants may have a different effect on murine or human DC-subsets [19].<br />

Further investigations on human DCs with these <strong>for</strong>mulations are there<strong>for</strong>e appropriate to<br />

clarify these findings.<br />

In summary, we found that TMC-RA w<strong>as</strong> degraded to a higher extent in n<strong>as</strong>al w<strong>as</strong>hes than<br />

TMC. In vivo fluorescence imaging did not reveal a significant difference in n<strong>as</strong>al clearance of<br />

TMC-WIV and TMC-RA-WIV. In n<strong>as</strong>al cross-sections both TMC-WIV <strong>for</strong>mulations were<br />

similarly distributed in the n<strong>as</strong>al cavity, indicating that enhancing the interaction between the<br />

antigen and the epithelial barrier alone is not sufficient <strong>for</strong> improving the immune response.<br />

The uptake and binding of fluorescent WIV in Calu-3 cells w<strong>as</strong> less pronounced when coated<br />

with TMC than with TMC-RA or uncoated and both TMCs significantly decre<strong>as</strong>e the transport<br />

of WIV over a Calu-3 monolayer. Furthermore, experiments with murine BM-DCs indicated<br />

that WIV is the main inducer of DC maturation, me<strong>as</strong>ured <strong>as</strong> CD40 expression, and that TMC<br />

and TMC-RA do not have an additive effect. Our data suggest that the loss of adjuvanticity by<br />

reacetylation of TMC might be due to the lower stability of TMC-RA-WIV in the n<strong>as</strong>al cavity,<br />

rather than by any of the other factors studied in this paper.<br />

Acknowledgement. This research w<strong>as</strong> partially per<strong>for</strong>med under the framework of <strong>TI</strong><br />

<strong>Pharma</strong> project number D5-106-1; Vaccine delivery: alternatives <strong>for</strong> conventional multiple<br />

injection vaccines. E. Kaijzel and C. Löwik are acknowledged <strong>for</strong> their help with the n<strong>as</strong>al<br />

residence time experiment.<br />

108


Influence of N-Acetylation on the <strong>Adjuvant</strong>icity of TMC<br />

References<br />

1. S. Chadwick, C. Kriegel, and M. Amiji. Nanotechnology solutions <strong>for</strong> mucosal immunization.<br />

Adv Drug Deliv Rev 62: 394-407 (2010).<br />

2. N. Csaba, M. Garcia-Fuentes, and M. J. Alonso. Nanoparticles <strong>for</strong> n<strong>as</strong>al vaccination. Adv Drug<br />

Deliv Rev 61: 140-157 (2009).<br />

3. B. Slütter, N. Hagenaars, and W. Jiskoot. Rational design of n<strong>as</strong>al vaccines. J Drug Target 16: 1-17<br />

(2008).<br />

4. N. Hagenaars, E. M<strong>as</strong>trobattista, H. Glansbeek, J. Heldens, H. van den Bosch, V. Schijns, D.<br />

Betbeder, H. Vromans, and W. Jiskoot. Head-to-head comparison of four nonadjuvanted<br />

inactivated cell culture-derived influenza vaccines: Effect of composition, spatial organization<br />

and immunization route on the immunogenicity in a murine challenge model. Vaccine 26: 6555-<br />

6563 (2008).<br />

5. N. Mishra, A. K. Goyal, S. Tiwari, R. Paliwal, S. R. Paliwal, B. Vaidya, S. Mangal, M. Gupta, D.<br />

Dube, A. Mehta, and S. P. Vy<strong>as</strong>. Recent advances in mucosal delivery of vaccines: Role of<br />

mucoadhesive/biodegradable polymeric carriers. Exp Opin Ther Patents 20: 661-679 (2010).<br />

6. M. Amidi, E. M<strong>as</strong>trobattista, W. Jiskoot, and W. E. Hennink. Chitosan-b<strong>as</strong>ed delivery systems<br />

<strong>for</strong> protein therapeutics and antigens. Adv Drug Deliv Rev 62: 59-82 (2010).<br />

7. H. C. Arca, M. Günbeyaz, and S. Şenel. Chitosan-b<strong>as</strong>ed systems <strong>for</strong> the delivery of vaccine<br />

antigens. Exp Rev Vaccines 8: 937-953 (2009).<br />

8. L. Illum, I. Jabbal-Gill, M. Hinchcliffe, A. N. Fisher, and S. S. Davis. Chitosan <strong>as</strong> a novel n<strong>as</strong>al<br />

delivery system <strong>for</strong> vaccines. Adv Drug Deliv Rev 51: 81-96 (2001).<br />

9. D. Snyman, J. H. Hamman, and A. F. Kotze. Evaluation of the mucoadhesive properties of N-<br />

trimethyl chitosan chloride. Drug Develop Ind Pharm 29: 61-69 (2003).<br />

10. G. Di Colo, S. Burgal<strong>as</strong>si, Y. Zambito, D. Monti, and P. Chetoni. Effects of different N-<br />

trimethyl chitosans on in vitro/in vivo ofloxacin transcorneal permeation. J Pharm Sci 93: 2851-<br />

2862 (2004).<br />

11. A. F. Kotze, H. L. Luessen, B. J. De Leeuw, B. G. De Boer, J. C. Verhoef, and H. E. Junginger.<br />

N-<strong>Trimethyl</strong> chitosan chloride <strong>as</strong> a potential absorption enhancer across mucosal surfaces: In<br />

vitro evaluation in intestinal epithelial cells (Caco-2). Pharm Res 14: 1197-1202 (1997).<br />

12. G. Sandri, S. Rossi, M. C. Bonferoni, F. Ferrari, Y. Zambito, G. Di Colo, and C. Caramella.<br />

Buccal penetration enhancement properties of N-trimethyl chitosan: Influence of quaternization<br />

degree on absorption of a high molecular weight molecule. Int J Pharm 297: 146-155 (2005).<br />

13. N. Hagenaars, E. M<strong>as</strong>trobattista, R. J. Verheul, I. Mooren, H. L. Glansbeek, J. G. M. Heldens,<br />

H. Van Den Bosch, and W. Jiskoot. Physicochemical and immunological characterization of<br />

N,N,N-trimethyl chitosan-coated whole inactivated influenza virus vaccine <strong>for</strong> intran<strong>as</strong>al<br />

administration. Pharm Res 26: 1353-1364 (2009).<br />

14. R. J. Verheul, M. Amidi, S. van der Wal, E. van Riet, W. Jiskoot, and W. E. Hennink. Synthesis,<br />

characterization and in vitro biological properties of O-methyl free N,N,N-trimethylated<br />

chitosan. Biomaterials 29: 3642-3649 (2008).<br />

15. R. J. Verheul, M. Amidi, M. J. van Steenbergen, E. van Riet, W. Jiskoot, and W. E. Hennink.<br />

Influence of the degree of acetylation on the enzymatic degradation and in vitro biological<br />

properties of trimethylated chitosans. Biomaterials 30: 3129-3135 (2009).<br />

16. N. Hagenaars, R. J. Verheul, I. Mooren, P. H. J. L. F. de Jong, E. M<strong>as</strong>trobattista, H. L.<br />

Glansbeek, J. G. M. Heldens, H. van den Bosch, W. E. Hennink, and W. Jiskoot. Relationship<br />

between structure and adjuvanticity of N,N,N-trimethyl chitosan (TMC) structural variants in a<br />

n<strong>as</strong>al influenza vaccine. J Control Rele<strong>as</strong>e 140: 126-133 (2009).<br />

17. X. Jiang, A. Van Der Horst, M. J. Van Steenbergen, N. Akeroyd, C. F. Van Nostrum, P. J.<br />

Schoenmakers, and W. E. Hennink. Molar-m<strong>as</strong>s characterization of cationic polymers <strong>for</strong> gene<br />

delivery by aqueous size-exclusion chromatography. Pharm Res 23: 595-603 (2006).<br />

18. N. Hagenaars, M. Mania, P. de Jong, I. Que, R. Nieuwland, B. Slütter, H. Glansbeek, J. Heldens,<br />

H. van den Bosch, C. Löwik, E. Kaijzel, E. M<strong>as</strong>trobattista, and W. Jiskoot. Role of<br />

109


Chapter 5A<br />

trimethylated chitosan (TMC) in n<strong>as</strong>al residence time, local distribution and toxicity of an<br />

intran<strong>as</strong>al influenza vaccine. J Control Rele<strong>as</strong>e 144: 17-24 (2010).<br />

19. S. K. Singh, J. Stephani, M. Schaefer, H. Kalay, J. J. García-Vallejo, J. den Haan, E. Saeland, T.<br />

Sparw<strong>as</strong>ser, and Y. van Kooyk. Targeting glycan modified OVA to murine DC-SIGN<br />

transgenic dendritic cells enhances MHC cl<strong>as</strong>s I and II presentation. Mol Immunol 47: 164-174<br />

(2009).<br />

20. A. M. Cole, H. I. Liao, O. Stuchlik, J. Tilan, J. Pohl, and T. Ganz. Cationic polypeptides are<br />

required <strong>for</strong> antibacterial activity of human airway fluid. J Immunol 169: 6985-6991 (2002).<br />

21. M. Amidi, S. G. Romeijn, J. C. Verhoef, H. E. Junginger, L. Bungener, A. Huckriede, D. J. A.<br />

Crommelin, and W. Jiskoot. N-<strong>Trimethyl</strong> chitosan (TMC) nanoparticles loaded with influenza<br />

subunit antigen <strong>for</strong> intran<strong>as</strong>al vaccination: Biological properties and immunogenicity in a mouse<br />

model. Vaccine 25: 144-153 (2007).<br />

22. S. M. Van Der Merwe, J. C. Verhoef, J. H. M. Verheijden, A. F. Kotze, and H. E. Junginger.<br />

<strong>Trimethyl</strong>ated chitosan <strong>as</strong> polymeric absorption enhancer <strong>for</strong> improved peroral delivery of<br />

peptide drugs. Eur J Pharm Biopharm 58: 225-235 (2004).<br />

23. I. M. Van der Lubben, J. C. Verhoef, M. M. Fretz, O. Van, I. Mesu, G. Kersten, and H. E.<br />

Junginger. <strong>Trimethyl</strong> chitosan chloride (TMC) <strong>as</strong> a novel excipient <strong>for</strong> oral and n<strong>as</strong>al<br />

immunisation against diphtheria. S.T.P. <strong>Pharma</strong> Sciences 12: 235-242 (2002).<br />

24. S. Mao, X. Shuai, F. Unger, M. Wittmar, X. Xie, and T. Kissel. Synthesis, characterization and<br />

cytotoxicity of poly(ethylene glycol)-graft-trimethyl chitosan block copolymers. Biomaterials 26:<br />

6343-6356 (2005).<br />

25. V. E. Schijns. Immunological concepts of vaccine adjuvant activity. Curr Opin Immunol 12: 456-<br />

463 (2000).<br />

26. M. Amidi, S. G. Romeijn, G. Borchard, H. E. Junginger, W. E. Hennink, and W. Jiskoot.<br />

Preparation and characterization of protein-loaded N-trimethyl chitosan nanoparticles <strong>as</strong> n<strong>as</strong>al<br />

delivery system. J Control Rele<strong>as</strong>e 111: 107-116 (2006).<br />

27. D. Teijeiro-Osorio, C. Remuñán-López, and M. J. Alonso. New generation of hybrid<br />

poly/oligosaccharide nanoparticles <strong>as</strong> carriers <strong>for</strong> the n<strong>as</strong>al delivery of macromolecules.<br />

Biomacromolecules 10: 243-249 (2009).<br />

28. K. A. Foster, M. L. Avery, M. Yazdanian, and K. L. Audus. Characterization of the Calu-3 cell<br />

line <strong>as</strong> a tool to screen pulmonary drug delivery. Int J Pharm 208: 1-11 (2000).<br />

29. B. I. Florea, M. Thanou, H. E. Junginger, and G. Borchard. Enhancement of bronchial<br />

octreotide absorption by chitosan and N-trimethyl chitosan shows linear in vitro/in vivo<br />

correlation. J Control Rele<strong>as</strong>e 110: 353-361 (2006).<br />

30. C. Witschi and R. J. Mrsny. In vitro evaluation of microparticles and polymer gels <strong>for</strong> use <strong>as</strong><br />

n<strong>as</strong>al plat<strong>for</strong>ms <strong>for</strong> protein delivery. Pharm Res 16: 382-390 (1999).<br />

31. S. K. Lai, Y.-Y. Wang, and J. Hanes. Mucus-penetrating nanoparticles <strong>for</strong> drug and gene<br />

delivery to mucosal tissues. Adv Drug Deliv Rev 61: 158-171 (2009).<br />

32. N. Hagenaars. Towards an intran<strong>as</strong>al influenza vaccine - B<strong>as</strong>ed on whole inactivated influenza<br />

virus with N,N,N-trimethyl chito<strong>as</strong>n <strong>as</strong> adjuvant. Utrecht University, Utrecht (2010).<br />

33. Y. Fujimura, M. Takeda, H. Ikai, K. Haruma, T. Akisada, T. Harada, T. Sakai, and M. Ohuchi.<br />

The role of M cells of human n<strong>as</strong>opharyngeal lymphoid tissue in influenza virus sampling.<br />

Virchows Archiv 444: 36-42 (2004).<br />

34. G. N. Rogers and J. C. Paulson. Receptor determinants of human and animal influenza virus<br />

isolates: Differences in receptor specificity of the H3 hemagglutinin b<strong>as</strong>ed on species of origin.<br />

Virology 127: 361-373 (1983).<br />

35. K. Nishimura, S. Nishimura, and H. Seo. Macrophage activation with multi-porous beads<br />

prepared from partially deacetylated chitin. J Biomed Mat Res 20: 1359-1372 (1986).<br />

36. M. J. Robinson, D. Sancho, E. C. Slack, S. LeibundGut-Landmann, and C. R. Sousa. Myeloid C-<br />

type lectins in innate immunity. Nat Immunol 7: 1258-1265 (2006).<br />

37. J. Nadesalingam, A. W. Dodds, K. B. M. Reid, and N. Palaniyar. Mannose-binding lectin<br />

recognizes peptidoglycan via the N-acetyl glucosamine moiety, and inhibits ligand-induced<br />

110


Influence of N-Acetylation on the <strong>Adjuvant</strong>icity of TMC<br />

proinflammatory effect and promotes chemokine production by macrophages. J Immunol 175:<br />

1785-1794 (2005).<br />

38. P. Zhang, S. Snyder, P. Feng, P. Azadi, S. Zhang, S. Bulgheresi, K. E. Sanderson, J. He, J.<br />

Klena, and T. Chen. Role of N-acetylglucosamine within core lipopolysaccharide of several<br />

species of Gram-negative bacteria in targeting the DC-SIGN (CD209). J Immunol 177: 4002-<br />

4011 (2006).<br />

39. G. Sandri, P. Poggi, M. C. Bonferoni, S. Rossi, F. Ferrari, and C. Caramella. Histological<br />

evaluation of buccal penetration enhancement properties of chitosan and trimethyl chitosan. J<br />

Pharm <strong>Pharma</strong>col 58: 1327 (2006).<br />

40. B. Forbes and C. Ehrhardt. Human respiratory epithelial cell culture <strong>for</strong> drug delivery<br />

applications. Eur J Pharm Biopharm 60: 193-205 (2005).<br />

41. B. Slütter, L. Plapied, V. Fievez, M. Alonso Sande, A. des Rieux, Y. J. Schneider, E. Van Riet,<br />

W. Jiskoot, and V. Préat. Mechanistic study of the adjuvant effect of biodegradable<br />

nanoparticles in mucosal vaccination. J Control Rele<strong>as</strong>e 138: 113-121 (2009).<br />

42. S. M. Bal, B. Slütter, E. van Riet, A. C. Kruithof, Z. Ding, G. F. A. Kersten, W. Jiskoot, and J.<br />

A. Bouwstra. Efficient induction of immune responses through intradermal vaccination with N-<br />

trimethyl chitosan containing antigen <strong>for</strong>mulations. J Control Rele<strong>as</strong>e 142: 374-383 (2010).<br />

43. B. Slütter, P. C. Soema, Z. Ding, R. Verheul, W. Hennink, and W. Jiskoot. Conjugation of<br />

ovalbumin to trimethyl chitosan improves immunogenicity of the antigen. J Control Rele<strong>as</strong>e 143:<br />

207-214 (2010).<br />

44. K. Nishimura, C. Ishihara, and S. Ukei. Stimulation of cytokine production in mice using<br />

deacetylated chitin. Vaccine 4: 151-156 (1986).<br />

111


CHAPTER 5B<br />

MATURA<strong>TI</strong>ON OF HUMAN MONOCYTE<br />

DERIVED DENDRI<strong>TI</strong>C CELLS BY TRIMETHYL<br />

CHITOSAN IS CORRELATED WITH ITS N-<br />

ACETYL GLUCOSAMINE (GLCNAC) CONTENT<br />

Rolf J. Verheul, Niels Hagenaars, Thom<strong>as</strong> van Es, Sven Bruins,<br />

Bram Slütter, Wim E. Hennink, Wim Jiskoot.<br />

Manuscript submitted


Chapter 5B<br />

Abstract<br />

N-acetylated glucosamine units or GlcNAcs present in trimethyl chitosan (TMC) have been<br />

described to interact with several pathogen <strong>as</strong>sociated molecular pattern receptors involved in<br />

the human innate immune response. The aim of this study w<strong>as</strong> to compare the immunomodulatory<br />

effects of TMC with a low (17%) and high (54%, TMC-RA) degree of acetylation or<br />

GlcNAc content on the uptake and maturation of human monocyte derived dendritic cells (MO-<br />

DCs) in vitro, using whole inactivated influenza virus (WIV) <strong>as</strong> antigen. The GlcNAc content of<br />

TMC had no effect on the uptake of TMC(-RA) coated WIV by MO-DCs. Interestingly, TMC-RA,<br />

either <strong>as</strong> a solution or when <strong>for</strong>mulated with WIV, induced a much stronger DC maturation<br />

than any of the other <strong>for</strong>mulations, <strong>as</strong> judged from CD86 expression and cytokine rele<strong>as</strong>e (IL-<br />

10, TNF-α and IL-12p40 and IL-12p70). Since both IL-10 and IL-12p70 levels were elevated, no<br />

polarization towards a Th1 or Th2 type immune response could be established. In conclusion,<br />

<strong>as</strong> compared to TMC, TMC-RA h<strong>as</strong> strong immuno-stimulatory effects in vitro on human MO-<br />

DCs. This implies that the degree of N-acetylation or GlcNAc content may be critical <strong>for</strong> the<br />

adjuvant effect of TMC in humans.<br />

114


Maturation of Human DCs is Correlated with the GlcNAc Content<br />

Introduction<br />

Subunit or inactivated vaccines are generally safer but less effective than live attenuated<br />

vaccines and there<strong>for</strong>e require potent adjuvants to elicit adequate immune responses [1, 2]. As<br />

few adjuvants are approved <strong>for</strong> use in humans [3], there is a need <strong>for</strong> novel compounds. N,N,Ntrimethyl<br />

chitosan (TMC), a quaternized chitosan derivative, h<strong>as</strong> successfully been used <strong>for</strong><br />

mucosal and intradermal vaccination [4-7] and the structural properties of TMC like the<br />

degree of quaternization (DQ), the degree of O-methylation (DOM), and the degree of N-<br />

acetylation (DAc) can be varied during synthesis [8, 9]. In vitro evaluation showed that<br />

incre<strong>as</strong>ing the DAc or N-acetyl glucosamine (GlcNAc) content of TMC resulted in a decre<strong>as</strong>e in<br />

cell-toxicity and better enzymatic degradation by lysozyme [9]. GlcNAc moieties are, <strong>as</strong><br />

pathogen-<strong>as</strong>sociated molecular patterns (PAMPs), known to interact with C-type lectins [10],<br />

like DC-SIGN [11, 12], mannose receptors [13, 14], and toll-like receptor type 2 (TLR2) [15]<br />

present on the surface of several antigen presenting cells (APCs), such <strong>as</strong> DCs and<br />

macrophages [10]. However, after intran<strong>as</strong>al (i.n.) immunization with whole inactivated<br />

influenza virus (WIV) in mice, a striking loss in adjuvanticity of TMC w<strong>as</strong> observed when WIV<br />

w<strong>as</strong> <strong>for</strong>mulated with reacetylated TMC (TMC-RA; DAc of 54%), compared to all other tested<br />

TMCs with a low DAc (17%) [16]. Since no differences in the physico-chemical properties of<br />

the <strong>for</strong>mulations were observed, this lack of adjuvanticity had to be attributed to the higher<br />

GlcNAc content of TMC-RA, however the precise mechanism still remains to be elucidated.<br />

A recent study by our group showed no significant difference in activation of murine bone<br />

marrow derived DCs after exposure to regular TMC or TMC-RA (Chapter 5A). Interestingly,<br />

human DCs may be different from mouse DCs <strong>as</strong> murine and human DCs express different<br />

receptors on their surface [11, 17]. There<strong>for</strong>e, in this study we further explored the differences<br />

in immuno-stimulatory effect between reacetylated TMC-RA (GlcNAc content 54%) and<br />

conventional TMC (GlcNAc content 17%) using human monocyte derived DCs (MO-DCs).<br />

Materials and Methods<br />

Materials. Chitosan with a DAc of 17% (determined with 1 H-NMR <strong>as</strong> described in [9] and a<br />

number average molecular weight (M n) and weight average molecular weight (M w) of 28 and<br />

43 kDa, respectively, <strong>as</strong> determined by gel permeation chromatography (GPC) <strong>as</strong> described in<br />

[8], w<strong>as</strong> purch<strong>as</strong>ed from Primex (Siglufjordur, Iceland). Acetic anhydride, sodium borohydrate,<br />

<strong>for</strong>mic acid, <strong>for</strong>maldehyde 37% (stabilized with methanol), deuterium oxide, sodium acetate,<br />

115


Chapter 5B<br />

acetic acid (anhydrous), sodium hydroxide and hydrochloric acid were obtained from Sigma-<br />

Aldrich Chemical Co (Zwijndrecht, The Netherlands). Iodomethane 99% stabilized with copper<br />

w<strong>as</strong> obtained from Acros Organics (Geel, Belgium). Purified, cell culture-grown (Madin-Darby<br />

Canine Kidney (MDCK) cells, β-propiolacton (BPL)-inactivated influenza virus, strain<br />

A/PR/8/34 (WIV) w<strong>as</strong> from Nobilon International BV (Boxmeer, The Netherlands). Alexa<br />

Fluor® 488 labeling kit w<strong>as</strong> purch<strong>as</strong>ed from Invitrogen (Breda, the Netherlands).<br />

Trypsin/EDTA 10x, Plain DMEM (Dulbecco’s modification of Eagle’s medium, with 3.7 g/l<br />

sodium bicarbonate, 1 g/l L-glucose, L-glutamine) and antibiotics/antimycotics (penicillin,<br />

streptomycin sulphate, amphotericin B) were from PAA Laboratories GmbH (P<strong>as</strong>ching,<br />

Austria) and MEM with 0.292 g/L L-glutamine, 1 g/L glucose, 2.2 g/L NaHCO 3 w<strong>as</strong> from<br />

(Invitrogen, Breda, The Netherlands). All other chemicals used were of analytical grade.<br />

Synthesis and characterization of methylated chitosans. O-methyl free N,N,Ntrimethylated<br />

chitosans (TMCs) with a DQ of about 45% and varying DAc were synthesized<br />

from chitosan <strong>as</strong> described previously [9]. The DQ and DAc of the TMCs were determined with<br />

1H NMR on a Varian INOVA 500 MHz NMR spectrometer (Varian Inc., Palo Alto, Ca, USA) at 80<br />

°C in D 2O [9]. Furthermore, M n and M w of the TMCs were determined, <strong>as</strong> described previously<br />

[8], by GPC on a Viscotek system detecting refractive index, viscosity and light scattering. A<br />

Shodex OHPak SB-806 column (30 cm) w<strong>as</strong> used with 0.3 M sodium acetate pH 4.4 (adjusted<br />

with acetic acid) <strong>as</strong> running buffer. The structural characteristics of the synthesized TMCs are<br />

summarized in Table 1.<br />

Table 1. Polymer characteristics of TMC and TMC-RA.<br />

Abbreviation M n (kDa) M w (kDa) DQ (%) DAc (%)<br />

TMC 36 75 43 17<br />

TMC-RA 43 83 44 54<br />

In vitro studies on human MO-DCs. Immature DCs were cultured <strong>as</strong> described be<strong>for</strong>e [18].<br />

In short, human blood monocytes were isolated from buffy coats by use of a Ficoll gradient and<br />

subsequently a Percoll gradient. Purified monocytes were differentiated into immature DCs in<br />

the presence of interleukin-4 (IL-4, 500 U/ml) and granulocyte-macrophage colonystimulating<br />

factor (GM-CSF, 800 U/ml).<br />

Immature DCs (day 6) were seeded (50,000 cells/well) in a 96-well plate in a total volume of<br />

100 μl RPMI 10% FCS in the presence of TMC (0.001-30 μg/ml), TMC-WIV <strong>for</strong>mulations at 1:1<br />

and 5:1 (w/w) ratios (0.001-6 μg/ml WIV) or lipopolysaccharide (LPS) (0.01-1000 ng/ml) <strong>as</strong> a<br />

116


Maturation of Human DCs is Correlated with the GlcNAc Content<br />

positive control. As described elsewhere, TMC-WIV <strong>for</strong>mulations had similar physico-chemical<br />

properties (size and zeta potential) independent of the type or amount of TMC used (Chapter<br />

5A). There<strong>for</strong>e, it can be stated that all <strong>for</strong>mulations were physically similar, except <strong>for</strong> the<br />

amount of free polymer.<br />

Uptake and binding of TMC-coated WIV <strong>for</strong>mulations by human MO-DCs. The uptake<br />

and binding of (TMC coated) WIV by human MO-DCs were studied with AlexaFluor-488®<br />

labeled WIV. WIV w<strong>as</strong> labeled with the AlexaFluor-488® labeling kit using the procedure<br />

provided by Invitrogen. DCs were incubated <strong>for</strong> 1 h at 37 or 4 °C with the labeled WIV<br />

<strong>for</strong>mulations and flow cytometry w<strong>as</strong> used to determine cell fluorescence (n=2).<br />

Maturation and cytokine production of human MO-DCs. Human MO-DCs were incubated<br />

<strong>for</strong> 16 h at 37 °C with the various polymers with or without WIV. The cells were w<strong>as</strong>hed<br />

extensively with PBS and incubated <strong>for</strong> 30 min with anti-CD86-APC (BD, Breda, The<br />

Netherlands). DC maturation w<strong>as</strong> determined by analyzing cell-surface expression of costimulatory<br />

molecules (CD86) by flow cytometry. After 16 h culture supernatants were<br />

harvested and frozen at −80 °C until analysis. The supernatants were analyzed <strong>for</strong> the presence<br />

of the cytokines IL-10 and IL-12p40, IL12p70 and TNFα by ELISA (Biosource International,<br />

CA). For the DC maturation studies, 4 donors were used. For the cytokine experiments, 2<br />

donors were used.<br />

Results<br />

Uptake and binding of (TMC coated) WIV by human MO-DCs. The influence of TMC and<br />

TMC-RA on the uptake and binding of fluorescently labeled WIV by human MO-DCs w<strong>as</strong><br />

studied. After incubation <strong>for</strong> 1 h, all <strong>for</strong>mulations showed a concentration dependent uptake<br />

and binding (Figure 1). At 37 °C (Figure 1A) much higher intensities were observed than at 4<br />

°C (Figure 1B), indicating active uptake of WIV. Surprisingly, uncoated WIV showed a slightly<br />

higher binding and uptake than TMC or TMC-RA coated WIV.<br />

117


Chapter 5B<br />

WIV binding/uptake<br />

(MFI)<br />

A<br />

200<br />

150<br />

100<br />

50<br />

WIV<br />

TMC-WIV<br />

TMC-RA-WIV<br />

WIV binding/uptake<br />

(MFI)<br />

B<br />

20<br />

15<br />

10<br />

5<br />

WIV<br />

TMC-WIV<br />

TMC-RA-WIV<br />

0<br />

3.0 1.5 0.8 0.4 0.2<br />

Conc. WIV (μg/ml)<br />

0<br />

3.0 1.5 0.8 0.4 0.2<br />

Conc. WIV (μg/ml)<br />

Figure 1. WIV uptake and/or binding by human MO-DCs after incubation <strong>for</strong> 1 hour with fluorescently<br />

labeled (TMC or TMC-RA coated at w/w ratio 5:1) WIV at 37 °C (A) and 4°C (B). Results from one<br />

representative donor are shown (n=2). TMC(-RA)-WIV <strong>for</strong>mulations at w/w ratio 1:1 showed a similar<br />

outcome (results not shown).<br />

Maturation and cytokine production by human MO-DCs. The effect of TMC and TMC-RA<br />

on MO-DC maturation w<strong>as</strong> studied by quantifying maturation marker CD86 and production of<br />

cytokines IL-10, TNF-α and IL-12. Incubation of human MO-DCs with TMC solutions and TMC-<br />

WIV <strong>for</strong>mulations ((w/w) ratio 5:1) revealed a striking difference in CD86 expression between<br />

TMC-RA and TMC (Figure 2).<br />

CD 86 expression<br />

(MFI)<br />

A<br />

300<br />

200<br />

100<br />

TMC<br />

TMC-RA<br />

CD 86 expression<br />

(MFI)<br />

B<br />

500<br />

400<br />

300<br />

200<br />

100<br />

WIV<br />

TMC-WIV<br />

TMC-RA-WIV<br />

0<br />

15.0<br />

7.5<br />

3.8<br />

1.9<br />

0.9<br />

Medium<br />

1 μg/ml LPS<br />

0<br />

3.0<br />

1.5<br />

0.8<br />

0.4<br />

0.2<br />

Medium<br />

1 μg/ml LPS<br />

Conc. TMC (μg/ml)<br />

Conc. WIV (μg/ml)<br />

Figure 2. CD 86 expression by human MO-DCs after incubation with TMC and TMC-RA solutions (A) and<br />

by TMC-WIV and TMC-RA-WIV (5:1 w/w ratio) <strong>for</strong>mulations (B). Results from one representative donor<br />

are shown (n=4).<br />

118


Maturation of Human DCs is Correlated with the GlcNAc Content<br />

TMC-RA induced a much higher concentration dependent CD86 expression than TMC, <strong>for</strong><br />

both soluble TMC-RA (Figure 2A) and TMC-RA co-<strong>for</strong>mulated with WIV (Figure 2B).<br />

Furthermore, TMC-RA-WIV also stimulated, in a concentration dependent manner, the<br />

secretion of pro-inflammatory cytokines IL-10, TNF-α, IL12p40 and IL-12p70 more strongly<br />

than the other TMC-WIV <strong>for</strong>mulations or naked WIV did (Figure 3). As compared to TMC-RA-<br />

WIV, TMC-RA in solution induced similar cytokine levels (results not shown). Importantly, an<br />

endotoxin dose-effect calibration on DCs verified that the observed effects on DCs cannot be<br />

attributed to the relatively low endotoxin levels in the TMCs (≤0.16 EU/µg TMC), <strong>as</strong> me<strong>as</strong>ured<br />

with a LAL test. As TMC-RA showed very low in vitro cytotoxicity compared to TMC [9], it is<br />

highly unlikely that the maturation of DCs by TMC-RA is attributable to its toxicity. Since both<br />

IL-10 (<strong>as</strong> marker <strong>for</strong> a Th2 type response) and IL-12p70 (<strong>as</strong> marker <strong>for</strong> a Th1 type response)<br />

were elevated by TMC-RA (Figure 3), no polarization towards a Th1 or Th2 type of immune<br />

response could be established. These results suggest that TMC-RA h<strong>as</strong> much stronger intrinsic<br />

adjuvant effect than conventional TMC due to its higher GlcNAc content and that this enhanced<br />

activation does not seem to be the result of incre<strong>as</strong>ed uptake of antigen.<br />

Discussion<br />

In general, cationic nanoparticles are better taken up by APCs than their anionic<br />

counterparts [19, 20], likely due to improved interaction with the negatively charged cell<br />

membrane. Despite being anionic, uncoated WIV demonstrated slightly better uptake<br />

compared to the cationic, TMC-coated WIV, indicating that besides electrostatic interactions<br />

other interactions may play a role. It is known that sialic acid and galactose residues on the<br />

surface of influenza viruses can interact with C-type lectin receptors of APCs and enhance<br />

uptake of these particles [21, 22]. The coating of WIV with TMC or TMC-RA may hinder these<br />

interactions. Additionally, the GlcNAc units present in TMC(-RA) may have altered the uptake<br />

mechanism [10, 13]. This obviously resulted in a slightly decre<strong>as</strong>ed uptake of TMC(-RA)-coated<br />

WIV compared to plain WIV. This effect w<strong>as</strong> similar <strong>for</strong> TMC (DAc 17%) and TMC-RA (DAc<br />

54%), indicating that the GlcNAc content of TMC does not critically influence antigen uptake.<br />

In agreement with our present results, in previous studies the extent of antigen uptake did<br />

not appear to be a major determinant <strong>for</strong> DC maturation [5, 23, 24]. However, previous studies<br />

with TMC with low DAc (


6.0<br />

0.2<br />

3.0<br />

6.0<br />

3.0<br />

6.0<br />

3.0<br />

1.5<br />

0.2<br />

Chapter 5B<br />

(with a low DAc) h<strong>as</strong> only a limited intrinsic adjuvant effect on human MO-DCs compared to<br />

TMC-RA (high DAc). TMC, <strong>as</strong> coating of WIV, w<strong>as</strong> not capable to induce additional MO-DC<br />

maturation compared to WIV alone. The different effects of TMC on the maturation of MO-DCs<br />

may be explained by the nature of the antigens used <strong>for</strong> these studies: WIV h<strong>as</strong> a<br />

nanoparticulate structure (appr. 180 nm) and already induces (minor) DC activation without<br />

TMC. Ovalbumin, <strong>as</strong> a soluble antigen, h<strong>as</strong> no effect on DC maturation and mixing with TMC<br />

will lead to (minor) self-<strong>as</strong>sembly into larger structures due to charge interactions [5, 25]. In<br />

the latter c<strong>as</strong>e, DC activation by TMC, either due to changes in the particulate nature of a<br />

<strong>for</strong>mulation [1, 26] or due to a direct effect of the TMC [5, 24], will be more noticeable.<br />

IL-10 (pg/ml)<br />

A<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

WIV<br />

TMC-WIV<br />

TMC-RA-WIV<br />

1 μg/ml LPS<br />

TNFα (pg/ml)<br />

B<br />

10000<br />

7500<br />

5000<br />

2500<br />

WIV<br />

TMC-WIV<br />

TMC−RA-WIV<br />

1 μg/ml LPS<br />

0<br />

0<br />

0.8<br />

0.4<br />

6.0<br />

1.5<br />

0.8<br />

0.4<br />

0.2<br />

Conc. WIV (μg/ml)<br />

Conc. WIV (μg/ml)<br />

IL-12p40 (pg/ml)<br />

C<br />

10000<br />

7500<br />

5000<br />

2500<br />

WIV<br />

TMC-WIV<br />

TMC-RA-WIV<br />

1 μg/ml LPS<br />

IL-12p70 (pg/ml)<br />

D<br />

50<br />

40<br />

30<br />

20<br />

10<br />

WIV<br />

TMC-WIV<br />

TMC-RA-WIV<br />

1 μg/ml LPS<br />

0<br />

3.0<br />

1.5<br />

0.8<br />

0.4<br />

0<br />

1.5<br />

0.8<br />

0.4<br />

0.2<br />

Conc. WIV (μg/ml)<br />

Conc. WIV (μg/ml)<br />

Figure 3. Expression of IL-10 (A); TNFα (B); IL-12p40 (C) and IL12p70 (D) by human MO-DCs after<br />

incubation with TMC-WIV <strong>for</strong>mulations (w/w ratio 5:1). The figure shows results from one<br />

representative donor (n=2).<br />

TMC-RA, <strong>as</strong> coating of WIV or solution, lacked immuno-stimmulatory effects in murine bonemarrow<br />

derived DCs (Chapter 5A). In contr<strong>as</strong>t, our present data show that TMC-RA induces<br />

maturation of human monocyte derived DCs. This suggest that TMC-RA may activate human<br />

120


Maturation of Human DCs is Correlated with the GlcNAc Content<br />

DCs by binding to a receptor that is not present on murine DCs. GlcNAc moieties are known to<br />

interact with species-independent C-type lectins, like mannose receptors [10, 13, 14], and tolllike<br />

receptor type 2 (TLR2) [15] present on the surface of APCs. DC-SIGN is a human C-type<br />

lectin not present on murine DCs [11, 17], but experiments with DCs derived from DC-SIGN<br />

transgenic mice showed no immuno-stimulatory effect of TMC-RA (unpublished results). This<br />

suggests that other GlcNAc, human specific, receptors may play a role. Further studies should<br />

be done to investigate the mechanism of human MO-DC activation by TMC-RA in more detail.<br />

Conclusion<br />

As opposed to TMC with a low GlcNAc content, TMC-RA with a high GlcNAc content (54%),<br />

both <strong>as</strong> solution and <strong>as</strong> coating of WIV, w<strong>as</strong> shown to be capable of inducing maturation of<br />

human MO-DCs. This maturation w<strong>as</strong> not a result of incre<strong>as</strong>ed antigen uptake. Both IL-10 and<br />

IL-12p70 levels were elevated after stimulation with TMC-RA, indicating a mixed Th1/Th2<br />

type immune response. These immuno-stimulatory characteristics, together with its low<br />

toxicity profile, make TMC-RA a promising adjuvant <strong>for</strong> future vaccinations in humans.<br />

Acknowledgement. This research w<strong>as</strong> partially per<strong>for</strong>med under the framework of <strong>TI</strong><br />

<strong>Pharma</strong> project number D5-106-1; Vaccine delivery: alternatives <strong>for</strong> conventional multiple<br />

injection vaccines. Nobilon International BV (Boxmeer, The Netherlands) is acknowledged <strong>for</strong><br />

supplying the WIV.<br />

121


Chapter 5B<br />

References<br />

1. A.C. Rice-Ficht, A.M. Aren<strong>as</strong>-Gamboa, M.M. Kahl-McDonagh, T.A. Ficht. Polymeric particles<br />

in vaccine delivery. Curr Opin Microbiol 13: 106-112 (2010).<br />

2. V.E. Schijns. Immunological concepts of vaccine adjuvant activity. Curr Opin Immunol 12: 456-<br />

463 (2000).<br />

3. P. Nordly, H.B. Madsen, H.M. Nielsen, C. Foged. Status and future prospects of lipid-b<strong>as</strong>ed<br />

particulate delivery systems <strong>as</strong> vaccine adjuvants and their combination with<br />

immunostimulators. Exp Opin Drug Deliv 6:657-672 (2009).<br />

4. M. Amidi, S.G. Romeijn, J.C. Verhoef, H.E. Junginger, L. Bungener, A. Huckriede, D.J.A.<br />

Crommelin, W. Jiskoot. N-<strong>Trimethyl</strong> chitosan (TMC) nanoparticles loaded with influenza<br />

subunit antigen <strong>for</strong> intran<strong>as</strong>al vaccination: Biological properties and immunogenicity in a mouse<br />

model. Vaccine 25:144-153 (2007).<br />

5. S.M. Bal, B. Slütter, E. van Riet, A.C. Kruithof, Z. Ding, G.F.A. Kersten, W. Jiskoot, J.A.<br />

Bouwstra. Efficient induction of immune responses through intradermal vaccination with N-<br />

trimethyl chitosan containing antigen <strong>for</strong>mulations. J Control Rele<strong>as</strong>e 142:374-383 (2010).<br />

6. W. Boonyo, H.E. Junginger, N. Waranuch, A. Polnok, T. Pitaksuteepong. Chitosan and<br />

trimethyl chitosan chloride (TMC) <strong>as</strong> adjuvants <strong>for</strong> inducing immune responses to ovalbumin in<br />

mice following n<strong>as</strong>al administration. J Control Rele<strong>as</strong>e 121:168-175 (2007).<br />

7. N. Hagenaars, E. M<strong>as</strong>trobattista, R.J. Verheul, I. Mooren, H.L. Glansbeek, J.G.M. Heldens, H.<br />

Van Den Bosch, W. Jiskoot. Physicochemical and immunological characterization of N,N,Ntrimethyl<br />

chitosan-coated whole inactivated influenza virus vaccine <strong>for</strong> intran<strong>as</strong>al<br />

administration. Pharm Res 26:1353-1364 (2009).<br />

8. R.J. Verheul, M. Amidi, S. van der Wal, E. van Riet, W. Jiskoot, W.E. Hennink. Synthesis,<br />

characterization and in vitro biological properties of O-methyl free N,N,N-trimethylated<br />

chitosan. Biomaterials 29:3642-3649 (2008).<br />

9. R.J. Verheul, M. Amidi, M.J. van Steenbergen, E. van Riet, W. Jiskoot, W.E. Hennink. Influence<br />

of the degree of acetylation on the enzymatic degradation and in vitro biological properties of<br />

trimethylated chitosans. Biomaterials 30:3129-3135 (2009).<br />

10. M.J. Robinson, D. Sancho, E.C. Slack, S. LeibundGut-Landmann, C.R. Sousa. Myeloid C-type<br />

lectins in innate immunity. Nat Immunol 7:1258-1265 (2006).<br />

11. S.K. Singh, J. Stephani, M. Schaefer, H. Kalay, J.J. Garcia-Vallejo, J. den Haan, E. Saeland, T.<br />

Sparw<strong>as</strong>ser, Y. van Kooyk. Targeting glycan modified OVA to murine DC-SIGN transgenic<br />

dendritic cells enhances MHC cl<strong>as</strong>s I and II presentation. Mol Immunol 47:164-174 (2009).<br />

12. P. Zhang, S. Snyder, P. Feng, P. Azadi, S. Zhang, S. Bulgheresi, K.E. Sanderson, J. He, J. Klena,<br />

T. Chen. Role of N-acetylglucosamine within core lipopolysaccharide of several species of<br />

Gram-negative bacteria in targeting the DC-SIGN (CD209). J Immunol 177:4002-4011 (2006).<br />

13. J. Nadesalingam, A.W. Dodds, K.B.M. Reid, N. Palaniyar. Mannose-binding lectin recognizes<br />

peptidoglycan via the N-acetyl glucosamine moiety, and inhibits ligand-induced<br />

proinflammatory effect and promotes chemokine production by macrophages. J Immunol<br />

175:1785-1794 (2005).<br />

14. Y. Shibata, W. James Metzger, Q.N. Myrvik. Chitin particle-induced cell-mediated immunity is<br />

inhibited by soluble mannan: Mannose receptor-mediated phagocytosis initiates IL-12<br />

production. J Immunol 159:2462-2467 (1997).<br />

15. C.A. Da Silva, D. Hartl, W. Liu, C.G. Lee, J.A. Eli<strong>as</strong>. TLR-2 and IL-17A in chitin-induced<br />

macrophage activation and acute inflammation. J Immunol 181:4279-4286 (2008).<br />

16. N. Hagenaars, R.J. Verheul, I. Mooren, P.H.J.L.F. de Jong, E. M<strong>as</strong>trobattista, H.L. Glansbeek,<br />

J.G.M. Heldens, H. van den Bosch, W.E. Hennink, W. Jiskoot. Relationship between structure<br />

and adjuvanticity of N,N,N-trimethyl chitosan (TMC) structural variants in a n<strong>as</strong>al influenza<br />

vaccine. J Control Rele<strong>as</strong>e 140:126-133 (2009).<br />

17. K. Wethmar, Y. Helmus, K. Lühn, C. Jones, A. L<strong>as</strong>kowska, G. Varga, S. Grabbe, R. Lyck, B.<br />

Engelhardt, M.G. Bixel, S. Butz, K. Loser, S. Beissert, U. Ipe, D. Vestweber, M.K. Wild.<br />

122


Maturation of Human DCs is Correlated with the GlcNAc Content<br />

Migration of immature mouse DC across resting endothelium id mediated by ICAM-2 but<br />

independent ß2-integrins and murine DC-SIGN homologues. Eur J Immunol 36:2781-2794<br />

(2006).<br />

18. F. Sallusto A. Lanzavecchia. Efficient presentation of soluble antigen by cultured human<br />

dendritic cells is maintained by granulocyte/macrophage colony-stimulating factor plus<br />

interleukin 4 and downregulated by tumor necrosis factor α. J Exp Med 179:1109-1118 (1994).<br />

19. K.S. Jaganathan S.P. Vy<strong>as</strong>. Strong systemic and mucosal immune responses to surface-modified<br />

PLGA microspheres containing recombinant Hepatitis B antigen administered intran<strong>as</strong>ally.<br />

Vaccine 24:4201-4211 (2006).<br />

20. B. Sayin, S. Somavarapu, X.W. Li, M. Thanou, D. Sesardic, H.O. Alpar, S. Şenel. Mono-Ncarboxymethyl<br />

chitosan (MCC) and N-trimethyl chitosan (TMC) nanoparticles <strong>for</strong> non-inv<strong>as</strong>ive<br />

vaccine delivery. Int J Pharm 363:139-148 (2008).<br />

21. Y. Fujimura, M. Takeda, H. Ikai, K. Haruma, T. Akisada, T. Harada, T. Sakai, M. Ohuchi. The<br />

role of M cells of human n<strong>as</strong>opharyngeal lymphoid tissue in influenza virus sampling. Virch<br />

Archiv 444:36-42 (2004).<br />

22. G.N. Rogers, J.C. Paulson. Receptor determinants of human and animal influenza virus isolates:<br />

Differences in receptor specificity of the H3 hemagglutinin b<strong>as</strong>ed on species of origin. Virology<br />

127:361-373 (1983).<br />

23. B. Slütter, P.C. Soema, Z. Ding, R. Verheul, W. Hennink, and W. Jiskoot. Conjugation of<br />

ovalbumin to trimethyl chitosan improves immunogenicity of the antigen. J Control Rele<strong>as</strong>e 143<br />

:207-214 (2010).<br />

24. B. Slütter, S. Bal, C. Keijzer, R. Mallants, N. Hagenaars, I. Que, E. Kaijzel, W. van Eden, P.<br />

Augustijns, C. Löwik, J. Bouwstra, F. Broere, W. Jiskoot. N<strong>as</strong>al vaccination with N-trimethyl<br />

chitosan and PLGA b<strong>as</strong>ed nanoparticles: Nanoparticle characteristics determine quality and<br />

strength of the antibody response in mice against the encapsulated antigen. Vaccine 28:6282-<br />

6291 (2010).<br />

25. S. Bakowsky, A. Jintapattanakit, T. Kissel. Self-<strong>as</strong>sembled polyelectrolyte nanocomplexes<br />

between chitosan derivatives and insulin. J Pharm Sci 95:1035-1043 (2006).<br />

26. S.D. Xiang, A. Scholzen, G. Minigo, C. David, V. Apostolopoulos, P.L. Mottram, M. Plebanski.<br />

Pathogen recognition and development of particulate vaccines: Does size matter?. Methods 40:1-<br />

9 (2006).<br />

123


CHAPTER 6<br />

TAILORABLE THIOLATED TRIMETHYL<br />

CHITOSANS FOR COVALENTLY STABILIZED<br />

NANOPAR<strong>TI</strong>CLES<br />

Rolf J. Verheul, Steffen van der Wal, Wim E. Hennink.<br />

Biomacromolecules 2010, 11, 1965-1971


Chapter 6<br />

Abstract<br />

A novel four-step method is presented to synthesize partially thiolated trimethylated<br />

chitosan (TMC) with tailorable degree of quaternization and thiolation. First, chitosan w<strong>as</strong><br />

partially N-carboxylated with glyoxilic acid and sodium borohydride. Next, the remaining<br />

amines were quantitatively dimethylated with <strong>for</strong>maldehyde and sodium borohydride and<br />

then quaternized with iodomethane in NMP. Subsequently, these partially carboxylated TMCs<br />

dissolved in water were reacted with cystamine at pH 5.5 using EDC <strong>as</strong> coupling agent. After<br />

addition of DTT and dialysis, thiolated TMCs were obtained varying in degree of quaternization<br />

(25-54%) and degree of thiolation (5-7%) <strong>as</strong> determined with 1 H-NMR and Ellman’s <strong>as</strong>say. Gel<br />

permeation chromatography with light scattering detection indicated limited intermolecular<br />

crosslinking. All thiolated TMCs showed rapid oxidation to yield disulfide crosslinked TMC at<br />

pH 7.4 while the thiolated polymers were rather stable at pH 4.0. Using Calu-3 cells, XTT and<br />

LDH cell viability tests showed a slight reduction in cytotoxicity <strong>for</strong> thiolated TMCs <strong>as</strong><br />

compared to the non-thiolated polymers with similar DQs. Positively charged nanoparticles<br />

loaded with fluorescently labeled ovalbumin were made from thiolated TMCs and thiolated<br />

hyaluronic acid. The stability of these particles w<strong>as</strong> confirmed in 0.8 M NaCl, in contr<strong>as</strong>t to<br />

particles made from non-thiolated polymers which dissociated under these conditions<br />

demonstrating that the particles were held together by intermolecular disulfide bonds.<br />

Chitosan<br />

+<br />

Thiolated TMC<br />

N(CH 3 ) 3<br />

HS<br />

Thiolated<br />

Hyaluronic<br />

acid<br />

+<br />

―SH<br />

+<br />

+<br />

―SH<br />

+<br />

-<br />

-<br />

--<br />

+<br />

―SH<br />

―SH<br />

+<br />

+<br />

―SH<br />

Covalently stabilized<br />

nanoparticles<br />

126


Thiolated TMC <strong>for</strong> Covalently Stabilized Nanoparticles<br />

Introduction<br />

Chitosan, a polysaccharide consisting of β1→4-D-glucosamine and β1→4 N-acetyl D-<br />

glucosamine units, is under investigation <strong>for</strong> various biomedical and pharmaceutical<br />

applications [1-3]. However, its poor aqueous solubility and loss of penetration enhancing<br />

activity above pH 6 is a major drawback <strong>for</strong> its use at physiological conditions. N,N,N,-<br />

trimethylated chitosan (TMC), a partially quaternized derivative of chitosan, is water-soluble<br />

at physiological pH. TMC h<strong>as</strong> been widely studied in the biomedical field <strong>as</strong> drug, antigen and<br />

gene delivery vehicle. It h<strong>as</strong> been shown in several in vitro and in vivo studies that TMC h<strong>as</strong> low<br />

toxicity, possesses muco-adhesive properties and can facilitate the uptake of small drug<br />

molecules <strong>as</strong> well <strong>as</strong> proteins via various mucosal routes [4-11] Several studies have suggested<br />

an optimal degree of quaternization (DQ) of 40-50% <strong>for</strong> mucosal transport and gene delivery<br />

[12-16]. In addition, molecular weight [17, 18] and, <strong>as</strong> recently demonstrated, O-methylation<br />

[19] and degree of N-acetylation [20] also have a major impact on the physical and biological<br />

characteristics of TMC.<br />

Introduction of thiol-moieties will further broaden the potential pharmaceutical applications<br />

of TMC by enhancing its muco-adhesive potential and allowing further chemical derivatization<br />

reactions via reducible disulfide-bridges [11, 21-24]. Yin et al. [25] synthesized thiolated TMC<br />

by coupling cysteine to the remaining free –NH 2 units of the polymer. However, this method is<br />

dependent on available free amines remaining after quaternization, and there<strong>for</strong>e this<br />

synthetic route is only applicable <strong>for</strong> TMC with a low DQ (up to 30 %).<br />

Nanoparticulate systems have superior penetration enhancing characteristics in mucosal<br />

protein and/or antigen delivery over polymer solutions [1, 26, 27]. Although often<br />

tripolyphosphate (TPP) is used <strong>as</strong> a crosslinker to <strong>for</strong>m TMC nanoparticles via ionic gelation,<br />

recent results by Sayın et al. [28] showed that n<strong>as</strong>al immunization with tetanus toxoid loaded<br />

TMC:mono-N-carboxymethyl chitosan nanoparicles resulted in superior antibody titers<br />

compared to the TMC/TPP particles. The physico-chemical stability of these complexes is<br />

dependent on the characteristics of the polyelectrolytes used and they often have a limited<br />

stability in physiological salt conditions [29, 30] or at low pH [21]. Thiolated chitosans, and<br />

more recently, thiolated TMCs complexed with insulin or using TPP <strong>as</strong> crosslinker resulted in<br />

nanoparticles stabilized via intracellularly degradable disulfide bridges [21, 25, 31]. Combining<br />

the cationic, thiolated TMC polymer with the muco-adhesive [22], anionic, thiolated hyaluronic<br />

acid to <strong>for</strong>m stabilized nanoparticles may further improve these mucosal delivery vehicles.<br />

Further, it can be anticipated that the charge of the particles can be tuned by the<br />

127


Chapter 6<br />

TMC:hyaluronic acid ratio and that both anionically and cationically charged proteins can be<br />

loaded.<br />

In this paper, a novel synthetic method is described to yield thiolated TMCs with a high DQ.<br />

The different thiolated TMCs were physico-chemically characterized, evaluated in in vitro<br />

cytotoxicity <strong>as</strong>says and used <strong>for</strong> preparation of covalently stabilized nanoparticles with<br />

thiolated hyaluronic acid.<br />

Materials and Methods<br />

Materials. Chitosan with a residual degree of acetylation of 17% (determined with 1 H-NMR<br />

<strong>as</strong> described below) and a number average (M n) and weight average molecular weight (M w) of<br />

28 and 43 kDa, (determined with GPC-TD <strong>as</strong> below), respectively, w<strong>as</strong> purch<strong>as</strong>ed from Primex<br />

(Siglufjodur, Iceland). Sodium borohydride, <strong>for</strong>maldehyde 37% (stabilized with methanol),<br />

glyoxilic acid monohydrate, cystamine dihydrochloride, dithiotreitol (DTT), 1-ethyl-3-(3-<br />

dimethylaminopropyl) carbodiimide HCl (EDC), L-cysteine HCl monohydrate, deuterium oxide,<br />

sodium 3’-[1-(phenylaminocarbonyl)-3,4-tetrazolium]-bis(4-methoxy-6-nitro) benzene<br />

sulphonic acid hydrate (XTT), N-methyl dibenzopyrazine methylsulphate (PMS), sodium<br />

acetate, acetic acid (anhydrous), sodium hydroxide and hydrochloric acid were obtained from<br />

Sigma-Aldrich Chemical Co. Fluorescently labeled ovalbumin (OVA-FITC), minimal essential<br />

medium (MEM) and fetal calf serum (FCS) were obtained from Invitrogen (Breda, The<br />

Netherlands). Sicapent w<strong>as</strong> ordered from Merck (Darmstadt, Germany). Iodomethane 99%<br />

stabilized with copper w<strong>as</strong> obtained from Acros Organics (Geel, Belgium). 5,5-dithio-bis-(2-<br />

nitrobenzoic acid) (Ellman’s reagent) w<strong>as</strong> purch<strong>as</strong>ed from Pierce (Rock<strong>for</strong>d, IL, USA). Linear<br />

polyethylenimine (PEI) (22kDa) w<strong>as</strong> synthesized according to Thom<strong>as</strong> et al [32]. TMCs with<br />

DQs of 30% (M n 33 kDa, M w 59 kDa) and 56% (M n 37 kDa, M w 78 kDa) were synthesized and<br />

characterized <strong>as</strong> described previously [19]. HA-SH with a M n of 39.8 kDa, PDI of 1.26 and a<br />

degree of thiolation of 52% w<strong>as</strong> synthesized <strong>as</strong> described elsewhere [33]. All other chemicals<br />

used were of analytical grade.<br />

Synthesis of Partially N-Carboxylated Chitosan. Chitosan (degree of acetylation 17%) w<strong>as</strong><br />

used <strong>as</strong> obtained and selective, partial N-carboxylation w<strong>as</strong> carried out according to a<br />

previously described method with some adjustments [10, 13]. Briefly, chitosan (10 g) w<strong>as</strong><br />

dissolved in 1% (v/v) acetic acid (300 ml). Then, 2.3 g glyoxylic acid w<strong>as</strong> added and pH w<strong>as</strong><br />

raised to 4.5 with 1 M NaOH. Subsequently, 3 g of sodium borohydride dissolved in H 2O (5%<br />

128


Thiolated TMC <strong>for</strong> Covalently Stabilized Nanoparticles<br />

w/v) w<strong>as</strong> added drop wise over a period of 3 hours in portions of 500 mg re-adjusting the pH<br />

with 1 M HCl to 4.5 after each addition. Finally, the partially carboxylated chitosan (CS-COOH)<br />

w<strong>as</strong> precipitated by dropping the reaction mixture into an ethanol/diethyl ether mixture<br />

(50/50 v/v) and the product w<strong>as</strong> extensively w<strong>as</strong>hed with diethyl ether and dried overnight.<br />

Synthesis of Dimethylated, Partially N-Carboxylated Chitosan. N,N,-Dimethylated,<br />

partially N-carboxylated chitosan (DMC-COOH) w<strong>as</strong> synthesized from CS-COOH <strong>as</strong> described<br />

previously <strong>for</strong> the synthesis of N,N,-dimethylated chitosan [20, 34]. In short, CS-COOH (10 g)<br />

w<strong>as</strong> dissolved in 500 ml of 1% acetic acid (v/v). Next, 20 ml of 37% <strong>for</strong>maldehyde solution w<strong>as</strong><br />

added, the pH w<strong>as</strong> adjusted to 4.5 with 1M NaOH and the mixture w<strong>as</strong> stirred <strong>for</strong> 30 minutes at<br />

room temperature. Subsequently, sodium borohydride (5 g) w<strong>as</strong> added in portions of 500 mg<br />

over a period of 5 hours, adjusting the pH to 4.5 with 1M HCl after each addition. After<br />

precipitation with acetone, the product w<strong>as</strong> w<strong>as</strong>hed extensively with acetone and dried<br />

overnight. To obtain quantitative dimethylation a second methylation step w<strong>as</strong> per<strong>for</strong>med<br />

using the procedure described above.<br />

Synthesis of O-Methyl free, <strong>Trimethyl</strong>ated, Partially N-Carboxylated Chitosan. DMC-<br />

COOH w<strong>as</strong> reacted with iodomethane to yield trimethylated, partially N-carboxylated chitosan<br />

(TMC-COOH). To prevent O-methylation, the reaction of DMC with iodomethane w<strong>as</strong> done in<br />

NMP, without the addition of a b<strong>as</strong>e catalyst. TMC-COOH with a degree of quaternization of<br />

around 25% or 54% were obtained by the method described previously [19]. In detail, DMC-<br />

COOH (2.5 g) w<strong>as</strong> dissolved in 80 ml deionized water and the pH w<strong>as</strong> adjusted to 11 with a 1 M<br />

solution of NaOH, resulting in gel <strong>for</strong>mation. Then, the gel w<strong>as</strong> w<strong>as</strong>hed with water and<br />

subsequently with acetone. To remove residual solvents, the DMC w<strong>as</strong> dried under vacuum <strong>for</strong><br />

2 hours. Next, DMC-COOH w<strong>as</strong> suspended in 100 ml NMP followed by the addition of 5 or 15<br />

ml iodomethane. The dispersion w<strong>as</strong> stirred at 40°C <strong>for</strong> 40 hours and subsequently dropped<br />

into 400 ml of an ethanol/diethyl ether mixture (50/50 v/v) to precipitate the <strong>for</strong>med TMC-<br />

COOH, which w<strong>as</strong> collected by centrifugation and subsequently extensively w<strong>as</strong>hed with<br />

diethyl ether. After drying overnight at room temperature, the obtained TMC-COOH w<strong>as</strong><br />

dissolved in 100 ml of an aqueous 10% NaCl solution <strong>for</strong> ion-exchange. Finally, the TMC-COOH<br />

w<strong>as</strong> dialyzed against 1% NaCl in diluted HCl (pH 4) <strong>for</strong> 2 days followed by dialysis against<br />

diluted HCl (pH 4) <strong>for</strong> another 2 days (changing buffer twice daily). The polymer solution w<strong>as</strong><br />

filtered through a 0.8 µm filter and collected after freeze-drying.<br />

129


Chapter 6<br />

Synthesis of O-Methyl Free, <strong>Trimethyl</strong>ated, Partially Thiolated Chitosan. Thiolated TMC<br />

w<strong>as</strong> synthesized by first coupling cystamine to the carboxylic acid moieties of TMC-COOH<br />

using EDC <strong>as</strong> coupling agent followed by reduction of the S-S bonds of the coupled cystamine<br />

groups by dithiothreitol. In detail, TMC-COOH (1 g) w<strong>as</strong> dissolved in deionized water at a<br />

concentration of 100 mg/ml. Then, five hundred milligram of cystamine dihydrochloride w<strong>as</strong><br />

added (molar ratio COOH:cystamine approx. 1:4). After complete dissolution, EDC w<strong>as</strong> added<br />

to the reaction mixture (final concentration of 200 mM) to activate the carboxylic acid groups<br />

of the polymer and the pH w<strong>as</strong> adjusted to 5.5 with 1 M NaOH. After reacting <strong>for</strong> 6 hours at<br />

room temperature on a roller bank, pH w<strong>as</strong> raised to 8 with 1 M NaOH and DTT w<strong>as</strong> added in<br />

final concentration of 100 mM. After approximately 2 hours, NaCl w<strong>as</strong> added to the reaction<br />

mixture (final concentration 1% (g/v)), and the pH w<strong>as</strong> adjusted to 4 with 1M HCl. The<br />

resulting solution w<strong>as</strong> dialyzed against 1% NaCl in diluted HCl (pH 4) <strong>for</strong> 2 days changing<br />

buffer twice daily followed by dialysis against diluted HCl (pH 4) <strong>for</strong> another 2 days (changing<br />

buffer twice daily). Dialysis w<strong>as</strong> per<strong>for</strong>med at 4°C. Finally, the solution w<strong>as</strong> filtered through a<br />

0.8 µm filter and the polymers were collected after freeze-drying. To improve the reduction of<br />

remaining disulfide bonds, TMC-SH obtained after this first cycle w<strong>as</strong> subjected to a second<br />

reduction cycle with DTT <strong>as</strong> described above.<br />

Determination of the Degrees of Carboxylation, Dimethylation, Acetylation and<br />

Quaternization. The 1 H-NMR spectra of the various chitosan derivatives were recorded with<br />

a Varian INOVA 300MHz NMR spectrometer (Varian Inc., Palo Alto, Ca, USA) at 25°C in D 2O.<br />

The degree of carboxylation of CS-COOH w<strong>as</strong> calculation <strong>as</strong> follows:<br />

D Carb = [[CH 2]/[H2-H6] x 6/2] x 100%<br />

Here, [CH 2] is the integral of the two hydrogens of the N-carboxymethyl groups at 3.2 ppm [35,<br />

36] and [H2-H6] is the integral corresponding the six protons bound to the C-2 to C-6 ring<br />

carbons between 3.9 and 3.0 ppm (without the signal observed at 3.2 ppm).<br />

The degree of dimethylation (DDM) of the DMC-COOH w<strong>as</strong> calculated <strong>as</strong> follows [19]:<br />

DDM = [(CH 3) 2]/[H2-H6] x 100%<br />

130


Thiolated TMC <strong>for</strong> Covalently Stabilized Nanoparticles<br />

Here, [(CH 3) 2] is the integral of the hydrogens of the dimethyl amino groups at 2.9 ppm and<br />

[H2-H6] is the integral corresponding to the six protons bound to the C-2 to C-6 ring carbons<br />

between 3.9 and 3.0 ppm (without the signal observed at 3.2 ppm).<br />

The 1 H-NMR spectra of the carboxylated and thiolated TMCs were recorded with a Varian<br />

INOVA 500MHz NMR spectrometer (Varian Inc., Palo Alto, Ca, USA) at 80°C in D 2O. The DQ and<br />

degree of dimethylation (DDM) of the TMC-COOHs and TMC-SHs were calculated <strong>as</strong> previously<br />

described [4, 37, 38]:<br />

DQ = [[(CH 3) 3]/[H] × 1/9] × 100%<br />

DDM = [[(CH 3) 2]/[H] × 1/6] × 100%<br />

Here, [(CH 3) 3] and [(CH 3) 2] are the integrals of the hydrogens of the trimethylated amino<br />

groups at 3.3 ppm and the dimethylated amino groups at 2.9 ppm, respectively. [H] is the<br />

integral of the H-1 peaks between 4.7 and 5.7 ppm; the signal related to the hydrogen atoms<br />

bound to the C-1 ring carbon of TMC.<br />

The degree of acetylation (DAc) of the chitosan derivatives w<strong>as</strong> calculated <strong>as</strong> described<br />

previously [39]:<br />

DAc = [[CH 3]/[H] x 1/3] x 100<br />

Here, [CH 3] is the integral of the three hydrogens of the acetyl groups at 2.0 ppm and [H] is the<br />

integral of the H-1 peaks between 4.7 and 5.7 ppm; the signal related to the hydrogen atoms<br />

bound to the C-1 ring carbon of TMC.<br />

Determination of Free Thiol Group Content. The degree of thiolation w<strong>as</strong> quantified with<br />

5,5-dithio-bis-(2-nitrobenzoic acid) (DTNB, Ellman’s reagent). DTNB w<strong>as</strong> dissolved in 0.1 M<br />

sodium phosphate, pH 8.0 containing 1 mM EDTA in a final concentration of 100 μg/ml. TMC-<br />

SHs were dissolved in 0.1 M sodium phosphate, pH 8.0 containing 1 mM EDTA in a<br />

concentration of 2.5 mg/ml. Then, 20 μl of this solution w<strong>as</strong> mixed with 180 μl of DTNB<br />

reagent solution in a microplate and after incubating <strong>for</strong> 15 minutes at room temperature the<br />

absorbance w<strong>as</strong> me<strong>as</strong>ured with a microplate reader at a wavelength of 405 nm (Bio-Rad<br />

Novapath, Hercules, Ca, USA). Cysteine standards were used to calculate the amount of free<br />

thiol moieties in the polymer. The degree of thiolation (D thiol) w<strong>as</strong> calculated using the<br />

estimated average molecular weight per glucosamine unit.<br />

131


Chapter 6<br />

Determination of Total Thiol Content. The total thiol content (free thiol moieties and<br />

disulfides) on the polymers w<strong>as</strong> determined according to a slightly modified method described<br />

by Hombach et al. [40]. In detail, TMC-SHs were dissolved in 50 mM phosphate buffer pH 7.0 (1<br />

mg/ml). Subsequently, 400 μl of a freshly prepared 10% (w/v) sodium borohydride solution in<br />

deionized water w<strong>as</strong> added to 500 μl of polymer solution and the mixture w<strong>as</strong> slightly shaken<br />

at 37°C <strong>for</strong> 60 minutes. Thereafter, 1 ml of 1 M HCl w<strong>as</strong> added to decompose the remaining<br />

sodium borohydride and after 10 minutes the pH w<strong>as</strong> raised to 8 by adding 1 ml of 0.5 M<br />

phosphate buffer pH 8.0. Then, 100 μl of DTNB reagent solution (4 mg/ml in 0.5 M sodium<br />

phosphate, pH 8.0) w<strong>as</strong> added and after incubating <strong>for</strong> 15 minutes at room temperature, a 200<br />

μl sample w<strong>as</strong> transferred into a microplate. The absorbance w<strong>as</strong> me<strong>as</strong>ured with a microplate<br />

reader at a wavelength of 405 nm (Bio-Rad Novapath, Hercules, Ca, USA). Cysteine standards<br />

prepared in the same way <strong>as</strong> the samples were used to calculate the total amount of thiol<br />

moieties (<strong>as</strong> free thiols or disulfides) on the polymer.<br />

Thiol Oxidation. TMC-SH w<strong>as</strong> dissolved in 100 mM acetate buffer pH 4.0, 100 mM acetate<br />

buffer pH 5.0, 100 mM phosphate buffer pH 6.2, or 100 mM phosphate buffer pH 7.4, in a final<br />

concentration of 5 mg/ml. Samples were incubated at 37°C under permanent shaking and<br />

aliquots were taken at different time points. Immediately after collecting the aliquots, the<br />

amount of remaining thiol moieties w<strong>as</strong> determined with Ellman’s reagent <strong>as</strong> described be<strong>for</strong>e.<br />

Determination of M n and M w of the Different Polymers. Polymers were placed overnight<br />

in a vacuum oven at 40 o C in the presence of Sicapent to remove residual water and<br />

subsequently dissolved in 0.3 M sodium acetate pH 4.4. The number average weight (M n) and<br />

weight average weight (M w) of chitosan and the various TMCs were determined by gel<br />

permeation chromatography (GPC) on a Viscotek-triple detection system using a Shodex<br />

OHPak SB-806 column (30 cm) and 0.3 M sodium acetate pH 4.4 (adjusted with acetic acid) <strong>as</strong><br />

running buffer [41]. Data from the l<strong>as</strong>er photometer (λ = 670 nm) (right (90 0 ) and low (7 0 )<br />

angle light scattering), refractive index detector and viscosity detector were integrated using<br />

the provided Omnisec-software to calculate the M n, M w of the different samples. Pullulan (M n =<br />

102 kDa, M w = 106 kDa) obtained from Viscotek Benelux (Oss, the Netherlands) w<strong>as</strong> used <strong>for</strong><br />

calibration.<br />

XTT Cytotoxicity Assay. Calu-3 cells (ATCC, Teddington, UK) were seeded into a 96-well<br />

plate at a density of 2x10 4 cells per well and incubated <strong>for</strong> 2 days at 37°C, CO 2 5% in culture<br />

132


Thiolated TMC <strong>for</strong> Covalently Stabilized Nanoparticles<br />

medium (MEM with L-glutamine and sodium pyruvate supplemented with<br />

antibiotics/antimycotics and 10% FCS). The medium w<strong>as</strong> removed and the cells were<br />

incubated <strong>for</strong> 2.5 hours with 100 µl TMC solutions in MEM (TMC concentrations were 0.01, 0.1,<br />

1 and 10 mg/ml, pH set at 7 with 0.1 M NaOH). Linear PEI (0.1 and 0.01 mg/ml) w<strong>as</strong> used <strong>as</strong><br />

positive control and cells incubated with MEM were used <strong>as</strong> reference. Thereafter, the<br />

solutions were removed and the cells w<strong>as</strong>hed MEM. Then, 100 µl of MEM containing 10% FCS<br />

w<strong>as</strong> added to the cells together with 50 µl of a freshly prepared solution of 1 mg/ml XTT in<br />

MEM containing 25 µmol PMS <strong>as</strong> electron-coupling reagent. Cells were incubated <strong>for</strong> 1 hour at<br />

37°C in 5% CO 2 and the absorbance w<strong>as</strong> read at 490 nm with 655 nm <strong>as</strong> reference wavelength.<br />

Obtained values of the samples were related to the mitochondrial activity of Calu-3 cells<br />

incubated with MEM only [42].<br />

LDH Cytotoxicity Assay. Calu-3 cells were seeded into a 96-well plate at a density of 2x10 4<br />

cells per well and incubated <strong>for</strong> 2 days at 37°C, CO 2 5% in culture medium (MEM with L-<br />

glutamine and sodium pyruvate supplemented with antibiotics/antimycotics and 10% FCS).<br />

The medium w<strong>as</strong> removed and the cells were w<strong>as</strong>hed with MEM and incubated <strong>for</strong> 2.5 hours<br />

with 100 µl TMC solutions in MEM (TMC concentrations were 0.01, 0.1, 1 and 10 mg/ml, pH set<br />

at 7 with 0.1 M NaOH). After incubation, the concentration of lactate dehydrogen<strong>as</strong>e (LDH)<br />

present in the supernatant of the samples w<strong>as</strong> determined with the Cytotoxicity Detection Kit-<br />

Plus (Roche Diagnostics, Mannheim, Germany) by me<strong>as</strong>uring absorbance at 490 nm with 655<br />

nm <strong>as</strong> a reference wavelength. A calibration curve w<strong>as</strong> made with the lysis buffer provided by<br />

the manufacturer, setting the LDH concentration me<strong>as</strong>ured with the undiluted lysis buffer at<br />

100% LDH rele<strong>as</strong>e. Cells incubated with linear PEI (0.1 and 0.01 mg /ml) were used <strong>as</strong> control.<br />

One-way ANOVA with Tukey’s post-test w<strong>as</strong> used <strong>for</strong> comparison.<br />

Preparation and Characterization of Covalently Stabilized Nanoparticles.<br />

Covalently stabilized nanoparticles loaded with fluorescently labeled ovalbumin were<br />

prepared with TMC-SH and thiolated hyaluronic acid (HA-SH). TMC-SH with a DQ 25% and<br />

D thiol 5% or a DQ of 54% and D thiol 6% were dissolved in 10 mM HEPES pH 7.4 at 1 mg/ml and<br />

subsequently mixed with 2.5 mg/ml fluorescently labeled ovalbumin (OVA-FITC) in 10 mM<br />

HEPES pH 7.4 at a weight ratio of 10:1 (TMC-SH:OVA-FITC) under magnetic stirring. Then, HA-<br />

SH (0.5 mg/ml in 10 mM HEPES pH 7.4) w<strong>as</strong> added drop-wise yielding an opalescent<br />

nanoparticle dispersion which w<strong>as</strong> incubated at 37°C <strong>for</strong> 3 hours to allow disulfide <strong>for</strong>mation.<br />

As a control nanoparticles with non-thiolated TMC (DQ 30% or 56%), OVA-FITC and HA were<br />

133


Chapter 6<br />

prepared in a similar way. After 3 hours of incubation, the nanoparticle dispersions were<br />

mixed with either 10 mM HEPES pH 7.4 or 0.8 M NaCl in 10 mM HEPES pH 7.4 and<br />

subsequently incubated at 37°C <strong>for</strong> 1 hour.<br />

Particle size w<strong>as</strong> me<strong>as</strong>ured by dynamic light scattering (DLS) using a Malvern ALV CGS-3<br />

(Malvern Instruments, Malvern, UK). DLS results are given <strong>as</strong> a z-average particle size<br />

diameter and a polydispersity index (PDI). The zeta-potential of the nanoparticles w<strong>as</strong><br />

me<strong>as</strong>ured in 10 mM HEPES pH 7.4 using a Zet<strong>as</strong>izer Nano (Malvern Instruments, Malvern, UK).<br />

The protein loading w<strong>as</strong> determined by <strong>as</strong>saying the concentration of OVA-FITC in the<br />

supernatant obtained after centrifugation (10 min at 13000 rpm; Biofuge Pico, PP1/97 #3324;<br />

Heraeus Instruments, Osterode, Germany) of the nanoparticle dispersion using a Fluostar<br />

Optima with 488 nm <strong>as</strong> excitation wavelength and emission w<strong>as</strong> me<strong>as</strong>ured at 520 nm (BMG<br />

Labtech, Offenburg, Germany). The OVA-FITC <strong>as</strong>sociation efficiency (AE%) w<strong>as</strong> calculated<br />

using the following equation:<br />

AE% = (total OVA-FITC ― OVA-FITC in supernatant)/ total OVA-FITC x 100%<br />

134


Thiolated TMC <strong>for</strong> Covalently Stabilized Nanoparticles<br />

Scheme 1. Synthetic route <strong>for</strong> thiolated TMCs.<br />

135


Chapter 6<br />

Results and discussion<br />

Synthesis and Characterization of Thiolated TMC. Thiolated TMCs with tailorable degree<br />

of quaternization (DQ) and thiolation (D thiol) were obtained <strong>as</strong> depicted in Scheme 1. First,<br />

chitosan w<strong>as</strong> selectively N-carboxylated with glyoxylic acid and sodium borohydride to a<br />

degree of carboxylation of 12% <strong>as</strong> determined with 1 H-NMR. Higher degrees of carboxylation<br />

could be achieved by using larger quantities of glyoxylic acid (results not shown), however, a<br />

further incre<strong>as</strong>e in carboxylic acid moieties that ultimately leads to a higher degree of<br />

thiolation w<strong>as</strong> not considered useful <strong>for</strong> the <strong>for</strong>eseen applications; about 10-20 thiol groups<br />

per polymer chain is sufficient <strong>for</strong> intermolecular disulfide-bridge <strong>for</strong>mation. Partially<br />

carboxylated chitosan w<strong>as</strong> subsequently quantitatively dimethylated using <strong>for</strong>maldehyde and<br />

sodium borohydride, <strong>as</strong> confirmed with 1H-NMR. Reaction of dimethylated, partially<br />

carboxylated chitosan (DMC-COOH) with different amounts of iodomethane (5 or 15 ml) in<br />

NMP resulted in TMC-COOH with DQs of 25% and 54%, respectively, indicating tailorability of<br />

the DQ (Table 1). 1 H-NMR analysis also demonstrated that neither O-methylation nor loss of<br />

residual N-acetylated groups w<strong>as</strong> observed after quaternization, in contr<strong>as</strong>t to the synthesis<br />

method applied by others using a strong alkaline component to induce quaternization [38].<br />

TMC-COOH w<strong>as</strong> not soluble at pH 7.4 confirming the zwitterionic character of this polymer.<br />

Table 1. Characteristics of chitosan and the synthesized thiolated TMCs.<br />

DQ DAc<br />

Total thiol<br />

amount<br />

Amount of<br />

free –SH D thiol M n (kDa) M w (kDa)<br />

(μmol/g) (μmol/g)<br />

Chitosan - 17% - - - 28 43<br />

TMC-SH<br />

High DQ<br />

54% 17% 529 (±2) 283 (±21) 6% 66 174<br />

TMC-SH<br />

Low DQ<br />

25% 17% 478 (±10) 236 (±13) 5% 62 215<br />

TMC-SH<br />

Low DQ<br />

2 nd cycle<br />

25% 17% 478 (±10) 342 (±4) 7% 60 144<br />

The carboxylic acid groups of TMC-COOH were reacted with cystamine using EDC <strong>as</strong><br />

coupling agent. In agreement with previous studies [43] we found that this reaction proceeded<br />

most efficiently at pH 5.5. An excess of cystamine w<strong>as</strong> used to avoid the <strong>for</strong>mation of<br />

crosslinked products. Incubation of cystamine with TMC-COOH without EDC did not result in<br />

polymers with detectable thiol groups. Ellman’s <strong>as</strong>say showed that significant thiolation of<br />

TMC w<strong>as</strong> achieved with 200 mM EDC (Table 1). This table shows that about 529 μmol/g thiol-<br />

136


Thiolated TMC <strong>for</strong> Covalently Stabilized Nanoparticles<br />

moieties (<strong>as</strong> free thiols or disulfides) were introduced into TMC-COOH DQ 54% corresponding<br />

with approx. 12% of glucosamine-units and indicates quantitative substitution of the<br />

carboxylic acids with cystamine. Treatment with DTT resulted in 283 μmol/g free –SH groups<br />

in TMC-SH DQ 54% corresponding to a degree of thiolation (D thiol) of ~6%. For TMC-COOH<br />

with a DQ of 25% the conversion w<strong>as</strong> slightly less efficient (478 μmol/g thiol moieties <strong>as</strong> free<br />

thiols or disulfide groups, approx. 10% of glucosamine-units) and reduction with DTT resulted<br />

in 236 μmol/g free –SH groups (approx. D thiol of 5%). D thiol incre<strong>as</strong>ed up to 341 μmol/g polymer<br />

(about 7% degree of thiolation) after a second reduction cycle with DTT. GPC analysis<br />

demonstrated an incre<strong>as</strong>e in both M n and M w with incre<strong>as</strong>ing the substitution degree of the<br />

TMC-SHs. This suggests that some intermolecular crosslinking had occurred. A second<br />

reduction cycle with DTT resulted in a decre<strong>as</strong>e of M w (215 to 144 kDa) compared to TMC-SH<br />

DQ 25% with only one DTT treatment, indicating that the intermolecular crosslinks were<br />

partly broken.<br />

Remaining thiol content (%)<br />

A<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

0 1 2 3 4 5 6<br />

time (h)<br />

pH 4.0<br />

pH 5.0<br />

pH 6.2<br />

pH 7.4<br />

Remaining thiol content (%)<br />

B<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

0 1 2 3 4 5 6<br />

time (h)<br />

TMC-SH DQ 25% D thiol 5%<br />

TMC-SH DQ 54% D thiol 6%<br />

TMC-SH DQ 25% D thiol 7%<br />

Figure 1.Thiol group content of TMC-SH DQ 25%, D thiol 5% in a concentration of 5 mg/ml, at pH 4.0, 5.0,<br />

6.2 and 7.4 at 37°C (A). Thiol group content of TMC-SH DQ 25%, D thiol 5%, TMC-SH DQ 54%, D thiol 6% and<br />

TMC-SH DQ 25%, D thiol 7% in a concentration of 5 mg/ml, at pH 7.4 at 37°C (B). Error bars indicate<br />

standard deviation of three independent samples.<br />

The pH-dependent oxidation of thiol groups of TMC-SH DQ 25%, D thiol 5%, w<strong>as</strong> studied<br />

(Figure 1A) at 37°C. This figure shows that at pH 4.0 no reduction of free thiols w<strong>as</strong> observed<br />

while at pH 7.4 almost quantitative oxidation w<strong>as</strong> achieved after 6 hours. For oxidation the<br />

thiol anion is needed and thus oxidation occurred much f<strong>as</strong>ter at higher pH. It w<strong>as</strong> further<br />

observed that an incre<strong>as</strong>e in DQ lead to a slight reduction in oxidation rate likely due to<br />

137


Chapter 6<br />

incre<strong>as</strong>ed charge repulsion [25, 43] but this had no effect on the extent of oxidation: after 6<br />

hours TMC-SH DQ 54%, D thiol 6% w<strong>as</strong> quantitatively oxidized (Figure 1B).<br />

Evaluation of TMC-SH on Cell Viability. Figure 2A shows the effect of TMC-SH on the<br />

mitochondrial activity (XTT) of Calu-3 cells. In agreement with previous data [19], an incre<strong>as</strong>e<br />

in DQ of the TMCs resulted in a decre<strong>as</strong>e in mitochondrial activity, especially at 10 mg/ml. The<br />

thiolated TMCs with a DQ of 25% were less cytotoxic at 1 mg/ml (p < 0.001, one-way ANOVA),<br />

<strong>as</strong> me<strong>as</strong>ured with the XTT <strong>as</strong>say, than the non-thiolated TMC with DQ of 30%. This might be<br />

due to the lower charge density of the TMC-SHs. Importantly, all TMCs were relatively nontoxic<br />

compared to linear PEI at concentrations < 1 mg/ml (p < 0.001). The influence of the<br />

various TMCs on membrane permeability determined with the LDH <strong>as</strong>say resulted in similar<br />

trends <strong>as</strong> obtained with the XTT <strong>as</strong>say (Figure 2B). Here, TMC-SH with a DQ of 54% showed<br />

less LDH rele<strong>as</strong>e than TMC DQ 56% at 1 mg/ml (p < 0.001). The LDH <strong>as</strong>say also indicated that<br />

linear PEI w<strong>as</strong> much more cytotoxic than the different TMCs (p < 0.001).<br />

# cells (x 10 3 )<br />

A<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

#<br />

#<br />

* *<br />

TMC-SH DQ 25% D thiol 5%<br />

TMC-SH DQ 25% D thiol 7%<br />

TMC DQ 30%<br />

TMC-SH DQ 54% D thiol<br />

6%<br />

TMC DQ 56%<br />

linear PEI<br />

MEM<br />

10 mg/ml<br />

1.0 mg/ml<br />

0.1 mg/ml<br />

0.01 mg/ml<br />

LDH rele<strong>as</strong>e (%)<br />

B<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

ο<br />

TMC-SH DQ 25% D thiol<br />

5%<br />

ο<br />

TMC-SH DQ 25% D thiol<br />

10 mg/ml<br />

1.0 mg/ml<br />

0.1 mg/ml<br />

0.01 mg/ml<br />

7%<br />

TMC DQ 30%<br />

TMC-SH DQ 54% D thiol<br />

#<br />

6%<br />

TMC DQ 56%<br />

linear PEI<br />

Figure 2. Effect of TMCs with various DQs and with or without thiolated units on the viability of Calu-3<br />

cells (XTT <strong>as</strong>say) at different concentrations (A). Asterics (*) indicate that no reliable XTT values were<br />

obtained due to gel-<strong>for</strong>mation of the TMC-SHs with a DQ of 25% at 10 mg/ml and there<strong>for</strong>e no bars are<br />

depicted. # indicate that significantly (p


Thiolated TMC <strong>for</strong> Covalently Stabilized Nanoparticles<br />

Recently, Hagenaars et al. demonstrated that n<strong>as</strong>al administration of TMC with a DQ of 68%<br />

in a concentration of 1.25 mg/ml induced only minimal local toxicity in mice while a solution of<br />

PEI resulted in moderate local toxicity [44].<br />

Covalently Stabilized Nanoparticles. Thiolated and non-thiolated TMCs were mixed with<br />

OVA-FITC and thiolated or non-thiolated hyaluronic acid in low ionic strength buffer (10 mM<br />

HEPES pH 7.4) to yield positively charged nanoparticles. Since the thiolation of hyaluronic acid<br />

leads to a reduction of negatively charged carboxylic acid moieties [33], more HA-SH than HA<br />

w<strong>as</strong> needed <strong>for</strong> ionic cross-linking to obtain particles with a size around 200 nm. Similarly,<br />

when TMC with a higher DQ w<strong>as</strong> used, larger amounts of negatively charged HA-SH or HA<br />

were needed to obtain nanoparticles (Table 2). The polydispersity index (PDI) w<strong>as</strong> below 0.2<br />

<strong>for</strong> all <strong>for</strong>mulations indicating fairly narrow size distributions. After <strong>for</strong>mulation, particles<br />

were incubated at 37°C <strong>for</strong> 3 hours while shaking to allow disulfide-bridge <strong>for</strong>mation. Table 2<br />

shows that the zeta-potential of the particles w<strong>as</strong> dependent on the DQ of the TMC: TMCs with<br />

a DQ of ~55% gave zeta-potentials of around +20 mV while TMCs with DQs of 25-30% resulted<br />

in particles with zeta-potentials of +16 mV. Thiolation of the polymers had no effect on the<br />

zeta-potential.<br />

Interestingly, when non-thiolated particles were dispersed in high ionic strength buffer (0.8<br />

M NaCl), DLS me<strong>as</strong>urements showed a substantial drop in counts indicating that nanoparticles<br />

disintegrated in this high ionic strength buffer. Similar destabilization in high ionic strength<br />

buffer w<strong>as</strong> observed <strong>for</strong> particles prepared with one thiolated polymer (e.g. TMC-SH or HA-SH)<br />

and one non-thiolated polymer (e.g. TMC or HA). Immediately after preparation, particles<br />

composed of TMC-SH and HA-SH also destabilized after dispersion in 0.8 M NaCl (results not<br />

shown). Importantly, nanoparticles made with TMC-SH and HA-SH and stabilized <strong>for</strong> 3 hours<br />

at 37°C showed almost no loss of DLS counts when incubated in this high ionic strength buffer<br />

indicating that the particles are held together by the <strong>for</strong>med covalent disulfide bonds between<br />

TMC-SH and HA-SH. All TMC:HA particles showed considerable <strong>as</strong>sociation of OVA-FITC in 10<br />

mM HEPES pH 7.4 (AE% up to 30%). The non-thiolated particles lost all their OVA-FITC in 0.8<br />

M NaCl where<strong>as</strong> the TMC-SH:HA-SH particles retained significant amounts of OVA-FITC<br />

(approx. 30-40% of originally <strong>as</strong>sociated protein) in high ionic strength buffer likely losing<br />

only surface bound protein.<br />

139


Chapter 6<br />

Table 2. Characteristics of thiolated and non-thiolated TMC-HA nanoparticles after incubation <strong>for</strong> 3<br />

hours at 37°C and dispersing in 10 mM HEPES pH 7.4 or in 0.8 M NaCl 10mM HEPES pH 7.4. Values are<br />

presented <strong>as</strong> the average of three samples ± standard deviation.<br />

TMC-SH 25%:<br />

HA-SH<br />

TMC:HA<br />

ratio<br />

(w/w)<br />

TMC 30%:HA 10:2<br />

TMC-SH 54%:<br />

HA-SH<br />

Zetapotential<br />

(mV)<br />

HEPES<br />

HEPES<br />

Size<br />

(nm)<br />

+<br />

0.8 M NaCl HEPES<br />

OVA-FITC<br />

Association Efficiency<br />

(%)<br />

+<br />

0.8 M NaCl<br />

10:2.4 + 13 (±0.8) 188 (±7) 192 (±4) 19 (±7) 8 (±2)<br />

+ 16.5<br />

(±0.3)<br />

224 (±11)<br />

Low<br />

scattering<br />

intensity<br />

22 (±9) -0.3 (±1)<br />

10:5.5 + 21 (±0.5) 191 (±3) 178 (±2) 25 (±7) 7 (±3)<br />

TMC 56%:HA 10:4 + 19 (±0.2) 197 (±3)<br />

Conclusion<br />

Low<br />

scattering<br />

intensity<br />

30 (±2) 0 (±3)<br />

In this paper a synthetic method <strong>for</strong> partial thiolation of TMC is presented resulting in D thiol<br />

up to 7% and DQs varying from 25-54% allowing investigation of thiolated TMCs with a<br />

potentially optimal DQ <strong>for</strong> mucosal vaccination and/or protein delivery or other applications<br />

such <strong>as</strong> DNA or siRNA delivery is presented. These thiolated TMCs readily <strong>for</strong>med disulfides at<br />

pH 7.4 and 37°C. Cell viability <strong>as</strong>says indicated a minor reduction in cytotoxicity on Calu-3 cells<br />

of TMC-SHs compared to non-thiolated TMCs with a similar DQ. The functionality of the thiol<br />

moieties w<strong>as</strong> confirmed by preparing covalently stabilized nanoparticles with thiolated<br />

hyaluronic acid. The opportunity to modify the remaining free thiols on the surface of the<br />

particles with e.g. PEG and targeting ligands <strong>as</strong> well <strong>as</strong> the option to prepare both negatively<br />

and positively charged particles by varying the TMC:HA ratio in the <strong>for</strong>mulation opens up wide<br />

possibilities <strong>for</strong> future pharmaceutical applications.<br />

Acknowledgement. This research w<strong>as</strong> partially per<strong>for</strong>med under the framework of <strong>TI</strong><br />

<strong>Pharma</strong> project number D5-106-1; Vaccine delivery: alternatives <strong>for</strong> conventional multiple<br />

injection vaccines. Roberta Censi is acknowledged <strong>for</strong> the synthesis and characterization of the<br />

thiolated hyaluronic acid.<br />

Supporting In<strong>for</strong>mation Available. Copies of 1 H-NMR spectra of TMC-COOH and TMC-SH<br />

<strong>for</strong> both DQs are depicted in the Supporting In<strong>for</strong>mation.<br />

140


Thiolated TMC <strong>for</strong> Covalently Stabilized Nanoparticles<br />

References<br />

1. M. Amidi, E. M<strong>as</strong>trobattista, W. Jiskoot, and W. E. Hennink. Chitosan-b<strong>as</strong>ed delivery systems<br />

<strong>for</strong> protein therapeutics and antigens. Adv Drug Deliv Rev 62: 59-82 (2010).<br />

2. B. A. Kumar, M. C. Varadaraj, and R. N. Tharanathan. Low Molecular Weight Chitosan-<br />

Preparation with the Aid of Pepsin, Characterization, and Its Bactericidal Activity.<br />

Biomacromolecules 8: 566-572 (2007).<br />

3. S. Mao, W. Sun, and T. Kissel. Chitosan-b<strong>as</strong>ed <strong>for</strong>mulations <strong>for</strong> delivery of DNA and siRNA.<br />

Adv Drug Deliv Rev 62: 12-27 (2010).<br />

4. M. Amidi, S. G. Romeijn, G. Borchard, H. E. Junginger, W. E. Hennink, and W. Jiskoot.<br />

Preparation and characterization of protein-loaded N-trimethyl chitosan nanoparticles <strong>as</strong> n<strong>as</strong>al<br />

delivery system. J Control Rele<strong>as</strong>e 111: 107-116 (2006).<br />

5. F. Chen, Z. R. Zhang, F. Yuan, X. Qin, M. Wang, and Y. Huang. In vitro and in vivo study of<br />

N-trimethyl chitosan nanoparticles <strong>for</strong> oral protein delivery. Int J Pharm 349: 226-233 (2008).<br />

6. F. Chen, Z. R. Zhang, and Y. Huang. Evaluation and modification of N-trimethyl chitosan<br />

chloride nanoparticles <strong>as</strong> protein carriers. Int J Pharm 336: 166-173 (2007).<br />

7. D. Snyman, J. H. Hamman, and A. F. Kotze. Evaluation of the mucoadhesive properties of N-<br />

trimethyl chitosan chloride. Drug Develop Ind Pharm 29: 61-69 (2003).<br />

8. M. M. Thanou, A. F. Kotze, T. Scharringhausen, H. L. Lueßen, A. G. De Boer, J. C. Verhoef,<br />

and H. E. Junginger. Effect of degree of quaternization of N-trimethyl chitosan chloride <strong>for</strong><br />

enhanced transport of hydrophilic compounds across intestinal Caco-2 cell monolayers. J<br />

Control Rele<strong>as</strong>e 64: 15-25 (2000).<br />

9. I. M. Van der Lubben, J. C. Verhoef, M. M. Fretz, O. Van, I. Mesu, G. Kersten, and H. E.<br />

Junginger. <strong>Trimethyl</strong> chitosan chloride (TMC) <strong>as</strong> a novel excipient <strong>for</strong> oral and n<strong>as</strong>al<br />

immunisation against diphtheria. S.T.P. <strong>Pharma</strong> Sciences 12: 235-242 (2002).<br />

10. B. Sayin, S. Somavarapu, X. W. Li, M. Thanou, D. Sesardic, H. O. Alpar, and S. Şenel. Mono-<br />

N-carboxymethyl chitosan (MCC) and N-trimethyl chitosan (TMC) nanoparticles <strong>for</strong> noninv<strong>as</strong>ive<br />

vaccine delivery. Int J Pharm 363: 139-148 (2008).<br />

11. B. Slütter, L. Plapied, V. Fievez, M. Alonso Sande, A. des Rieux, Y. J. Schneider, E. Van Riet,<br />

W. Jiskoot, and V. Préat. Mechanistic study of the adjuvant effect of biodegradable<br />

nanoparticles in mucosal vaccination. J Control Rele<strong>as</strong>e 138: 113-121 (2009).<br />

12. W. Boonyo, H. E. Junginger, N. Waranuch, A. Polnok, and T. Pitaksuteepong. Chitosan and<br />

trimethyl chitosan chloride (TMC) <strong>as</strong> adjuvants <strong>for</strong> inducing immune responses to ovalbumin in<br />

mice following n<strong>as</strong>al administration. J Control Rele<strong>as</strong>e 121: 168-175 (2007).<br />

13. G. Di Colo, S. Burgal<strong>as</strong>si, Y. Zambito, D. Monti, and P. Chetoni. Effects of different N-<br />

trimethyl chitosans on in vitro/in vivo ofloxacin transcorneal permeation. J Pharm Sci 93: 2851-<br />

2862 (2004).<br />

14. J. H. Hamman, M. Stander, and A. F. Kotze. Effect of the degree of quaternisation of N-<br />

trimethyl chitosan chloride on absorption enhancement: In vivo evaluation in rat n<strong>as</strong>al epithelia.<br />

Int J Pharm 232: 235-242 (2002).<br />

15. G. Sandri, M. C. Bonferoni, S. Rossi, F. Ferrari, S. Gibin, Y. Zambito, G. Di Colo, and C.<br />

Caramella. Nanoparticles b<strong>as</strong>ed on N-trimethylchitosan: Evaluation of absorption properties<br />

using in vitro (Caco-2 cells) and ex vivo (excised rat jejunum) models. Eur J Pharm Biopharm 65:<br />

68-77 (2007).<br />

16. T. Kean, S. Roth, and M. Thanou. <strong>Trimethyl</strong>ated chitosans <strong>as</strong> non-viral gene delivery vectors:<br />

Cytotoxicity and transfection efficiency. J Control Rele<strong>as</strong>e 103: 643 (2005).<br />

17. S. Mao, X. Shuai, F. Unger, M. Wittmar, X. Xie, and T. Kissel. Synthesis, characterization and<br />

cytotoxicity of poly(ethylene glycol)-graft-trimethyl chitosan block copolymers. Biomaterials 26:<br />

6343-6356 (2005).<br />

18. Z. Guo, R. Xing, S. Liu, Z. Zhong, X. Ji, L. Wang, and P. Li. The influence of molecular weight<br />

of quaternized chitosan on antifungal activity. Carbohydr Polym 71: 694-697 (2008).<br />

141


Chapter 6<br />

19. R. J. Verheul, M. Amidi, S. van der Wal, E. van Riet, W. Jiskoot, and W. E. Hennink. Synthesis,<br />

characterization and in vitro biological properties of O-methyl free N,N,N-trimethylated<br />

chitosan. Biomaterials 29: 3642-3649 (2008).<br />

20. R. J. Verheul, M. Amidi, M. J. van Steenbergen, E. van Riet, W. Jiskoot, and W. E. Hennink.<br />

Influence of the degree of acetylation on the enzymatic degradation and in vitro biological<br />

properties of trimethylated chitosans. Biomaterials 30: 3129-3135 (2009).<br />

21. A. Bernkop-Schnüch, A. Weithaler, K. Albrecht, and A. Greimel. Thiomers: Preparation and in<br />

vitro evaluation of a mucoadhesive nanoparticulate drug delivery system. Int J Pharm 317: 76-81<br />

(2006).<br />

22. V. Grabovac, D. Guggi, and A. Bernkop-Schnürch. Comparison of the mucoadhesive<br />

properties of various polymers. Adv Drug Deliv Rev 57: 1713-1723 (2005).<br />

23. T. M<strong>as</strong>uko, A. Minami, N. Iw<strong>as</strong>aki, T. Majima, S. I. Nishimura, and Y. C. Lee. Thiolation of<br />

chitosan. Attachment of proteins via thioether <strong>for</strong>mation. Biomacromolecules 6: 880-884 (2005).<br />

24. F. Meng, W. E. Hennink, and Z. Zhong. Reduction-sensitive polymers and bioconjugates <strong>for</strong><br />

biomedical applications. Biomaterials 30: 2180-2198 (2009).<br />

25. L. Yin, J. Ding, C. He, L. Cui, C. Tang, and C. Yin. Drug permeability and mucoadhesion<br />

properties of thiolated trimethyl chitosan nanoparticles in oral insulin delivery. Biomaterials 30:<br />

5691-5700 (2009).<br />

26. M. Amidi, S. G. Romeijn, J. C. Verhoef, H. E. Junginger, L. Bungener, A. Huckriede, D. J. A.<br />

Crommelin, and W. Jiskoot. N-<strong>Trimethyl</strong> chitosan (TMC) nanoparticles loaded with influenza<br />

subunit antigen <strong>for</strong> intran<strong>as</strong>al vaccination: Biological properties and immunogenicity in a mouse<br />

model. Vaccine 25: 144-153 (2007).<br />

27. S. Chadwick, C. Kriegel, and M. Amiji. Nanotechnology solutions <strong>for</strong> mucosal immunization.<br />

Adv Drug Deliv Rev 62: 394-407 (2010).<br />

28. B. Sayin, S. Somavarapu, X. W. Li, D. Sesardic, S. Şenel, and O. H. Alpar. TMC-MCC (Ntrimethyl<br />

chitosan-mono-N-carboxymethyl chitosan) nanocomplexes <strong>for</strong> mucosal delivery of<br />

vaccines. Eur J Pharm Sci 38: 362-369 (2009).<br />

29. D. V. Pergushov, H. M. Buchhammer, and K. Lunkwitz. Effect of a low-molecular-weight salt<br />

on colloidal dispersions of interpolyelectrolyte complexes. Colloid Polym Sci 277: 101-107 (1999).<br />

30. S. Boddohi, N. Moore, P. A. Johnson, and M. J. Kipper. Polysaccharide-b<strong>as</strong>ed polyelectrolyte<br />

complex nanoparticles from chitosan, heparin, and hyaluronan. Biomacromolecules 10: 1402-1409<br />

(2009).<br />

31. S. Shu, X. Wang, X. Zhang, X. Zhang, Z. Wang, and C. Li. Disulfide cross-linked biodegradable<br />

polyelectrolyte nanoparticles <strong>for</strong> the oral delivery of protein drugs. New J Chem 33: 1882-1887<br />

(2009).<br />

32. M. Thom<strong>as</strong>, J. J. Lu, Q. Ge, C. Zhang, J. Chen, and A. M. Klibanov. Full deacylation of<br />

polyethylenimine dramatically boosts its gene delivery efficiency and specificity to mouse lung.<br />

PNAS 102: 5679-5684 (2005).<br />

33. X. Z. Shu, Y. Liu, Y. Luo, M. C. Roberts, and G. D. Prestwich. Disulfide cross-linked<br />

hyaluronan hydrogels. Biomacromolecules 3: 1304-1311 (2002).<br />

34. Z. Guo, R. Xing, S. Liu, Z. Zhong, X. Ji, L. Wang, and P. Li. Antifungal properties of Schiff<br />

b<strong>as</strong>es of chitosan, N-substituted chitosan and quaternized chitosan. Carbohydr Res 342: 1329-<br />

1332 (2007).<br />

35. X. G. Chen and H. J. Park. Chemical characteristics of O-carboxymethyl chitosans related to<br />

the preparation conditions. Carbohydr Polym 53: 355-359 (2003).<br />

36. G. Lu, L. Kong, B. Sheng, G. Wang, Y. Gong, and X. Zhang. Degradation of covalently crosslinked<br />

carboxymethyl chitosan and its potential application <strong>for</strong> peripheral nerve regeneration.<br />

Eur Polym J 43: 3807-3818 (2007).<br />

37. A. Polnok, G. Borchard, J. C. Verhoef, N. Sarisuta, and H. E. Junginger. Influence of<br />

methylation process on the degree of quaternization of N-trimethyl chitosan chloride. Eur J<br />

Pharm Biopharms 57: 77-83 (2004).<br />

142


Thiolated TMC <strong>for</strong> Covalently Stabilized Nanoparticles<br />

38. A. B. Sieval, M. Thanou, A. F. Kotze, J. C. Verhoef, J. Brussee, and H. E. Junginger.<br />

Preparation and NMR characterization of highly substituted N-trimethyl chitosan chloride.<br />

Carbohydr Polym 36: 157-165 (1998).<br />

39. M. Lavertu, Z. Xia, A. N. Serreqi, M. Berrada, A. Rodrigues, D. Wang, M. D. Buschmann, and<br />

A. Gupta. A validated 1H NMR method <strong>for</strong> the determination of the degree of deacetylation of<br />

chitosan. J Pharm Biomed Anal 32: 1149-1158 (2003).<br />

40. J. Hombach, H. Hoyer, and A. Bernkop-Schnürch. Thiolated chitosans: Development and in<br />

vitro evaluation of an oral tobramycin sulphate delivery system. Eur J Pharm Sci 33: 1-8 (2008).<br />

41. X. Jiang, A. Van Der Horst, M. J. Van Steenbergen, N. Akeroyd, C. F. Van Nostrum, P. J.<br />

Schoenmakers, and W. E. Hennink. Molar-m<strong>as</strong>s characterization of cationic polymers <strong>for</strong> gene<br />

delivery by aqueous size-exclusion chromatography. Pharm Res 23: 595-603 (2006).<br />

42. D. A. Scudiero, R. H. Shoemaker, K. D. Paull, A. Monks, S. Tierney, T. H. Nofziger, M. J.<br />

Currens, D. Seniff, and M. R. Boyd. Evaluation of a soluble tetrazolium/<strong>for</strong>mazan <strong>as</strong>say <strong>for</strong> cell<br />

growth and drug sensitivity in culture using human and other tumor cell lines. Cancer Res 48:<br />

4827-4833 (1988).<br />

43. C. E. K<strong>as</strong>t and A. Bernkop-Schnürch. Thiolated polymers - thiomers: Development and in vitro<br />

evaluation of chitosan-thioglycolic acid conjugates. Biomaterials 22: 2345-2352 (2001).<br />

44. N. Hagenaars, M. Mania, P. de Jong, I. Que, R. Nieuwland, B. Slütter, H. Glansbeek, J. Heldens,<br />

H. van den Bosch, C. Löwik, E. Kaijzel, E. M<strong>as</strong>trobattista, and W. Jiskoot. Role of<br />

trimethylated chitosan (TMC) in n<strong>as</strong>al residence time, local distribution and toxicity of an<br />

intran<strong>as</strong>al influenza vaccine. J Control Rele<strong>as</strong>e 144: 17-24 (2010).<br />

143


Chapter 6<br />

144


Thiolated TMC <strong>for</strong> Covalently Stabilized Nanoparticles<br />

Supporting In<strong>for</strong>mation:<br />

TAILORABLE THIOLATED TRIMETHYL CHITOSANS FOR<br />

COVALENTLY STABILIZED NANOPAR<strong>TI</strong>CLES<br />

Rolf J. Verheul, Steffen van der Wal, Wim E. Hennink<br />

Contents:<br />

1H-NMR (D 2O, 80°C, 500MHz), TMC-COOH DQ 25%<br />

1H-NMR (D 2O, 80°C, 500MHz), TMC-COOH DQ 54%<br />

1H-NMR (D 2O, 80°C, 500MHz), TMC-SH DQ 25%, D thiol 7%<br />

1H-NMR (D 2O, 80°C, 500MHz), TMC-SH DQ 54%, D thiol 6%<br />

S1<br />

S2<br />

S3<br />

S4<br />

145


Chapter 6<br />

ppm (f1)<br />

5.50<br />

5.00<br />

4.50<br />

S1. 1 H-NMR of TMC-COOH DQ 25%.<br />

4.00<br />

3.50<br />

3.00<br />

2.50<br />

2.00<br />

ppm (f1)<br />

5.50<br />

5.00<br />

4.50<br />

4.00<br />

3.50<br />

3.00<br />

2.50<br />

2.00<br />

S2. 1 H-NMR of TMC-COOH DQ 54%.<br />

146


Thiolated TMC <strong>for</strong> Covalently Stabilized Nanoparticles<br />

ppm (f1)<br />

5.50<br />

5.00<br />

4.50<br />

4.00<br />

S3. 1 H-NMR of TMC-SH DQ 25%, D thiol 7%.<br />

3.50<br />

3.00<br />

2.50<br />

2.00<br />

ppm (f1)<br />

5.50<br />

5.00<br />

4.50<br />

4.00<br />

S4. 1 H-NMR of TMC-SH DQ 54%, D thiol 6%.<br />

3.50<br />

3.00<br />

2.50<br />

2.00<br />

147


CHAPTER 7<br />

COVALENTLY STABILIZED<br />

TRIMETHYL CHITOSAN-HYALURONIC ACID<br />

NANOPAR<strong>TI</strong>CLES FOR NASAL AND<br />

INTRADERMAL VACCINA<strong>TI</strong>ON<br />

Rolf J. Verheul, Bram Slütter, Suzanne M. Bal,<br />

Joke A. Bouwstra, Wim Jiskoot, Wim E. Hennink.<br />

Manuscript submitted


Chapter 7<br />

Abstract<br />

The physical stability of polyelectrolyte nanocomplexes composed of trimethyl chitosan<br />

(TMC) and hyaluronic acid (HA) is limited in physiological conditions. This may minimize the<br />

favorable adjuvant effects <strong>as</strong>sociated with particulate systems <strong>for</strong> n<strong>as</strong>al and intradermal<br />

immunization. There<strong>for</strong>e, covalently stabilized nanoparticles loaded with ovalbumin (OVA)<br />

were prepared with thiolated TMC and thiolated HA via ionic gelation followed by<br />

spontaneous disulfide <strong>for</strong>mation after incubation at pH 7.4 and 37 °C. Also, maleimide PEG w<strong>as</strong><br />

coupled to the remaining thiol-moieties on the particles to shield their surface charge.<br />

OVA-loaded TMC/HA nanoparticles had a size of around 250-350 nm, a positive zeta<br />

potential and OVA loading efficiencies up to 60%. Reacting the thiolated particles with<br />

maleimide PEG resulted in a slight reduction of zeta potential (from +7 to +4 mV) and a minor<br />

incre<strong>as</strong>e in particle size. Stabilized TMC-S-S-HA particles (PEGylated or not) showed superior<br />

stability in saline solutions compared to non-stabilized particles (composed of nonthiolated<br />

polymers) but readily disintegrated upon incubation in a saline buffer containing 10 mM<br />

dithiothreitol. In both the n<strong>as</strong>al and intradermal immunization study, OVA loaded stabilized<br />

TMC-S-S-HA particles demonstrated superior immunogenicity compared to non-stabilized<br />

particles (indicated by higher IgG titers). Intran<strong>as</strong>al, PEGylation completely abolished the<br />

beneficial effects of stabilization and it induced no enhanced immune responses against OVA<br />

after intradermal administration. In conclusion, stabilization of the TMC/HA particulate<br />

system greatly enhances the immunogenicity of OVA in n<strong>as</strong>al and intradermal vaccination.<br />

+<br />

TMC/HA<br />

+<br />

+<br />

-<br />

+<br />

+<br />

-<br />

+<br />

TMC-S-S-HA<br />

+<br />

TMC-S-S-HA PEG<br />

+<br />

- +<br />

intran<strong>as</strong>al<br />

- -<br />

intradermal<br />

vaccination<br />

log IgG titres<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

intran<strong>as</strong>al<br />

prime<br />

boost<br />

6/8<br />

1/8<br />

1/8<br />

TMC/HA<br />

TMC-S-S-HA<br />

TMC-S-S-HA PEG<br />

log IgG titres<br />

intradermal<br />

5 prime boost<br />

4<br />

3<br />

2<br />

1<br />

0<br />

OVA<br />

TMC/HA<br />

TMC-S-S-HA<br />

TMC-S-S-HA PEG<br />

OVA i.m.<br />

150


Covalently Stabilized Nanoparticles <strong>for</strong> N<strong>as</strong>al and Intradermal Vaccination<br />

Introduction<br />

Most vaccines under development today are subunit vaccines b<strong>as</strong>ed on highly purified and<br />

well-characterized antigens derived from the respective pathogens against which one wants to<br />

protect. Although favorable because of their safety profile, these purified proteins generally<br />

show reduced immunogenicity compared to inactivated or attenuated pathogens [1-3].<br />

There<strong>for</strong>e, subunit vaccines have to be <strong>for</strong>mulated with adjuvants, i.e. delivery systems and/or<br />

immune potentiators that improve the immunogenicity of the antigens to elicit adequate,<br />

protective immune responses [4]. For instance, when co-<strong>for</strong>mulated in micro- or nanoparticles,<br />

<strong>for</strong>eign proteins are much more effective in eliciting immune responses than <strong>as</strong> plain protein<br />

solution [5-9], most likely because of the particle’s resemblance to the original pathogen, their<br />

multimeric antigen presentation and improved protection of the antigen against degradation<br />

[10]. Furthermore, particles are better taken up by antigen presenting cells (APCs), they may<br />

prolong the residence time of the antigen at the site of action and can co-deliver antigen and<br />

adjuvant to the same cell [8, 11, 12]. As a consequence, several types of particulate systems<br />

have been studied <strong>for</strong> vaccine delivery, including liposomes, oil-in-water emulsions, virus like<br />

particles, ISCOMs and polymeric carriers [1, 8, 9, 13].<br />

Alternative vaccine administration routes to conventional intramuscular immunization have<br />

been widely studied [2, 3, 5, 10, 14]. Mucosal vaccination offers several advantages over<br />

inv<strong>as</strong>ive (intramuscular or subcutaneous) immunization routes, like needle-free<br />

administration, potentially less adverse effects and the induction of local mucosal immune<br />

responses [15]. However, adequate antigen delivery via the n<strong>as</strong>al route is challenging, because<br />

of intran<strong>as</strong>al degradation and poor antigen uptake through the n<strong>as</strong>al epithelium. As trimethyl<br />

chitosan (TMC, a quaternized, water-soluble derivative of chitosan) and hyaluronic acid (HA)<br />

have muco-adhesive properties [16, 17], both polymers have been investigated in particulate<br />

<strong>for</strong>m in mucosal vaccine delivery [5, 18-23]. In n<strong>as</strong>al vaccine delivery, TMC nanoparticles have<br />

proven to have excellent adjuvant properties, most likely due to improved antigen delivery [5,<br />

24], but also immunostimulatory effects of TMC on monocyte derived dendritic cells (DCs)<br />

were observed [12, 23]. In these studies tripolyphosphate (TPP) w<strong>as</strong> used <strong>as</strong> a physical<br />

crosslinker to <strong>for</strong>m TMC nanoparticles via ionic gelation. Interestingly, results by Sayın et al.<br />

[25] showed that n<strong>as</strong>al immunization with tetanus toxoid loaded TMC:mono-N-carboxymethyl<br />

chitosan nanoparticles resulted in superior antibody titers compared to the TMC/TPP<br />

particles, indicating that combining TMC with an anionic polymer may further improve the<br />

adjuvant activity. However, the physical stability of such polyelectrolyte complexes may be<br />

limited in physiological conditions [26, 27] or at low pH [28].<br />

151


Chapter 7<br />

Intradermal (ID) immunization is another interesting immunization route. While muscular<br />

and subcutaneous tissues contain only limited numbers of APCs [29], skin tissue is abundant in<br />

DCs that play a central role in eliciting an immune response [12]. Recently, TMC nanoparticles<br />

were studied <strong>as</strong> intradermal vaccine carrier system, showing superior antibody titers against<br />

ovalbumin and diphtheria toxoid (DT) compared to the plain antigens. For DT encapsulated in<br />

TMC nanoparticles comparable immunopotency <strong>as</strong> subcutaneously administered DT-alum w<strong>as</strong><br />

observed [12].<br />

Recently, we developed covalently stabilized nanoparticles made from two oppositely<br />

charged, partially thiolated polymers, namely, thiolated trimethyl chitosan (TMC-SH) and<br />

thiolated hyaluronic acid (HA-SH) [30]. Polyelectrolyte complexes prepared with these<br />

polymers had a size of about 200-300 nm, a positive zeta potential and showed antigen<br />

encapsulation capacity up to 30%. The intermolecular disulfide bonds resulted in incre<strong>as</strong>ed<br />

stability of the TMC-S-S-HA particles in saline <strong>as</strong> compared to particles made with their<br />

nonthiolated counterparts (TMC/HA particles), which were kept together only by electrostatic<br />

interactions. Importantly, these covalently stabilized particles still allowed simple, aqueous<br />

and low stress preparation conditions <strong>as</strong> used with the preparation of conventional<br />

polyelectrolyte complexes [20, 27].<br />

It can be expected that both n<strong>as</strong>al and ID immunization antigen-loaded covalently stabilized<br />

TMC-S-S-HA particles may show enhanced immunogenicity compared to non-stabilized<br />

particles because superior particle integrity in the external environment may result in<br />

improved antigen delivery to, and activation of, APCs.<br />

Furthermore, remaining thiol groups present on the surface of TMC-S-S-HA particles allow<br />

post-particle modifications [31, 32]. Selective PEGylation of the free thiol moieties with PEGmaleimide<br />

may prove an interesting strategy <strong>as</strong> PEGylation of chitosan led to improved<br />

antibody titers in a n<strong>as</strong>al vaccination study with diphtheria toxoid possibly due to improved<br />

stability [1]. Furthermore, shielding of cationic charges may be beneficial in ID vaccination by<br />

limiting interactions with the negatively charged extracellular matrix. There<strong>for</strong>e, PEGylation<br />

may result in incre<strong>as</strong>ed mobility and uptake by APCs [33].<br />

In the present study we investigated covalently stabilized TMC-S-S-HA nanoparticles, with<br />

and without PEG coating, <strong>as</strong> potential n<strong>as</strong>al and intradermal vaccine delivery systems and<br />

compared them with non-stabilized TMC/HA nanoparticles. The particles contained ovalbumin<br />

<strong>as</strong> a model antigen. The <strong>for</strong>mulations were physico-chemically characterized and their stability<br />

in buffered saline and in the presence or absence of a reducing agent w<strong>as</strong> studied.<br />

152


Covalently Stabilized Nanoparticles <strong>for</strong> N<strong>as</strong>al and Intradermal Vaccination<br />

Furthermore, the extent and type of immune response elicited after n<strong>as</strong>al and intradermal<br />

administration of the <strong>for</strong>mulations in mice w<strong>as</strong> determined.<br />

Materials and Methods<br />

Materials. Chitosan with a residual degree of acetylation of 17% (determined with 1 H-NMR)<br />

and a number average (M n) and weight average molecular weight (M w) of 28 and 43 kDa,<br />

(determined with GPC-TD <strong>as</strong> described below), respectively, w<strong>as</strong> purch<strong>as</strong>ed from Primex<br />

(Siglufjodur, Iceland). Sodium borohydride, <strong>for</strong>maldehyde 37% (stabilized with methanol),<br />

glyoxilic acid monohydrate, cystamine dihydrochloride, dithiotreitol (DTT), 1-ethyl-3-(3-<br />

dimethylaminopropyl) carbodiimide HCl (EDC), L-cysteine HCl monohydrate, hen egg-white<br />

ovalbumin (OVA, grade V), deuterium oxide, sodium acetate, acetic acid (anhydrous), sodium<br />

hydroxide and hydrochloric acid were obtained from Sigma-Aldrich Chemical Co.<br />

Fluorescently labeled ovalbumin (OVA-FITC) w<strong>as</strong> obtained from Invitrogen (Breda, The<br />

Netherlands). Horseradish peroxid<strong>as</strong>e (HRP) conjugated goat anti-mouse IgG (γ chain specific),<br />

IgG1 (γ1 chain specific) and IgG2a (γ2a chain specific) were purch<strong>as</strong>ed from Southern Biotech<br />

Birmingham, USA). Chromogen 3, 3′, 5, 5′-tetramethylbenzidine (TMB) and the substrate buffer<br />

were purch<strong>as</strong>ed from Invitrogen. Iodomethane 99% stabilized with copper w<strong>as</strong> obtained from<br />

Acros Organics (Geel, Belgium). 5,5-dithio-bis-(2-nitrobenzoic acid) (Ellman’s reagent) w<strong>as</strong><br />

purch<strong>as</strong>ed from Pierce (Rock<strong>for</strong>d, IL, USA). Hyaluronic acid (HA, molecular weight 17 kDa <strong>as</strong><br />

determined by manufacturer) w<strong>as</strong> obtained from Lifecore (Ch<strong>as</strong>ka, USA). Methoxy<br />

polyethylene glycol (mPEG) maleimide (M w 2000) w<strong>as</strong> purch<strong>as</strong>ed from JenKem Technology<br />

(Beijing, China). TMCs with DQs of 30% (M n 33 kDa, M w 59 kDa) and 56% (M n 37 kDa, M w 78<br />

kDa) were synthesized and characterized <strong>as</strong> described previously [34]. Thiolated hyaluronic<br />

acid (HA-SH) with a M n of 15 kDa, PDI of 2.6 and a degree of thiolation of 21% w<strong>as</strong> synthesized<br />

and characterized according to the procedure by Shu et al.[35, 36]. All other chemicals used<br />

were of analytical grade.<br />

Synthesis and characterization of O-methyl free, trimethylated, partially thiolated<br />

chitosan (TMC-SH). Thiolated TMCs with different degrees of quaternization (DQ) were<br />

synthesized <strong>as</strong> described be<strong>for</strong>e [30]. The degree of thiolation and the total thiol content (free<br />

thiol moieties and disulfides) of the TMC-SHs were quantified with 5,5-dithio-bis-(2-<br />

nitrobenzoic acid) (DTNB, Ellman’s reagent) <strong>as</strong> described elsewhere [30, 37]. The M n and M w of<br />

the TMC-SHs were determined, <strong>as</strong> described previously [38], by GPC on a Viscotek system<br />

153


Chapter 7<br />

detecting refractive index, viscosity and light scattering. A Shodex OHPak SB-806 column (30<br />

cm) w<strong>as</strong> used with 0.3 M sodium acetate pH 4.4 (adjusted with acetic acid) <strong>as</strong> running buffer.<br />

Preparation of covalently stabilized nanoparticles with or without post-PEGylation.<br />

Covalently stabilized nanoparticles loaded with ovalbumin (OVA) were prepared with TMC-SH<br />

and HA-SH essentially <strong>as</strong> described be<strong>for</strong>e [30]. TMC-SH with a DQ 25% and D thiol 5% or a DQ<br />

of 54% and D thiol 6% were dissolved in 10 mM HEPES pH 7.4 at 1 mg/ml and subsequently<br />

mixed with 2.5 mg/ml ovalbumin in 10 mM HEPES pH 7.4 at a weight ratio of 10:1 (TMC-<br />

SH:OVA) under magnetic stirring. Then, HA-SH (0.5 mg/ml in 10 mM HEPES pH 7.4) w<strong>as</strong> added<br />

drop-wise yielding an opalescent nanoparticle dispersion which w<strong>as</strong> incubated at 37 °C <strong>for</strong> 3<br />

hours to allow disulfide <strong>for</strong>mation. In c<strong>as</strong>e of PEGylated particles, after 30 minutes of<br />

incubation at 37 °C mPEG maleimide (M w 2000 Da) w<strong>as</strong> added to the TMC-S-S-HA particles in a<br />

TMC:PEG w/w ratio of 2/1 and particles were incubated <strong>for</strong> an additional 2.5 hours at 37 °C.<br />

Remaining thiol-moieties on the surface of the particles were used to react with the maleimide<br />

group on the PEG at pH 7.4 (post PEGylation of the TMC-S-S-HA particles). Also, nanoparticles<br />

with non-thiolated TMC (DQ 30% or 56%), OVA and hyaluronic acid (HA) were prepared in a<br />

similar way to obtain ‘conventional’ particles only kept together by charge interactions.<br />

After incubation, the nanoparticle dispersions were centrifuged <strong>for</strong> 10 min at 10000 rpm<br />

(Biofuge Pico, PP1/97 #3324; Heraeus Instruments, Osterode, Germany) to remove the free<br />

polymers and unbound OVA. The obtained nanoparticle pellets were resuspended in 10 mM<br />

HEPES pH 7.4 and diluted to obtain a final OVA concentration of 0.5 mg/ml.<br />

Physical characterization of prepared nanoparticles. Particles were diluted in 10 mM<br />

HEPES until a slightly opalescent dispersion w<strong>as</strong> obtained. Particle size w<strong>as</strong> me<strong>as</strong>ured by<br />

dynamic light scattering (DLS) using a Malvern ALV CGS-3 (Malvern Instruments, Malvern,<br />

UK). DLS results are given <strong>as</strong> a z-average particle size diameter and a polydispersity index<br />

(PDI). The PDI can vary from 0 (indicating monodisperse particles) to 1 (indicating a<br />

completely heterodisperse system). The zeta potential of the nanoparticles w<strong>as</strong> me<strong>as</strong>ured in<br />

10 mM HEPES pH 7.4 using a Zet<strong>as</strong>izer Nano (Malvern Instruments, Malvern, UK).<br />

Determination of remaining thiol moieties on surface of particles. Remaining thiol<br />

groups on the surface of the particles after preparation were <strong>as</strong>sessed with 5,5-dithio-bis-(2-<br />

nitrobenzoic acid) (DTNB, Ellman’s reagent). DTNB w<strong>as</strong> dissolved in 0.1 M sodium phosphate,<br />

pH 8.0 containing 1 mM EDTA in a final concentration of 100 μg/ml. Then, 10 μl of particle<br />

154


Covalently Stabilized Nanoparticles <strong>for</strong> N<strong>as</strong>al and Intradermal Vaccination<br />

solution (corresponding to 0.5 mg/ml OVA) w<strong>as</strong> mixed with 400 μl of DTNB reagent solution.<br />

After incubation at room temperature <strong>for</strong> 10 min, the mixtures were centrifuged <strong>for</strong> 10 min at<br />

13000 rpm (Biofuge Pico, PP1/97 #3324; Heraeus Instruments, Osterode, Germany). The<br />

supernatant w<strong>as</strong> transferred in a microplate and the absorbance w<strong>as</strong> me<strong>as</strong>ured with a<br />

microplate reader at a wavelength of 405 nm (Bio-Rad Novapath, Hercules, Ca, USA). Cysteine<br />

standards were used to calculate the amount of remaining thiol moieties on the particles and<br />

non-incubated particles were used <strong>as</strong> a control.<br />

Loading efficiency of prepared nanoparticles. The protein loading w<strong>as</strong> determined by<br />

<strong>as</strong>saying the concentration of fluorescently labeled OVA (OVA-FITC) in the supernatant<br />

obtained after centrifugation of non-purified nanoparticles prepared in exactly the same way<br />

<strong>as</strong> OVA loaded particles (10 min at 10000 rpm; Biofuge Pico, PP1/97 #3324; Heraeus<br />

Instruments, Osterode, Germany). After resuspension of the particle pellet in 10 mM HEPES pH<br />

7.4, weakly <strong>as</strong>sociated OVA-FITC w<strong>as</strong> determined by centrifuging the purified particles <strong>for</strong> 10<br />

min 13000 rpm and then me<strong>as</strong>uring the OVA-FITC content in the supernatant. Fluorescence<br />

intensity w<strong>as</strong> determined using a Fluostar Optima with 488 nm <strong>as</strong> excitation wavelength and<br />

emission w<strong>as</strong> me<strong>as</strong>ured at 520 nm (BMG Labtech, Offenburg, Germany). The OVA loading<br />

efficiency (LE%) w<strong>as</strong> calculated using the following equation:<br />

LE% = (total OVA-FITC ― OVA-FITC in supernatant)/ total OVA-FITC x 100%<br />

Stability of nanoparticles To <strong>as</strong>sess the stability of the prepared nanoparticles, changes in<br />

size, PDI and number of nanoparticles (by optical density) were determined by dispersing the<br />

nanoparticles in physiological and high saline solutions buffered with 10 mM HEPES pH 7.4,<br />

with or without a disulfide reducing agent present. In detail, non-stabilized, stabilized and<br />

PEGylated nanoparticles prepared with TMC with high or low DQ <strong>as</strong> described be<strong>for</strong>e were<br />

diluted 50 fold in 10 mM HEPES pH 7.4 and dispersed 1:1 (v/v) in buffer or buffered saline<br />

solutions to final concentrations of 150 or 800 mM NaCl and with or without 10 mM DTT. After<br />

incubation <strong>for</strong> 30 minutes at room temperature, size and PDI were determined using DLS and<br />

samples were transferred into a 96-well plate to me<strong>as</strong>ure the optical density at 450 nm<br />

(microplate reader, Bio-Rad Novapath, Hercules, Ca, USA) to quantify the amount (turbidity) of<br />

nanoparticles in the various (saline) buffers [27]. The optical density of the particles me<strong>as</strong>ured<br />

in 10 mM HEPES w<strong>as</strong> set at 100%.<br />

155


Chapter 7<br />

N<strong>as</strong>al immunization. Groups of eight female Balb/c mice (Charles River, Boxmeer, The<br />

Netherlands), 6-8 weeks old, received two n<strong>as</strong>al doses of 10 μg OVA in the different<br />

<strong>for</strong>mulations with an interval of 3 weeks. Injections of 20 μg plain OVA were administered<br />

intramuscularly (i.m.) <strong>as</strong> control (n=5). For n<strong>as</strong>al administration, <strong>for</strong>mulations were applied in<br />

a volume of 10 μl 10 mM HEPES pH 7.4, 5 μl per nostril in two sessions. Blood samples were<br />

taken 3 weeks after the prime and booster dose. After sacrificing the animals, spleens were<br />

harvested and n<strong>as</strong>al w<strong>as</strong>hes collected.<br />

Intradermal immunization. The immunogenicity of intradermally (ID) administered OVA<br />

<strong>for</strong>mulations w<strong>as</strong> <strong>as</strong>sessed in female Balb/c mice (Charles River, Boxmeer, The Netherlands),<br />

6-8 weeks old. The mice were vaccinated twice with 3 weeks intervals. Groups of five mice<br />

were injected ID with a Hamilton syringe equipped with a 30-Gauge needle. A total volume of<br />

30 μl containing 2 μg OVA dissolved in PBS or co-<strong>for</strong>mulated with TMC/HA nanoparticles w<strong>as</strong><br />

injected into the abdominal skin under anaesthesia (by intraperitoneal injection of 150 mg/kg<br />

ketamine and 10 mg/kg xylazine). As a control 20 μg plain OVA w<strong>as</strong> injected i.m.. Blood<br />

samples were collected from the tail vein 3 weeks after the prime dose and three weeks after<br />

boost vaccination the mice were sacrificed. Just be<strong>for</strong>e euthan<strong>as</strong>ia total blood w<strong>as</strong> collected<br />

from the femoral artery. Blood samples were collected in MiniCollect® tubes (Greiner Bio-one,<br />

Alphen a/d Rijn, The Netherlands) till clot <strong>for</strong>mation and centrifuged 10 minutes at 10,000 g to<br />

obtain cell-free sera.<br />

Determination of serum IgG, IgG1, IgG2a and secretory IgA. Micro titer plates (Nunc,<br />

Roskilde, Denmark) were coated with OVA, by incubation of 1 μg/ml OVA in 40 mM sodium<br />

carbonate buffer pH 9.6 <strong>for</strong> 24 hours at 4°C. To reduce <strong>as</strong>pecific binding, wells were blocked<br />

with 1% (w/v) BSA in PBS <strong>for</strong> 1 hour at room temperature. After extensive w<strong>as</strong>hing with PBS<br />

serial dilutions of serum ranging from 10 to 2*10 6 were applied, where<strong>as</strong> n<strong>as</strong>al w<strong>as</strong>hes were<br />

added undiluted. After incubation <strong>for</strong> 1.5 hours at room temperature and extensive w<strong>as</strong>hing,<br />

OVA specific antibodies were detected using HRP conjugated goat anti mouse IgG, IgG1, IgG2a<br />

or IgA (1 hour room temperature) and by incubating with 0.1 mg/ml TMB and 30 μg/ml H 2O 2<br />

in 110 mM sodium acetate buffer pH 5.5 <strong>for</strong> 15 min at room temperature. Reaction w<strong>as</strong><br />

stopped with 2 M H 2SO 2 and absorbance w<strong>as</strong> determined at 450 nm with an EL808 microplate<br />

reader (Bio-Tek Instruments, Bad Friedrichshall, Germany).<br />

156


Covalently Stabilized Nanoparticles <strong>for</strong> N<strong>as</strong>al and Intradermal Vaccination<br />

Statistical analysis. Statistical analysis w<strong>as</strong> per<strong>for</strong>med with Prism 5 <strong>for</strong> Windows<br />

(Graphpad, San Diego, USA). Data are presented <strong>as</strong> mean ± standard deviation. Statistical<br />

significance w<strong>as</strong> determined by a one way analysis of variance (ANOVA) with a Bonferroni<br />

post-test.<br />

Results and discussion<br />

Characteristics of nanoparticle <strong>for</strong>mulations. The structural characteristics of the<br />

synthesized TMC-SHs are summarized in Table 1.<br />

D thiol M n (kDa) M w (kDa)<br />

Table 1. Characteristics of the synthesized thiolated TMCs.<br />

DQ DA<br />

Total thiol<br />

amount<br />

Amount of<br />

free –SH<br />

(μmol/g) a (μmol/g) a<br />

TMC-SH<br />

High DQ<br />

TMC-SH<br />

Low DQ<br />

54% 17% 529 (±2) 283 (±21) 6% 66 174<br />

25% 17% 478 (±10) 236 (±13) 5% 62 215<br />

a values given <strong>as</strong> mean (± standard deviation) (n=3).<br />

Thiolated and nonthiolated TMCs with high and low DQ were mixed with OVA and thiolated<br />

or nonthiolated hyaluronic acid in low ionic strength buffer (10 mM HEPES pH 7.4) to yield<br />

positively charged nanoparticles. As w<strong>as</strong> observed be<strong>for</strong>e [30], more HA-SH than HA w<strong>as</strong><br />

needed <strong>for</strong> ionic cross-linking to obtain particles and, similarly, when TMC with a higher DQ<br />

w<strong>as</strong> used, larger amounts of negatively charged HA-SH or HA were needed to obtain<br />

nanoparticles (Table 2). Additionally, mPEG maleimide w<strong>as</strong> added to thiolated particles after<br />

30 min of incubation at 37 °C to obtain PEGylation via selective maleimide coupling to the<br />

remaining free thiol moieties on the surface of the particle. Importantly, when mPEG<br />

maleimide w<strong>as</strong> added to the thiolated particles without this prior incubation step, complete<br />

disintegration of the particles w<strong>as</strong> observed; likely this is due to shielding of the chargeinteractions<br />

between TMC-SH and HA-SH by PEGylation. Apparently, some disulfide crosslinking<br />

connecting TMC-SH and HA-SH is required to allow PEGylation while maintaining<br />

particle integrity. The particle <strong>for</strong>mulations were incubated at 37 °C <strong>for</strong> 3 hours while shaking<br />

to allow disulfide-bridge <strong>for</strong>mation. Previously we demonstrated that after 3 hours incubation<br />

at 37 °C, pH 7.4 more than 80% of the thiol groups of TMC-SH polymers were converted into<br />

157


Chapter 7<br />

disulfides [30]. Nanoparticle <strong>for</strong>mulations prepared with TMC with a low DQ resulted in higher<br />

OVA loading (50-60%) than <strong>for</strong>mulations with a high DQ (ca. 30%). However, loading<br />

efficiencies were highly dependent on TMC:crosslinker ratio and the type of TMC used [39], so<br />

this may explain the differences between the <strong>for</strong>mulations. Post PEGylation of the particles had<br />

no effect on the <strong>as</strong>sociation efficiency.<br />

Table 2. Characteristics of nonthiolated, thiolated and post-PEGylated nanoparticles in 10 mM HEPES<br />

pH 7.4. Values are presented <strong>as</strong> the average of three samples ± standard deviation.<br />

OVA-FITC<br />

TMC:OVA TMC:HA TMC:mPEG<br />

Zeta<br />

Loading<br />

Size<br />

ratio ratio ratio<br />

potential<br />

Efficiency<br />

(nm)<br />

(w/w) (w/w) (w/w)<br />

(mV)<br />

(%)<br />

TMC-S-S-HA<br />

+ 7.3<br />

10:1 10:3.7 n.a. 58 (±1)<br />

338 (±73)<br />

low DQ<br />

TMC-S-S-HA<br />

low DQ PEG<br />

TMC/HA<br />

low DQ<br />

TMC-S-S-HA<br />

high DQ<br />

TMC-S-S-HA<br />

high DQ PEG<br />

TMC/HA<br />

high DQ<br />

n.a. = not applicable<br />

10:1 10:3.7 10:5 59 (±1)<br />

(±1.2)<br />

+ 4.4<br />

(±1.2)<br />

352 (±27)<br />

10:1 10:2.8 n.a. 52 (±1) + 13 (±1.6) 321 (±39)<br />

10:1 10:9.3 n.a. 30 (±3) + 16 (±0.7) 226 (±16)<br />

10:1 10:9.3 10:5 30 (±2) + 10 (±0.6) 251 (±16)<br />

10:1 10:6 n.a. 29 (±1) + 18 (±0.9) 290 (±26)<br />

Free polymers and un<strong>as</strong>sociated OVA were removed by centrifugation and resuspension of<br />

the obtained particle-pellet in 10 mM HEPES pH 7.4. All particle <strong>for</strong>mulations showed limited<br />

burst-rele<strong>as</strong>e of <strong>as</strong>sociated OVA (< 10%) after resuspension in 10 mM HEPES pH 7.4. The zeta<br />

potential of the non PEGylated particles varied from +18 mV to +7 mV (Table 2) most likely<br />

due to differences in the DQ of TMC and amount of cross-linker used [39]. Importantly,<br />

PEGylation of the thiolated particles reduced the zeta potential <strong>for</strong> both PEGylated<br />

<strong>for</strong>mulations compared to the non PEGylated TMC-S-S-HA particles (from +7 mV to +4 mV and<br />

+16 mV to +10 mV <strong>for</strong> TMC-S-S-HA with low and high DQ, respectively. When mPEG maleimide<br />

w<strong>as</strong> added to nonthiolated particles, no reduction in zeta potential w<strong>as</strong> observed, indicating<br />

that available thiol groups are required <strong>for</strong> PEGylation. Similarly, when mPEG maleimide w<strong>as</strong><br />

added to TMC-S-S-HA particles after three hours of incubation at 37 °C, no significant<br />

reduction in zeta potential w<strong>as</strong> observed, indicating that almost all thiols were converted into<br />

disulfides. This w<strong>as</strong> confirmed with Ellman’s <strong>as</strong>say <strong>as</strong> neglectable amounts of remaining thiol<br />

moieties on the surface of TMC-S-S-HA particles were found (< 4 % of amount of thiol groups<br />

on the surface of non-incubated particles). DLS analysis showed that the size of the <strong>for</strong>med<br />

158


Covalently Stabilized Nanoparticles <strong>for</strong> N<strong>as</strong>al and Intradermal Vaccination<br />

nanoparticle varied from 226 to 352 nm. TMCs with a low DQ yielded particles with a bigger<br />

size and higher PDI compared those prepared with TMC with a high DQ (Figure 1A). Also,<br />

PEGylation resulted in a slight incre<strong>as</strong>e in particle diameter by 15-25 nm.<br />

Particle stability in saline and in presence of a disulfide-reducing agent. In general, the<br />

stability of polyelectrolyte complexes is dependent on the type of polymers used and may be<br />

limited solutions of physiological salt concentrations [26, 27]. We there<strong>for</strong>e studied the<br />

physical stability of nonthiolated, thiolated and PEGylated nanoparticles in buffers with<br />

different sodium chloride concentrations. Figure 1A shows that non-stabilized TMC/HA high<br />

DQ particles demonstrated a large incre<strong>as</strong>e in size in 150 mM NaCl implying aggregation while<br />

an incre<strong>as</strong>e in polydispersity is seen in 800 mM NaCl. On the other hand, the particles made of<br />

thiolated polymers, with and without PEG-coating showed that particle size and PDI w<strong>as</strong><br />

hardly affected by salt in the buffer. The turbidity of particle solutions w<strong>as</strong> me<strong>as</strong>ured at 450<br />

nm and w<strong>as</strong> used to quantify the amount of particles remaining in physiological or high saline<br />

conditions and with or without DTT (Figure 1B). After dispersion of the non-stabilized<br />

TMC/HA particles in 150 mM NaCl, a reduction of more than 75% in optical density w<strong>as</strong><br />

observed which became even more dramatic in 800 mM NaCl. In contr<strong>as</strong>t, stabilized TMC-S-S-<br />

HA particles showed a much less reduction in OD 450 in 800 mM NaCl indicating superior<br />

stability in high ionic strength buffers. However, the TMC-S-S-HA low DQ particles showed less<br />

stability in 800 mM NaCl than w<strong>as</strong> observed in the previous study where almost no reduction<br />

in DLS counts w<strong>as</strong> observed [30]. This may be attributed to the lower degree of thiolation and<br />

lower molecular weight of the HA-SH used in this study (21 vs 52% and M n of 14 vs. 39 kDa).<br />

Both altered characteristics will likely decre<strong>as</strong>e the probability of the <strong>for</strong>mation of disulfideb<strong>as</strong>ed<br />

networks between TMC-SH and HA-SH polymers in a nanoparticle. Interestingly, TMC-S-<br />

S-HA high DQ showed less reduction of OD 450 nm in the saline solutions compared to TMC-S-<br />

S-HA low DQ. This higher stability of TMC-S-S-HA high DQ particles may be due to the higher<br />

amount of HA-SH incorporated in the particles (Table 2); more HA-SH present in a particle<br />

incre<strong>as</strong>es the chances of TMC-S-S-HA <strong>for</strong>mation which is crucial <strong>for</strong> the particle stability [30].<br />

When a disulfide-reducing agent (10 mM DTT) w<strong>as</strong> present in a physiological saline solution,<br />

the turbidity of the stabilized particles dramatically decre<strong>as</strong>ed and this shows that the<br />

stabilization of the particles is due to disulfide <strong>for</strong>mation. PEGylation of the thiolated<br />

nanoparticles had no effect on the stability in presence of salt and particles were still sensitive<br />

to DTT.<br />

159


Chapter 7<br />

OD 450 nm (% in HEPES)<br />

B<br />

100<br />

75<br />

50<br />

25<br />

0<br />

TMC/HA low DQ<br />

TMC-S-S-HA low DQ<br />

TMC-S-S-HA low DQ PEG<br />

TMC/HA high DQ<br />

TMC-S-S-HA high DQ<br />

TMC-S-S-HA high DQ PEG<br />

HEPES<br />

150 mM NaCl<br />

800 mM NaCl<br />

150 mM NaCl<br />

+10 mM DTT<br />

Figure 1. Physical characteristics and stability of the nanoparticle <strong>for</strong>mulations in presence of<br />

physiological or high saline concentrations after 30 minutes incubation at room temperature; Z-average<br />

diameter (bars) and PDI (dots) (A). Optical density at 450 nm (OD 450 nm) of the nanoparticle<br />

<strong>for</strong>mulations in presence of physiological or high saline concentrations and with or without DTT (B).<br />

Error bars represent the standard deviation (n=3).<br />

160


Covalently Stabilized Nanoparticles <strong>for</strong> N<strong>as</strong>al and Intradermal Vaccination<br />

Immunization studies. N<strong>as</strong>al and ID immunization studies were carried out with<br />

nanoparticles prepared with TMC with a low DQ because we previously demonstrated that the<br />

DQ of TMC plays only a minor role in its (n<strong>as</strong>al) adjuvanticity [22].<br />

Serum antigen-specific antibodies after n<strong>as</strong>al vaccination. The OVA specific total IgG<br />

antibody titers elicited after n<strong>as</strong>al administration of the antigen and the various TMC-HA<br />

<strong>for</strong>mulations were determined three weeks after prime and boost vaccination (Figure 2). After<br />

prime vaccination hardly any IgG titers were detected <strong>for</strong> all n<strong>as</strong>al <strong>for</strong>mulations. Interestingly,<br />

after the boost vaccination the stabilized particles showed significantly higher total IgG titers<br />

compared to the non-stabilized or the PEGylated system and six out of eight mice had a<br />

detectable IgG titer after intran<strong>as</strong>al vaccination with the stabilized particles. These results<br />

suggest that stabilization does indeed <strong>as</strong> anticipated improve the adjuvanticity of the TMC-HA<br />

nanoparticulate system and that PEGylation of these particles abolishes this effect. The<br />

administered dose w<strong>as</strong> apparently too low to (significantly) discriminate between plain OVA<br />

and the TMC-S-S-HA particles and this also led to hardly detectable IgG1 and IgG2a after boost<br />

vaccination and secretory IgA (sIgA) levels in the n<strong>as</strong>al w<strong>as</strong>h (results not shown).<br />

5<br />

prime<br />

boost<br />

5/5<br />

10 log IgG titres<br />

4<br />

3<br />

2<br />

1<br />

3/8<br />

1/8<br />

6/8<br />

* *<br />

1/8<br />

0<br />

OVA<br />

TMC/HA low DQ<br />

TMC-S-S-HA low DQ<br />

TMC-S-S-HA low DQ PEG<br />

OVA i.m.<br />

Figure 2. Antigen-specific total IgG antibody titers after n<strong>as</strong>al vaccination with OVA <strong>for</strong>mulations. Error<br />

bars indicate standard deviation (n=8 (n<strong>as</strong>al) or 5 (i.m.)). * p


Chapter 7<br />

Serum antigen-specific total IgG after intradermal vaccination. The antigen specific total<br />

IgG antibody titers elicited after ID administration of the various OVA <strong>for</strong>mulations are shown<br />

in Figure 3. After prime vaccination the stabilized particles (with or without PEGylation)<br />

showed significantly higher antibody titers compared to plain OVA (up to 75-fold) and the nonstabilized<br />

TMC/HA particles (up to 20-fold). Interestingly, the non-stabilized nanoparticles did<br />

not significantly improve immune responses compared to plain OVA. Also, after boost<br />

vaccination the stabilized systems showed incre<strong>as</strong>ed total IgG antibody titers compared to<br />

OVA, in contr<strong>as</strong>t to the non-stabilized <strong>for</strong>mulations. The immunogenicity of ID administered<br />

PEGylated stabilized particles w<strong>as</strong> similar to that of the non-PEGylated ones. These results<br />

indicate that stabilization, but not PEGylation of TMC/HA particles is essential <strong>for</strong> their ID<br />

adjuvanticity.<br />

10 log IgG titres<br />

5<br />

4<br />

3<br />

2<br />

1<br />

prime<br />

boost<br />

** **<br />

OOO<br />

OO<br />

***<br />

***<br />

***<br />

0<br />

OVA<br />

TMC/HA low DQ<br />

TMC-S-S-HA low DQ<br />

TMC-S-S-HA low DQ PEG<br />

OVA i.m.<br />

Figure 3. Antigen-specific total IgG antibody titers after intradermal vaccination with OVA <strong>for</strong>mulations.<br />

Error bars indicate standard deviation (n=5). *** p


Covalently Stabilized Nanoparticles <strong>for</strong> N<strong>as</strong>al and Intradermal Vaccination<br />

immune response. Log IgG1/IgG2a ratio <strong>for</strong> each individual mouse did not reveal a significant<br />

shifting of the type of immune response compared to plain OVA (results not shown). This is in<br />

line with previous results observed in an ID vaccination study with TMC/OVA/TPP particles<br />

[12].<br />

10 log IgG1/IgG2a titres<br />

4<br />

3<br />

2<br />

1<br />

0<br />

OVA<br />

IgG1 IgG2a<br />

*<br />

*<br />

TMC/HA low DQ<br />

TMC-S-S-HA low DQ<br />

TMC-S-S-HA low DQ PEG<br />

Figure 4. Antigen-specific IgG1 and IgG2a antibody titers after boost intradermal vaccination with OVA<br />

and nanoparticle <strong>for</strong>mulations. Error bars indicate standard deviation (n=5). * p


Chapter 7<br />

2). However, previously we found that <strong>for</strong> n<strong>as</strong>al vaccination with TMC-coated whole<br />

inactivated influenza virus, similar differences in zeta potential did not result in altered<br />

adjuvant effects [22]. This indicates that it is unlikely that the lower zeta potential of the<br />

stabilized particles is responsible <strong>for</strong> the observed beneficial effects in the present n<strong>as</strong>al<br />

vaccination study. Thus, it can be concluded that the stabilization of TMC-S-S-HA nanoparticles<br />

resulted in improved immunogenicity, however, the exact mechanism(s) need to be<br />

determined.<br />

PEGylation of the stabilized particles inhibited the beneficial effects of stabilization (n<strong>as</strong>al<br />

administration), while <strong>for</strong> ID administration the PEGylated particles had similar effects <strong>as</strong> the<br />

non-PEGylated ones. This indicates that PEGylation of particles decre<strong>as</strong>ed n<strong>as</strong>al antigen<br />

delivery. Although it h<strong>as</strong> been suggested that PEGylation of TMC improves muco-adhesion<br />

[41], possibly, PEGylation will also hinder electrostatic interactions between the cationic<br />

particle and the epithelial barrier thereby reducing particle transport over epithelial cells [42]<br />

or particle uptake by microfold (M)-cells. Van den Berg et al. showed that PEGylation of<br />

cationic polyplexes improves DNA transfection efficiencies after ID tattooing [33]. Although<br />

PEGylation may improve mobility in the intracellular matrix, reduced electrostatic interactions<br />

with DCs may lead to lower uptake [43]. Better understanding of these conflicting effects may<br />

result in optimal use of PEGylation <strong>for</strong> ID vaccination. Importantly, we showed that post<br />

particle modifications of the stabilized TMC-S-S-HA polyelectrolyte system are possible and<br />

this opens up a variety of options such <strong>as</strong> specific targeting towards DCs or M-cells.<br />

Conclusion<br />

In this paper we showed that stabilized nanoparticles can be prepared using TMC-SH with<br />

high and low DQ together with HA-SH and that these particles can be post PEGylated. These<br />

particles showed adequate OVA <strong>as</strong>sociation efficiency, preserved their particle integrity under<br />

saline conditions but readily disintegrated when a disulfide reducing agent w<strong>as</strong> present.<br />

Stabilized particles showed enhanced adjuvanticity in n<strong>as</strong>al and ID vaccination compared to<br />

non-stabilized particles. PEGylation abolished the beneficial effects of stabilization in n<strong>as</strong>al<br />

vaccine administration and showed similar immunogenicity <strong>as</strong> stabilized particles in ID<br />

immunization. Our results imply that the stabilized TMC-S-S-HA nanoparticles <strong>for</strong>m a highly<br />

versatile and promising vaccine carrier system while also offering options <strong>for</strong> post particle<br />

modifications. Further studies should be per<strong>for</strong>med to elaborate the exact mechanism by<br />

which stabilization results in improved immunogenicity.<br />

164


Covalently Stabilized Nanoparticles <strong>for</strong> N<strong>as</strong>al and Intradermal Vaccination<br />

Acknowledgement. This research w<strong>as</strong> partially per<strong>for</strong>med under the framework of <strong>TI</strong><br />

<strong>Pharma</strong> project number D5-106-1; Vaccine delivery: alternatives <strong>for</strong> conventional multiple<br />

injection vaccines.<br />

165


Chapter 7<br />

References<br />

1. N. Csaba, M. Garcia-Fuentes, and M. J. Alonso. Nanoparticles <strong>for</strong> n<strong>as</strong>al vaccination. Adv Drug<br />

Deliv Rev 61: 140-157 (2009).<br />

2. N. Mishra, A. K. Goyal, S. Tiwari, R. Paliwal, S. R. Paliwal, B. Vaidya, S. Mangal, M. Gupta, D.<br />

Dube, A. Mehta, and S. P. Vy<strong>as</strong>. Recent advances in mucosal delivery of vaccines: Role of<br />

mucoadhesive/biodegradable polymeric carriers. Exp Opin Ther Patents 20: 661-679 (2010).<br />

3. D. T. O'Hagan and R. Rappuoli. Novel approaches to vaccine delivery. Pharm Res 21: 1519-1530<br />

(2004).<br />

4. V. E. Schijns. Immunological concepts of vaccine adjuvant activity. Curr Opin Immunol 12: 456-<br />

463 (2000).<br />

5. M. Amidi, E. M<strong>as</strong>trobattista, W. Jiskoot, and W. E. Hennink. Chitosan-b<strong>as</strong>ed delivery systems<br />

<strong>for</strong> protein therapeutics and antigens. Adv Drug Deliv Rev 62: 59-82 (2010).<br />

6. L. Feng, X. J. Zhou, and X. R. Qi. Preparation, rele<strong>as</strong>e and immunogenicity evaluation of<br />

HBsAg-PLGA microspheres. Beijing da xue xue bao. Yi xue ban = J Peking University. Health sci 37:<br />

527-531 (2005).<br />

7. D. T. O'Hagan, M. Singh, and J. B. Ulmer. Microparticle-b<strong>as</strong>ed technologies <strong>for</strong> vaccines.<br />

Methods 40: 10 (2006).<br />

8. A. C. Rice-Ficht, A. M. Aren<strong>as</strong>-Gamboa, M. M. Kahl-McDonagh, and T. A. Ficht. Polymeric<br />

particles in vaccine delivery. Curr Opin Microbiol 13: 106-112 (2010).<br />

9. S. D. Xiang, A. Scholzen, G. Minigo, C. David, V. Apostolopoulos, P. L. Mottram, and M.<br />

Plebanski. Pathogen recognition and development of particulate vaccines: Does size matter?<br />

Methods 40: 1-9 (2006).<br />

10. M. R. Neutra and P. A. Kozlowski. Mucosal vaccines: The promise and the challenge. Nat Rev<br />

Immunol 6: 148-158 (2006).<br />

11. T. Akagi, X. Wang, T. Uto, M. Baba, and M. Ak<strong>as</strong>hi. Protein direct delivery to dendritic cells<br />

using nanoparticles b<strong>as</strong>ed on amphiphilic poly(amino acid) derivatives. Biomaterials 28: 3427-<br />

3436 (2007).<br />

12. S. M. Bal, B. Slütter, E. van Riet, A. C. Kruithof, Z. Ding, G. F. A. Kersten, W. Jiskoot, and J.<br />

A. Bouwstra. Efficient induction of immune responses through intradermal vaccination with N-<br />

trimethyl chitosan containing antigen <strong>for</strong>mulations. J Control Rele<strong>as</strong>e 142: 374-383 (2010).<br />

13. A. Pattani, V. B. Patravale, L. Panicker, and P. D. Potdar. Immunological Effects and<br />

Membrane Interactions of Chitosan Nanoparticles. Mol Pharm 6: 345-352 (2009).<br />

14. P. Kuo-Haller, Y. Cu, J. Blum, J. A. Appleton, and W. M. Saltzman. Vaccine Delivery by<br />

Polymeric Vehicles in the Mouse Reproductive Tract Induces Sustained Local and Systemic<br />

Immunity. Mol Pharm (2010).<br />

15. B. Slütter, N. Hagenaars, and W. Jiskoot. Rational design of n<strong>as</strong>al vaccines. J Drug Target 16: 1-17<br />

(2008).<br />

16. V. Grabovac, D. Guggi, and A. Bernkop-Schnürch. Comparison of the mucoadhesive<br />

properties of various polymers. Adv Drug Deliv Rev 57: 1713-1723 (2005).<br />

17. D. Snyman, J. H. Hamman, and A. F. Kotze. Evaluation of the mucoadhesive properties of N-<br />

trimethyl chitosan chloride. Drug Develop Ind Pharm 29: 61-69 (2003).<br />

18. P. A. O' Hagan D. Use of hyaluronic acid polymers <strong>for</strong> mucosal delivery of vaccine and<br />

adjuvants., 2004.<br />

19. M. Amidi, H. C. Pellikaan, H. Hirschberg, A. H. de Boer, D. J. A. Crommelin, W. E. Hennink,<br />

G. Kersten, and W. Jiskoot. Diphtheria toxoid-containing microparticulate powder <strong>for</strong>mulations<br />

<strong>for</strong> pulmonary vaccination: Preparation, characterization and evaluation in guinea pigs. Vaccine<br />

25: 6818-6829 (2007).<br />

20. M. Amidi, S. G. Romeijn, J. C. Verhoef, H. E. Junginger, L. Bungener, A. Huckriede, D. J. A.<br />

Crommelin, and W. Jiskoot. N-<strong>Trimethyl</strong> chitosan (TMC) nanoparticles loaded with influenza<br />

subunit antigen <strong>for</strong> intran<strong>as</strong>al vaccination: Biological properties and immunogenicity in a mouse<br />

model. Vaccine 25: 144-153 (2007).<br />

166


Covalently Stabilized Nanoparticles <strong>for</strong> N<strong>as</strong>al and Intradermal Vaccination<br />

21. W. Boonyo, H. E. Junginger, N. Waranuch, A. Polnok, and T. Pitaksuteepong. Chitosan and<br />

trimethyl chitosan chloride (TMC) <strong>as</strong> adjuvants <strong>for</strong> inducing immune responses to ovalbumin in<br />

mice following n<strong>as</strong>al administration. J Control Rele<strong>as</strong>e 121: 168-175 (2007).<br />

22. N. Hagenaars, R. J. Verheul, I. Mooren, P. H. J. L. F. de Jong, E. M<strong>as</strong>trobattista, H. L.<br />

Glansbeek, J. G. M. Heldens, H. van den Bosch, W. E. Hennink, and W. Jiskoot. Relationship<br />

between structure and adjuvanticity of N,N,N-trimethyl chitosan (TMC) structural variants in a<br />

n<strong>as</strong>al influenza vaccine. J Control Rele<strong>as</strong>e 140: 126-133 (2009).<br />

23. B. Slütter, L. Plapied, V. Fievez, M. Alonso Sande, A. des Rieux, Y. J. Schneider, E. Van Riet,<br />

W. Jiskoot, and V. Préat. Mechanistic study of the adjuvant effect of biodegradable<br />

nanoparticles in mucosal vaccination. J Control Rele<strong>as</strong>e 138: 113-121 (2009).<br />

24. N. Hagenaars, M. Mania, P. de Jong, I. Que, R. Nieuwland, B. Slütter, H. Glansbeek, J. Heldens,<br />

H. van den Bosch, C. Löwik, E. Kaijzel, E. M<strong>as</strong>trobattista, and W. Jiskoot. Role of<br />

trimethylated chitosan (TMC) in n<strong>as</strong>al residence time, local distribution and toxicity of an<br />

intran<strong>as</strong>al influenza vaccine. J Control Rele<strong>as</strong>e 144: 17-24 (2010).<br />

25. B. Sayin, S. Somavarapu, X. W. Li, D. Sesardic, S. Şenel, and O. H. Alpar. TMC-MCC (Ntrimethyl<br />

chitosan-mono-N-carboxymethyl chitosan) nanocomplexes <strong>for</strong> mucosal delivery of<br />

vaccines. Eur J Pharm Sci 38: 362-369 (2009).<br />

26. S. Boddohi, N. Moore, P. A. Johnson, and M. J. Kipper. Polysaccharide-b<strong>as</strong>ed polyelectrolyte<br />

complex nanoparticles from chitosan, heparin, and hyaluronan. Biomacromol 10: 1402-1409<br />

(2009).<br />

27. D. V. Pergushov, H. M. Buchhammer, and K. Lunkwitz. Effect of a low-molecular-weight salt<br />

on colloidal dispersions of interpolyelectrolyte complexes. Coll Polym Sci 277: 101-107 (1999).<br />

28. A. Bernkop-Schnürch, A. Weithaler, K. Albrecht, and A. Greimel. Thiomers: Preparation and in<br />

vitro evaluation of a mucoadhesive nanoparticulate drug delivery system. Int J Pharm 317: 76-81<br />

(2006).<br />

29. E. Raz, D. A. Carson, S. E. Parker, T. B. Parr, A. M. Abai, G. Aichinger, S. H. Gromkowski, M.<br />

Singh, D. Lew, M. A. Yankauck<strong>as</strong>, S. M. Baird, and G. H. Rhodes. Intradermal gene<br />

immunization: The possible role of DNA uptake in the induction of cellular immunity to<br />

viruses. PNAS 91: 9519-9523 (1994).<br />

30. R. J. Verheul, S. van der Wal, and W. E. Hennink. <strong>Tailorable</strong> Thiolated <strong>Trimethyl</strong> <strong>Chitosans</strong> <strong>for</strong><br />

Covalently Stabilized Nanoparticles. Biomacromol (2010).<br />

31. T. M<strong>as</strong>uko, A. Minami, N. Iw<strong>as</strong>aki, T. Majima, S. I. Nishimura, and Y. C. Lee. Thiolation of<br />

chitosan. Attachment of proteins via thioether <strong>for</strong>mation. Biomacromol 6: 880-884 (2005).<br />

32. B. Slütter, P. C. Soema, Z. Ding, R. Verheul, W. Hennink, and W. Jiskoot. Conjugation of<br />

ovalbumin to trimethyl chitosan improves immunogenicity of the antigen. J Control Rele<strong>as</strong>e 143:<br />

207-214 (2010).<br />

33. J. H. van den Berg, K. Oosterhuis, W. E. Hennink, G. Storm, L. J. van der Aa, J. F. J.<br />

Engbersen, J. B. A. G. Haanen, J. H. Beijnen, T. N. Schumacher, and B. Nuijen. Shielding the<br />

cationic charge of nanoparticle-<strong>for</strong>mulated dermal DNA vaccines is essential <strong>for</strong> antigen<br />

expression and immunogenicity. J Control Rele<strong>as</strong>e 141: 234-240 (2010).<br />

34. R. J. Verheul, M. Amidi, S. van der Wal, E. van Riet, W. Jiskoot, and W. E. Hennink. Synthesis,<br />

characterization and in vitro biological properties of O-methyl free N,N,N-trimethylated<br />

chitosan. Biomaterials 29: 3642-3649 (2008).<br />

35. R. Censi, P. J. Fieten, P. di Martino, W. E. Hennink, and T. Vermonden. In Situ Forming<br />

Hydrogels by Tandem Thermal Gelling and Michael Addition Reaction between<br />

Thermosensitive Triblock Copolymers and Thiolated Hyaluronan. Macromol 43: 5771-5778<br />

(2010).<br />

36. X. Z. Shu, Y. Liu, Y. Luo, M. C. Roberts, and G. D. Prestwich. Disulfide cross-linked<br />

hyaluronan hydrogels. Biomacromol 3: 1304-1311 (2002).<br />

37. J. Hombach, H. Hoyer, and A. Bernkop-Schnürch. Thiolated chitosans: Development and in<br />

vitro evaluation of an oral tobramycin sulphate delivery system. Eur J Pharm Sci 33: 1-8 (2008).<br />

167


Chapter 7<br />

38. X. Jiang, A. Van Der Horst, M. J. Van Steenbergen, N. Akeroyd, C. F. Van Nostrum, P. J.<br />

Schoenmakers, and W. E. Hennink. Molar-m<strong>as</strong>s characterization of cationic polymers <strong>for</strong> gene<br />

delivery by aqueous size-exclusion chromatography. Pharm Res 23: 595-603 (2006).<br />

39. A. Jintapattanakit, V. B. Junyapr<strong>as</strong>ert, S. Mao, J. Sitterberg, U. Bakowsky, and T. Kissel. Peroral<br />

delivery of insulin using chitosan derivatives: A comparative study of polyelectrolyte<br />

nanocomplexes and nanoparticles. Int J Pharm 342: 240-249 (2007).<br />

40. L. Yin, J. Ding, C. He, L. Cui, C. Tang, and C. Yin. Drug permeability and mucoadhesion<br />

properties of thiolated trimethyl chitosan nanoparticles in oral insulin delivery. Biomaterials 30:<br />

5691-5700 (2009).<br />

41. A. Jintapattanakit, V. B. Junyapr<strong>as</strong>ert, and T. Kissel. The role of mucoadhesion of trimethyl<br />

chitosan and PEGylated trimethyl chitosan nanocomplexes in insulin uptake. J Pharm Sci 98:<br />

4818-4830 (2009).<br />

42. S. Mao, O. Germershaus, D. Fischer, T. Linn, R. Schnepf, and T. Kissel. Uptake and transport<br />

of PEG-graft-trimethyl-chitosan copolymer-insulin nanocomplexes by epithelial cells. Pharm Res<br />

22: 2058 (2005).<br />

43. Y. Sheng, Y. Yuan, C. Liu, X. Tao, X. Shan, and F. Xu. In vitro macrophage uptake and in vivo<br />

biodistribution of PLA-PEG nanoparticles loaded with hemoglobin <strong>as</strong> blood substitutes: Effect<br />

of PEG content. J Mat Sci: Mat Med 20: 1881-1891 (2009).<br />

168


CHAPTER 8<br />

SUMMARY AND FUTURE PERSPEC<strong>TI</strong>VES


Summary and Future Perspectives<br />

Summary<br />

Active vaccination h<strong>as</strong> proven to be the most (cost) effective tool in the fight against<br />

infectious dise<strong>as</strong>es. Vaccines are usually made of live attenuated or inactivated pathogens (e.g.<br />

viruses or bacteria) or purified immunogenic protein(conjugate)s derived from these<br />

pathogens. Nowadays, most vaccines are administered via parenteral injection however, the<br />

risk <strong>for</strong> contaminated needles and need <strong>for</strong> trained personnel have risen interest <strong>for</strong><br />

alternative immunization routes.<br />

Chapter 1 discusses the (dis)advantages of intramuscular and alternative administration<br />

routes, in particular intran<strong>as</strong>al immunization that allows relatively simple, needle-free<br />

administration, reduces the need <strong>for</strong> trained professionals and may, importantly, elicit both<br />

mucosal and systemic immune responses. On the other hand, antigen degradation and poor<br />

delivery to antigen presenting cells (APCs) are major drawbacks of n<strong>as</strong>al vaccination. As live<br />

attenuated vaccines raise considerable safety issues, (n<strong>as</strong>al) vaccine development now mainly<br />

focuses on purified, well-characterized antigenic proteins. However, these subunit vaccines are<br />

generally less immunogenic and need potent adjuvant(system)s to elicit an adequate immune<br />

response. The use of muco-adhesive polymers like N,N,N-trimethylchitosan (TMC), a partially<br />

quaternized, water-soluble chitosan derivative, can enhance the immune responses against<br />

antigens, presumably by incre<strong>as</strong>ing the n<strong>as</strong>al residence time and/or improving the contact<br />

area between the antigen and the mucosal surface. TMC’s chemical structure can vary in the<br />

degree of quaternization (DQ), giving the polymer its cationic charge, but also in extent of O-<br />

methylation (DOM), degree of acetylation (DAc) and polymer molecular weight. Tailorability of<br />

these structural elements will allow better understanding of the contribution of these moieties<br />

to the physico-chemical and biological properties of TMC. Additionally, the introduction of side<br />

groups such <strong>as</strong> thiol-moieties may further be applied to improve and optimize TMC’s<br />

properties.<br />

Chitosan and its derivatives are much more effective in eliciting immune responses in microor<br />

nanoparticulate <strong>for</strong>m than <strong>as</strong> plain polymer solution. Ionic gelation b<strong>as</strong>ed upon electrostatic<br />

interactions is a simple, commonly used method to yield nanoparticulate systems. The<br />

complexation between the positively charged TMC and oppositely charged (macro)molecules<br />

added drop-wise under stirring in low ionic strength buffer results in the spontaneous<br />

<strong>for</strong>mation of nanoparticles. However, the physico-chemical stability and the immunogenicity of<br />

these antigen-loaded complexes are dependent on the characteristics of the crosslinker used,<br />

and the currently applied carrier-systems may be sub-optimal. Furthermore, although TMC’s<br />

171


Chapter 8<br />

mode of action is usually attributed to its muco-adhesive and penetration enhancing<br />

properties, detailed knowledge of the mechanism behind the favorable adjuvant<br />

characteristics is lacking (e.g., the effect of TMC on APCs is currently unknown). Concluding,<br />

many opportunities <strong>for</strong> optimizing TMC structure and TMC-b<strong>as</strong>ed nanoparticles remain and<br />

mechanistic insights into the mode of action of TMC have to be obtained.<br />

The aim of this thesis is to develop synthetic routes to synthesize TMC structural variants in<br />

a controllable and tailorable manner and introduce substitutions such <strong>as</strong> thiol-moieties that<br />

may further improve TMC’s properties. In this way, structure-activity relationships can be<br />

investigated exploiting in vitro <strong>as</strong>says and in in vivo (n<strong>as</strong>al) vaccination studies.<br />

Chapter 2 challenges the current synthetic method to synthesize TMC that is <strong>as</strong>sociated with<br />

several side-reactions such <strong>as</strong> O-methylation and chain scission. Since these side reactions may<br />

affect the polymer characteristics, there is a need <strong>for</strong> TMCs without O-methylation and<br />

disparities in chain lengths while varying the DQ. In this chapter, O-methyl free TMCs with<br />

varying DQs were successfully synthesized by using a two-step method. First, chitosan w<strong>as</strong><br />

quantitatively dimethylated using <strong>for</strong>mic acid and <strong>for</strong>maldehyde. Then, in presence of an<br />

excess amount of iodomethane, TMC w<strong>as</strong> obtained with different DQs (22-68%) by varying the<br />

reaction time. TMC obtained by this two-step method showed no detectable O-methylation<br />

( 1 H-NMR) and a slight incre<strong>as</strong>e in molecular weight with incre<strong>as</strong>ing DQ (GPC), implying that no<br />

chain scission occurred during synthesis. The solubility in aqueous solutions at pH 7 of O-<br />

methyl free TMC with DQ < 22% w<strong>as</strong> less <strong>as</strong> compared to O-methylated TMC with the same DQ.<br />

On the other hand, O-methyl free TMC with DQ > 30% had an excellent aqueous solubility. On<br />

Caco-2, cells O-methyl free TMCs demonstrated a larger decre<strong>as</strong>e in trans-epithelial electrical<br />

resistance (TEER) than O-methylated TMCs. Also, with incre<strong>as</strong>ing DQ, an incre<strong>as</strong>e in<br />

cytotoxicity (MTT) and membrane permeability (LDH) w<strong>as</strong> observed. This chapter clearly<br />

demonstrates that the DQ <strong>as</strong> well <strong>as</strong> O-methylation substantially influence the physicochemical<br />

and in vitro biological properties of TMC.<br />

The influence of another structural variant of TMC, the degree of acetylation (DAc), on in<br />

vitro degradation and biological properties w<strong>as</strong> evaluated in Chapter 3. TMCs with a DAc<br />

ranging from 11 to 55% were synthesized by using a three-step method. First, chitosan w<strong>as</strong><br />

partially re-acetylated using acetic anhydride followed by quantitative dimethylation using<br />

<strong>for</strong>maldehyde and sodium borohydride. Then, in presence of an excess amount of<br />

iodomethane, TMC w<strong>as</strong> synthesized. The TMCs obtained by this method showed neither<br />

172


Summary and Future Perspectives<br />

detectable O-methylation nor loss in acetyl groups ( 1 H-NMR) and a slight incre<strong>as</strong>e in molecular<br />

weight (GPC) with incre<strong>as</strong>ing degree of substitution, implying that no chain scission occurred<br />

during synthesis. The extent of lysozyme-catalyzed degradation of TMC, and that of its<br />

precursors chitosan and dimethyl chitosan, w<strong>as</strong> highly dependent on the DAc; polymers with<br />

the highest DAc showed the largest decre<strong>as</strong>e in molecular weight. On Caco-2 cells, TMCs with a<br />

high DAc (~50%), a DQ of around 44% and with or without O-methylated groups, were not<br />

able to open tight junctions in the trans-epithelial electrical resistance (TEER) <strong>as</strong>say. This in<br />

contr<strong>as</strong>t to TMCs (both O-methylated and O-methyl free; concentration 2.5 mg/ml) with a<br />

similar DQ but a lower DAc which were able to reduce the TEER with 30 and 70%,<br />

respectively. Additionally, TMCs with a high DAc (~50%) demonstrated no cell toxicity (MTT,<br />

LDH rele<strong>as</strong>e) up to a concentration of 10 mg/ml. This chapter shows that the degree of N-<br />

acetylation dramatically influences the enzymatic degradation and in vitro biological<br />

properties of TMC.<br />

The results of Chapter 2 and 3 demonstrate that the DQ, DOM and DAc all influence the in<br />

vitro biological properties of TMC. There<strong>for</strong>e we investigated the influence of the structural<br />

properties of TMC on its adjuvanticity in an in vivo intran<strong>as</strong>al (i.n.) immunization study. In<br />

Chapter 4, TMCs with varying degrees of quaternization (DQ, 22-86%), O-methylation (DOM,<br />

0-76%) and acetylation (DAc 9-54%) were <strong>for</strong>mulated with whole inactivated influenza virus<br />

(WIV). Simple mixing of the TMCs with WIV at a 1:1 (w/w) ratio resulted in comparable<br />

positively charged nanoparticles, indicating coating of the negatively charged WIV with TMC.<br />

The amount of free TMC in solution w<strong>as</strong> comparable <strong>for</strong> all TMC-WIV <strong>for</strong>mulations. After i.n.<br />

immunization of mice with WIV and TMC-WIV on day 0 and 21, all TMC-WIV <strong>for</strong>mulations<br />

induced stronger total IgG, IgG1 and IgG2a/c responses than WIV alone, except WIV<br />

<strong>for</strong>mulated with re-acetylated TMC with a DAc of 54% and a DQ of 44% (TMC-RA44). No<br />

significant differences in antibody titers were observed <strong>for</strong> TMCs that varied in DQ or DOM,<br />

indicating that these structural characteristics play a minor role in their adjuvant properties.<br />

TMC with a DQ of 56% (TMC56) <strong>for</strong>mulated with WIV at a ratio of 5:1 (w/w) resulted in<br />

significantly lower IgG2a/c:IgG1 ratio’s compared to TMC56 mixed in ratios of 0.2:1 and 1:1,<br />

implying a shift towards a Th2 type immune response. Challenge of vaccinated mice with<br />

aerosolized virus demonstrated protection <strong>for</strong> all TMC-WIV <strong>for</strong>mulations with the exception of<br />

TMC-RA44-WIV. This chapter demonstrates that coating of WIV with TMCs strongly enhances<br />

the immunogenicity and induced protection after i.n. vaccination with WIV. The adjuvant<br />

173


Chapter 8<br />

properties of TMCs <strong>as</strong> i.n. adjuvant are strongly decre<strong>as</strong>ed by re-acetylation of TMC, where<strong>as</strong><br />

the DQ and DOM hardly affect the adjuvanticity of TMC.<br />

The aim of Chapter 5A w<strong>as</strong> to elucidate the re<strong>as</strong>on <strong>for</strong> the lack of adjuvanticity of reacetylated<br />

TMC (TMC-RA) by comparing TMC-RA (degree of acetylation 54%) with TMC<br />

(degree of acetylation 17%) at six potentially critical steps in the induction of an immune<br />

response after intran<strong>as</strong>al (i.n.) administration in mice: chemical stability of the polymer in<br />

murine n<strong>as</strong>al w<strong>as</strong>hings, local i.n. distribution of WIV, n<strong>as</strong>al residence time of WIV, cellular<br />

uptake of WIV by epithelial cells, transport of WIV by epithelial cells, and capacity of the<br />

<strong>for</strong>mulation to induce maturation of murine bone marrow derived dendritic cells (DCs). TMC-<br />

RA w<strong>as</strong> degraded in a n<strong>as</strong>al w<strong>as</strong>h to a slightly larger extent than TMC. The local i.n. distribution<br />

and n<strong>as</strong>al clearance were similar <strong>for</strong> both TMC types. Fluorescently labeled WIV w<strong>as</strong> taken up<br />

more efficiently by Calu-3 cells when <strong>for</strong>mulated with TMC-RA compared to TMC and both<br />

TMCs significantly reduced transport of WIV over a Calu-3 monolayer. Murine bone-marrow<br />

derived dendritic cell activation w<strong>as</strong> similar <strong>for</strong> plain WIV, <strong>as</strong> well <strong>as</strong> <strong>for</strong> WIV <strong>for</strong>mulated with<br />

TMC-RA or TMC. The inferior adjuvant effect of TMC-RA over that of TMC might be caused by a<br />

slightly lower stability of TMC-RA in the n<strong>as</strong>al cavity, rather than by any of the other factors<br />

studied in this paper.<br />

Interestingly, N-acetylated glucosamine units or GlcNAcs present in TMC have been<br />

described to bind several human C-type lectins, a family of lectins involved in the human<br />

innate immune response. There<strong>for</strong>e the effect of TMC and re-acetylated TMC with a degree of<br />

acetylation of 54% (TMC-RA) on the uptake and maturation of human dendritic cells (DCs) w<strong>as</strong><br />

<strong>as</strong>sessed in Chapter 5B using whole inactivated influenza virus (WIV) <strong>as</strong> antigen. Studies on<br />

monocyte-derived human DCs indicated that the uptake of TMC(-RA) coated WIV w<strong>as</strong> slightly<br />

lower than plain WIV. TMC-RA, <strong>as</strong> a solution or when <strong>for</strong>mulated with WIV, induced much<br />

stronger DC maturation, <strong>as</strong> me<strong>as</strong>ured with CD86 expression, than any of the other<br />

<strong>for</strong>mulations. Also, only TMC-RA(-WIV) induced high IL-10, TNF-α and IL-12p40 and IL-12p70<br />

rele<strong>as</strong>e by DCs. Since both IL-10 and IL-12p70 levels were elevated, no polarization towards a<br />

Th1 or Th2 type immune response could be established. Concluding, TMC-RA h<strong>as</strong> strong<br />

immuno-stimulatory effects in vitro on human monocyte derived dendritic cells which<br />

indicates that the degree of N-acetylation is critical <strong>for</strong> the adjuvant effect of TMC on human<br />

DCs, but not on mice DCs.<br />

174


Summary and Future Perspectives<br />

Introduction of thiol-moieties will further broaden the potential pharmaceutical applications<br />

of TMC by enhancing its muco-adhesive properties, allowing further chemical derivatization<br />

reactions via reducible disulfide-bridges and opening up possibilities <strong>for</strong> covalently linked<br />

nanoparticles together with a thiolated anionic polymer. In Chapter 6, a novel four-step<br />

method is presented to synthesize partially thiolated TMC with tailorable degrees of<br />

quaternization and thiolation. First, chitosan w<strong>as</strong> partially N-carboxylated with glyoxylic acid<br />

and sodium borohydride. Next, the remaining amines were quantitatively dimethylated with<br />

<strong>for</strong>maldehyde and sodium borohydride and then quaternized with iodomethane in N-<br />

methylpyrrolidone. Subsequently, these partially carboxylated TMCs, dissolved in water, were<br />

reacted with cystamine at pH 5.5 using EDC <strong>as</strong> coupling agent. After addition of dithiothreitol<br />

<strong>for</strong> disulfide reduction and dialysis, thiolated TMCs were obtained varying in degree of<br />

quaternization (25-54%) and degree of thiolation (5-7%) <strong>as</strong> determined with 1 H-NMR and<br />

Ellman’s <strong>as</strong>say. Gel permeation chromatography with light scattering detection indicated<br />

limited intermolecular crosslinking. All thiolated TMCs showed rapid oxidation to yield<br />

disulfide crosslinked TMC at pH 7.4 while the thiolated polymers were rather stable at pH 4.0.<br />

Using Calu-3 cells, XTT and LDH cell viability tests showed a slight reduction in cytotoxicity <strong>for</strong><br />

thiolated TMCs <strong>as</strong> compared to the non-thiolated polymers with similar DQs. Positively<br />

charged nanoparticles loaded with fluorescently labeled ovalbumin were made from thiolated<br />

TMCs and thiolated hyaluronic acid. The stability of these particles w<strong>as</strong> confirmed in 0.8 M<br />

NaCl, in contr<strong>as</strong>t to particles made from non-thiolated polymers which dissociated under these<br />

conditions demonstrating that the particles were held together by intermolecular disulfide<br />

bonds.<br />

As shown in Chapter 6, the physical stability of polyelectrolyte nanoparticles composed of<br />

trimethyl chitosan (TMC) and hyaluronic acid (HA) is limited in physiological conditions. This<br />

may adversely affect the favorable adjuvant effects of particulate systems <strong>for</strong> n<strong>as</strong>al and<br />

intradermal immunization. There<strong>for</strong>e, in Chapter 7 the effect of covalent stabilization of<br />

antigen-loaded TMC/HA nanoparticles by disulfides on their immunogenicity is evaluated.<br />

Furthermore, remaining thiols on the surface of these particles allow PEGylation which may<br />

result in additional beneficial effects by hindering interactions with the extracellular matrix<br />

(intradermal) and/or enhanced muco-adhesion (intran<strong>as</strong>al). Ovalbumin (OVA) loaded,<br />

covalently stabilized nanoparticles were prepared with thiolated TMC and thiolated HA via<br />

ionic gelation followed by spontaneous disulfide <strong>for</strong>mation after incubation at pH 7.4 and 37°C.<br />

175


Chapter 8<br />

Also, PEG maleimide w<strong>as</strong> coupled to the remaining thiol-moieties on the particles to shield the<br />

surface charge.<br />

TMC/HA nanoparticles had a size of around 250-350 nm, a positive zeta potential and OVA<br />

<strong>as</strong>sociation efficiencies up to 60%. PEGylation resulted in a slight reduction of zeta potential<br />

and a minor incre<strong>as</strong>e in particle size. Stabilized TMC-S-S-HA particles (PEGylated or not)<br />

showed superior stability in saline solutions compared to non-stabilized TMC/HA particles<br />

(composed of nonthiolated polymers), but readily disintegrated when a disulfide reducing<br />

agent w<strong>as</strong> introduced. In both the n<strong>as</strong>al and intradermal immunization study, OVA-loaded<br />

stabilized TMC-S-S-HA particles demonstrated superior immunogenicity compared to nonstabilized<br />

particles. For intran<strong>as</strong>al immunization, PEGylation completely abolished the positive<br />

effects of stabilization and it had no additional effect when used <strong>for</strong> intradermal vaccination. In<br />

conclusion, stabilization of the TMC/HA particulate system enhances its immunogenicity in<br />

n<strong>as</strong>al and intradermal vaccination, however, PEGylation of these stabilized particles had no<br />

beneficial effects.<br />

The novel synthetic methods to obtain structural variants of TMC presented in this thesis<br />

allowed tailorability of the degrees of quaternization and acetylation and excluded the<br />

introduction of other alterations such <strong>as</strong> O-methylation and polymer chain scission.<br />

Additionally, thiol-moieties were introduced in a controllable manner. This tailorability of TMC<br />

provided the possibility to properly establish structure-activity relationships in in vitro<br />

biological <strong>as</strong>says and in in vivo (n<strong>as</strong>al) vaccination studies. Also, mechanistic insight into the<br />

mode of action of TMC w<strong>as</strong> acquired by investigating several potentially crucial steps in n<strong>as</strong>al<br />

vaccination. Finally, a promising new, covalently stabilized, polymeric carrier system w<strong>as</strong><br />

introduced b<strong>as</strong>ed upon the <strong>for</strong>mation of intracellularly degradable disulfides.<br />

Discussion and Future perspectives<br />

TMC-b<strong>as</strong>ed particulate systems <strong>for</strong> n<strong>as</strong>al vaccine delivery. Although already synthesized<br />

from chitosan in the mid-80s by Muzzarelli [1] and later by Domard [2], trimethyl chitosan<br />

(TMC) h<strong>as</strong> not been used <strong>for</strong> mucosal vaccination until the beginning of this millennium [3].<br />

Even more recently, in 2007, Amidi et al. demonstrated <strong>for</strong> the first time the potency of TMCtripolyphosphate<br />

ionically crosslinked nanoparticulate systems in an intran<strong>as</strong>al vaccination<br />

study with influenza subunit antigen [4]. Although this system showed promising results <strong>as</strong><br />

176


Summary and Future Perspectives<br />

n<strong>as</strong>al vaccine delivery system, it w<strong>as</strong> anticipated that much could be gained by optimizing, in a<br />

controllable way, the chemical structure of TMC. In particular the degree of quaternization<br />

(DQ) of TMC seemed important since this is the polymer’s main determinant <strong>for</strong> charge<br />

density. Also, in vitro and in vivo studies with <strong>for</strong>mulation of TMC and both protein and low<br />

molecular weight compounds suggested an optimal DQ of 40-50% <strong>for</strong> transepithelial delivery<br />

[5-10].<br />

In this thesis we show that the DQ, but also the exclusion of O-methylation, h<strong>as</strong> no effect on<br />

the immunogenicity of TMC when used <strong>as</strong> coating on whole inactivated influenza virus (WIV).<br />

Incre<strong>as</strong>ing another variable, the degree of acetylation (DAc), resulted even in a reduction of the<br />

beneficial effects of coating WIV with TMC in n<strong>as</strong>al immunization in mice. So, looking from an<br />

immunological perspective, despite all ef<strong>for</strong>ts put in the development of novel synthetic routes<br />

to obtain structural variants of TMC, no improvements in immunological efficacy were<br />

achieved. However, looking from a pharmaceutical perspective, the availability of<br />

reproducible, well-characterized TMCs that can be structurally tailored in a controllable<br />

manner without introducing side reactions, may be considered <strong>as</strong> a major step <strong>for</strong>ward.<br />

Additionally, these structural variants (DQ, DOM, DAc) also influence others physicochemical<br />

characteristics and superior stability of O-methyl free TMC-WIV <strong>for</strong>mulations w<strong>as</strong> found [11].<br />

The covalently stabilized polyelectrolyte nanocomplexes of TMC-SH and HA-SH offer a novel<br />

vaccine delivery plat<strong>for</strong>m with high versatility. To mention, positively charged antigens may be<br />

entrapped using HA-SH <strong>for</strong> initial complexation and TMC-SH <strong>as</strong> crosslinker and also covalent<br />

coupling of antigens, adjuvants or targeting ligands is e<strong>as</strong>ily realizable on remaining free thiol<br />

moieties (<strong>as</strong> is already demonstrated <strong>for</strong> PEG maleimide). Active targeting to lectin receptors<br />

or integrins present on microfold (M)-cells can improve the immunogenicity of a <strong>for</strong>mulation<br />

[12] and thus coupling of sialic acid, galactose residues or specific antibodies (like IgA) on the<br />

particles’ surface may be an attractive next step to enhance binding and uptake of the particles<br />

by antigen presenting cells. Interestingly, these and similar sugar-moieties like mannose and<br />

N-acetyl glucosamine (GlucNAc) are also <strong>as</strong>sociated with targeting to dendritic cells (DCs).<br />

Indeed, the activation of DCs via GlucNAc is described in this thesis with TMC-RA (containing<br />

high GlucNAc) showing promising results on human DCs (Chapter 5B). Furthermore the type<br />

of immune response may be steered by incorporation of CpG motifs (usually towards Th1) and<br />

lipopolysaccharides (LPS, usually towards Th2) or other immuno-modulators. Finally, <strong>as</strong><br />

thiolation of TMC enhances its muco-adhesiveness [13], the use of TMC-SHs in ‘conventional’<br />

TMC-TPP particles may further incre<strong>as</strong>e n<strong>as</strong>al residence time of the particles resulting in<br />

better antigen delivery compared to nonthiolated TMCs.<br />

177


Chapter 8<br />

This thesis further shows the limitations of in vitro predictability and/or correlation <strong>for</strong> in<br />

vivo outcome <strong>as</strong> large differences were observed in vitro <strong>for</strong> toxicity and capacity to open<br />

cellular tight junctions <strong>for</strong> polymers that showed no differences in vivo n<strong>as</strong>al vaccination<br />

studies. Additionally, extensive investigation in various potentially crucial steps <strong>for</strong> n<strong>as</strong>al<br />

vaccination with both in vitro and in vivo models did not reveal major differences between an<br />

effective (TMC-WIV) and ineffective (TMC-RA-WIV) <strong>for</strong>mulation. Apparently, displaying<br />

cationic charges, muco-adhesiveness and improving interaction between the <strong>for</strong>mulation and<br />

the epithelial barrier are not sufficient <strong>for</strong> effective n<strong>as</strong>al vaccination. This emph<strong>as</strong>izes the<br />

need <strong>for</strong> even more detailed knowledge of TMC’s (and that of other adjuvants) mode of action.<br />

Only then, in a rational way, in vitro models can be developed that adequately predict in vivo<br />

outcome and ultimately minimize the use of animal models.<br />

Another issue that is raised by this thesis is the discrepancy in results between human and<br />

murine in vitro models (<strong>as</strong> w<strong>as</strong> observed <strong>for</strong> differences in the immuno-stimulatory effects of<br />

TMC-RA in human and murine DCs). Again, in depth knowledge on the mode of action of TMC<br />

is crucial <strong>for</strong> the predictability of murine in vivo/in vitro results <strong>for</strong> human outcome. E.g. if<br />

TMC’s adjuvant effect is achieved through improved muco-adhesion, minor differences may be<br />

anticipated between mice and men. In contr<strong>as</strong>t, M-cells are more abundant in murine n<strong>as</strong>al<br />

cavity and murine and human M-cells may express different receptors <strong>as</strong> may antigen<br />

presenting cells [14]; if TMC-b<strong>as</strong>ed systems accomplish their effect mainly via these<br />

mechanisms, limited predictability can be expected from studies with mice models.<br />

As can be concluded from the above, to choose just one type of TMC(-system) <strong>for</strong> further<br />

research <strong>for</strong> n<strong>as</strong>al vaccine development is not recommended: in particular, the effects of<br />

incre<strong>as</strong>ing the GlucNAc content are not fully understood yet and the full potential of the<br />

stabilized TMC-S-S-HA particles needs further investigation. Animal models (such <strong>as</strong> ferrets or<br />

transgenic mice) that mimic the human immune responses in a better extent may elaborate<br />

whether a future application of TMC <strong>for</strong> n<strong>as</strong>al vaccination in man is fe<strong>as</strong>ible.<br />

Other potential applications <strong>for</strong> novel TMCs. Next to mucosal vaccination, TMCs are<br />

frequently studied <strong>for</strong> mucosal delivery of pharmaceutically active proteins and low molecular<br />

weight compounds. Interestingly, <strong>for</strong> these applications, good correlations are found between<br />

the capacity of a polymer to open tight junctions (<strong>as</strong> me<strong>as</strong>ured in a transepithelial electrical<br />

resistance (TEER) <strong>as</strong>say) and in vivo uptake. This implies that, in particular, O-methyl free<br />

TMCs can improve the mucosal uptake compared to ‘conventional’ TMCs <strong>as</strong> they showed a<br />

better, reversible reduction of the TEER. Further studies should be carried out to confirm this.<br />

178


Summary and Future Perspectives<br />

TMCs have been used <strong>for</strong> DNA delivery with some encouraging results. However,<br />

introductory studies with O-methyl free TMCs <strong>for</strong> DNA transfection showed low transfection<br />

efficiencies probably due to poor endosomal escape. Interestingly, thiolation of TMC may not<br />

only enhance the transfection potential of DNA [15] but also gene silencing via delivery of<br />

siRNA into the cytosol (Varhouchi et al, manuscript submitted). Moreover, the stabilized TMC-<br />

S-S-HA carrier system shows excellent properties <strong>for</strong> DNA/siRNA delivery demonstrating<br />

enhanced stability under physiological conditions but they readily dissociate in a reductive<br />

environment (like in the cytosol). Depending on the target cell or the application route<br />

(systemic or local) PEG or targeting ligands (such <strong>as</strong> nano- or antibodies) can be attached. This<br />

makes TMC-SH and in particular the TMC-S-S-HA carrier systems highly interesting <strong>for</strong><br />

siRNA/DNA delivery and they are currently being evaluated <strong>for</strong> this purpose in our<br />

Department.<br />

PEGylated nanoparticles (e.g. PEG-liposomes) are widely studied <strong>for</strong> systemic delivery of<br />

proteins, cytokines and anti-cancer and immunomodulatory drugs to area’s of tumor growth<br />

or inflammation <strong>as</strong> a result of the enhanced permeability and retention effect in those tissues.<br />

The PEGylated, stabilized TMC-S-S-HA system may be a valuable alternative to currently used<br />

systems because of its simple preparation method and its potentially high loading capacity. As<br />

TMCs with a high DAc are readily degraded by lysozyme, an interesting option may be the<br />

incorporation of this enzymatically-degradable TMC into particles. In environments with high<br />

lysozyme concentrations, <strong>for</strong> example in inflammated or infectioned tissues (<strong>as</strong> lysozyme is<br />

excreted by macrophages), these particles should disintegrate and rele<strong>as</strong>e their contents.<br />

Finally, the thiolated polymers can be used <strong>for</strong> covalently linked layer-by-layer technologies<br />

[16, 17] and/or the preparation of covalently stabilized hydrogels with acrylated or<br />

methacrylated polymers (b<strong>as</strong>ed upon Michael addition) [18].<br />

TMC h<strong>as</strong> only been used <strong>for</strong> a decade or so <strong>for</strong> mucosal vaccination and other applications<br />

and the l<strong>as</strong>t couple of years interest h<strong>as</strong> risen rapidly due to its promising results. However, to<br />

look into the future of TMC much can be learned from its older precursor, chitosan. This<br />

polymer h<strong>as</strong> been investigated <strong>for</strong> over <strong>for</strong>ty years in various biomedical, nutritional and<br />

technological fields and widely studied <strong>for</strong> toxicity, biocompatibility and biodegradation. In<br />

contr<strong>as</strong>t to TMC, it h<strong>as</strong> only two (major) variables, the molecular weight and the DAc and their<br />

effects on the physicochemical and biological properties of chitosan have been thoroughly<br />

studied. Despite all this scientific ef<strong>for</strong>t, the FDA h<strong>as</strong> not yet approved chitosan <strong>as</strong> GRAS<br />

(generally regarded <strong>as</strong> safe) material which hinders its application in (biomedical) products<br />

179


Chapter 8<br />

[19]. Likely, in the p<strong>as</strong>t manufacturers disliked the difficulty of patentability <strong>for</strong> chitosanproducts<br />

and there<strong>for</strong>e little stress on regulatory authorities w<strong>as</strong> applied. Recently, however,<br />

pressure is build by researchers and clinicians to revise this regulatory perspective <strong>as</strong><br />

chitosan-like substances are highly awaited in the various biomedical fields. For instance, a<br />

chitosan-b<strong>as</strong>ed delivery plat<strong>for</strong>m is exploited by Archimedes (UK-b<strong>as</strong>ed pharmaceutical<br />

company) and this system is currently evaluated in ph<strong>as</strong>e I clinical trails <strong>for</strong> the n<strong>as</strong>al delivery<br />

of granisetron <strong>as</strong> anti-emetic (stated on the Archimedes website by December 2009). If<br />

encouraging results are obtained, it can be anticipated that approval by regulatory authorities<br />

will follow. As TMC h<strong>as</strong> superior characteristics to chitosan, TMC may shortly go behind.<br />

Further convincing studies showing TMC’s potential may speed up this process <strong>as</strong> may the<br />

availability of well-characterized and defined polymers described in this thesis.<br />

180


Summary and Future Perspectives<br />

References<br />

1. R. A. A. Muzzarelli and F. Tanfani. The N-permethylation of chitosan and the preparation of<br />

N-trimethyl chitosan iodide. Carbohydr Polym 5: 297-307 (1985).<br />

2. A. Domard, M. Rinaudo, and C. Terr<strong>as</strong>sin. New method <strong>for</strong> the quaternization of chitosan. Int J<br />

Bio Macromol 8: 105-107 (1986).<br />

3. I. M. Van der Lubben, J. C. Verhoef, M. M. Fretz, O. Van, I. Mesu, G. Kersten, and H. E.<br />

Junginger. <strong>Trimethyl</strong> chitosan chloride (TMC) <strong>as</strong> a novel excipient <strong>for</strong> oral and n<strong>as</strong>al<br />

immunisation against diphtheria. S.T.P. <strong>Pharma</strong> Sciences 12: 235-242 (2002).<br />

4. M. Amidi, S. G. Romeijn, J. C. Verhoef, H. E. Junginger, L. Bungener, A. Huckriede, D. J. A.<br />

Crommelin, and W. Jiskoot. N-<strong>Trimethyl</strong> chitosan (TMC) nanoparticles loaded with influenza<br />

subunit antigen <strong>for</strong> intran<strong>as</strong>al vaccination: Biological properties and immunogenicity in a mouse<br />

model. Vaccine 25: 144-153 (2007).<br />

5. F. Chen, Z. R. Zhang, F. Yuan, X. Qin, M. Wang, and Y. Huang. In vitro and in vivo study of<br />

N-trimethyl chitosan nanoparticles <strong>for</strong> oral protein delivery. Int J Pharm 349: 226-233 (2008).<br />

6. G. Di Colo, S. Burgal<strong>as</strong>si, Y. Zambito, D. Monti, and P. Chetoni. Effects of different N-<br />

trimethyl chitosans on in vitro/in vivo ofloxacin transcorneal permeation. J Pharm Sci 93: 2851-<br />

2862 (2004).<br />

7. J. H. Hamman, C. M. Schultz, and A. F. Kotze. N-trimethyl chitosan chloride: Optimum degree<br />

of quaternization <strong>for</strong> drug absorption enhancement across epithelial cells. Drug Develop Ind<br />

Pharm 29: 161-172 (2003).<br />

8. J. H. Hamman, M. Stander, and A. F. Kotze. Effect of the degree of quaternisation of N-<br />

trimethyl chitosan chloride on absorption enhancement: In vivo evaluation in rat n<strong>as</strong>al epithelia.<br />

Inter J Pharm 232: 235-242 (2002).<br />

9. A. F. Kotze, M. M. Thanou, H. L. Luessen, A. B. G. De Boer, J. C. Verhoef, and H. E.<br />

Junginger. Effect of the degree of quaternization of N-trimethyl chitosan chloride on the<br />

permeability of intestinal epithelial cells (Caco-2). Eur J Pharm Biopharm 47: 269-274 (1999).<br />

10. M. M. Thanou, A. F. Kotze, T. Scharringhausen, H. L. Lueßen, A. G. De Boer, J. C. Verhoef,<br />

and H. E. Junginger. Effect of degree of quaternization of N-trimethyl chitosan chloride <strong>for</strong><br />

enhanced transport of hydrophilic compounds across intestinal Caco-2 cell monolayers. J<br />

Control Rele<strong>as</strong>e 64: 15-25 (2000).<br />

11. N. Hagenaars. Towards an intran<strong>as</strong>al influenza vaccine - B<strong>as</strong>ed on whole inactivated influenza<br />

virus with N,N,N-trimethylchitosan <strong>as</strong> adjuvant. Utrecht University, Utrecht (2010).<br />

12. B. Slütter, N. Hagenaars, and W. Jiskoot. Rational design of n<strong>as</strong>al vaccines. J Drug Target 16: 1-17<br />

(2008).<br />

13. L. Yin, J. Ding, C. He, L. Cui, C. Tang, and C. Yin. Drug permeability and mucoadhesion<br />

properties of thiolated trimethyl chitosan nanoparticles in oral insulin delivery. Biomaterials 30:<br />

5691-5700 (2009).<br />

14. S. K. Singh, J. Stephani, M. Schaefer, H. Kalay, J. J. García-Vallejo, J. den Haan, E. Saeland, T.<br />

Sparw<strong>as</strong>ser, and Y. van Kooyk. Targeting glycan modified OVA to murine DC-SIGN<br />

transgenic dendritic cells enhances MHC cl<strong>as</strong>s I and II presentation. Mol Immunol 47: 164-174<br />

(2009).<br />

15. X. Zhao, L. Yin, J. Ding, C. Tang, S. Gu, C. Yin, and Y. Mao. Thiolated trimethyl chitosan<br />

nanocomplexes <strong>as</strong> gene carriers with high in vitro and in vivo transfection efficiency. J Control<br />

Rele<strong>as</strong>e 144: 46-54 (2010).<br />

16. B. G. De Geest, G. B. Sukhorukov, and H. Möhwald. The pros and cons of polyelectrolyte<br />

capsules in drug delivery. Exp Opin Drug Deliv 6: 613-624 (2009).<br />

17. A. Szarpak, D. Cui, F. Dubreuil, B. G. De Geest, L. J. De Cock, C. Picart, and R. Auzély-Velty.<br />

Designing hyaluronic acid-b<strong>as</strong>ed layer-by-layer capsules <strong>as</strong> a carrier <strong>for</strong> intracellular drug<br />

delivery. Biomacromolecules 11: 713-720 (2010).<br />

18. R. Censi, P. J. Fieten, P. di Martino, W. E. Hennink, and T. Vermonden. In Situ Forming<br />

Hydrogels by Tandem Thermal Gelling and Michael Addition Reaction between<br />

181


Chapter 8<br />

Thermosensitive Triblock Copolymers and Thiolated Hyaluronan. Macromolecules 43: 5771-5778<br />

(2010).<br />

19. T. Kean and M. Thanou. Biodegradation, biodistribution and toxicity of chitosan. Adv Drug<br />

Deliv Rev 62: 3-11 (2010).<br />

182


APPENDICES


Affiliations of Collaborating Authors<br />

Affiliations of collaborating authors:<br />

Maryam Amidi<br />

Department of <strong>Pharma</strong>ceutics, Utrecht Institute <strong>for</strong> <strong>Pharma</strong>ceutical Sciences, Utrecht<br />

University, Utrecht, The Netherlands<br />

Suzanne Bal<br />

Division of Drug Delivery Technology, Leiden/Amsterdam Center <strong>for</strong> Drug Research, Leiden<br />

University, Leiden, The Netherlands<br />

Han van den Bosch<br />

Nobilon, part of Merck, Boxmeer, The Netherlands<br />

Joke Bouwstra<br />

Division of Drug Delivery Technology, Leiden/Amsterdam Center <strong>for</strong> Drug Research, Leiden<br />

University, Leiden, The Netherlands<br />

Sven Bruins<br />

Department of Molecular Cell Biology and Immunology, Vrije University Medical Center,<br />

Amsterdam, The Netherlands<br />

Thom<strong>as</strong> van Es<br />

Department of Molecular Cell Biology and Immunology, Vrije University Medical Center,<br />

Amsterdam, The Netherlands<br />

Ethlinn van Gaal<br />

Department of <strong>Pharma</strong>ceutics, Utrecht Institute <strong>for</strong> <strong>Pharma</strong>ceutical Sciences, Utrecht<br />

University, Utrecht, The Netherlands<br />

Harrie Glansbeek<br />

Nobilon, part of Merck, Boxmeer, The Netherlands<br />

Niels Hagenaars<br />

Department of <strong>Pharma</strong>ceutics, Utrecht Institute <strong>for</strong> <strong>Pharma</strong>ceutical Sciences, Utrecht<br />

University, Utrecht, The Netherlands<br />

Jacco Heldens<br />

Nobilon, part of Merck, Boxmeer, The Netherlands<br />

Wim Hennink<br />

Department of <strong>Pharma</strong>ceutics, Utrecht Institute <strong>for</strong> <strong>Pharma</strong>ceutical Sciences, Utrecht<br />

University, Utrecht, The Netherlands<br />

Wim Jiskoot<br />

Division of Drug Delivery Technology, Leiden/Amsterdam Center <strong>for</strong> Drug Research, Leiden<br />

University, Leiden, The Netherlands<br />

P<strong>as</strong>cal de Jong<br />

Department of <strong>Pharma</strong>ceutics, Utrecht Institute <strong>for</strong> <strong>Pharma</strong>ceutical Sciences, Utrecht<br />

University, Utrecht, The Netherlands<br />

185


Affiliations of Collaborating Authors<br />

Enrico M<strong>as</strong>trobattista<br />

Department of <strong>Pharma</strong>ceutics, Utrecht Institute <strong>for</strong> <strong>Pharma</strong>ceutical Sciences, Utrecht<br />

University, Utrecht, The Netherlands<br />

Imke Mooren<br />

Animal Service Department, Intervet Animal Health part of Merck, Boxmeer, The Netherlands<br />

Ivo Que<br />

Department of Endocrinology and Metabolic Dise<strong>as</strong>es, Leiden University Medical Center,<br />

Leiden, The Netherlands<br />

Elly van Riet<br />

Division of Drug Delivery Technology, Leiden/Amsterdam Center <strong>for</strong> Drug Research, Leiden<br />

University, Leiden, The Netherlands<br />

Bram Slütter<br />

Division of Drug Delivery Technology, Leiden/Amsterdam Center <strong>for</strong> Drug Research, Leiden<br />

University, Leiden, The Netherlands<br />

Mies van Steenbergen<br />

Department of <strong>Pharma</strong>ceutics, Utrecht Institute <strong>for</strong> <strong>Pharma</strong>ceutical Sciences, Utrecht<br />

University, Utrecht, The Netherlands<br />

Steffen van der Wal<br />

Department of Medicinal Chemistry and Chemical Biology, Utrecht Institute <strong>for</strong> <strong>Pharma</strong>ceutical<br />

Sciences, Utrecht University, Utrecht, The Netherlands<br />

186


List of Abbreviations<br />

List of abbreviations:<br />

APC<br />

CpG<br />

CS<br />

CS-RA<br />

DAc<br />

DC<br />

D carb<br />

DDM<br />

DLS<br />

DMC<br />

DOM<br />

D thiol<br />

DTT<br />

DQ<br />

EDC<br />

ELISA<br />

FACS<br />

FCS<br />

FDA<br />

FITC<br />

GPC<br />

HA<br />

HA-SH<br />

HBSS<br />

HEPES<br />

ID<br />

IgG<br />

IL<br />

i.m.<br />

i.n.<br />

kDa<br />

LDH<br />

antigen presenting cell<br />

cytosine guanine dinucleotide<br />

chitosan<br />

re-acetylated chitosan<br />

degree of acetylation<br />

dendritic cell<br />

degree of carboxylation<br />

degree of dimethylation<br />

dynamic light scattering<br />

dimethyl chitosan<br />

degree of O-methylation<br />

degree of thiolation<br />

dithiothreitol<br />

degree of quaternization<br />

1-ethyl-3-(3-dimethylaminopropyl) carbodiimide)<br />

enzyme-linked immunosorbent <strong>as</strong>say<br />

fluorescence activated cell sorter<br />

fetal calf serum<br />

food and drug administration<br />

fluorescein isothiocyanate<br />

gel permeation chromatography<br />

hyaluronic acid<br />

thiolated hyaluronic acid<br />

Hank’s balanced salt solution<br />

N-2-hydroxyethylpiperazine-N’-2-ethanesulfonic acid<br />

intradermal<br />

immunoglobulin G<br />

interleukin<br />

intramuscular<br />

intran<strong>as</strong>al<br />

kilodalton<br />

lactate dehydrogen<strong>as</strong>e<br />

187


List of Abbreviations<br />

LPS<br />

lipopolysaccharide<br />

MALT<br />

mucosal <strong>as</strong>sociated lymphoid tissue<br />

MTT<br />

3-(4,5-Dimethylthiazol-2-yl)-2,5-diphenyltetrazolium<br />

bromide<br />

M n<br />

number average molecular weight<br />

M w<br />

weight average molecular weight<br />

NALT<br />

n<strong>as</strong>al <strong>as</strong>sociated lymphoid tissue<br />

NMP<br />

N-methyl-2-pyrrolidone<br />

NMR<br />

nuclear magnetic resonance<br />

OD<br />

optical density<br />

OVA<br />

ovalbumin<br />

PBS<br />

phosphate buffer saline<br />

PDI<br />

polydispersity index<br />

PEG<br />

poly(ethylene glycol)<br />

PEI<br />

polyethylenimine<br />

SDS<br />

sodium dodecyl sulphate<br />

sIgA<br />

secretory immunoglobulin A<br />

siRNA<br />

small interfering ribonucleic acid<br />

TEER<br />

trans epithelial electrical resistance<br />

Th1/Th2 T helper cell type 1/2<br />

TMC<br />

trimethyl chitosan<br />

TMC-COOH<br />

carboxylated trimethyl chitosan<br />

TMC-OM<br />

O-methylated trimethyl chitosan<br />

TMC-RA<br />

re-acetylated trimethyl chitosan<br />

TMC-SH<br />

thiolated trimethyl chitosan<br />

TNF-α<br />

tumor necrosis factor alfa<br />

TPP<br />

tripolyphosphate<br />

WIV<br />

whole inactivated influenza virus<br />

XTT<br />

sodium 3´-[1-(phenylaminocarbonyl)- 3,4-tetrazolium]-<br />

bis (4-methoxy-6-nitro) benzene sulfonicacid hydrate<br />

188


Curriculum Vitae<br />

Curriculum Vitae<br />

Rolf Verheul w<strong>as</strong> born on the 8 th of March 1980 in Oss, The<br />

Netherlands. After finishing pre-university education<br />

(Gymn<strong>as</strong>ium) at the Titus Brandsma Lyceum in Oss in 1998,<br />

he started studying pharmacy at Utrecht University. In 2003<br />

he obtained his M<strong>as</strong>ter’s degree in pharmaceutical sciences<br />

cum laude, followed by his pharmacist degree (PharmD) in<br />

2005. During his study, he completed a research traineeship<br />

at the University of British Columbia, Vancouver, Canada under the supervision of prof. dr.<br />

Cullis <strong>as</strong> well <strong>as</strong> internships at the National Institute <strong>for</strong> Public Health and the Environment<br />

(RIVM) and various public and hospital pharmacies. Also, under the framework of the honors<br />

program in pharmacy, he followed a minor in medical anthropology and sociology at the<br />

University of Amsterdam. Thereafter, he worked <strong>as</strong> a laboratory pharmacist in the hospital<br />

pharmacy of the Academic Medical Center, Amsterdam, <strong>for</strong> one year. In November 2006 he<br />

started his PhD research program at the department of <strong>Pharma</strong>ceutics, Utrecht University<br />

under the supervision of prof. dr. ir. W.E. Hennink and prof. dr. W. Jiskoot. The results of his<br />

work are presented in this thesis.<br />

189


List of Publications<br />

List of publications:<br />

RJ Verheul, M Amidi, S van der Wal, E van Riet, W Jiskoot, WE Hennink. Synthesis,<br />

characterization and in vitro biological properties of O-methyl free N,N,N,-trimethylated<br />

chitosan. Biomaterials 29, 3642-3649 (2008).<br />

RJ Verheul, M Amidi, M van Steenbergen, E van Riet, W Jiskoot, WE Hennink. Influence of the<br />

degree of acetylation on the enzymatic degradation and in vitro biological properties of<br />

trimethylated chitosans. Biomaterials 30, 3129-3135 (2009).<br />

N Hagenaars, E M<strong>as</strong>trobattista, RJ Verheul, I Mooren, HL Glansbeek, JGM Heldens, H van den<br />

Bosch, W Jiskoot. Physicochemical and immunological characterization of N,N,N-trimethyl<br />

chitosan-coated whole inactivated influenza virus vaccine <strong>for</strong> intran<strong>as</strong>al administration.<br />

<strong>Pharma</strong>ceutical Research 26,1353-1364 (2009).<br />

RJ Verheul, N Hagenaars, I Mooren, PH de Jong, E M<strong>as</strong>trobattista, HL Glansbeek, JGM Heldens,<br />

H van den Bosch, WE Hennink, W Jiskoot. Relationship between structure and adjuvanticity<br />

of N,N,N-trimethyl chitosan (TMC) structural variants in a n<strong>as</strong>al influenza vaccine. Journal of<br />

Controlled Rele<strong>as</strong>e 140, 126-133 (2009).<br />

B Slütter, PC Soema, Z Ding, R Verheul, W Hennink, W Jiskoot. Conjugation of ovalbumin to<br />

trimethyl chitosan improves immunogenicity of the antigen. Journal of Controlled Rele<strong>as</strong>e<br />

143, 207-214 (2010).<br />

RJ Verheul, S van der Wal, WE Hennink. <strong>Tailorable</strong> thiolated trimethyl chitosans <strong>for</strong><br />

covalently stabilized nanoparticles. Biomacromolecules 11, 1965-1971 (2010).<br />

RJ Verheul, N Hagenaars, T van Es, E van Gaal, P de Jong, S Bruins, I Que, B Slütter, E<br />

M<strong>as</strong>trobattista, HL Glansbeek, JGM Heldens, H van den Bosch, WE Hennink, W Jiskoot. A stepby-step<br />

approach to study the influence of N-acetylation on the adjuvanticity of N,N,Ntrimethyl<br />

chitosan (TMC) in an intran<strong>as</strong>al whole inactivated influenza virus vaccine.<br />

Submitted.<br />

RJ Verheul¸ N Hagenaars, T van Es, S Bruins, B Slütter, WE Hennink, W Jiskoot. Maturation of<br />

human monocyte derived dendritic cells by trimethyl chitosan is correlated with its N-acetyl<br />

glucosamine (GlcNAc) content. Submitted <strong>as</strong> short communication.<br />

RJ Verheul, B Slütter, SM Bal, JA Bouwstra, W Jiskoot, WE Hennink. Covalently stabilized<br />

trimethyl chitosan-hyaluronic acid nanoparticles <strong>for</strong> n<strong>as</strong>al and intradermal vaccination.<br />

Submitted.<br />

SM Bal, B Slütter, RJ Verheul, JA Bouwstra, W Jiskoot. <strong>Adjuvant</strong>ed, antigen loaded N-trimethyl<br />

chitosan nanoparticles <strong>for</strong> n<strong>as</strong>al and intradermal vaccination: adjuvant- and site-dependent<br />

immunogenicity in mice. Submitted.<br />

AK Varkouhi, RJ Verheul, RM Schiffelers, T Lammers, G Storm, WE Hennink. Gene silencing<br />

activity of siRNA polyplexes b<strong>as</strong>ed on thiolated N,N,N-trimethylated chitosan. Submitted.<br />

190


NEDERLANDSE SAMENVAT<strong>TI</strong>NG<br />

EN<br />

DANKWOORD


Nederlandse Samenvatting<br />

Nederlandse Samenvatting<br />

Vaccinatie is het actief opwekken van een afweerreactie tegen een bepaalde<br />

lichaamsvreemde stof of eiwit (antigeen) van een ziekteverwekker. Hierdoor zal het<br />

immuunsysteem in het vervolg deze indringer snel herkenen en het kunnen verwijderen<br />

voordat het daadwerkelijk schade kan veroorzaken. Over de jaren heeft vaccinatie bewezen<br />

om het meest (kosten)effectieve instrument te zijn in de strijd tegen besmettelijke ziekten.<br />

Vaccins worden meestal gemaakt van levende, verzwakte of geïnactiveerde ziekteverwekkers<br />

(bijv. virussen of bacteriën) of gezuiverde immunogene eiwitten die afkomstig zijn van deze<br />

pathogenen. Hierdoor zullen vaccins zelf geen ziekte induceren, maar kunnen ze wel het<br />

immuunsysteem activeren. Tegenwoordig worden de meeste vaccins toegediend via een<br />

injectienaald. Echter, vanwege het risico van besmette naalden, met name in derde wereld<br />

landen, en de noodzaak van geschoold personeel voor een dergelijke toediening, is de<br />

belangstelling voor alternatieve vaccinatieroutes toegenomen.<br />

Hoofdstuk 1 bediscussieert de voor- en nadelen van de intramusculaire en alternatieve<br />

toedieningsroutes voor vaccins. In het bijzonder wordt immunisatie via de neus (intran<strong>as</strong>ale<br />

vaccinatie) besproken, waarmee relatief eenvoudig met bijvoorbeeld een neusspray of<br />

neusdruppels, naald-vrij en zonder getraind personeel een vaccin toegediend kan worden.<br />

Bovendien kan intran<strong>as</strong>ale vaccinatie zowel mucosale (op de slijmvliezen) als systemische (in<br />

het lichaam) immuunresponsen veroorzaken wat een verbeterde bescherming zou kunnen<br />

betekenen t.o.v. intramusculaire toediening. Aan de andere kant zijn, vanwege de natuurlijke<br />

barrières in de neus, antigeenafbraak en het bereiken van de antigeen presenterende cellen<br />

(APC's) over de slijm- en cellagen heen, belangrijke nadelen van n<strong>as</strong>ale vaccinatie (Figuur 1).<br />

Aangezien verzwakte, levende vaccins een aanzienlijk veiligheidsrisico met zich meebrengen<br />

voor mensen met een suboptimaal immuun systeem (b.v. kinderen, ouderen, patiënten met<br />

bepaalde ziektes als HIV), richt de ontwikkeling van vaccins zich nu vooral op het gebruik van<br />

gezuiverde, goed gekarakteriseerde antigene eiwitten. Echter, deze subunit vaccins zijn over<br />

het algemeen minder immunogeen en hebben een krachtig adjuvans (systeem) nodig om een<br />

adequate afweerreactie op te roepen. Door een muco-adhesief polymeer (dat ‘plakt’ aan de<br />

slijmlaag), zoals N,N,N-trimethylchitosan (TMC), te gebruiken bij intran<strong>as</strong>ale vaccinatie kan de<br />

immuunrespons tegen een antigeen worden verhoogd.<br />

193


Nederlandse Samenvatting<br />

Figuur 1. Schematische weergave van de cruciale stappen bij intran<strong>as</strong>ale vaccinatie. Na toediening zal<br />

het antigeen in contact komen met de oppervlaktes van het neusslijmvlies (1) waarna het vervolgens<br />

getransporteerd wordt over het n<strong>as</strong>ale epitheel (2) en opgenomen wordt door de antigeen<br />

presenterende cellen (APCs) zoals dendritische cellen (DCs) (3). Deze APCs zullen na activering o.a. tot<br />

de aanmaak van antilichamen overgaan (4).<br />

Vermoedelijk werkt TMC, een in water-oplosbaar chitosan derivaat, door de verblijftijd in de<br />

neus te verlegen en/of het contact tussen het antigeen en het mucosale oppervlak te<br />

verbeteren. TMC (Figuur 2) is eigenlijk een verzamelnaam voor een groep van polymeren met<br />

veel overeenkomsten in de chemische structuur maar met specifieke verschillen (zoals de<br />

ketenlengte, ladingsdichtheid en mate van methylering en acetylering). Het gecontroleerd<br />

kunnen variëren van deze structurele elementen maakt het mogelijk beter inzicht te krijgen in<br />

de afzonderlijke bijdrage van deze groepen op de fysisch-chemische en biologische<br />

eigenschappen van TMC. Ook kunnen nieuwe funtionaliteiten worden geïntroduceerd, zoals<br />

194


Nederlandse Samenvatting<br />

thiolgroepen die kunnen bijdragen aan het verbeteren en optimaliseren van de eigenschappen<br />

van TMC.<br />

Chitosan, en zijn derivaten als TMC, zijn in de vorm van nano- of microdeeltjes effectiever in<br />

het induceren van een immuunrespons dan een ‘gewone’ polymeeroplossing. Nano-gelering,<br />

gebruikmakend van electrostatische interacties, is een eenvoudige, veel gebruikte methode om<br />

nanodeeltjes van TMC te maken. Door negatief geladen (macro)moleculen druppelsgewijs<br />

onder roeren aan de TMC oplossing toe te voegen, vormen zich spontaan nanodeeltjes in een<br />

buffer met lage ionsterkte. Echter, de fysisch-chemische stabiliteit en de immunogeniciteit van<br />

deze met antigeen beladen complexen zijn afhankelijk van de eigenschappen van de gebruikte<br />

componenten. Mogelijk kunnen de momenteel toegep<strong>as</strong>te systemen nog worden verbeterd.<br />

Bovendien, hoewel het werkingsmechanisme van TMC doorgaans wordt toegeschreven aan de<br />

muco-adhesieve en penetratie verbeterende eigenschappen, ontbreekt er gedetailleerde<br />

kennis over het mechanisme (zo is bijvoorbeeld het directe effect van TMC op APCs niet<br />

bekend). Concluderend kan gesteld worden dat er zijn vele mogelijkheden voor het<br />

optimaliseren van de structuur van TMC en de op TMC geb<strong>as</strong>eerde nanodeeltjes. Tegelijkertijd<br />

zal een beter inzicht in het werkingsmechanisme van TMC tot een rationeel ontwerp van<br />

dergelijke adjuvantia kunnen leiden.<br />

Figuur 2. Schematische weergave van de structurele variaties van TMC. TMC kan variëren in de mate<br />

van acetylering (blok x), quaternisering (blok y) en O-methylering (blok z). Deze variaties kunnen<br />

willekeurig over het polymeer verdeeld zijn.<br />

Het doel van dit proefschrift is om synthese-routes te ontwikkelen waarmee op een<br />

controleerbare wijze de chemische structuur van TMC kan worden gevarieerd. Verder worden<br />

er nieuwe groepen geïntroduceerd, zoals thiolen, die mogelijk de eigenschappen van TMC nog<br />

195


Nederlandse Samenvatting<br />

verder kunnen verbeteren. Op deze manier kunnen structuur-activiteit-relaties worden<br />

onderzocht in zowel in vitro studies als in in vivo (n<strong>as</strong>ale) vaccinatie studies.<br />

De huidige methode om TMC te synthetiseren veroorzaakt, na<strong>as</strong>t de gewenste trimethylering<br />

van de vrije amines, ook O-methylering op de hydroxylgroepen en ketenbreuk van het<br />

polymer. Omdat deze nevenreacties van invloed kunnen zijn op de eigenschappen van het<br />

polymeer, is er een behoefte aan een syntheseroute waarbij deze ongewenste nevenreacties<br />

niet optreden en waarbij de mate van quaternisering (oftewel het percentage amine groepen<br />

dat getrimethyleerd is (DQ)) vrijelijk gevarieerd en gecontroleerd kan worden.<br />

In Hoofdstuk 2 werden met behulp van een twee-staps methode O-methyl vrije TMCs met<br />

variërende DQs gesynthetiseerd. Eerst werd chitosan kwantitatief gedimethyleerd met behulp<br />

van mierenzuur en <strong>for</strong>maldehyde. Vervolgens kon TMC worden verkregen door dit<br />

dimethylchitosan te laten reageren met een overmaat aan iodomethaan. Door de reactietijd te<br />

variëren, konden TMCs met verschillende DQs (22-68%) worden gesynthetiseerd. Deze TMCs<br />

bleken geen waarneembare O-methylering (bepaald met 1 H-NMR) te hebben en uit een geringe<br />

verhoging van het molecuulgewicht met toenemende DQ (bepaald met gel permeatie<br />

chromatografie of GPC) kon geconcludeerd worden dat zich geen noemenswaardige<br />

ketenbreuk had voorgedaan tijdens de synthese. De oplosbaarheid in water met een pH van 7<br />

w<strong>as</strong> voor O-methyl vrij TMC met een DQ 30% losten uitstekend op in water met<br />

een pH van 7. O-methyl vrije TMCs gaven een grotere daling in de trans-epitheliale elektrische<br />

weerstand (TEER) van een caco-2 monolaag dan O-gemethyleerde TMCs. Door het meten van<br />

de TEER wordt een idee verkregen in hoeverre een polymeer in staat is de intercellulaire<br />

ruimtes tussen epitheelcellen te openen waardoor het transport van stoffen over deze cellaag<br />

gemakkelijker wordt. Een daling in de TEER geeft aan dat de weerstand afneemt en dus dat de<br />

cellaag permeabeler wordt. Ook werd met de verhoging van de DQ, een stijging van de<br />

cytotoxiciteit (m.b.v. een MTT test) en celmembraan-permeabiliteit (m.b.v. een LDH test)<br />

waargenomen. Dit hoofdstuk toont duidelijk aan dat zowel de DQ als de aan/afwezigheid van<br />

O-methylering een belangrijke invloed hebben op de fysische en in vitro biologische<br />

eigenschappen van TMC.<br />

De invloed van een andere variatie in de moleculaire structuur, de mate van acetylering<br />

(DAc), op de in vitro enzymatische afbraak en biologische eigenschappen van TMC is<br />

onderzocht in hoofdstuk 3. TMCs met een DAc variërend van 11 tot 55% werden<br />

196


Nederlandse Samenvatting<br />

gesynthetiseerd met behulp van een drie-stappen-methode. Eerst werden de amines van<br />

chitosan gedeeltelijk geacetyleerd met azijnzuuranhydride en natrium borohydride, gevolgd<br />

door de kwantitatieve dimethylering van de resterende amines met <strong>for</strong>maldehyde en natrium<br />

borohydride. Vervolgens werd TMC gesynthetiseerd in aanwezigheid van een overmaat aan<br />

iodomethaan. De TMCs verkregen via deze methode bleken geen O-methylering, noch een<br />

verlies in acetylgroepen ( 1 H-NMR) of ketenbreuk (GPC) te vertonen. De mate van afbraak van<br />

TMC, en dat van haar voorgangers chitosan en dimethyl chitosan, door lysozyme (een<br />

lichaamseigen enzym) w<strong>as</strong> sterk afhankelijk van de DAc; polymeren met de hoogste DAc<br />

vertoonde de grootste daling in het molecuulgewicht. TMCs met een hoge DAc (~ 50%), een<br />

DQ van ongeveer 44% en met of zonder O-gemethyleerde groepen, waren niet in staat om de<br />

intercellulaire ruimtes van een caco-2 monolaag te openen zoals onderzocht in de transepitheliale<br />

elektrische weerstand (TEER) test. Dit in tegenstelling tot TMCs (zowel O-<br />

gemethyleerde en O-methyl vrij; concentratie van 2.5 mg/ml) met een soortgelijke DQ maar<br />

een lagere DAc; deze polymeren waren in staat de TEER tot 70 en 30% te verminderen,<br />

respectievelijk voor de O-methyl vrij en de O-gemethyleerde TMCs. Daarna<strong>as</strong>t vertoonden<br />

TMCs met een hoge DAc (~ 50%) geen noemenswaardige celtoxiciteit (MTT, LDH afgifte) tot<br />

een concentratie van 10 mg/ml. Dit hoofdstuk toont aan dat de mate van N-acetylering van<br />

TMC een grote invloed heeft op de in vitro enzymatische afbraak en biologische eigenschappen<br />

van TMC.<br />

De resultaten uit hoofdstukken 2 en 3 tonen aan dat de DQ, DAc en mate van O-methylering<br />

van invloed zijn op de in vitro biologische eigenschappen van TMC. Daarom onderzochten we<br />

de effecten van deze structurele variaties van TMC op de werkzaamheid als adjuvans in een<br />

intran<strong>as</strong>ale (i.n.) immunisatie studie in muizen.<br />

In hoofdstuk 4 werden TMCs met variërende mate van quaternisering (DQ, 22-86%), O-<br />

methylering (DOM, 0-76%) en N-acetylering (DAc 9-54%) ge<strong>for</strong>muleerd met geïnactiveerd<br />

influenza virus (WIV). Eenvoudig mengen van het WIV met de TMCs in een 1:1<br />

gewichtsverhouding resulteerde in vergelijkbare, positief geladen nanodeeltjes door coating<br />

van het negatief geladen WIV met het positief geladen TMC. De hoeveelheid vrij TMC in<br />

oplossing w<strong>as</strong> vergelijkbaar voor alle TMC-WIV <strong>for</strong>muleringen en er is dus een overmaat aan<br />

TMC aanwezig in de <strong>for</strong>muleringen.<br />

Na intran<strong>as</strong>ale immunisatie van muizen met WIV en TMC-WIV op dag 0 en 21, vertoonden de<br />

TMC-WIV <strong>for</strong>muleringen een sterkere respons (gemeten als hogere totaal IgG, IgG1 en IgG2a/c<br />

titers) dan WIV zonder TMC, met uitzondering van WIV gecoat met gereacetyleerd TMC (DAc<br />

197


Nederlandse Samenvatting<br />

54 %, DQ 44%). Er werden geen significante verschillen in de antistof-titers waargenomen<br />

voor TMCs die varieerden in DQ of DOM, wat aangeeft dat deze structurele kenmerken een<br />

ondergeschikte rol spelen in de i.n. adjuvans eigenschappen. TMC met een DQ van 56%<br />

(TMC56) ge<strong>for</strong>muleerd met WIV in een gewichtsverhouding van 5:1 resulteerde in significant<br />

lagere IgG2a/c:IgG1 ratio's ten opzichte van TMC56 gemengd in de verhouding van 0.2:1 en<br />

1:1. De IgG2a/c:IgG1 ratio geeft een indicatie voor het type immuunrespons dat opgewekt<br />

wordt: een hoge IgG1 spiegel wijst op een type 2 (of humorale, met antilichamen,<br />

afweerreactie) terwijl een hoge IgG2a/c titer een type 1 (of cellulaire, met cytotoxische T-<br />

cellen, afweerreactie) indiceert. Deze resultaten impliceren dus een verschuiving naar een<br />

Th2-type immuunrespons bij een hogere TMC dosis. Blootstelling van de gevaccineerde<br />

muizen aan een actief, ziekmakend, virus bewezen dat alle TMC-WIV <strong>for</strong>muleringen<br />

bescherming boden met uitzondering van WIV gecoat met gereacetyleerd TMC. Dit hoofdstuk<br />

toont aan dat de immunogeniciteit van WIV sterk verbetert door het gebruik van TMC en dat<br />

bescherming tegen een actief virus na i.n. vaccinatie met TMC-WIV kan worden bewerkstelligd.<br />

De werkzaamheid van TMC als adjuvans wordt sterk verminderd door het acetyleren van TMC,<br />

terwijl de DQ en DOM nauwelijks van invloed zijn.<br />

Het doel van hoofdstuk 5A w<strong>as</strong> om de oorzaak te achterhalen voor het gebrek aan<br />

werkzaamheid van gereacetyleerd TMC (TMC-RA) als adjuvans. TMC-RA (mate van acetylering<br />

54%) werd vergeleken met TMC (mate van acetylering 17%) op zes potentieel kritieke<br />

stappen in de inductie van een immuunrespons na intran<strong>as</strong>ale (i.n.) toediening bij muizen:<br />

chemische stabiliteit van het polymeer in een neusw<strong>as</strong>sing, de locatie van WIV in de neus, de<br />

n<strong>as</strong>ale verblijftijd van WIV, de cellulaire opname van WIV door calu-3 epitheelcellen, het<br />

transport van WIV over een monolaag van epitheel cellen, en de capaciteit van de verschillende<br />

<strong>for</strong>muleringen om activering te induceren van muizen dendritische cellen (DCs).<br />

TMC-RA werd in een grotere mate afgebroken in de neusw<strong>as</strong>sing dan TMC. De locatie van het<br />

antigeen in de neus en de n<strong>as</strong>ale verblijftijd waren hetzelfde voor beide types TMC.<br />

Fluorescent gelabeld WIV <strong>as</strong>socieerde beter met calu-3 cellen indien gecoat met TMC-RA dan<br />

met TMC, en beide TMCs verminderden aanzienlijk het transport van WIV over een calu-3<br />

monolaag. Muizen DCs werden door alle <strong>for</strong>muleringen in gelijke mate geactiveerd.<br />

Samenvattend, de inferieure werkzaamheid van TMC-RA als adjuvans zou kunnen worden<br />

verklaard door een lagere stabiliteit van de TMC-RA-WIV in de neusholte, en het is in ieder<br />

geval onwaarschijnlijk dat een van de andere cruciale stappen die zijn onderzocht in deze<br />

studie, een rol spelen.<br />

198


Nederlandse Samenvatting<br />

Het is bekend dat de N-geacetyleerde glucosamine blokken (of GlcNAcs) aanwezig in TMC<br />

kunnen binden aan verschillende humane C-type lectines, een familie van cel-receptoren die<br />

betrokken zijn bij de immuunrespons. Mogelijk is hierdoor het effect van TMC en<br />

gereacetyleerd TMC (met DAc van 54%, TMC-RA) op opname en activering van menselijke<br />

dendritische cellen (DCs) verschillend. Dit werd onderzocht in hoofdstuk 5B met behulp van<br />

het geïnactiveerde influenza virus (WIV) als antigeen. Studies met monocyt-afgeleide humane<br />

DCs lieten zien dat de opname van TMC(-RA) gecoat WIV iets lager w<strong>as</strong> dan ‘gewoon’ WIV.<br />

TMC-RA, als oplossing of als coating van WIV, veroorzaakte een veel sterkere activering van<br />

DCs, zoals gemeten met CD86 expressie, dan de andere <strong>for</strong>muleringen. Verder induceerde<br />

alleen TMC-RA(-WIV) een hoge afgifte van de cytokines IL-10, TNF- α en IL-12p40 en IL-12p70<br />

door de DCs. Omdat zowel de afgifte van IL-10 en IL-12p70 w<strong>as</strong> verhoogd, konden er geen<br />

uitspaken worden gedaan over het mogelijk induceren van een Th1 of Th2 type<br />

immuunrespons. Concluderend, TMC-RA heeft sterke immuno-stimulerende effecten in vitro<br />

op humane dendritische cellen. Dit impliceert dat de mate van N-acetylering mogelijk van<br />

cruciaal belang is voor het adjuvans effect van TMC in mensen.<br />

Door thiol-groepen te introduceren in de structuur zullen de (farmaceutische) toep<strong>as</strong>singen<br />

van TMC nog verder verbreed kunnen worden. Thiolen kunnen de muco-adhesieve<br />

eigenschappen verbeteren, vergemakkelijken verdere derivatiseringsreacties via o.m.<br />

intracellulair afbreekbare disulfide-bruggen, en maken het mogelijk om covalent vernette<br />

nanodeeltjes te produceren met een gethioleerd anionisch polymeer als crosslinker. In<br />

Hoofdstuk 6 werd een nieuwe vier-stappen methode gepresenteerd om gedeeltelijk<br />

gethioleerd TMC te synthetiseren waarmee de mate van quaternisering en thiolering vrijelijk<br />

kunnen worden gevarieerd. Eerst werden de amines van chitosan gedeeltelijk gecarboxyleerd<br />

met glyoxylzuur en natrium borohydride. Daarna werden de resterende amines kwantitatief<br />

gedimethyleerd met <strong>for</strong>maldehyde en natrium borohydride en vervolgens konden<br />

gecarboxyleerde TMCs verkregen worden door dit polymeer te laten reageren met een<br />

overmaat aan iodomethaan in N-methylpyrrolidon. Hierna werden de gedeeltelijk<br />

gecarboxyleerde TMCs opgelost in water en werd cystamine aan de carboxylzuren gekoppeld<br />

bij een pH van 5.5 met behulp van EDC. Na toevoeging van dithiotreïtol disulfide om de<br />

disulfide-bruggen te breken, werden gethioleerde TMCs verkregen die varieerden in mate van<br />

quaternisering (25-54%) en thiolering (5-7%) zoals bepaald met 1 H-NMR en Ellman's test. Gel<br />

permeatie chromatografie toonde aan dat er een beperkt aantal intermoleculaire bindingen<br />

w<strong>as</strong> ontstaan. De thiol-groepen in de TMCs bleken bij pH 7.4 snel te oxideren tot disulfides,<br />

199


Nederlandse Samenvatting<br />

terwijl ze bij een pH van 4.0 redelijk intact bleven na 8 uur op 37 graden. Studies met calu-3<br />

cellen lieten zien dat er een lichte daling in cytotoxiciteit is voor de gethioleerde TMCs in<br />

vergelijking met de niet-gethioleerde polymeren met eenzelfde DQ.<br />

Positief geladen nanodeeltjes werden gemaakt van de gethioleerde TMCs en gethioleerd<br />

hyaluronzuur (HA-SH) en hierin werd fluorescent gelabeld ovalbumine als model-eiwit<br />

ingesloten. Deze covalent gestabiliseerde nanodeeltjes bleven stabiel in 0.8 M NaCl, in<br />

tegenstelling tot de deeltjes gemaakt van niet-gethioleerde polymeren die in deze sterk<br />

ionische buffer uit elkaar vielen. Dit toonde aan dat de gestabiliseerde deeltjes bij elkaar<br />

gehouden werden door intermoleculaire zwavelbruggen tussen TMC-SH en HA-SH.<br />

In hoofdstuk 6 werd aangetoond dat de fysische stabiliteit van polyelectroliet nanodeeltjes<br />

van trimethyl chitosan (TMC) en hyaluronzuur (HA) beperkt is onder fysiologische<br />

omstandigheden. Vooral in aanwezigheid van zout kan de stabiliteit verminderen aangezien zij<br />

de ladingsinteracties tussen het positief geladen TMC en het negatief geladen HA geringer<br />

worden. Dit kan de werkzaamheid van het adjuvans effect van de nanodeeltjes voor de<br />

intran<strong>as</strong>ale en mogelijk ook intradermale immunisatie verminderen. Zodoende werd in<br />

hoofdstuk 7 het effect van de stabilisatie van TMC/HA nanodeeltjes door disulfide-bruggen op<br />

hun immunogeniciteit geëvalueerd. Ook werden, als alternatief, de resterende thiol-groepen op<br />

het oppervlak van deze deeltjes gepegyleerd waardoor de positieve lading van deze deeltjes<br />

(enigszins) werd afgeschermd. Mogelijk heeft dit een additioneel gunstig effect aangezien PEG<br />

(poly(ethyleen glycol)) de interacties met de extracellulaire matrix zou kunnen verminderen<br />

(intradermaal) en/of de muco-adhesie zou kunnen verbeteren (intran<strong>as</strong>aal). Covalent<br />

gestabiliseerde nanodeeltjes beladen met ovalbumine (OVA) werden gemaakt met gethioleerd<br />

TMC en gethioleerd HA via gelering op b<strong>as</strong>is van ladings-interactie gevolgd door spontane<br />

vorming van disulfide-bruggen na een incubatie van 3 uur bij pH 7.4 en 37 °C. Tevens werd<br />

maleïmide PEG gekoppeld aan de resterende thiol-groepen aan het oppervlak van de deeltjes<br />

om de positieve lading af te schermen.<br />

De nanodeeltjes hadden een grootte van circa 250-350 nm, een positieve zeta potentiaal en<br />

OVA ladings-efficiënties tot 60%. Pegylering resulteerde in een kleine verlaging van de zeta<br />

potentiaal en een geringe toename in deeltjesgrootte. Gestabiliseerde TMC-S-S-HA deeltjes<br />

(gepegyleerd of niet) hadden een superieure stabiliteit in zoutoplossingen in vergelijking met<br />

niet-gestabiliseerde TMC/HA deeltjes (bestaande uit niet-gethioleerde polymeren). Echter, de<br />

TMC-S-S-HA deeltjes vielen momentaan uiteen wanneer een reagens werd toegevoegd dat<br />

disulfide-bruggen breekt. In zowel de intran<strong>as</strong>ale als intradermale vaccinatie-studie,<br />

200


Nederlandse Samenvatting<br />

vertoonden de gestabiliseerde TMC-S-S-HA deeltjes beladen met OVA superieure<br />

immunogeniciteit in vergelijking met niet-gestabiliseerde deeltjes. Pegylering resulteerde<br />

intran<strong>as</strong>aal in een volledige tenietdoening van de positieve effecten van stabilisatie en het had<br />

geen additioneel effect wanneer deze gepegyleerde deeltjes werden gebruikt voor<br />

intradermale vaccinatie. Concluderend, de stabilisatie van de TMC/HA deeltjes verbeterde de<br />

immunogeniciteit na intran<strong>as</strong>ale en intradermale vaccinatie, echter pegylering van deze<br />

gestabiliseerde deeltjes had geen gunstig effect.<br />

De nieuwe synthese-routes die in dit proefschrift worden beschreven, maken het mogelijk<br />

voor elke specifieke toep<strong>as</strong>sing de gewenste structurele variant van TMC te produceren. Zo<br />

kunnen de mate van quaternisering en acetylering worden gevarieerd zonder dat er andere<br />

wijzigingen, zoals O-methylering of ketenbreuk, in het polymeer optreden. Daarna<strong>as</strong>t wordt<br />

een methode beschreven om op een controleerbare manier thiol-groepen in TMC te<br />

introduceren. Het gecontroleerd kunnen variëren van de structurele eigenschappen van TMC<br />

geeft de mogelijkheid om op de juiste manier structuur-activiteit-relaties te onderzoek in in<br />

vitro biologische <strong>as</strong>says en in vivo (n<strong>as</strong>ale) vaccinatie-studies. Verder werd het inzicht in het<br />

werkingsmechanisme van TMC verdiept door verschillende, potentieel cruciale, stappen in<br />

intran<strong>as</strong>ale vaccinatie op een systematische manier te onderzoeken. Ten slotte werd een<br />

veelbelovend nieuw, covalent gestabiliseerd, drager systeem onderzocht. De bemoedigende<br />

resultaten met deze nieuwe TMC-varianten geven meer dan voldoende aanleiding tot verdere<br />

pre-klinische en eventueel klinische ontwikkeling.<br />

201


Dankwoord<br />

Dankwoord<br />

Beste mensen, mijn mooie periode als AIO zit erop! Natuurlijk zijn er wel eens wat<br />

tegenslagen geweest en heb ik af en toe hard moeten werken, maar al met al is het project<br />

soepel verlopen en heb ik me erg vermaakt op het Went. Dit is dan ook naar mijn mening het<br />

grootste compliment voor iedereen met wie ik, op welke manier dan ook, te maken mee heb<br />

gehad de afgelopen jaren. Zonder jullie hulp, enthousi<strong>as</strong>me en kritische noot zou dit boekje<br />

nooit op deze prettige wijze zijn voltooid. Heel erg bedankt! Een aantal mensen wil ik graag in<br />

het bijzonder bedanken:<br />

In den beginne is er Wim H, mijn promotor: ‘Wat een biologische meuk dat TMC (Thai<br />

M<strong>as</strong>sage Center)! Kun je er geen hydrogel van maken?’ Waarna de voetbaluitslagen van het<br />

weekend werden besproken, vaak met enig leedvermaak aangezien Ajax de afgelopen 4 jaar<br />

niet de beste periode uit haar geschiedenis kende. Beste Wim, bedankt dat je deze PSV-er<br />

zoveel vrijheid en vertrouwen heb gegeven maar ook voor me klaar stond als er even spijkers<br />

met koppen geslagen moesten worden.<br />

Beste Wim J, mijn andere promotor. Hoewel officieel p<strong>as</strong> op het einde aan boord geklommen,<br />

w<strong>as</strong> je eigenlijk vanaf het begin al de grote coördinator van het TMC (Toxiciteit Misschien<br />

Cruciaal) project. Mooie discussies en een gezonde rivaliteit maakten de tripjes naar Leiden<br />

een leuk vooruitzicht. Dank voor je wetenschappelijke hulp en tekstuele creativiteit!<br />

Dan zijn er natuurlijk nog mijn alter-AIOs uit Leiden, Suzanne en Bram. Toch mooi dat we na<br />

3 jaar vergaderen onze projecten hebben kunnen afsluiten met een gezamenlijk in vivo<br />

experiment! Heel veel succes in de toekomst en bedankt voor de leuke samenwerking. Ook Elly<br />

mag hier niet ontbreken voor alle hulp bij de cel-experimenten en Joke voor de bijdrage op het<br />

gebied van de intradermale vaccinaties.<br />

Mijn studenten, Sara de Madrid, much<strong>as</strong> graci<strong>as</strong> por todo en Francesco di Camarino, grazie<br />

mille! Steffen, jouw vrijdagmiddag-experimentje heeft mij een pikstart bezorgd en ook daarna<br />

heb je me regelmatig de weg gewezen uit het organisch-chemische moer<strong>as</strong>. Bijna w<strong>as</strong> ik in een<br />

chemicus veranderd maar gelukkig heeft het niet zover hoeven komen, dank. Natuurlijk mag<br />

ook de man die in elk dankwoord staat vermeld ook hier niet ontbreken: Mies, zonder jou ziet<br />

het leven er toch een stuk minder rooskleurig uit, crying at the Viscotek!<br />

203


Dankwoord<br />

Thom<strong>as</strong> en Sven van het VUmc, geweldig dat jullie als immunologen zo enthousi<strong>as</strong>t met onze<br />

polymeren aan de gang wilden gaan. Uiteindelijk heeft het nog mooie resultaten opgeleverd.<br />

De mensen van Nobilon en Intervet, de twee uur durende treinreis w<strong>as</strong> snel vergeten door de<br />

stimulerende en prettige sfeer die er in Boxmeer heerste. Heel veel succes in deze tijd.<br />

Dr. ‘so you think you can dance’ Ethlinn, dat je hebt tijdens je eigen lange weg nog tijd hebt<br />

gevonden om me te helpen met experimenten en tips en trucs te geven, heb ik zeer kunnen<br />

waarderen. Ook anderen die me op de een of andere manier met de experimenten hebben<br />

geholpen mogen natuurlijk niet ontbreken: Maryam, Enrico, Roel, P<strong>as</strong>cal en Roberta. Dank<br />

voor jullie hulp!<br />

Het sterrenensemble van Z605, Marina (sorry if I te<strong>as</strong>ed you too much), Hajar (sorry if I<br />

te<strong>as</strong>ed you too much) and Amir V (sorry if I te<strong>as</strong>ed you too much), thanks <strong>for</strong> the good times on<br />

and off the lab. De oude garde Sophie, Marjan, Marion, Frankie, Sabrina, Marcel, Jozef ‘je<br />

gedraagt je wel hè? Ik hoef maar dìt te zien...’ en de rijpe garde Wouter, Joris en Ellen bedankt<br />

voor de gezelligheid en de wijsheden. De jonge hindes van biofarmacie: Inge, Roy, DaMarkus,<br />

Bart, speedy Maria, Alex, Pieter, Emmy, Albert, Melody, Kimberley, Audrie, Luis, Kim, Barbara,<br />

Lydia, Peter en de anderen, heel veel plezier en succes met het afronden!<br />

Gelukkig waren er nog voldoende mensen die me na<strong>as</strong>t het werk bezig hielden. Mijn<br />

teamgenoten van de Ody-zondag, we gaan er weer een frank en vrij seizoen van maken! De<br />

oud-huisgenoten van de ple<strong>as</strong>uredome, Sindre & Torkel, Japus, de Farmacie-gang, dank voor<br />

jullie vriendschap. Gert, zonder je voedzame maaltijden en de waardevolle gesprekken zou er<br />

slechts een skelet resteren, got to get behind the mule in the morning and plow!<br />

Amir-idjan, alle (oude) wijze mannen komen uit het Oosten. Vaak hebben we Utrecht onveilig<br />

gemaakt en nog vaker heb je me uitgelegd hoe de dingen werkelijk in elkaar zitten. Over vijf<br />

jaar zitten we bij je op de veranda ergens in Toronto met een lekkere whisky maar eerst heb ik<br />

je nog hier even nodig. Merci!<br />

Nelis, met jou iets ondernemen staat garant voor een succesvol avontuur. Met -5 ⁰C bananen<br />

roosteren aan de Sunshine Co<strong>as</strong>t, raven op surrealistisch Roskilde of afreizen naar Boxmeer<br />

om muizen te pesten, het w<strong>as</strong> allemaal enerverend. Je bent een unieke gozer en ik weet zeker<br />

dat wij elkaar blijven zien.<br />

204


Dankwoord<br />

Lieve Femke, het avontuur met jou is eigenlijk p<strong>as</strong> net begonnen maar hoe meer ik van je<br />

ontdek hoe meer ik er van zeker van ben dat we een mooie toekomst voor ons hebben.<br />

En als laatste wil ik nog mijn familie bedanken. Oma, die afspraak van vier jaar geleden<br />

komen we allebei na, geweldig dat je hier bij kan zijn. Lieve Marc en Nienke, nog even wachten<br />

en dan is het jullie dag. Lieve ‘spap en ‘smam, dank voor de liefde, interesse en vertrouwen.<br />

Zonder jullie w<strong>as</strong> dit allemaal niet mogelijk geweest dus geniet ervan.<br />

Tijd voor een nieuw hoofdstuk!<br />

Rolf<br />

205

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!