01.06.2013 Views

Polymers and Drug Delivery Systems

Polymers and Drug Delivery Systems

Polymers and Drug Delivery Systems

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>Polymers</strong> <strong>and</strong> <strong>Drug</strong> <strong>Delivery</strong> <strong>Systems</strong><br />

Gemma Vilar a,b , Judit Tulla-Puche a,b <strong>and</strong> Fern<strong>and</strong>o Albericio a,b,c,*<br />

Current <strong>Drug</strong> <strong>Delivery</strong>, 2012, 9, 000-000 1<br />

a Institute for Research in Biomedicine, Barcelona Science Park, Baldiri Reixac 10, 08028-Barcelona, Spain; b CIBER-<br />

BBN, Networking Centre on Bioengineering, Biomaterials <strong>and</strong> Nanomedicine, Barcelona Science Park, Baldiri Reixac<br />

10, 08028-Barcelona, Spain; c Department of Organic Chemistry, University of Barcelona, Martí i Franqués 1-11,<br />

08028-Barcelona, Spain<br />

Abstract: In the treatment of health related dysfunctions, it is desirable that the drug reaches its site of action at a particular<br />

concentration <strong>and</strong> that this therapeutic dose range remains constant over a sufficiently long period of time to alter the<br />

process. However, the action of pharmaceutical agents is limited by various factors, including their degradation, their interaction<br />

with other cells, <strong>and</strong> their incapacity to penetrate tissues as a result of their chemical nature. For these reasons,<br />

new formulations are being studied to achieve a greater pharmacological response; among these, polymeric systems of<br />

drug carriers are of high interest. These systems are an appropriate tool for time- <strong>and</strong> distribution-controlled drug delivery.<br />

The mechanisms involved in controlled release require polymers with a variety of physicochemical properties. Thus, several<br />

types of polymers have been tested as potential drug delivery systems, including nano- <strong>and</strong> micro-particles, dendrimers,<br />

nano- <strong>and</strong> micro-spheres, capsosomes, <strong>and</strong> micelles. In all these systems, drugs can be encapsulated or conjugated in<br />

polymer matrices. These polymeric systems have been used for a range of treatments for antineoplastic activity, bacterial<br />

infections <strong>and</strong> inflammatory processes, in addition to vaccines.<br />

Keywords: Dendrimers, drug delivery, micelles, nano-micro-particle, nano-micro-spheres, polymers.<br />

1. INTRODUCTION<br />

1.1. Conventional Pharmacotherapy<br />

Human health is threatened by autoimmune, neurodegenerative,<br />

metabolic <strong>and</strong> cancer diseases, just to mention a few,<br />

which are difficult to treat with systemically, delivered<br />

drugs. Conventional pharmacotherapy involves the use of<br />

drugs whose absorption <strong>and</strong> therefore bioavailability depends<br />

on many factors, such as solubility, pKa, molecular<br />

weight, number of bonds per hydrogen atom of the molecule,<br />

<strong>and</strong> chemical stability, all of which can hinder the achievement<br />

of a therapeutic response.<br />

In general, the nature of conventional therapeutics, especially<br />

their low molecular weight, confers them the capacity<br />

to cross various body compartments <strong>and</strong> access numerous<br />

cell types <strong>and</strong> subcellular organelles. Thus these drugs are<br />

suitable for the treatment of diseases. However, this form of<br />

indiscriminate distribution leads to the occurrence of side<br />

effects <strong>and</strong> to the need for higher doses of the drug to elicit a<br />

satisfactory pharmacological response. Rapid renal clearance<br />

as a result of the low molecular weight of these compounds,<br />

among other factors such as protein binding, lipophilicity,<br />

ionizability etc., implies frequent administration <strong>and</strong>/or a<br />

high dose to achieve a therapeutic effect.<br />

Research is being conducted into new formulations that<br />

ensure a greater pharmacological response, which in turn<br />

would lead to lower doses <strong>and</strong> therefore the minimization of<br />

side effects. Thus, it is necessary to improve the<br />

*Address correspondence to this author at the Institute for Research in Biomedicine,<br />

Barcelona Science Park, Baldiri Reixac 10, 08028-Barcelona,<br />

Spain; Tel:/Fax: ??????????; E-mail: ????????????????????????<br />

bioavailability of drugs. Bioavailability is affected by several<br />

factors, including the physical <strong>and</strong> chemical characteristics<br />

of the drug, the dose <strong>and</strong> concentration, the frequency of<br />

dosing, <strong>and</strong> the administration route.<br />

Therefore, research into drug delivery systems seeks to<br />

improve the pharmacological activity of drugs by enhancing<br />

pharmacokinetics (absorption, distribution, metabolism <strong>and</strong><br />

excretion) <strong>and</strong> also by amending pharmacodynamic properties,<br />

such as the mechanism of action, pharmacological response,<br />

<strong>and</strong> affinity to the site of action.<br />

Active pharmaceutical ingredients (APIs) are almost<br />

never administered alone but in dosage forms that generally<br />

include other substances called excipients. The latter are<br />

added to formulations in order to improve the bioavailability<br />

<strong>and</strong> the acceptance of the drug on the part of patients. Excipients<br />

come in numerous forms, such as emulsifiers, dyes,<br />

lubricants, diluents, supporters, <strong>and</strong> chemical stabilizers.<br />

These substances were initially considered inert because they<br />

do not exert therapeutic action per se nor do they modify the<br />

biological action of a drug. However, it is currently upheld<br />

that excipients influence the speed <strong>and</strong> extent of drug absorption,<br />

<strong>and</strong> therefore the pharmaceutical form of these substances<br />

affects drug bioavailability.<br />

In this regard, recent years have witnessed intense research<br />

on the modification of drug release <strong>and</strong> absorption.<br />

The development of new drug delivery systems will offer<br />

additional advantages to those mentioned above <strong>and</strong> may<br />

facilitate the launch of poorly soluble drugs. They may also<br />

allow an extension of patent protection for an API. And finally,<br />

<strong>and</strong> possibly most importantly, these systems will facilitate<br />

more patient-friendly administration, thus resulting in<br />

patient increased compliance <strong>and</strong> satisfaction.<br />

1567-2018/12 $58.00+.00 © 2012 Bentham Science Publishers


2 Current <strong>Drug</strong> <strong>Delivery</strong>, 2012, Vol. 9, No. 4 Vilar et al.<br />

The integral components of drug delivery systems are<br />

usually high molecular weight carriers, such as nano- <strong>and</strong><br />

micro-particles, nano- <strong>and</strong> micro-capsules, capsosomes, micelles,<br />

<strong>and</strong> dendrimers, in which the drug is embedded or<br />

covalently bound.<br />

Hence, the main function of polymeric carriers is to<br />

transport drugs to the site of action. <strong>Drug</strong>s are protected from<br />

interacting with other molecules which could cause a change<br />

in the chemical structure of the active ingredient causing it to<br />

loose its pharmaceutical action. Moreover, polymeric carriers<br />

avoid the interaction of the drug with macromolecules<br />

such as proteins, which could sequester the active ingredient<br />

preventing its arrival at the action place.<br />

If a polymeric carrier is to be used, the next step is to<br />

design a type of polymeric structure that will permit obtaining<br />

the desired release conditions. Therefore, the polymeric<br />

structure should be: i) biodegradable, because the chemical<br />

bonds that make up its chemical structure break; ii) disassemblable,<br />

because the various pieces forming the polymer<br />

disassemble but the chemical bonds do not break; <strong>and</strong> iii)<br />

undisassemblable, because the chemical bonds do not disassemble<br />

or break, that is, the polymer remains unchanged. ..<br />

In the first two cases, micro-sized polymeric carriers could<br />

be used. However, if the polymeric structure of the polymer<br />

is not biodegradable or disassembable, then nano-sized<br />

polymers must be used for renal elimination.<br />

A crucial feature of these polymers is the mechanism by<br />

which they are removed from the body. They may be excreted<br />

directly via kidneys (renal clearance) or biodegraded<br />

(metabolic clearance) into smaller molecules, which are then<br />

excreted. Passage through the renal glomerular membrane is<br />

limited to substances with a molecular weight under 50 KDa,<br />

although this value varies depending on the chemical structure<br />

of the molecule [1]. Molecular weight is especially relevant<br />

for substances that are not biodegradable, <strong>and</strong> macromolecules<br />

with a molecular weight lower than the glomerular<br />

limit can be safely removed from the body by preventing<br />

their accumulation <strong>and</strong> therefore their potential toxicity.<br />

In the case of biodegradable polymers, another option to<br />

consider in their design is the chemical structure of the<br />

polymer (degree of hydrophobicity, covalent bonds between<br />

monomers, etc.), since the speed <strong>and</strong> degradation condition,<br />

<strong>and</strong> therefore, the rate <strong>and</strong> site of drug release, can be modulated<br />

depending on the chemical structure of the polymer<br />

used. .<br />

If the polymer is not biodegradable, the drug can be covalently<br />

attached to the polymeric structure by a linker which<br />

can be degraded under different conditions such as in an<br />

acidic medium or by different enzymes. On the other h<strong>and</strong>,<br />

targets can be bound covalently to the surface which will<br />

help the directionality of the vector to the site of action.<br />

Another strategy to follow would be to use smart polymers.<br />

These are polymers that release drugs when they are<br />

induced by a stimulation which causes a change in their<br />

structure. Therefore, the drug is released at the appropriate<br />

time <strong>and</strong> place.<br />

Therefore, these new formulations seek to improve the<br />

pharmaceutical profile <strong>and</strong> stability of a drug, ensure its correct<br />

concentration, achieve maximum biocompatibility,<br />

minimize side effects, stabilize the drug in vivo <strong>and</strong> in vitro,<br />

facilitate the accumulation of the drug at a specific site of<br />

action, <strong>and</strong> increase exposure time in the target cell.<br />

1.2. Controlled Release Methods<br />

Controlled release systems aim to improve the effectiveness<br />

of drug therapy [2]. These systems modify several parameters<br />

of the drug: the release profile <strong>and</strong> capacity to cross<br />

biological carriers (depending on the size of the particle),<br />

biodistribution, clearance, <strong>and</strong> stability (metabolism), among<br />

others. In other words, the pharmacokinetics <strong>and</strong> the pharmacodynamics<br />

of the drug are modified by these formulations.<br />

Controlled release offers numerous advantages over conventional<br />

dosage forms. This approach increases therapeutic<br />

activity <strong>and</strong> decreases side effects, thus reducing the number<br />

of drug dosages required during treatment.<br />

Controlled release methods offer an appropriate tool for<br />

site-specific <strong>and</strong> time-controlled drug delivery.<br />

There are two main situations in which the distribution<br />

<strong>and</strong> time-controlled delivery of a drug can be beneficial. The<br />

first is when the natural distribution of the drug causes major<br />

side effects due to its interaction with other tissues, while the<br />

second is when the natural distribution of the drug does not<br />

allow it to reach its molecular site of action due to degradation.<br />

Many different kinds of drugs can benefit from distribution<br />

or time-controlled delivery, such as anti-inflammatory<br />

agents [3], antibiotics [4], chemotherapeutic drugs [5], immunosuppressants<br />

[6], anesthetics [7] <strong>and</strong> vaccines [8].<br />

1.2.1. Time-Controlled: Modified-Release Formulation<br />

This formulation refers to the release of the drug through<br />

a system that protects <strong>and</strong> retains it. Through controlled release<br />

over time, this formulation allows the drug to reach a<br />

therapeutic concentration in tissues.<br />

The usual dose administration of a drug can achieve immediate<br />

(short-term) therapeutic concentrations of the active<br />

ingredient [9] (Fig. 1.1).<br />

Fig. (1.1). Curve of drug concentration in plasma over time after a<br />

single dose [10].<br />

The figure shows that drug absorption [11] does not usually<br />

stop abruptly when it reaches the maximum concentration,


<strong>Polymers</strong> <strong>and</strong> <strong>Drug</strong> <strong>Delivery</strong> <strong>Systems</strong> Current <strong>Drug</strong> <strong>Delivery</strong>, 2012, Vol. 9, No. 4 3<br />

but may continue for some time along the downstream portion<br />

of the curve. <strong>Drug</strong> absorption ceases with time once the<br />

bioavailable dose has been absorbed. Therefore, the elimination<br />

phase of plasma drug concentration depends on the rate<br />

of removal by metabolism or excretion.<br />

The effective (therapeutic) concentration is the minimal<br />

drug concentration required to achieve the desired therapeutic<br />

effect. In contrast, the maximum safe concentration is<br />

drug concentration in plasma above which toxic or side effects<br />

occur. The interval between these two concentrations is<br />

the therapeutic window. The concentrations that yield the<br />

desired new therapeutic effect in the absence of toxic effects<br />

fall within this window.<br />

In a chronic treatment, doses are administered at regular<br />

intervals (multiple doses, shown in Figs. (1.2)). This approach<br />

can lead to variations in API concentrations, reaching<br />

levels that are lower or higher than therapeutic concentrations,<br />

in other words, concentrations that are neither effective<br />

nor toxic, respectively.<br />

Fig. (1.2). <strong>Drug</strong> concentrations in plasma after delivery by conventional<br />

injection in a chronic disease. Representation of the therapeutic<br />

window between the minimal therapeutic concentration (therapeutic<br />

level, dotted line) <strong>and</strong> the maximum therapeutic concentration<br />

(toxic level, solid line). Adapted from reference 10.<br />

Chronic treatment calls for adherence to a pattern of drug<br />

administration because an incorrect number of dosages may<br />

exceed the limit of therapeutic concentration <strong>and</strong> thus result<br />

in a toxic response. In general, in chronic treatments drug<br />

doses are separated by two times the half-life of the drug in<br />

blood. The time required for a compound that has reached<br />

the systemic circulation of the body to reduce the maximum<br />

concentration by half is called the half-life.<br />

In contrast, a modified-release formulation is designed to<br />

deliver the active ingredient at a predetermined speed or at a<br />

site other than that of administration. These formulations can<br />

be classified on the basis of the systems that regulate drug<br />

release.<br />

1.2.1.1. Extended-Release Formulation<br />

The release of the active ingredient in these formulations<br />

is initially produced in a sufficient amount to produce a<br />

therapeutic effect <strong>and</strong> then to continue with a release, which<br />

is not necessarily constant but extends the therapeutic action.<br />

There are two types of extended-release formulations. One is<br />

a polymeric matrix that contains the active ingredient, the<br />

delivery of which is controlled by diffusion through the polymeric<br />

system network. These formulations are presented in<br />

the form of hydrogels <strong>and</strong> nano- <strong>and</strong> micro-particles. The<br />

other type of extended-release formulation comprises API<br />

capsules with polymer coatings, forming a reserve deposit.<br />

The permeability or the degradation of the polymer coating<br />

regulates the release of the drug.<br />

There are some commercial products such as Plenidil Er<br />

or Kapanol. Plenidil or Kapanol tablets provide extended<br />

release of felopine or morphine surfate, respectively. Both<br />

tablets are administered orally <strong>and</strong> the drug is released by<br />

diffusion through the system [12].<br />

There are other commercially available products in which<br />

the degradation of the polymer coating regulates the release<br />

of the drug. An example of this formulation is Sinemet CR.<br />

The newest is a polymeric-based drug delivery system that<br />

controls the release of carbidopa <strong>and</strong> levodopa by slowly<br />

eroding the polymer´s coating, making the system particularly<br />

suitable in the treatment of Parkinson’s disease <strong>and</strong><br />

syndrome [13].<br />

Finally, it is important to make a few comments about the<br />

<strong>Drug</strong>-Eluting Stent (DES), which consists of supports coated<br />

with drug-laden polymers. Generally, a stent is a metal support<br />

wire whose function is to maintain arteries, blood vessels<br />

or other anatomical ducts open. One interesting example<br />

is the cardiovascular stent, which is a device that is inserted<br />

into the coronary arteries when they are blocked to keep the<br />

arteries open <strong>and</strong> normal blood flow is resumed. Sometimes,<br />

this stent can be coated with a drug or polymer which controls<br />

the release of the drug, thus preventing the arteries from<br />

reclosing. Nowadays, several DES platforms have been developed<br />

<strong>and</strong> evaluated for clinical use. The difference between<br />

them has to do with the type of stent, anti-proliferative<br />

drug <strong>and</strong> polymer used to control drug release.<br />

Depending on the type of polymer used, different drugrelease<br />

kinetics is obtained. San Juan et al. [14] functionalized<br />

a biocompatible polymer for stent coating for local arterial<br />

therapy. Therefore, metal stents were coated with a<br />

cationized pullulan hydrogel which was loaded with small<br />

interfering RNA for gene silencing in vascular cells. It was<br />

observed that the release of siRNA from the polymer was<br />

modulated by the presence of the cationic groups. Consequently,<br />

the release of the drug can be modulated depending<br />

on the type of polymer used.<br />

There are some DES which have been approved for clinical<br />

use such as Cypher [15] <strong>and</strong> Taxus [16] containing sirolimus<br />

<strong>and</strong> paclitaxel, respectively (Figs. 1.3 <strong>and</strong> 1.4).<br />

1.2.1.2. Sustained-Release Formulation<br />

The active substance is steadily released with a kinetic<br />

profile of zero-order. The therapeutic effect is maintained<br />

over a long period of time Fig. (1.5). The closest pharmaceutical<br />

formulations to this type of system are osmotic pumps.<br />

These devices control the outflow of drug solutions through<br />

osmotic potential gradients across semi-permeable polymer<br />

barriers. The pressurized chambers contain the aqueous solution


4 Current <strong>Drug</strong> <strong>Delivery</strong>, 2012, Vol. 9, No. 4 Vilar et al.<br />

Fig. (1.3). A). Nano- or micro-capsules whose degradation allows the release of the drug. B) Nano- or micro-particles of hydrogels whose<br />

swelling structure allows the diffusion of the drug through the pores of the structure. Adapted from reference 10a.<br />

Fig. (1.4). Representation of the concentration of active ingredient versus time for a prolonged release formulation. The graph shows the<br />

therapeutic window between the minimal therapeutic concentration (therapeutic level, dotted line) <strong>and</strong> the maximum therapeutic concentration<br />

(toxic level, solid line). Adapted from reference 10a.<br />

Fig. (1.5). Osmotic pumps. Adapted from reference 10b.<br />

of the drug <strong>and</strong> the polymeric osmotic system. Upon immersion<br />

in the water, the osmotic system is hydrated <strong>and</strong> swollen,<br />

thus causing an increase in pressure, which is relieved<br />

by the flow of the solution out of the delivery device through<br />

an orifice in the upper part [17].<br />

<strong>Drug</strong> delivery through osmotic systems is affected by a<br />

number of factors such as solubility, osmotic pressure, size<br />

of the delivery orifice <strong>and</strong> membrane type <strong>and</strong> characteristics.<br />

An example of this type of product is Adalat OROS.<br />

The OROS is a 24-hour controlled release oral drug delivery<br />

system in the form of a tablet, which acts as an osmotic<br />

pump releasing nifedipine [18]. The tablet has a rigid water-<br />

permeable jacket with one or more laser drilled small holes.<br />

As the tablet passes through the body, the osmotic pressure<br />

of water entering the tablet pushes the active drug through<br />

the opening in the tablet. Once the active ingredient in the<br />

Oros tablet has been released, the remnant is excreted in the<br />

feces (Fig. 1.6)<br />

1.2.1.3. Pulsatile-Release Formulation<br />

Another form of time- controlled release is a responsive<br />

delivery system in which drugs are released in a pulsatile<br />

manner only when required by the body. Thus, the drug is<br />

released after a well-defined lag time. In general, these drug


<strong>Polymers</strong> <strong>and</strong> <strong>Drug</strong> <strong>Delivery</strong> <strong>Systems</strong> Current <strong>Drug</strong> <strong>Delivery</strong>, 2012, Vol. 9, No. 4 5<br />

delivery systems comprise two components, namely a sensor<br />

that detects the environmental parameter that stimulates drug<br />

release, <strong>and</strong> a delivery device that releases the drug.<br />

Fig. (1.6). Representation of the concentration of active ingredient<br />

versus time for a sustained-release formulation. The graph represents<br />

the therapeutic window between the minimal therapeutic concentration<br />

(therapeutic level, dotted line) <strong>and</strong> the maximum therapeutic<br />

concentration (toxic level, solid line). Adapted from reference<br />

10a.<br />

Pulsatile drug delivery systems release the drug at the<br />

right time <strong>and</strong> place <strong>and</strong> in accurate amounts, <strong>and</strong> they are<br />

convenient for patients suffering from chronic problems such<br />

as diabetes, arthritis, asthma, <strong>and</strong> hypertension [19].<br />

The pulsatile-release formulation implies the release of<br />

the drug only when required by the body, thus preventing it<br />

from being continuously in the biophase.<br />

Insulin formulations currently require repeated injections<br />

daily, <strong>and</strong> hence responsive drug delivery systems that release<br />

insulin in response to increased blood glucose levels<br />

have been designed. For the treatment of diabetes, a pulsatile-release<br />

formulation that uses the enzyme glucose oxidase<br />

as a sensor has been proposed [20]. When blood sugar levels<br />

rise, glucose oxidase converts glucose to gluconic acid,<br />

thereby lowering the pH. This decrease in pH causes insulin<br />

release because the pH-sensitive polymers (smart polymers)<br />

either swell or degrade in acidic environments.<br />

Another example of the above is the formulation developed<br />

<strong>and</strong> studied by Sadaphal et al. [21]. This formulation is<br />

constituted by an impermeable anionic polymer called Eudragit<br />

S100 which encapsulates theophylline. This system<br />

attempts to mimic the circadian rhythm of the disease by<br />

releasing the drug at the appropriate times (Fig. 1.7).<br />

1.2.1.4. Delayed-Release Formulation<br />

In this case, the release of the active ingredient does not<br />

coincide with the time of administration <strong>and</strong> its therapeutic<br />

action is not extended.<br />

Normally, these formulations are enteric coatings that<br />

protect the active ingredient against adverse physiological<br />

media or protect certain parts of the body against the hostile<br />

action of the active ingredient. Typically, these coatings are<br />

sensitive to pH <strong>and</strong> allow drug release into the small intestine,<br />

thus preventing its delivery to the stomach.<br />

Fig. (1.7). Representation of the concentration of the active ingredient<br />

versus time for a sustained-release formulation. The graph<br />

shows the therapeutic window between the minimal therapeutic<br />

concentration (therapeutic level, dotted line) <strong>and</strong> the maximum<br />

therapeutic concentration (toxic level, solid line). Adapted from<br />

reference 10a.<br />

Therefore, modified-release formulations provide alternatives<br />

to the routine administration of the drug for improving<br />

its management <strong>and</strong> optimizing its therapeutic action. Thus,<br />

controlled release is highly beneficial for drugs that show<br />

poor bioavailability, in other words, those that are rapidly<br />

metabolized <strong>and</strong> eliminated from the body after administration<br />

<strong>and</strong> thus have a short life, or those with a narrow therapeutic<br />

window. Some drugs are unstable in certain environments,<br />

such as in plasma or in acidic conditions. Therefore,<br />

in these conditions, it is of interest to use a polymer to protect<br />

the drug. These formulations are also used in pathologies<br />

in which the degree of compliance is low because of difficult<br />

administration or unpleasant taste of the medicine.<br />

Given that modified-release formulations have a series of<br />

drawbacks over st<strong>and</strong>ard systems, such as higher production,<br />

cost, lack of reproducibility, unpredictable in vitro/in vivo<br />

correlation <strong>and</strong> problems associated with improper h<strong>and</strong>ling,<br />

their development for an active ingredient is not justified<br />

unless they show a clear advantage over the st<strong>and</strong>ard formulation.<br />

1.2.2. Controlled Distribution<br />

The treatment of most diseases requires the distribution<br />

of the drug at a specific site. It is therefore important to use<br />

drug carriers with a controlled distribution system. Several<br />

approaches are currently available to overcome the obstacles<br />

posed by non-specific drug delivery.<br />

One method is to functionalize the surface of nano- <strong>and</strong><br />

micro-particles with receptor recognition elements that are<br />

present in diseased tissue. Thanks to molecular biology, a<br />

comprehensive database of molecular targets has been built<br />

<strong>and</strong> it is now known that some specific receptors are overexpressed<br />

in malignant tissue. Therefore, the identification of<br />

these overexpressed targets is used to functionalize polymers<br />

with tracking devices, such as antibodies [22], carbohydrates<br />

[23], or simply an electrically charged species, which interact<br />

with these targets [24]. Consequentlty, these tracking<br />

devices help the polymer reach the tissues <strong>and</strong> regions of the<br />

body where the drug is to be released [25] (Fig. 1.8).


6 Current <strong>Drug</strong> <strong>Delivery</strong>, 2012, Vol. 9, No. 4 Vilar et al.<br />

Fig. (1.8). Polymer functionalized to interact with specific targets.<br />

Adapted from reference10c.<br />

Another example of a controlled distribution system is<br />

drug release by diffusion through a hydrogel. To this end, the<br />

drug is encapsulated or absorbed in the hydrogel <strong>and</strong> released<br />

in response to an external stimulus.<br />

The stimuli that can be used in various systems to trigger<br />

drug release have been reviewed by Qiu et al. <strong>and</strong> Gupta et<br />

al. [26] One interesting model is called the pH-dependent<br />

system, which is characterized by ionizable groups in its<br />

structure. These groups are ionized depending on the pH, <strong>and</strong><br />

therefore an attractive interaction may occur between them<br />

through hydrogen bonding, or a repulsive interaction may be<br />

induced because of ionic groups with the same charge. Thus,<br />

a change in the structure of these groups opens or closes the<br />

route through which the drug is released. Other systems are<br />

controlled by electrical signals. This stimulus can change<br />

features in the structure of the formulation, such as pore size<br />

<strong>and</strong> permeability. Another approach is the use of hydrogels,<br />

the behavior of which differs in response to temperature.<br />

Also of note are those systems that are triggered by external<br />

stimuli in the form of variations in the concentration of certain<br />

substances, such as glucose, urea <strong>and</strong> morphine. However,<br />

these will be discussed further in the section on smart<br />

polymers (Table 1.1).<br />

Table 1.1. Some Examples of Smart <strong>Polymers</strong><br />

The controlled release of an active ingredient can also be<br />

achieved by connecting the drug to the polymer through a<br />

specific linker [27] that can be degraded in an acidic environment<br />

or by a specific enzyme. Thus, the drug is released<br />

from the polymer when the entire system is exposed to either<br />

of these two conditions, such as in tumor tissues where the<br />

pH is lower than in healthy tissues, in an endosomal acidic<br />

environment, <strong>and</strong> in a lysosomal environment, which contains<br />

proteolytic <strong>and</strong> hydrolytic enzymes.<br />

The introduction of macromolecules into cells by endocytosis<br />

[28] involves the invagination of the plasmatic membrane<br />

to form a small vesicle called the endosome. This early<br />

endosome matures <strong>and</strong> finally fuses with the lysosome,<br />

which contains hydrolytic <strong>and</strong> proteolytic enzymes in an<br />

acidic environment, thus forming what is known as a secondary<br />

lysosome. Therefore, in this environment within the<br />

lysosome, the linkers that are labile to acids or digestive enzymes<br />

are cleaved, causing drug release [29].<br />

It is important to highlight that these linkers are stable in<br />

plasma <strong>and</strong> are degraded only in tumor tissues or in lysosomal<br />

or endosomal [30] microenvironments. Tumor tissues or<br />

early endosomes have a slightly acidic pH, where labile acid<br />

linkers such as cis-aconityl or hydrazone can be degraded<br />

[31]. For example, the proteins carried by the polymer <strong>and</strong><br />

conjugated in it with an acid-labile linker are released into<br />

the endosome because peptide bonds are degraded in the<br />

lysosome. Once the linker has been biodegraded, the drug is<br />

released from the endosome or lysosome by diffusion to the<br />

cytosol (Fig. 1.9).<br />

The linkers most frequently used are peptides <strong>and</strong> oligosaccharides.<br />

Linker libraries have been established by VectraMed,<br />

based on enzyme cleavage specificity of a selected<br />

disease. One example is the enzyme family of Cathepsin,<br />

which includes lysosomal endoproteases, Cathepsin B being<br />

among these [32], a molecule localized at the invasive edges<br />

of human tumors <strong>and</strong> therefore overexpressed in many types<br />

of cancer. Other examples of proteins include Prostate Specific<br />

Antigen (PSA), which is expressed by prostate tumors,<br />

<strong>and</strong> Proline peptidases, which hydrolyze proline sites. These<br />

peptidases are overexpressed in pulmonary hypertension <strong>and</strong><br />

during inflammation.<br />

2. POLYMERS USED FOR CONTROLLED-DRUG<br />

RELEASE<br />

<strong>Polymers</strong> used for controlled drug release are often called<br />

therapeutic polymers. However, these systems do not have<br />

Stimulus Polymer <strong>Drug</strong> Released<br />

pH Poly(methacrylic-g-ethylene glicol) (p(MMA-g-EG) Insulin<br />

Electric field Poly(methacrylic acid) (PMA) Pilocarpine <strong>and</strong> raffinose<br />

Glucose concentration Poly(methacrylic acid-co-butyl methacrylate) Insulin<br />

Temperature Layer of Chitosan Pluronic on PLGA microparticles Indomethacin<br />

Morphine concentration Methyl vinyl ether-Co-anhydride maleic copolymer Naltrexone<br />

Urea concentration Methyl vinyl ehter-Co-anhydride maleic copolymer Hydrocortisone


<strong>Polymers</strong> <strong>and</strong> <strong>Drug</strong> <strong>Delivery</strong> <strong>Systems</strong> Current <strong>Drug</strong> <strong>Delivery</strong>, 2012, Vol. 9, No. 4 7<br />

specific therapeutic activity <strong>and</strong> are considered excipients in<br />

pharmaceutical formulations.<br />

The main interest in polymers for drug delivery purposes<br />

is in controlled-release systems. In these applications, drugs<br />

are embedded or conjugated in polymer matrices.<br />

Fig. (1.9). Endocytotic pathway for the cellular uptake of macromolecules<br />

<strong>and</strong> nanocarriers for drug delivery. Adapted from ref.<br />

10c.<br />

From a polymer chemistry perspective, it should be noted<br />

that distinct mechanisms of controlled release require polymers<br />

with a variety of physicochemical properties [33].<br />

Therefore, formulation for drug delivery includes numerous<br />

architectural designs in terms of size, shape, <strong>and</strong> materials.<br />

Several types of polymers have been tested as potential drug<br />

delivery systems, including nano- <strong>and</strong> micro-particles, dendrimers,<br />

nano- <strong>and</strong> micro-spheres, capsosomes <strong>and</strong> micelles.<br />

The difference between nano- <strong>and</strong> micro-formulation is<br />

the size. The size of the former confers this approach a<br />

greater capacity to reach the smallest capillary vessels <strong>and</strong> to<br />

penetrate tissues either through paracellular (transport of<br />

substances between cells of an epithelium) or transcellular<br />

pathways (substances travel through the cell). Nanoformulations<br />

can be administered through various routes, including<br />

oral, pulmonary, nasal, parenteral, <strong>and</strong> intraocular: however,<br />

some routes cannot be used with microformulations, such as<br />

the intravenous route, because of possible obstruction of<br />

veins.<br />

2.1. <strong>Drug</strong>-Polymer Conjugates<br />

As mentioned earlier, therapeutic agents can be covalently<br />

bound to the polymer backbone. In these types of<br />

drug-polymer conjugates, the goal is to improve the mechanism<br />

of cellular internalization <strong>and</strong> cell specificity to achieve<br />

optimal release of the drug at the proposed target. In contrast,<br />

the systems in which the drugs are embedded or encapsulated<br />

in the polymer seek to enhance the distribution <strong>and</strong><br />

serum stability <strong>and</strong> to attain a decrease in drug immuno-<br />

genicity through a controlled distribution at the correct site<br />

<strong>and</strong> over time.<br />

Polysaccharides, for instance, contain hydroxyl groups<br />

that allow direct reaction with drugs with carboxylic acid<br />

functions, thereby producing ester linkages that are biodegradable<br />

<strong>and</strong> thus facilitate the release of the drug in the<br />

body.<br />

<strong>Drug</strong> polymer conjugates are of interest when these drugs<br />

are attached to the polymer with linkers that are labile to<br />

certain digestive enzymes or acidic conditions (see controlled<br />

distribution). This type of release is used by nanoparticles,<br />

which have high mobility in the smallest capillaries,<br />

allowing for efficient uptake <strong>and</strong> selective drug accumulation<br />

at the target sites [34]. This type of nanoparticle is common<br />

in cancer therapy.<br />

Depending on the desired application, nanodelivery systems<br />

must have a certain particle size, surface charge <strong>and</strong><br />

hydrophobicity in order to overcome physiological barriers.<br />

These barriers comprise biological structures or physiological<br />

mechanisms that prevent nanoparticles from reaching<br />

their targets, thus compromising therapeutic efficacy. These<br />

biological barriers can be overcome when the nanoparticles<br />

show efficient extravasation through the vasculature, prolonged<br />

vascular circulation time, improved cellular uptake,<br />

<strong>and</strong> endosomal or lysosomal escape.<br />

In recent years, polymeric carriers of antineoplastic drugs<br />

have been extensively studied [35]. These polymers can accumulate<br />

passively in cancerous tissue as a result of certain<br />

differences in the biochemical <strong>and</strong> physiological characteristics<br />

of healthy <strong>and</strong> malignant tissue. In general, small molecules<br />

diffuse through the endothelial cell wall to healthy <strong>and</strong><br />

malignant tissues. However, macromolecules can reach only<br />

the latter because the walls of the endothelial cells of these<br />

tissues are fenestrated, resulting in pores that allow macromolecules<br />

to cross. Other features of malignant tissue include<br />

poor lymphatic drainage <strong>and</strong> increased vascularity<br />

(angiogenic process), which both lead to the retention of<br />

macromolecules. This passive accumulation of macromolecules<br />

in tumor tissue, known as the EPR effect (enhanced<br />

permeability <strong>and</strong> retention effect), was first reported by<br />

Maeda et al. [36] (Fig. 1.10).<br />

Therefore, nanoparticle-mediated drug delivery can be<br />

either an active or a passive process. The latter refers to<br />

transport through leaky capillary fenestrations into tumor<br />

cells by passive diffusion. Thus, there is an accumulation of<br />

nanoparticles in these cells due to the EPR effect. In contrast,<br />

an active process involves the use of peripherally conjugated<br />

targeting moieties for enhanced delivery to a specific site.<br />

Once macromolecules reach the malignant tissue, they<br />

are endocytosed (receptor-mediated endocytosis) or phagocytosed<br />

(internalization without the mediation of receptor)<br />

by malignant tissues cells to form a vesicle called the endosome.<br />

Given that the pH in these vesicles is lower (6.5-<br />

5.0) than that of the extracellular matrix (7.2-7.4), acid-labile<br />

linkers can be hydrolyzed, thus releasing the drug. Finally,<br />

the endosome is absorbed by a lysosome, resulting in a secondary<br />

lysosome that contains hydrolytic <strong>and</strong> digestive enzymes<br />

that are active in acidic environments (lysosome pH:<br />

4). Some studies show overexpression of certain enzymes in


8 Current <strong>Drug</strong> <strong>Delivery</strong>, 2012, Vol. 9, No. 4 Vilar et al.<br />

Fig. (1.10). Schematic representation of the anatomical <strong>and</strong> physiological characteristics of healthy <strong>and</strong> tumor tissue with respect to the vascular<br />

permeability <strong>and</strong> retention of small <strong>and</strong> large molecules (EPR effect). Adapted from reference 26.<br />

malignant tissue; this overexpression could be used to obtain<br />

a controlled release when a linker labile to these enzymes is<br />

used. Furthermore, the pH of tumor tissue is slightly lower<br />

than that of healthy tissue <strong>and</strong> therefore controlled-drug release<br />

can also be obtained when an acid-labile linker is used;<br />

however, in this case the drug is released at the extracellular<br />

level.<br />

Several drug-polymer conjugates have entered the clinical<br />

development stage [37].<br />

Some of these compounds, such as HPMA [38] copolymer-doxorubicin<br />

galactosamine (PK2), HPMA copolymer<br />

platinate (AP 5346), carboxymethyldextran polyalcoholbound<br />

exatecan (DE-310), poly(L-glutamic acid)-paclitaxel<br />

(Xyotac), PEG-SN38 (EZN-2208), linear cyclodextrinbound<br />

camptothecin (IT-101) <strong>and</strong> poly(L-glutamic acid)camptothecin<br />

(CT-2106) have progressed further to the clinical<br />

phase.<br />

However, clinical studies for several other polymer conjugates,<br />

such as HPMA copolymer camptothecin (PNU<br />

166148) <strong>and</strong> HPMA copolymer paclitaxel (PNU 166945),<br />

have been discontinued or suspended [34-51].<br />

Developed by Kopecek <strong>and</strong> Duncan, HPMA copolymerdoxorubicin<br />

(PK1) was the first cancer drug conjugated with<br />

a polymer to be tested clinically [39]. Its structure contains a<br />

peptide linker (-Gly-Phe-Leu-) that allows doxorubicin release<br />

through the action of lysozyme. Clinical studies<br />

showed that the toxicity of the conjugate was lower than<br />

doxorubicin alone. These studies also demonstrated prolonged<br />

circulation of the conjugate in plasma. Moreover, the<br />

results of phase II trials showed antitumor activity for the<br />

treatment of breast <strong>and</strong> non-small cell lung cancer but no<br />

response in patients with colorectal cancer [40].<br />

Later, HPMA copolymer-doxorubicin galactosamine<br />

(PK2) was synthesized Fig. (1.11) [41]. This compound has<br />

a similar structure to PK1 but incorporates an additional targeting<br />

lig<strong>and</strong>, namely galactosamine, which specifically targets<br />

liver tissues (controlled distribution). The drawback is<br />

that the receptor of the galactosamine lig<strong>and</strong> is expressed on<br />

both malignant <strong>and</strong> healthy hepatocytes.<br />

HPMA copolymer platinate (AP 5346) [42] is another<br />

drug-polymer conjugate that has progressed into clinical<br />

trials. This compound consists of a cytotoxic diaminocyclohexane<br />

(DACH)-platinum moiety coupled to a biocompatible<br />

hydroxypropylmethacrylamide copolymer (HPMA).<br />

Platinum drugs (carboplatin <strong>and</strong> cisplatin) are used to treat a<br />

wide range of cancers <strong>and</strong> they react in vivo, binding to <strong>and</strong><br />

causing crosslinking of DNA, thus triggering apoptosis. Furthermore,<br />

DACH has activity against cisplatin-resistant cancer<br />

cells. Therefore, the DACH-platinum moiety was bound<br />

to HPMA via a pH-sensitive linker, <strong>and</strong> consequently released<br />

in the extracellular space of tumors <strong>and</strong>/or the intracellular<br />

lysosomal compartment. Another interesting feature<br />

of this drug-polymer conjugate is its size (approx. 25 kDa),<br />

which not only allows its renal clearance but also enables it<br />

to reach tumor cells through fenestrated endothelial cells.<br />

Therefore, AP 5346 has a longer half-life relative to small<br />

platinum drugs that enables it to reach tumor cells by a passive<br />

process <strong>and</strong> then release the drug. Partial responses in<br />

antitumor activity in metastatic melanoma <strong>and</strong> ovarian cancer<br />

to AP 5346 were observed in a Phase I study. Thus, this<br />

drug has been shown to have a higher therapeutic index in<br />

multiple preclinical tumor models <strong>and</strong> it is entering phase II<br />

clinical testing under the new name of Prolindac [43].<br />

Another drug-polymer conjugate in clinical trials is DE-<br />

310. This conjugate is composed of exatecan (DX-8951) <strong>and</strong><br />

carboxymethyldextran polyalcohol (CM-Dex-PA) as a carrier.<br />

Used in cancer chemotherapy, exatecan is camptothecin<br />

analog, a cytotoxic quinoline alkaloid that inhibits the DNA<br />

enzyme topoisomerase I. Camptothecin binds to topoisomerase<br />

I <strong>and</strong> DNA, resulting in a stable ternary complex,<br />

thus preventing DNA re-ligation. This drug causes DNA<br />

damage, thus causing apoptosis. The structures of exatecan<br />

<strong>and</strong> camptothecin hold a lactone ring that is highly susceptible<br />

to hydrolysis <strong>and</strong> is inactive in its open form. In other<br />

words, the open form does not inhibit topoisomerase I. To


<strong>Polymers</strong> <strong>and</strong> <strong>Drug</strong> <strong>Delivery</strong> <strong>Systems</strong> Current <strong>Drug</strong> <strong>Delivery</strong>, 2012, Vol. 9, No. 4 9<br />

H 2<br />

C<br />

O NH<br />

HN<br />

Fig. (1.11). Structure of PK2, which has a targeting lig<strong>and</strong> (R): galactosamine (targeting lig<strong>and</strong>).<br />

protect the closed form (active form), this drug is conjugated<br />

to a polymer such as DE-310. In this polymer, exatecan is<br />

covalently linked via a peptidyl spacer (Gly-Gly-Phe-Gly).<br />

This spacer is labile to the activity of cathepsin <strong>and</strong> is thus<br />

used to release the drug inside the cell. This release was<br />

studied by Shiose et al. [44], who addressed the relationship<br />

between cathepsin activity <strong>and</strong> the release of DE-310 in several<br />

types of tumor. Moreover, clinical trials in patients with<br />

advanced solid tumors were studied by Soepenberg et al.<br />

[45]. Those clinical trials determined drug toxicity <strong>and</strong><br />

pharmacokinetics <strong>and</strong> therefore, the maximum tolerated dose<br />

in phase II. Exatecan was observed to show a slow release<br />

from DE-310, thus achieving prolonged exposure to this<br />

topoisomerase I inhibitor.<br />

Another interesting drug polymer conjugate that has been<br />

studied in numerous clinical trials, including phase III trials,<br />

is poly(L-glutamic acid)-paclitaxel, PG-TXL, also called<br />

XYOTAX (CT-2103) [46]. Paclitaxel is used in cancer chemotherapy.<br />

This drug stimulates the assembly of microtubules<br />

from tubulin dimers, thus preventing depolymerization.<br />

Thus, it inhibits the formation of mitotic cell division by<br />

blocking mitosis. Paclitaxel is used to treat ovarian, breast<br />

<strong>and</strong> lung cancer. Paclitaxel is conjugated via an ester linkage<br />

through its 2’hydroxyl group to the carboxylic acid of<br />

poly(L-glutamic acid). PG-paclitaxel is a highly promising<br />

anticancer drug polymer conjugate because it has the characteristics<br />

of an ideal polymeric drug carrier, such as biodegradability,<br />

high drug loading, stability in circulation, <strong>and</strong><br />

C<br />

H 2<br />

O<br />

O<br />

O NH<br />

Ph<br />

HN<br />

O<br />

O<br />

C<br />

H 2<br />

NH O NH<br />

R<br />

OH<br />

HN<br />

O<br />

O<br />

O O OH<br />

O<br />

OH<br />

NH<br />

Ph<br />

HN<br />

O<br />

O<br />

O<br />

NH<br />

OH<br />

O OH<br />

OH<br />

HPMA<br />

Linker<br />

Doxorubicine<br />

apparent lack of immunogenicity. In vivo, the biodistribution<br />

[47], in other words, the area under the curve obtained in the<br />

graph on tissue tumor concentration versus time (AUC), is<br />

greater when the drug is administered as PG-TXL than when<br />

it is administered in other formulations of paclitaxel. This<br />

enhanced distribution is caused by prolonged tumor exposure<br />

<strong>and</strong> limited systemic exposure compared to the active drug.<br />

Consequently, PG-TXL shows less toxicity. This conjugate<br />

can be administered by short infusion into peripheral veins,<br />

<strong>and</strong> preclinical studies have demonstrated that it is well tolerated<br />

in this mode. The maximum tolerable dose (MTD)<br />

<strong>and</strong> pharmacology of PG-paclitaxel administered weekly to<br />

patients with solid tumors were studied [48]. Results of<br />

phase II studies [49] showed that the MTD of PG-TXL was<br />

higher than that of other formulations of paclitaxel <strong>and</strong> that<br />

PG-paclitaxel has activity in patients who have not responded<br />

to taxol therapy. Phase III trials are currently underway<br />

to test the efficacy of this compound in patients with<br />

advanced non-small-cell lung cancer.<br />

PEG-SN38 (EZN-2208) [50] is currently entering phase<br />

II clinical testing. This compound is composed by SN38<br />

linked by a covalent bound to poly(ethylene glycol). SN38 is<br />

a camptothecin analog <strong>and</strong> therefore has the same function as<br />

camptothecin (topoisomerase I inhibitor) <strong>and</strong> a similar structure,<br />

containing a lactone ring. Thus, SN38 has to be administrated<br />

with a formulation, such as by conjugation to a PEG<br />

polymer that protects the lactone ring <strong>and</strong> prevents its hydrolysis.<br />

The results in animal experiments showed a higher


10 Current <strong>Drug</strong> <strong>Delivery</strong>, 2012, Vol. 9, No. 4 Vilar et al.<br />

half-life of the drug-polymer conjugate than the free drug, a<br />

higher drug exposure to tumor cells, <strong>and</strong> an increased solubility.<br />

The dose-limiting toxicity <strong>and</strong> dosing schedules were<br />

determined in phase I clinical trials.<br />

IT-101 [51] is a novel approach to cancer chemotherapy<br />

that is currently under study in human trials. IT-101 is composed<br />

by camptothecin covalently attached through a glycine<br />

link to a cyclodextrin-based polymer (CDP), which in turn<br />

consists of -cyclodextrin <strong>and</strong> poly(ethylene glycol). This<br />

compound is administered by intravenous injection <strong>and</strong><br />

demonstrates high antitumor activity because of extended<br />

circulation times, which favors its access to tumor tissues.<br />

Results from a phase I clinical trial showed a decrease in the<br />

side effects of patients who had been administered this drug<br />

<strong>and</strong> thus an enhancement in their quality of life. These favorable<br />

results prompted the initiation of a phase II trial designed<br />

to determine the side effects of this compound’s profile<br />

as maintenance therapy in ovarian cancer patients.<br />

Another polymer conjugate is poly(L-glutamic acid)camptothecin<br />

(CT-2106). This compound is composed by<br />

camptothecin covalently attached to poly(L-glutamic acid)<br />

through a glycine linkage. CT- 2106 has entered phase I/II<br />

trials <strong>and</strong> clinical studies have been conducted by Honsi et<br />

al. <strong>and</strong> Springett et. al. CT- 2106 was studied in patients with<br />

solid tumor malignancies [52].<br />

In addition, as mentioned previously, clinical studies of<br />

other polymer conjugates, PNU 166148 <strong>and</strong> PNU 166945,<br />

have been discontinued or suspended.<br />

PNU 166945 [53] is a HPMA polymer-conjugated derivative<br />

of paclitaxel. It was studied with the aim to improve<br />

drug solubility with the subsequent controlled release of paclitaxel.<br />

Toxicity associated with paclitaxel, such as neurotoxicity,<br />

was observed in phase I trials; however, this drug<br />

showed antitumor activity.<br />

PNU 166148 [54] is composed by N-(2hydroxypropyl)methacrylamide<br />

(HPMA) conjugated with<br />

camptothecin. The latter is modified at the C20 position of<br />

hydroxyl group with glycine <strong>and</strong> was conjugated with Gly-<br />

C6-Gly. In phase I trials, PNU 166148 was administered to<br />

patients with solid tumors. The results showed toxicity <strong>and</strong><br />

absence of tumor targeting. Thus, clinical development of<br />

this compound was ab<strong>and</strong>oned.<br />

2.2. Carriers in which the <strong>Drug</strong> is Embedded or Encapsulated<br />

Chemists working in synthesis seek to develop polymeric<br />

structures with physical <strong>and</strong> chemical properties that allow<br />

the specific release of the therapeutic agent under predetermined<br />

conditions.<br />

O<br />

O<br />

O<br />

O<br />

O O<br />

O<br />

O<br />

O<br />

O<br />

The polymers used in the development of drug delivery<br />

systems are designed with the capacity to form polymeric<br />

supramolecular structures (matrices <strong>and</strong> capsules), which are<br />

suitable for retaining a therapeutic agent <strong>and</strong> allow controlled<br />

release. Interactions between the polymer chains,<br />

which can be non-covalent (electrostatic <strong>and</strong> hydrogen<br />

bonds) or have covalent bonds (cross-linked polymers), confer<br />

stability to these structures. The integrity <strong>and</strong> the behavior<br />

of these structures in physiological conditions allow the<br />

controlled release of the therapeutic agent. The literature<br />

describes several mechanisms of drug release [55].<br />

One such mechanism is controlled release by swelling.<br />

Hydration of the polymer causes an increase in the volume<br />

of the polymeric structure <strong>and</strong> the resulting increase in pore<br />

size allows diffusion of the aqueous medium within the polymeric<br />

structure <strong>and</strong> thus the release of the drug.<br />

<strong>Drug</strong> release can also be achieved through polymer degradation.<br />

Gradual hydration of the polymeric structure may<br />

cause hydrolytic degradation with release of its contents.<br />

This degradation depends on the stability of the polymer<br />

chain linkages in body fluids Fig. (1.12).<br />

The release of a therapeutic agent depends on the rate of<br />

polymer degradation. Heterogeneous degradation occurs<br />

when the polymer is not fully degraded. This system requires<br />

two types of polymers, one degradable <strong>and</strong> located on the<br />

surface of the material to allow it to be in contact with the<br />

physiological environment, <strong>and</strong> the other one non-degradable<br />

<strong>and</strong> located inside the external polymer. In this case, the degradation<br />

rate is constant <strong>and</strong> the non-degraded material (material<br />

inside the external biodegradable scaffold) maintains<br />

its chemical integrity during the process. Thus, the therapeutic<br />

agent is released once the material surface has been degraded<br />

by diffusion through the non-degradable material. In<br />

contrast, homogeneous degradation is a r<strong>and</strong>om deterioration<br />

that occurs in the entire mass of the polymer matrix. While<br />

the molecular weight of the polymer decreases continuously,<br />

the material maintains its original shape <strong>and</strong> retains the drug.<br />

When the polymer mass is highly degraded <strong>and</strong> reaches a<br />

critical molecular weight, solubilization of the polymer <strong>and</strong><br />

the drug begins.<br />

Finally, another drug release mechanism is through pure<br />

diffusion. The drug slowly diffuses through the voids of the<br />

polymeric device. These polymers are fabricated as matrices<br />

in which the drug is uniformly distributed.<br />

Modulation of the chemical architecture of polymers<br />

(functional groups, interactions between chains…) can regulate<br />

the release of the therapeutic agent <strong>and</strong> thus the profile<br />

of the final formulation (delayed or immediate release, pulsatile,<br />

etc.). However, the release occurs through a combination<br />

of the three basic mechanisms mentioned previously<br />

Fig. (1.12). Linkages depicted from highest to lowest lability against hydrolysis in aqueous media, published by W. R. Gombotz, et al. [56].<br />

N H<br />

O<br />

O O<br />

Polyanhydrides > Polycarbonates > Polyesters > Polyurethanes > Polyorthoesters > Polyamides<br />

O<br />

O<br />

O<br />

N<br />

H


<strong>Polymers</strong> <strong>and</strong> <strong>Drug</strong> <strong>Delivery</strong> <strong>Systems</strong> Current <strong>Drug</strong> <strong>Delivery</strong>, 2012, Vol. 9, No. 4 11<br />

(release by swelling control, by pure diffusion <strong>and</strong> by polymer<br />

degradation).<br />

There are several pharmaceutical formulations in which<br />

the drug is embedded or encapsulated in the polymer.<br />

Among these, we find the micro- <strong>and</strong> nano-particles, micro<strong>and</strong><br />

nano-capsules, capsosomes <strong>and</strong> micelles. In the first<br />

case, the therapeutic agent is normally absorbed into the particles<br />

<strong>and</strong> the release is caused by diffusion through the pores<br />

of the polymer. In contrast, in the case of microcapsules, the<br />

drug is encapsulated within these structures <strong>and</strong> generally its<br />

release requires the degradation of the polymeric matrix.<br />

Other interesting carrier polymers are capsosomes, which<br />

consist of a drug carrier with an external core <strong>and</strong> some<br />

liposomes inside, thus allowing various compartments inside<br />

its structure. Finally, micelles are composed of amphipathic<br />

linear polymers that are spontaneously formed by selfassembly<br />

in water <strong>and</strong> the drug is encapsulated in the hydrophobic<br />

core.<br />

Furthermore, dendrimers are polymers with interesting<br />

properties as drug carriers because of their specific surface<br />

functional groups, which allow the attachment of hydrophilic<br />

drugs by covalent bonding, electrostatic interactions or hydrogen<br />

bonding.<br />

2.2.1. Gels <strong>and</strong> Hydrogels<br />

Gels <strong>and</strong> hydrogels are hydrophilic polymers that vary in<br />

their structures. The former have a linear structure while the<br />

latter are polymeric networks with three-dimensional covalent<br />

crosslinking.<br />

Therefore, linear polymer gels are formed by long-chain<br />

monomer units linked by covalent bonds such as amides,<br />

esters, orthoesters, <strong>and</strong> glycosidic bonds. Once the linear<br />

polymers are formed, other types of interactions (hydrogen<br />

bonds or Van der Waals interactions) contribute to achieve<br />

the three-dimensional structure. In contrast, hydrogels involve<br />

monomers of distinct chains that are covalently linked.<br />

Covalent bonds between chains affect polymer properties<br />

<strong>and</strong> thus make these polymers convenient for use as drug<br />

carriers in the form of micro- or nano-particles.<br />

One of the main characteristics of hydrophilic polymers<br />

is their capacity to absorb large amounts of water. The affinity<br />

for water is attributed to the presence of hydrophilic<br />

groups in their structure. This interesting property has led to<br />

the use of these polymers in several fields of biotechnology,<br />

such as biosensors <strong>and</strong> drug carriers.<br />

As an example, Pishko et al. [57] developed a fluorescent<br />

biosensor consisting of seminapthofluorescein conjugated<br />

with the enzyme organophosphorus hydrolase (SNALF-1)<br />

which is encapsulated in a PEG hydrogel. This enzyme catalyzes<br />

the hydrolysis of neurotoxins, thus leading to an increase<br />

in pH. In addition, the fluorescent compound,<br />

seminapthofluorescein, which is conjugated to the enzyme, is<br />

labile to changes in pH, thereby causing a change in its emission<br />

spectrum. Therefore, when the biosensor is incubated in<br />

an aqueous medium containing a neurotoxin such as paraoxon,<br />

at first the polymer swells, thus allowing diffusion of<br />

the water <strong>and</strong> toxin inside the polymer. Once the organophosphorus<br />

hydrolase enters into contact with toxin, the latter<br />

is hydrolyzed, resulting in a change in the pH of the me-<br />

dium <strong>and</strong> thus a change in the emission spectrum of the<br />

seminapthofluorescein.<br />

Another interesting example is the study carried out by<br />

Syu et al. [58] on the design <strong>and</strong> synthesis of biosensors for<br />

urea concentration measurement. They immobilized urease<br />

enzymes in a core of carbon paper on which a polypyrrole<br />

matrix was deposited forming an external layer. Urea concentration<br />

measurement is based on ammonia concentration,<br />

which is the product obtained from the catalysis reaction of<br />

urease enzymes.<br />

Gels <strong>and</strong> hydrogels are also used as drug carriers. These<br />

are classified as matrices or capsules, depending on the way<br />

in which the drug is found within the polymer, either embedded<br />

(matrix) or encapsulated (capsules).<br />

2.2.2. Nano- <strong>and</strong> Micro-particles<br />

Nano- <strong>and</strong> micro-particles are formulations in which the<br />

drug is trapped or embedded within the polymer. The polymer<br />

serves to transport <strong>and</strong> protect the drug <strong>and</strong> to allow its<br />

controlled release. The drug carrier consists of hydrophilic<br />

polymers, gels <strong>and</strong> hydrogels being among these.<br />

Hydrogels are the most widely used hydrophilic polymers<br />

for this kind of pharmaceutical formulation because<br />

they are non-toxic <strong>and</strong> their three-dimensional structure allows<br />

controlled drug release.<br />

Since hydrogels were first introduced into the field of<br />

biomedicine, they have consistently shown excellent characteristics<br />

of biocompatibility, which is attributed to their<br />

physical properties that make them similar to living tissue,<br />

especially regarding their high water content <strong>and</strong> also soft<br />

consistency <strong>and</strong> elasticity.<br />

2.2.2.1. Synthetic Hydrogels<br />

Hydrogels are obtained by simultaneous polymerization<br />

<strong>and</strong> crosslinking of several polyfunctional monomers. Therefore,<br />

a crosslinking agent is required to obtain the hydrogel<br />

structure. Different types of chemistry have been used to<br />

generate cross-linked polymers, among these the attack of<br />

the nucleophile to an electrophile, or radical chemistry.<br />

The characteristics of the monomers used <strong>and</strong> the degree<br />

of crosslinking determine hydrogel properties, such as swelling,<br />

<strong>and</strong>, therefore, its applicability.<br />

A number of synthetic hydrogels have also been studied<br />

for drug delivery purposes, such as PHEMA (2-hydroxyethyl<br />

methacrylate), PMA (poly((metha)acrylic acid)) <strong>and</strong> copolymers<br />

of PVA (Poly(vinyl alcohol)) or PEG (polyethylene<br />

glycol) with acrylamides.<br />

2.2.2.2. <strong>Drug</strong> Loading into a Polymer<br />

Therapeutic agents can be physically entrapped in the<br />

polymeric matrix or covalently bound to the polymer backbone.<br />

Given the capacity of hydrogels to absorb large<br />

amounts of water <strong>and</strong> thus their capacity to swell, many studies<br />

have been devoted to this type of polymer, in which the<br />

therapeutic agent is embedded in the gel <strong>and</strong> released when<br />

the gel structure swells. The rate of drug release depends on<br />

crosslinking density (highly cross-linked systems provide<br />

slow drug release because of their small mesh size) <strong>and</strong> drug<br />

solubility.


12 Current <strong>Drug</strong> <strong>Delivery</strong>, 2012, Vol. 9, No. 4 Vilar et al.<br />

The drug can be loaded by in situ polymerization or can<br />

be embedded after the polymer has been synthesized.<br />

In the first case, microparticles of PEG hydrogels were<br />

synthesized by Peppas et al. [59]. These hydrogels were prepared<br />

by free radical copolymerization of the macromonomers<br />

poly (ethylene glycol) dimethacrylate (PEGDMA) <strong>and</strong><br />

poly (ethylene glycol) monomethacrylate (PEGMA) in a<br />

range of proportions, thus obtaining several degrees of<br />

crosslinking. This radical reaction required the use of a photoinitiator,<br />

a UV lamp, <strong>and</strong> distilled water as a solvent of the<br />

reaction. The therapeutic agent, diltiazem, was loaded by<br />

adding it to the reaction mixture prior to exposure to the UV<br />

light source.<br />

As mentioned earlier, a drug can also be absorbed into<br />

preformed polymers. This requires the incubation of a concentrated<br />

drug solution with the preformed polymer carrier.<br />

Varsohaz et al. [60] used theophylline as a model drug to<br />

be encapsulated in a polymer cross-linked PVA. In this case,<br />

the polymer was synthesized by the nucleophilic attack of<br />

the alcohol groups in the PVA structure <strong>and</strong> the aldehyde<br />

groups in the glutaraldehyde molecule, thus forming a crosslinked<br />

structure. Encapsulation of the drug was achieved by<br />

exposing the polymer previously synthesized with a drug<br />

solution in NaOH. Several tests of drug encapsulation in the<br />

polymer were carried out by varying the drug concentration<br />

<strong>and</strong> polymer composition (proportion of glutaraldehyde).<br />

The results showed that the greater the proportion of glutaraldehyde,<br />

the greater the extent of drug loading.<br />

An article published by Nakamura et al. [61] cites other<br />

interesting loaded preformed microparticles. In this case,<br />

Nakamura et al. addressed the behavior <strong>and</strong> kinetics of<br />

budenoside release from microparticles of P(MAA-g-EG).<br />

The latter is a copolymer of polymethacrylic acid <strong>and</strong> polyethylene<br />

glycol. They studied various methods of absorption<br />

because budesonide is soluble in ethanol but this solvent<br />

hinders the swelling of the resin <strong>and</strong> therefore the internalization<br />

of the drug within the polymer. Thus, the loading of<br />

the drug into the polymer was examined using a number of<br />

aqueous solutions with varying amounts of ethanol. It was<br />

observed that the smaller the amount of ethanol used in the<br />

solution, the greater the incorporation of the drug into the<br />

polymer.<br />

We can see from previous articles that absorption conditions<br />

used differ because of the distinct physicochemical<br />

properties of each polymer <strong>and</strong> drug. Thus, a drug can be<br />

soluble or unsoluble in an aqueous medium. The polymer<br />

may show distinct degrees of swelling depending on the pH<br />

of the medium or depending on the degree of crosslinking in<br />

the structure of the polymer. Therefore, all these characteristics<br />

may cause a variation in the partition coefficient of the<br />

drug between the polymer <strong>and</strong> the solution. Therefore, in<br />

these cases, the ideal is to obtain optimum conditions of polymer<br />

swelling <strong>and</strong> an acceptable degree of drug solubility.<br />

2.2.3. Nano- <strong>and</strong> Micro-Capsules<br />

Nano- <strong>and</strong> micro-capsules are formed by an outer shell<br />

that encapsulates the active agent. This kind of pharmaceutical<br />

formulation is commonly used to transport drugs, such as<br />

proteins, that are rapidly degraded in body fluids. Thus the<br />

drug is encapsulated in the particle <strong>and</strong> is released by degradation<br />

of the polymer coating.<br />

In general, micro- <strong>and</strong> nano-capsules are synthesized<br />

with hydrophilic polymers. Gels are the most widely used<br />

hydrophilic polymers for this kind of pharmaceutical formulation;<br />

however, example of micro- <strong>and</strong> nano-capsule synthesis<br />

with hydrogels can be found.<br />

Gels are made either from natural or synthetic polymers.<br />

Among the former, are the most widely used. These are<br />

polymers composed of repeated monomer units, monosaccharides,<br />

joined together by glycosidic bonds. Polysaccharide<br />

chitosan has been extensively used in controlled drug<br />

delivery. In addition, poly(D, L-lactic acid) PLA,<br />

poly(glycolic acid) PGA <strong>and</strong> their copolymer poly(D, Llactic-co-glycolic<br />

acid) PLGA are among the most commonly<br />

used synthetic polymers.<br />

Another interesting point concerns drug entrapment.<br />

Physical drug loading can be performed by incorporating the<br />

drug while producing the particles or by incubating a concentrated<br />

drug solution with the preformed polymer carrier.<br />

The optimum incorporation strategy for efficient entrapment<br />

should be selected on the basis of the physicochemical<br />

characteristics of the drug-carrier pair. Thus, entrapment<br />

efficiency depends on drug solubility in the polymeric matrix,<br />

which is in turn related to the composition <strong>and</strong> molecular<br />

weight of the polymer, drug-polymer interactions, <strong>and</strong> the<br />

presence of functional groups [62].<br />

The bibliography cites many examples in which the drug<br />

is physically incorporated after the micro- or nano-capsule<br />

has been synthesized. One instance is the study published by<br />

Z. Liu et al. [63] who addressed the use of sufopropyl dextran<br />

ion-exchange microparticles as carriers for drug delivery.<br />

These microparticles are hydrophilic polymers whose<br />

sulfopropyl group confers a net negative charge. They were<br />

loaded with the drug doxorubicin (Dox) to evaluate the anticancer<br />

activity of the drug released in vitro. Dox was incorporated<br />

into the microparticles by incubating these with an<br />

aqueous solution at room temperature. With the loaded microparticles,<br />

studies revealed that the drug release was dependent<br />

on the salt concentration of the medium.<br />

Furthermore, there are many methods for incorporating<br />

drugs during polymer production but the three most commonly<br />

used processes are phase separation, solvent evaporation<br />

<strong>and</strong> spray drying [64].<br />

Solvent evaporation methods have been widely used to<br />

prepare polymers loaded with various drugs. Solvent evaporation<br />

methods can be classified into two types, single emulsion<br />

or double emulsion, depending on the number of emulsions<br />

produced during the preparation of the drug-polymer.<br />

The single emulsion method (o/o or o/w) involves mixing<br />

a solution of the polymer dissolved in organic solvent with<br />

the drug. This solution is added on a continuous phase (immiscible<br />

with the first), which can be mineral oil (o/o) or an<br />

aqueous solution (o/w) <strong>and</strong> that contains an emulsifier whose<br />

function is to stabilize the emulsion produced. This emulsion<br />

is stirred or sonicated <strong>and</strong> the organic solvent is then removed<br />

by solvent evaporation or extraction. The coating material<br />

then shrinks around the core material, encapsulating it.


<strong>Polymers</strong> <strong>and</strong> <strong>Drug</strong> <strong>Delivery</strong> <strong>Systems</strong> Current <strong>Drug</strong> <strong>Delivery</strong>, 2012, Vol. 9, No. 4 13<br />

In double emulsion (w/o/w), an aqueous drug solution is<br />

first emulsified in a solution of the polymer in organic solvent<br />

<strong>and</strong> this emulsion (w/o) is added to an aqueous phase<br />

that contains an emulsifier, forming w/o/w emulsion. The<br />

organic solvent is then removed by extraction into an external<br />

aqueous phase <strong>and</strong> evaporated.<br />

Solvent evaporation is a method used to encapsulate hydrophobic<br />

(single emulsion) or hydrophilic (double emulsion)<br />

drugs in hydrophilic polymers such as PLA <strong>and</strong> PLGA.<br />

One example of the use of this method is the study by<br />

Jalón et al. [65] who synthesized PLGA microparticles<br />

loaded with rhodamine using solvent evaporation techniques<br />

for topical drug delivery. Specifically, these microparticles<br />

were prepared by the o/w solvent evaporation technique.<br />

Wada et al. [66] developed a new approach for the preparation<br />

of biodegradable lactic acid oligomer microspheres.<br />

These were obtained by the solvent evaporation method o/o<br />

(oil in oil). The solvent used for the dispersed phase solution<br />

was a mixture of acetonitrile <strong>and</strong> water, while the continuous<br />

phase medium was cottonseed oil. Dox <strong>and</strong> insulin were encapsulated<br />

in microspheres with an efficiency of 80-90%.<br />

Iwata et al. [67] synthesized PLA/PLGA microspheres<br />

containing water-soluble drugs, using a multiple emulsion<br />

solvent evaporation technique. Some drugs cannot be encapsulated<br />

with traditional solvent evaporation methods because<br />

of their high water solubility. This new method leads to<br />

greater drug loading efficiency into the polymer <strong>and</strong> to the<br />

formation of microparticles with a heterogeneous drug distribution.<br />

They studied the drug release behavior of the microparticles<br />

synthesized with the new method <strong>and</strong> compared<br />

it with that of the microparticles produced with the traditional<br />

method. A burst of drug release was detected with<br />

both methods, but the microparticles synthesized with this<br />

new method showed a larger amount of drug released than<br />

corresponding conventional microparticles.<br />

Phase separation <strong>and</strong> coacervation is a widely used<br />

method for the formation of microcapsules. Basically, this<br />

method consists of three steps. In the first, the drug is dispersed<br />

in a solution of the coating polymer, which is in the<br />

liquid phase. Then, a phase separation of the polymer coating,<br />

caused by a change in temperature or by the addition of<br />

salt or a non-compatible polymer with the first, creates coacervate<br />

droplets of the polymer which are absorbed on the<br />

surface of the drug. Finally, in the last step, these droplets of<br />

polymer are solidified <strong>and</strong> stabilized, forming microcapsules<br />

with the drug inside them (Fig. 1.13).<br />

Fig. (1.13). Method of coacervation.<br />

This method was used by Nihant et al., [68] who studied<br />

the microencapsulation of a protein by coacervation of<br />

poly(lactide-co-glycolide).<br />

As discussed above, this coating technique proceeds<br />

through three steps. The first consists of a phase separation<br />

of the coating polyesters (dissolved in CH2Cl2) induced by<br />

silicone oil, in the second step adsorption of the coacervate<br />

droplets around the protein phase occurs, <strong>and</strong> finally, the<br />

third step involves the hardening of microparticles. The researchers<br />

explained that the viscosity of the coacervate, in<br />

direct connection with the CH2Cl2 content, determines the<br />

size distribution, surface morphology <strong>and</strong> internal porosity of<br />

the final microspheres. Specifically, internal porosity occurs<br />

when the coacervate droplets harden so fast that solvent<br />

droplets are retained within the polymer matrix. Thus, these<br />

solvent droplets will create the pores.<br />

The spray drying technique is used in the pharmaceutical<br />

industry to encapsulate active ingredients but it is also used<br />

to wrap food, oils, etc.<br />

This technique initially involves dissolving the hydrophobic<br />

or hydrolytic polymer in a solvent. Core particles are<br />

then dispersed in the polymer solution <strong>and</strong> the mixture is<br />

sprayed into a hot chamber through a nozzle of a spray dry.<br />

Consequently, the polymer solidifies onto the core particles<br />

when the solvent evaporates, resulting in microspheres that<br />

are precipitated into the bottom collector (Fig. 1.14).<br />

This technique is widely used for the synthesis of microspheres,<br />

such as those that contain chitosan <strong>and</strong> gelatin as<br />

polymer coating <strong>and</strong> budesonide as the encapsulated drug.<br />

This research was performed by Naikwade et al., who mixed<br />

a polymeric phase with a drug solution that contained the<br />

drug dissolved in methanol <strong>and</strong> water. This mixture was<br />

spray-dried to achieve microspheres [69].<br />

2.2.3.1. <strong>Polymers</strong> used in the Formulation of Nano- <strong>and</strong><br />

Micro-spheres<br />

Several polymers have been exploited for the formulation<br />

of capsule carrier systems.<br />

2.2.3.1.1. Polyesters<br />

Polyesters based on PLA, PGA, their copolymers PLGA<br />

<strong>and</strong> poly (-caprolactone) (PCL) [70] have been extensively<br />

employed because of their biocompatibility <strong>and</strong> biodegradability<br />

(Fig. 1.15).


14 Current <strong>Drug</strong> <strong>Delivery</strong>, 2012, Vol. 9, No. 4 Vilar et al.<br />

Fig. (1.14). Spray drying technique<br />

Polyesters are a large group of polymers characterized by<br />

the presence of esters bonds in the main chain. Their interest<br />

as biomaterials resides in the fact that the ester groups are<br />

hydrolytically degradable. Among the most commonly used<br />

are PLA <strong>and</strong> PLGA, which have been approved by the FDA<br />

for the development of drug delivery systems <strong>and</strong> other biomedical<br />

applications. These polyesters have both advantages<br />

<strong>and</strong> disadvantages when used for the encapsulation of therapeutic<br />

macromolecules. Degradation by non-enzymatic hydrolysis<br />

of PLA <strong>and</strong> PLGA may lead to an accumulation of<br />

acidic monomers, causing a decrease in local pH. Thus,<br />

when the encapsulated therapeutic agent is a protein it may<br />

denaturalize [71]. However, these two polyesters are the<br />

most widely used because of their non-toxic degradation<br />

products <strong>and</strong> adjustable degradation rate.<br />

In the case of PCL polymers, their biodegradation is<br />

slower than that of PLGA. This slowness leads to an increased<br />

O<br />

O<br />

Fig. (1.15). Polyesters.<br />

PGA PLA<br />

n<br />

O<br />

O<br />

O<br />

half-life of encapsulated molecules, thereby allowing a sustained<br />

drug release over prolonged periods.<br />

Many studies have used combinations of polymers [72]<br />

to modify surface properties <strong>and</strong> to change the speed of<br />

polymer degradation <strong>and</strong> thus the drug release rate.<br />

Considerable attention has been given to micro- <strong>and</strong><br />

nano-spheres as drug carriers. The synthesis, characterization<br />

<strong>and</strong> in vitro <strong>and</strong> in vivo behavior of these carriers, <strong>and</strong> therefore<br />

the profile of the drug release, have been extensively<br />

studied.<br />

Polyesters have been used for the encapsulation of many<br />

types of therapeutic agents.<br />

• Cancer<br />

Cytostatic drugs inhibit the disordered growth of cells,<br />

disrupting cell division <strong>and</strong> destroying the cells that are dividing<br />

rapidly. However, their toxic effect is not confined to<br />

malignant cells as they also exert their action on rapidly proliferating<br />

tissues such as skin, mucous membranes, bone<br />

narrow <strong>and</strong> so on. Therefore, to reduce the side effects of<br />

these drugs, their encapsulation in controlled release polymers<br />

is of great interest.<br />

Huo et al. [73] summarized the encapsulation of cisplatin,<br />

a potent anticancer agent, by using a biodegradable<br />

polymer such as PLGA. A solvent evaporation method<br />

(commented on earlier) was used to perform this encapsulation.<br />

Recently, another interesting work was published by Li<br />

et al. [74]. They proposed to reduce the toxicity of cisplatin<br />

by loading it onto Fe5 (Fe3+ doped hydroxyapatite at atomic<br />

ratio of Feadded/Caadded=5%) <strong>and</strong>/or encapsulating the<br />

latter or cysplatin alone into PLGA microspheres using oil in<br />

water single emulsion. Once the nanoparticles were obtained,<br />

studies about drug release profile <strong>and</strong> degradation behaviors<br />

were performed <strong>and</strong> the results demonstrated that Fe5/PLGA<br />

composite has a favorable drug release profile with a short<br />

initial burst release <strong>and</strong> a long release time.<br />

Another relevant intercalating agent is doxorubicin,<br />

which blocks the synthesis <strong>and</strong> transcription of DNA <strong>and</strong><br />

O<br />

O<br />

n<br />

O<br />

O<br />

x y<br />

PCL PHB<br />

O<br />

n<br />

O<br />

O<br />

PLGA<br />

n<br />

O<br />

O<br />

O<br />

O


<strong>Polymers</strong> <strong>and</strong> <strong>Drug</strong> <strong>Delivery</strong> <strong>Systems</strong> Current <strong>Drug</strong> <strong>Delivery</strong>, 2012, Vol. 9, No. 4 15<br />

also inhibits the enzyme topoisomerase II. This therapeutic<br />

agent has been encapsulated in PLA microspheres in order to<br />

obtain lower drug release <strong>and</strong> sustained action. Experiments<br />

of these formulations achieved a significant reduction of side<br />

effects in mice [75]. Ike et al. [76] conducted a study addressing<br />

the treatment of malignant pleural effusion with this<br />

type of formulation <strong>and</strong> observed that local administration of<br />

microspheres produced an low systemic drug concentration,<br />

therefore improving patient quality of life.<br />

5-Fluorouracil is a potent anti-metabolite that is used in<br />

some cancer therapies. This drug acts on DNA synthesis <strong>and</strong><br />

causes cell death. Nagarwal et al. [77] synthesized nanospheres<br />

of PLA polymer encapsulating 5-flourouracil. These<br />

carriers were synthesized to topical ocular applications.<br />

• Bacterial <strong>and</strong> Parasitic infections<br />

Nanotechnology offers drug delivery systems that are<br />

easily endocytosed by phagocytic cells because of the nature<br />

of the polymer [78]. This idea is relevant when the target of<br />

the transported drug is the pathogen inside the infected<br />

phagocytic cell. Polymeric nanoparticles are efficient carriers<br />

<strong>and</strong> after their in vivo administration they are endocytosed by<br />

phagocytic cells, thereby releasing their drug load.<br />

Therefore, this feature has been used in various studies<br />

such as those performed by Italia et al. <strong>and</strong> Espuelas et al. on<br />

the in vitro anti-leishmanial activity of biodegradable<br />

nanoparticles. Leishmaniasis is a disease caused by obligate<br />

intracellular parasites. To minimize the adverse effects associated<br />

with amphotericin B, it was encapsulated into biodegradable<br />

nanoparticles of PLGA [79] <strong>and</strong> PCL [80] for release<br />

at the intracellular level in macrophages. Another study<br />

on the treatment of the same disease performed by Nahar et<br />

al. [81] involved encapsulating amphotericin into PLGA<br />

particles with anchored mannose, which is a macrophage<br />

target.<br />

Nanoparticles have been examined for the treatment of<br />

other intracellular bacterial infections, such as those caused<br />

by mycobacteria. These types of bacteria include pathogens<br />

that cause serious diseases in mammals, including tuberculosis<br />

<strong>and</strong> leprosy. In classical treatments of mycobacterial diseases,<br />

the drug reaches the macrophages infected with mycobacteria<br />

in low amounts <strong>and</strong> sometimes it does not persist<br />

long enough to develop the desired antimycobacterial effect.<br />

Therefore, the treatment involves prolonged large systemic<br />

doses, which are associated with unwanted side effects. As a<br />

result, some studies have addressed the treatment of mycobacterial<br />

diseases using drug carriers. For example, P<strong>and</strong>ey<br />

et al. [82] encapsulated the antibiotic streptomycin in nanospheres<br />

of poly-lactide-co-glycolide (PLG) using the multiple<br />

emulsion technique. This formulation was administered<br />

orally to mice <strong>and</strong> an increase in the bioavailability of the<br />

encapsulated antibiotic was observed. Moreover, Suarez et<br />

al. [83] synthesized PLGA microspheres with rifampicin,<br />

encapsulated via the spray drying technique. These microspheres<br />

were administered into the lungs of an animal model<br />

<strong>and</strong> resulted in a reduction of bacteria in the inoculated animals.<br />

Other antibiotics, such as cefazolin [84] <strong>and</strong> nafcillin<br />

[85], have been encapsulated into biodegradable nano- <strong>and</strong><br />

micro-spheres. The former was encapsulated into PLGA<br />

microspheres <strong>and</strong> the latter into PLGA nanospheres. Their<br />

activity was evaluated with pathogenic staphylococcus bacteria.<br />

Other drugs, such as ciprofloxacin [86] <strong>and</strong> rifampicin<br />

[87], have also been encapsulated into this kind of PLGA<br />

system.<br />

• Peptide Hormones<br />

Numerous studies have focused on polymer carriers that<br />

carry peptide hormones.<br />

The administration of peptides <strong>and</strong> proteins for therapeutic<br />

purposes is a challenge because of the enzymatic barriers,<br />

drastic chemical conditions in parts of the digestive system<br />

(such as the stomach), <strong>and</strong> low absorption of these compounds<br />

in the gut. Thus, almost all drugs of a peptide nature<br />

are given frequently via parenteral route because of susceptibility<br />

to degradation by enzymes. Therefore, as a result of<br />

the short in vivo half-lives of proteins, multiple injections are<br />

required to achieve a desirable therapeutic effect.<br />

The low bioavailability of peptide compounds has<br />

prompted the pharmaceutical industry to introduce new formulations<br />

to minimize the problems involved in administration.<br />

In recent years, considerable efforts have been made<br />

towards the encapsulation of proteins <strong>and</strong> peptides into polymeric<br />

particles such as PLGA <strong>and</strong> PLG because these<br />

polymers maintain the integrity <strong>and</strong> activity of the peptide or<br />

protein.<br />

Nutropin Depot® [88] is a polymeric formulation that is<br />

available in the market. This formulation consists of micronized<br />

particles of rhGH (growth hormone) embedded in<br />

biocompatible <strong>and</strong> biodegradable PLG microspheres. This<br />

system is indicated for the long term treatment of growth<br />

failure caused by the lack of adequate endogenous GH secretion<br />

<strong>and</strong> it is administered by subcutaneous injection. After<br />

administration, bioactive rhGH is released from the microspheres<br />

into the subcutaneous environment initially by diffusion<br />

<strong>and</strong> then by both polymer degradation <strong>and</strong> diffusion.<br />

Therefore, the in vivo profile is characterized by an initial<br />

rapid release followed by a slow decline in GH concentration.<br />

Recently, Rafi et al. [89] encapsulated rhGH into PLGA<br />

microparticles. The researchers studied the release behavior<br />

of the therapeutic agent <strong>and</strong> the results showed that the in<br />

vitro release of the hormone in porous particles is quicker<br />

than in non-porous particles; on the other h<strong>and</strong>, the hormone<br />

maintained a high average life in the body, as shown by the<br />

presence of bioactive rhGH in the serum of animals 30 days<br />

after one single injection.<br />

Nafarelin, another therapeutic agent of interest, was encapsulated<br />

into polymeric microspheres. Nafarelin is a potent<br />

hormone responsible for releasing gonadotropin. The study<br />

was conducted by S<strong>and</strong>ers et al. [90] who encapsulated the<br />

therapeutic agent into PLGA microspheres. Once these were<br />

administered as an injectable suspension, the drug was released<br />

by a diffusional mechanism, which resulted in a rapid<br />

pharmacological response followed by an abrupt termination.<br />

Another objective is to improve the bioavailability of<br />

peptides <strong>and</strong> proteins that are administered orally. For example,<br />

an overview of the work on PLGA formulations highlights<br />

some studies devoted to enhancing the bioavailability


16 Current <strong>Drug</strong> <strong>Delivery</strong>, 2012, Vol. 9, No. 4 Vilar et al.<br />

of hormones such as insulin [91]. However, most studies on<br />

this hormone are carried out with smart polymers, which will<br />

be discussed in the next section.<br />

Other polymers, such as polyanhydrides, polyamides,<br />

poly(ortho esters) <strong>and</strong> natural or modified polysaccharides,<br />

are also used as drug carriers.<br />

2.2.3.1.2. Polyanhydrides<br />

Polyanhydrides are bioabsorbable [92] materials that are<br />

useful in drug delivery because they are biocompatible [93]<br />

<strong>and</strong> degraded in vivo into their non-toxic diacid counterparts,<br />

which are easily removed from the body.<br />

The structure of polymers is characterized by repeated<br />

units linked by anhydride groups. These repeated units,<br />

called monomers, are diacids, which can have the same or a<br />

different chemical structure; in the latter case, they are referred<br />

to as copolymers.<br />

Most studies on polyanhydrides are based on sebacic<br />

acid (SA), p- (carboxyphenoxy)propane (CPP), p-<br />

(carboxyphenoxy)hexane (CHP) <strong>and</strong> their copolymers.<br />

Kumar et al. [94] reviewed the development, synthetic<br />

methods, structures <strong>and</strong> characterization of polyanhydrides<br />

(Fig. 1.16).<br />

Depending on their chemical structure, the monomers<br />

confer the polymer differing degrees of hydrophilic behavior,<br />

thereby affecting their degradation. The more hydrophobic<br />

the monomer, the more stable the anhydride bond is to<br />

hydrolysis. Also, by changing the ratio of monomers in the<br />

copolymers, their physical properties can be altered, thus<br />

polymer degradation <strong>and</strong> drug release behavior can be modified.<br />

In general, polyanhydrides are hydrolytically unstable<br />

polymers. The degradation of microspheres is pH-sensitive,<br />

being enhanced at high pH <strong>and</strong> becoming more stable in<br />

acidic conditions. Studies of polyanhydride degradation were<br />

carried out by Domb et al. [95] who used the co-polymer<br />

poly[1,3- bis(p-carboxyphenoxypropane):sebacic acid] in a<br />

ratio of 20:80 [P(CPP:SA)] 20:80] for this purpose. These<br />

O<br />

Fig. (1.16). Polyanhydrides.<br />

O<br />

CPP<br />

O O<br />

O<br />

O<br />

O<br />

n<br />

O<br />

authors used two types of radioactive labeling, one with<br />

[ 14 C]SA <strong>and</strong> unlabelled CPP, <strong>and</strong> the other co-polymer with<br />

[ 14 C]CPP <strong>and</strong> unlabelled SA implanted in the brain of rats.<br />

These studies showed the rate of degradation of the copolymer<br />

<strong>and</strong> the metabolic deposition of the degradation products.<br />

Therefore, the copolymer is biodegradable <strong>and</strong> can be<br />

used for drug delivery, resulting in a sustained release as a<br />

result of polymer degradation over a period of time depending<br />

on the composition of the polymer or co-polymer.<br />

Many studies have addressed the influence of the physico-chemical<br />

characteristics of polyanhydrides on drug release.<br />

It has been demonstrated that aliphatic polyanhydrides<br />

are degraded within days or weeks while aromatic polyanhydrides<br />

take months or even years. Kipper et al. [96] used<br />

microspheres of three bioerodible polyanhydrides, namely<br />

poly[1,6-bis (p-carboxyphenoxy) hexane] (poly (CHP)),<br />

poly(sebacic anhydride) (poly(SA)) <strong>and</strong> the copolymer<br />

poly(CHP-co- SA), to study the release of p-nitroaniline. The<br />

release profiles shown by the polymers differed. These differences<br />

are attributed to polymer erosion rates <strong>and</strong> to the<br />

different distribution of the drug in the polymer.<br />

On the other h<strong>and</strong>, the release profile also depends on the<br />

polarity of the drug, <strong>and</strong> therefore, on its different solubility<br />

in aqueous solution. Berkl<strong>and</strong> et al. [97] examined the release<br />

of drugs such as rhodamine B, p-nitroaniline <strong>and</strong> piroxicom,<br />

each of these with different water solubility <strong>and</strong><br />

encapsulated them into the surface of erodible polymer poly<br />

(sebacic anhydride) PSA. Researchers observed that each<br />

drug exhibited different release mechanisms <strong>and</strong> therefore<br />

distinct release profiles.<br />

It is important to indicate that this kind of polymer shows<br />

high reactivity, thus limiting its application for the encapsulation<br />

of nucleophilic drugs. Polyanhydrides can react with<br />

drugs containing nucleophilic functional groups, such as free<br />

amino groups, especially when a high temperature is required<br />

for encapsulation, thus limiting the type of drug that<br />

can be successfully incorporated into a polyanhydride matrix.<br />

CHP<br />

O<br />

O<br />

O<br />

SA<br />

O<br />

n<br />

O<br />

n


<strong>Polymers</strong> <strong>and</strong> <strong>Drug</strong> <strong>Delivery</strong> <strong>Systems</strong> Current <strong>Drug</strong> <strong>Delivery</strong>, 2012, Vol. 9, No. 4 17<br />

A wide variety of proteins [98] have been incorporated<br />

into polyanhydride matrices to improve in vivo protein stability.<br />

Recent studies have addressed the encapsulation of a<br />

range of proteins, such as insulin [99], human serum albumin<br />

(HSA) [100], <strong>and</strong> bovine serum albumin labeled with fluorescein<br />

isothiocynate (BSA-FITC) [101] into polyanhydride<br />

microspheres.<br />

Furthermore, antigens have been encapsulated. Thus,<br />

Carrillo-Conde et al. [102] encapsulated Yersinia pestis antigen<br />

into the co-polymers CHP with SA <strong>and</strong> co-polymer 1,8bis<br />

(p-carboxyphenoxy) -3,6-dioxooctane (CPTEG) with<br />

CHP . They studied the effect of the anhydride monomers,<br />

which are the degradation products, on the protein structures<br />

(antigen). The results showed that the environment provided<br />

by the CPTEG monomer was crucial for the preservation of<br />

the structure <strong>and</strong> antigenicity of the protein. Tetanus Toxoid<br />

(TT) [103] has also been used as model antigen for encapsulation<br />

into biodegradable polyanhydride microspheres to<br />

modulate the immune response mechanism. The copolymer<br />

CHP <strong>and</strong> SA were used for this purpose. In vivo studies in<br />

mice demonstrated that microspheres provided a prolonged<br />

exposure to antigen for sufficient time to induce primary <strong>and</strong><br />

secondary immune response.<br />

2.2.3.1.3. Polyamides<br />

Polyamides are polymers whose repeated units contain<br />

amide groups. Polyamides have some very attractive characteristics<br />

with respect to their use as biodegradable materials.<br />

First, these molecules have excellent mechanical properties<br />

afforded by the presence of amide groups <strong>and</strong> hydrogen<br />

bonds, <strong>and</strong> consequently display a highly polar behavior.<br />

Second, polyamides are metabolized into non-toxic products.<br />

However, the amide bond is hydrolyzed much more slowly<br />

than the ester linkages, so in general, the polyamide degradation<br />

rate is too slow for their practical use as biodegradable<br />

polymers. Therefore, the chemical structure of some polyamides<br />

is modified to accelerate degradation. Acceleration can<br />

be achieved through the copolymerization of monomers with<br />

other monomers to introduce a hydrolysable bond in the<br />

main chain. Thus, the amide bonds are combined in different<br />

proportions with other groups that are more susceptible to<br />

degradation.<br />

<strong>Polymers</strong> that contain ester <strong>and</strong> amide bonds in the backbone<br />

chain are the most interesting class of polyamides for<br />

controlled release. Generally, polyamides have been used to<br />

deliver low molecular weight drugs. These polymers are<br />

generally hydrophilic <strong>and</strong> their degradation rates depend on<br />

the hydrophilicity of the amino acids [104]. They are metabolized<br />

to non-toxic products.<br />

These polymers have been used to transport drugs encapsulated<br />

into a micro- or nano-sphere. Moreover, because<br />

their main structure is formed by amino acids, these can have<br />

a side chain that can provide sites for the attachment of drugs<br />

or pendant groups <strong>and</strong> can thus be used to modify the physico-chemical<br />

properties of the polymer or to transport the<br />

conjugated drug. For example, the amino acid lysine of the<br />

copolymer poly(lactic acid-co-lysine) (PLAL) provides an<br />

amino group that allows modifications of the PLAL systems.<br />

Capponetti et al. [105] synthesized <strong>and</strong> characterized biode-<br />

gradable microparticles with a PLAL backbone. Furthermore,<br />

in order to modify the physico-chemical properties of<br />

the polymer, pendant groups such as poly (L-lysine), poly<br />

(D, L, alanine) or poly (L-aspartic acid) were attached to the<br />

lysine residues on the backbone. Researchers studied the<br />

encapsulation <strong>and</strong> release of rhodamine B, a low molecular<br />

weight drug model, to demonstrate the capacity of microparticles<br />

to serve as carriers in controlled drug delivery devices.<br />

Hrkach et al. synthesized other types of copolymers such<br />

as poly(L-lactic acid-co-L-lysine), poly (L-lactic acid -co-Laspartate)<br />

<strong>and</strong> poly(L-lactic acid-co-D, L -alanine) [106].<br />

Furthermore, Vera et al. synthesized some biodegradable<br />

poly(ester amide) composed of SA, dodecanodiol <strong>and</strong> different<br />

ratios of the stereoisomers of L- <strong>and</strong> D-alanine. The microspheres<br />

were loaded with dicofenac sodium salt, triclosan<br />

<strong>and</strong> clofazimine to study the release rate in vitro [107].<br />

2.2.3.1.4. Poly (Ortho Esters)<br />

Over the last 40 years, poly(ortho esters) have been developed<br />

for drug delivery purposes. These esters comprise<br />

four families: POE I, POE II, POE III <strong>and</strong> POE IV [108].<br />

These differ in the type of monomers used <strong>and</strong> the synthesis<br />

of the polymer, which affect the physico-chemical features<br />

of the polymer <strong>and</strong> thus drug release.<br />

POE I polymers were prepared by transesterification reaction<br />

of diols <strong>and</strong> diethoxy-tetrahydrofuran. The drawback<br />

with this family is that degradation products are acids <strong>and</strong><br />

these catalyze polymer degradation (Fig. 1.17).<br />

Fig. (1.17). POE I.<br />

In contrast, POE II polymers were synthesized by addition<br />

of diol or polyol to diketene acetal. Cross-linked structures<br />

(hydrogels) can be obtained when the polymer is synthesized<br />

from a polyol with diketene acetal (Fig. 1.18).<br />

Fig. (1.18). POE II.<br />

O<br />

O O R<br />

POE II polymers were developed in order to prevent<br />

autocatalytic degradation because the degradation products<br />

are not acidic <strong>and</strong> therefore do not react with the ortho ester<br />

groups catalyzing their decomposition. These polymers were<br />

achieved by changing the initial acids to neutral products<br />

(diketene acetals). The problem in this case is that the degra-<br />

O<br />

O O O O R<br />

O<br />

n<br />

n


18 Current <strong>Drug</strong> <strong>Delivery</strong>, 2012, Vol. 9, No. 4 Vilar et al.<br />

dation of these polymers is very slow due to their hydrophobicity.<br />

Therefore, acid additives were added to the polymer<br />

matrix to accelerate degradation. Heller published a paper on<br />

the synthesis <strong>and</strong> use of various excipients to obtain controlled<br />

drug release of therapeutic agents that are physically<br />

dispersed in polymers. Heller [109] developed a polymeric<br />

system which is stable at physiological pH (7.4), but unstable<br />

at lower pH as a result of the excipients carried in the polymer.<br />

Therefore, polymer degradation can be controlled by<br />

these excipients <strong>and</strong> will take place only when the polymer<br />

<strong>and</strong> excipients are in contact with water, which causes a decrease<br />

in pH. Then, if the polymer is highly hydrophobic,<br />

only excipients that are on the surface layer will be exposed<br />

to water <strong>and</strong> thus polymer degradation will occur only on the<br />

surface layer.<br />

However, this approach proved only partially successful<br />

<strong>and</strong> POE II polymers were not developed for commercial<br />

use.<br />

Later on, the transesterificacion reaction was used again,<br />

but in this case, using trimethylorthoacetate <strong>and</strong> 1,2,6 hexanetriol,<br />

thus obtaining POE III polymers. However, the<br />

development of this family was ab<strong>and</strong>oned due to the difficulty<br />

in obtaining polymers with a specific molecular weight<br />

(Fig. 1.19).<br />

Fig. (1.19). POE III.<br />

Finally, the last family studied was POE IV [110], which<br />

consists of polymers synthesized from polyols <strong>and</strong> diketene<br />

acetals. However, in this case, latent acid monomers, such as<br />

fast hydrolysable glycolic or lactic acid, are introduced into<br />

the polymer backbone to obtain ester linkages. These linkages<br />

are rapidly hydrolyzed <strong>and</strong> their acid reaction products<br />

can catalyze autocatalytic degradation (Fig. 1.20).<br />

The amount of latent acid is small in order to maintain<br />

the hydrophobic nature of the polymer <strong>and</strong> it allows control<br />

over erosion rates. Therefore, POE IV polymers are autocatalyzed<br />

compounds that contain latent acid in the backbone,<br />

where erosion rates can be adjusted by varying the<br />

concentration of this acid. The synthesis of POE IV polymers<br />

is straightforward <strong>and</strong> reproducible <strong>and</strong> their structure<br />

Fig. (1.20). POE IV.<br />

O<br />

O O<br />

R'<br />

O<br />

O<br />

O<br />

O<br />

O<br />

R<br />

n<br />

CH 3<br />

O CH<br />

C<br />

O<br />

gives good mechanical <strong>and</strong> thermal properties, <strong>and</strong> allows<br />

control over the rate of degradation <strong>and</strong> therefore a regulated<br />

drug release rate. As a result of these features, this family of<br />

polymers has been commercialized.<br />

A number of studies have addressed the encapsulation of<br />

various drugs in poly (ortho esters) <strong>and</strong> their release kinetics.<br />

Bai et al. [111] encapsulated bovine serum albumin into a<br />

range of self-catalyzed poly (ortho esters). The experimental<br />

results suggested that drug release rates can vary depending<br />

on the composition of the polymer. Also, the researchers<br />

reported that the more hydrophilic the polymer, the faster the<br />

drug release. This is because the release occurs through two<br />

mechanisms: surface erosion <strong>and</strong> diffusion of the therapeutic<br />

agent. On the other h<strong>and</strong>, Chia et al. [112] proposed that the<br />

degradation of the polymer <strong>and</strong> therefore the release of the<br />

drug, which is encapsulated into microspheres of autocatalyzed<br />

poly (ortho esters), depends on the composition of<br />

the polymer (its hydrophobicity), its molecular weight <strong>and</strong> its<br />

morphology.<br />

Microspheres of poly (ortho esters) have been designed<br />

for enhancing the in vivo efficacy of newly developed DNA<br />

vaccine. This DNA vaccine is a plasmid that contains the<br />

DNA fragment of interest <strong>and</strong> an expression vector. This<br />

DNA encodes a protein antigen of interest that induces activation<br />

of the immune system. Therefore, it can induce a humoral<br />

response (antibody-mediated) <strong>and</strong> also cellular responses<br />

(cytotoxic response mediated by Tc lymphocytes).<br />

Therefore, this vaccine works by inserting the DNA of bacteria<br />

or viruses into human or animal cells. In many cases, the<br />

DNA is encapsulated into polymers such as polymeric microspheres<br />

[113] or liposomes in order to protect it from degradation.<br />

The composite polymer-DNA is phagocytosed by<br />

the cells <strong>and</strong> DNA is released into the cell as a result of the<br />

lower pH of the phagosome. Therefore, cells then begin to<br />

express proteins of foreign DNA <strong>and</strong> the immune system<br />

recognizes the protein <strong>and</strong> initiates a response. Wang et al.<br />

[114] described genetic vaccination, which consists of the<br />

encapsulation of a plasmid DNA into poly(ortho esters) microspheres<br />

in order to protect DNA from degradation. The<br />

results of the tests performed in mice indicated an activation<br />

of adaptive immunity with humoral <strong>and</strong> cytotoxic responses,<br />

leading to the suppression of the growth of tumor cells,<br />

which had phagocytosed the DNA-polymer system <strong>and</strong><br />

therefore expressed protein antigen.<br />

Other therapeutic agents, such as 5-fluorouracil [115]<br />

(antineoplastic agent) <strong>and</strong> dehydroepi<strong>and</strong>rosterone [116]<br />

(steroid hormone), have been encapsulated into microspheres<br />

of poly(ortho esters). Their encapsulation <strong>and</strong> release has<br />

been studied.<br />

O<br />

O R' O O<br />

m<br />

O<br />

O<br />

O R<br />

n


<strong>Polymers</strong> <strong>and</strong> <strong>Drug</strong> <strong>Delivery</strong> <strong>Systems</strong> Current <strong>Drug</strong> <strong>Delivery</strong>, 2012, Vol. 9, No. 4 19<br />

2.2.3.1.5. Natural or Modified Polysaccharides<br />

Natural <strong>and</strong> modified polysaccharides are biodegradable<br />

<strong>and</strong> biocompatible polymers. Polysaccharide polymers differ<br />

in their structure <strong>and</strong> conformation. The polymers can be<br />

formed by several monomers or by various types of Oglycoside<br />

bonds. Therefore, there are polymers with the<br />

same monosaccharide that differ in their O-glycoside bond<br />

configuration, or isomers, while others differ in the hydroxyl,<br />

which is part of the O-glycoside bond. Thus distinct<br />

configurational <strong>and</strong> constitutional isomers are obtained respectively.<br />

These polymers also differ in the length of the chain <strong>and</strong><br />

the number of branches in it.<br />

Chitosan is an interesting polymeric carrier. Properties<br />

such as biodegradability, low toxicity <strong>and</strong> good biocompatibility<br />

make it suitable for use in biomedical <strong>and</strong> pharmaceutical<br />

formulations. Its relevance is due to its cationic nature,<br />

which favors the encapsulation of compounds with opposite<br />

charge, therefore making it a good c<strong>and</strong>idate as a gene carrier.<br />

Cationic polymers condense DNA into a nanosized<br />

polymer by a self-assembly process, which consists of the<br />

electrostatic interaction of the positively charged polymer<br />

with the negatively charged DNA. These drug carriers may<br />

interact with the negatively charged cellular membrane, being<br />

internalized via endocytosis. In the intracellular environment,<br />

the carriers are located in the endosomes. These<br />

gene carriers have to leave the endosome before it fuses with<br />

lysosomes <strong>and</strong> before exogenous DNA is degraded by<br />

lysosomal enzymes. Once out of the endosome, the gene<br />

carrier will be translocated to the nucleus if the free DNA is<br />

located near this structure.<br />

Masotti et al. [117] studied the encapsulation of DNA<br />

into chitosan, obtaining nanoparticles of a definite size <strong>and</strong><br />

shape. The release of DNA from nanoparticles showed an<br />

initial burst release within 3 hours of incubation <strong>and</strong> then,<br />

DNA was released constantly up to 72 hours. On the other<br />

h<strong>and</strong>, Özbas-Turan et al. [118] discussed the possibility of<br />

encapsulating two plasmids, pMK3 <strong>and</strong> pPGL2, containing<br />

-galactosidase <strong>and</strong> luciferase as reporter genes, in the same<br />

microsphere structure as chitosan. Double plasmid-loaded<br />

chitosan microspheres were prepared by complex coacervation<br />

to obtain high encapsulation efficiency. They conducted<br />

studies on the in vitro <strong>and</strong> in vivo release of the plasmid. The<br />

in vitro release of the plasmid was studied spectrophotometrically,<br />

observing continuous drug release. In contrast, in<br />

vivo release was determined by the expression of the galactosidase<br />

<strong>and</strong> luciferase proteins. After transfections,<br />

these microspheres showed a high production of these proteins.<br />

Another key property of the polysaccharide chitosan is its<br />

mucoadhesivity; for this reason, it has been studied as a drug<br />

carrier in the pulmonary route [119].<br />

2.2.4. Smart <strong>Polymers</strong><br />

Recent years have witnessed increasing attention to smart<br />

polymers as drug carriers because their properties can be<br />

tailored to fit the requirements required for drug delivery.<br />

Hydrogels are a good choice for their use in controlled<br />

drug release because they are biocompatible <strong>and</strong> display satisfactory<br />

swelling properties in aqueous media.<br />

The main interest in studying these materials is to control<br />

the swelling by specific stimuli that will allow the controlled<br />

release of bioactive molecules. Stimuli such as pH, electrical<br />

signals, <strong>and</strong> temperature can change the features of the system,<br />

such as pore size <strong>and</strong> permeability. Therefore, once this<br />

stimulus causes polymer swelling, the drug is released<br />

through a diffusion mechanism.<br />

The volume of the hydrogel or the degree of swelling<br />

depends on the balance between the attractive <strong>and</strong> repulsive<br />

interactions present in the network. Thus, the combination of<br />

molecular interactions such as van der Waals forces, hydrogen<br />

bonds, hydrophobic interactions <strong>and</strong> electrostatic interactions,<br />

determine the degree of hydrogel swelling at equilibrium.<br />

Researchers modify these structural properties to design<br />

these intelligent polymers.<br />

The smart polymers that received most attention are those<br />

that respond to stimuli of temperature <strong>and</strong> pH because these<br />

are inherent variables of physiological systems.<br />

2.2.4.1. pH-Sensitive Hydrogels<br />

These polymers are characterized by the presence of acid<br />

groups, such as carboxylic acids or sulfonic acids, or also<br />

basic groups, such as amines, in their structures. These<br />

groups can accept or donate protons depending on the pH of<br />

the medium, leading to ionization variations in their structure.<br />

In the case of polymers [120] that contain acidic groups,<br />

when the pH of the medium rises, the degree of polymer<br />

ionization increases. In contrast, in the case of polymers that<br />

have amino groups in their structure, when the pH of the<br />

medium rises, the degree of polymer ionization decreases.<br />

This change in the degree of ionization causes an alteration<br />

in the swelling as a result of electrostatic repulsion forces<br />

between the charges present. Therefore, a way to decrease<br />

the repulsion between charges is to ensure polymer swelling,<br />

thereby maintaining the charges as far apart as possible.<br />

pH-sensitive hydrogels have been used to develop controlled<br />

release formulations for oral administration. Lowman<br />

et al. [121] studied the use of a hydrogel that responds to a<br />

change in pH for the transport of orally administered insulin.<br />

This drug is a peptide labile to proteolytic degradation in the<br />

acidic stomach. Therefore, the researchers examined a drug<br />

carrier that would protect the peptide in this organ <strong>and</strong> allow<br />

its release in the intestine where the pH is slightly alkaline.<br />

This drug carrier was poly(methacrylic-g-ethylene glycol)<br />

(P(MMA-g-EG)) <strong>and</strong> the peptide was absorbed in the preformed<br />

polymer. Thus this pH-responsive carrier protected<br />

insulin in the acidic environment of the stomach as a result<br />

of the intermolecular interaction that prevented the hydrogel<br />

from swelling. But once the microparticles reached alkaline<br />

<strong>and</strong> neutral environments, namely the intestine, the interactions<br />

that occurred previously were lost <strong>and</strong> the pore size of<br />

the hydrogel increased, thus allowing insulin release.<br />

pH-sensitive hydrogels have also been used to produce<br />

biosensors. Generally, hydrogels are loaded with enzymes<br />

that change the local pH of the microenvironment inside the


20 Current <strong>Drug</strong> <strong>Delivery</strong>, 2012, Vol. 9, No. 4 Vilar et al.<br />

gels. For example, there is an enzyme that catalyzes the conversion<br />

of glucose to gluconic acid. The formation of gluconic<br />

acid leads to a decrease in local pH, thereby resulting<br />

in a change in the swelling of the hydrogel. This mechanism<br />

is widely used for the release of insulin [122].<br />

2.2.4.2. Temperature-Sensitive Hydrogels<br />

In the last decade, phase transition in temperaturedependent<br />

hydrophilic hydrogels, has been studied. These<br />

phase transitions are manifested in significant changes in the<br />

volume of hydrogels.<br />

Temperature sensitive hydrogels are classified into negatively<br />

thermosensitive, positively thermosensitive <strong>and</strong> thermally<br />

reversible gels. Some polymers increase their hydrophilicity<br />

as the temperature increases, leading to an expansion<br />

in their structure (positively thermosensitive) while others<br />

decrease their hydrophilicity when there is a temperature<br />

increase, leading to a collapse of the polymer <strong>and</strong> polymer<br />

shrinkage (negatively thermosensitive).<br />

To obtain thermosensitivity properties, the polymer must<br />

have a correct balance of hydrophilic <strong>and</strong> hydrophobic<br />

chains in its structure. For instance, a polymer at room temperature<br />

in water forms hydrogen bonds; therefore, the<br />

polymer is swollen because its structure is hydrated. When<br />

the temperature increases, hydrogen bonds are weakened,<br />

while hydrophobic interactions are intensified <strong>and</strong> the polymer<br />

network shrinks. These observations lead to the conclusion<br />

that if the proportion of hydrophobic groups on the<br />

structure of the polymer rises, the temperature of transition<br />

between the shrunken to the swollen state increases.<br />

Negative temperature-sensitive hydrogels have a lower<br />

critical solution temperature (LCST) <strong>and</strong> they show an on/off<br />

drug release with on at a low temperature <strong>and</strong> off at a high<br />

temperature, thus allowing pulsatile drug release. One example<br />

of these hydrogels are copolymers of Nisopropylacrylamide<br />

(PNIAAm) whose value of LCST is<br />

around 32 °C. However, a very useful temperature for biomedical<br />

applications is close to that of body temperature, <strong>and</strong><br />

therefore an adjustment of the LCST of PNIPAAm can be<br />

achieved by copolymerizing with other monomers which<br />

will change the degree of hydrophobicity <strong>and</strong> consequently<br />

the LCST value of the resulting polymer. Other polymers<br />

which can be used by delivery of therapeutic molecules are<br />

poly(N, N-dietylacrylamide) (PDEAAm) <strong>and</strong> poly(Nvinylcaprolactam)<br />

(PVCL).<br />

In contrast, positive temperature-sensitive hydrogels have<br />

an upper critical solution temperature (UCST) <strong>and</strong> swell as a<br />

result of increased temperature. Some examples of these are<br />

poly(acrylic acid) (PAA), polyacrylamide (PAAm) <strong>and</strong><br />

poly(acrylamode-co-butyl methacrylate).<br />

Other temperature sensitive hydrogels are thermoreversible.<br />

Some of these are Pluronics, Tetronics <strong>and</strong> poloxamer).<br />

This type of polymer lends itself to many applications,<br />

one of the most important one being as drug carriers. Na et<br />

al. [123] conducted a study on the changes in the structure of<br />

a hydrogel versus temperature <strong>and</strong> how this change affects<br />

drug release. Specifically, they studied the changes in porosity<br />

on the surface of elastin polypeptide (ELP) microspheres.<br />

The transition temperature (Tt) of the polymer was deter-<br />

mined by differential scanning calorimetry (DSC) <strong>and</strong> this<br />

coincided with a change in volume of the polymer. The morphology<br />

of the microparticles was examined by field emission<br />

scanning electron microscopy. Finally, in drug delivery<br />

studies, it was observed that when the temperature was increased<br />

above the transition temperature, drug release was<br />

much faster. Therefore, when the temperature is higher than<br />

the transition temperature, the polymer structure, <strong>and</strong> thus<br />

the pores, change, resulting in drug release.<br />

Temperature-sensitive hydrogel systems are widely studied<br />

for use in body-temperature control. These systems depend<br />

on the patient’s temperature when the polymer is introduced.<br />

When the body temperature increases <strong>and</strong> exceeds a<br />

certain value, the hydrogel swells, leading to the release of<br />

the drug, whose function is to reduce the patient’s temperature.<br />

Once the temperature decreases to normal values, the<br />

output of the drug from the polymer matrix decreases dramatically<br />

as a result of the shrinkage of the polymer structure.<br />

Recent work by Curcio et al. [124] described the synthesis<br />

of a thermosensitive polymer based on hydrolyzed gelatin.<br />

One of the interesting features of this polymer is its transition<br />

temperature, which is slightly above body temperature.<br />

Diclofenac sodium salt, a non-steroidal anti-inflammatory,<br />

was encapsulated <strong>and</strong> used as a model drug to study the behavior<br />

of these microspheres.<br />

Another example of this application is found in a study<br />

by Yoon Ki Joung et al. [125], who encapsulated indomethacin<br />

into temperature-sensitive polymers. This drug reduces<br />

fever, pain <strong>and</strong> swelling. The researchers encapsulated the<br />

therapeutic agent into PLGA microparticles, which were<br />

prepared by the oil-in water (O/W) emulsion method. Once<br />

the microparticles were synthesized, they were embedded<br />

into a chitosan-pluronic hydrogel matrix, a temperaturesensitive<br />

hydrogel. One objective of introducing a layer of<br />

chitosan-pluronic hydrogel into PLGA microparticles was to<br />

reduce the drug-release rate of microparticles synthesized<br />

exclusively with PLGA, while the other was to control this<br />

release by the stimulus of temperature.<br />

2.2.5. Capsosomes<br />

Capsosomes are polymeric structures that have a biodegradable<br />

outer layer <strong>and</strong> contain liposomes, which compartmentalize<br />

the interior of the capsule polymer.<br />

These systems show permeability of the outer membrane,<br />

thereby allowing small molecules to cross by passive diffusion.<br />

Moreover, the liposomes in these systems divide the<br />

interior of the capsule into compartments, which can contain<br />

the same molecule or different ones. At the same time, these<br />

encapsulated molecules can be hydrophobic, <strong>and</strong> therefore<br />

will bind to the liposome membrane, or hydrophilic, which<br />

will be located in the aqueous core of liposomes.<br />

Capsosome synthesis is complex. First, a structure called<br />

the core is required, on which the capsosome will be synthesized.<br />

A layer of a polymer called the precursor is added<br />

onto this core <strong>and</strong> the first layer of liposomes is immobilized<br />

on it. After adding this first layer of liposomes, a second<br />

polymeric layer called the separation layer is added, <strong>and</strong> onto<br />

this, a second layer of liposomes is added again. After add-


<strong>Polymers</strong> <strong>and</strong> <strong>Drug</strong> <strong>Delivery</strong> <strong>Systems</strong> Current <strong>Drug</strong> <strong>Delivery</strong>, 2012, Vol. 9, No. 4 21<br />

ing the last layer of liposomes, the outermost layer of the<br />

capsosomes is built, which consists of biodegradable polymers.<br />

Finally, the structure around which the core was synthesized<br />

is removed by degradation to obtain the capsosome<br />

structure.<br />

One of the most studied subjects in capsosomes is the<br />

enzyme catalysis reaction. For this purpose, the enzymes<br />

were encapsulated into the different compartments<br />

(liposomes). The semi-permeable nature of the outermost<br />

layer of the capsosomes allows small substrates <strong>and</strong> products<br />

to diffuse inside these capsules. Furthermore, the lipidic<br />

membrane of the liposomes has a transition temperature<br />

phase that is normally moderately higher than the physiological<br />

temperature, so the structure of the lipidic membrane<br />

changes slightly at this temperature <strong>and</strong> facilitates the entry<br />

<strong>and</strong> exit of compounds into <strong>and</strong> from the liposome respectively.<br />

Therefore, encapsulated enzymes are accessible to<br />

their substrates; however, they cannot leave the liposomes<br />

because they are high molecular weight molecules <strong>and</strong> cannot<br />

diffuse through membranes.<br />

Enzymes such as glucose oxidase [126], peroxidase,<br />

chymotrypsin [127], catalase [128] <strong>and</strong> urease [129] have<br />

been encapsulated into capsosomes <strong>and</strong> these systems have<br />

been reported to show catalytic activity. Kreft <strong>and</strong> colleagues<br />

[130] published a paper about the synthesis of a capsosome<br />

with two coupled enzymatic reactions, in which the two enzymes<br />

were found in separate compartments. The catalyst<br />

system was composed of glucose oxidase <strong>and</strong> horseradish<br />

peroxidase.<br />

Finally, in a study by Hosta [131] <strong>and</strong> colleagues, a<br />

model peptide (thiocoraline) was introduced into the membrane<br />

of liposomes. This was the first step towards the possibility<br />

of incorporating transmembrane proteins that can act<br />

as specific transport proteins or as enzymes that must be attached<br />

to the membrane to retain their function.<br />

Therefore, the development of systems with compartments<br />

<strong>and</strong> the success in the encapsulation of various enzymes or<br />

proteins with diverse functions illustrate the potential of<br />

these systems as drug carriers <strong>and</strong> also as microreactors.<br />

2.2.6. Micelles<br />

These copolymers are composed of individual linear<br />

polymers that contain a hydrophobic <strong>and</strong> hydrophilic segment,<br />

which confers them the capacity to form micellar<br />

structures in aqueous media. These amphipathic polymers<br />

are arranged so that the hydrophobic part is located inside<br />

the structure, while the hydrophilic part is located outside, in<br />

contact with the aqueous medium. The block-copolymer<br />

micelles form spontaneously by self-assembly in water when<br />

the concentration of the amphiphilic block copolymer is<br />

above the critical micellar concentration (CMC) [132].<br />

Micelles are used as drug carriers, in which low molecular<br />

weight drugs are incorporated inside them, thereby improving<br />

therapeutic efficacy. Therefore, the hydrophobic<br />

core of polymeric micelles facilitates the incorporation of<br />

hydrophobic drugs either through covalent or non-covalent<br />

interactions.<br />

<strong>Drug</strong> loading into micelles can be achieved several ways<br />

[133]. One approach is to add the drug <strong>and</strong> the copolymer in<br />

a solvent miscible with water, <strong>and</strong> then to dialyze the mixture<br />

in aqueous medium. The drug-loaded polymeric micelles<br />

are formed as the solvent is replaced by water. Another<br />

method is to introduce the drug into the polymer by a<br />

method of oil in water emulsion (o/w). First, an aqueous solution<br />

of copolymer micelles is prepared, <strong>and</strong> an o/w emulsion<br />

is formed by adding the drug solution in chloroform.<br />

The copolymer emulsion stabilized o/w is stirred until the<br />

chloroform is evaporated <strong>and</strong> the micelles are loaded with<br />

the drug.<br />

Micellar systems have several advantages as drug carriers.<br />

Among these is their capacity to encapsulate hydrophobic<br />

drugs, thereby allowing circulation through the blood as<br />

a result of the high solubility <strong>and</strong> stability of the complex in<br />

physiological medium. Moreover, the drug is protected from<br />

metabolization by enzymes or other bioactive species of the<br />

physiological medium. Finally, the release rate is controlled<br />

by the stability of the micelles, the hydrophobic micellar<br />

core, <strong>and</strong> the links or interactions used to attach the drug to<br />

the polymer backbone.<br />

Another advantage of micelles is their long half-life in<br />

the body because they cannot be removed via kidney filtration<br />

cause of their large size. As a result of long-term circulation,<br />

the drug activity continues for a long period after a<br />

single injection. These polymeric structures are slowly excreted<br />

by the renal route as a result of their composition <strong>and</strong><br />

structure. The micellar structure, which is formed by noncovalent<br />

intermolecular interactions between individual linear<br />

polymers, is in equilibrium with free-form linear ones.<br />

Thus, the polymer can be removed as an individual linear<br />

polymer from the micelle structure by the kidney because the<br />

polymer chain molecular weight is lower than the critical<br />

value for kidney filtration.<br />

The outer layer plays a crucial role in polymeric micelles.<br />

It is responsible for interactions with biostructures, such as<br />

proteins <strong>and</strong> cells, <strong>and</strong> this interaction determines the pharmacokinetics<br />

<strong>and</strong> biodistribution of the drug. The outer layer<br />

often consists of a polar poly(ethylene oxide) (PEO) that<br />

forms the shell of the carriers <strong>and</strong> protects its core.<br />

Therefore, in vivo drug release is controlled by the segments<br />

of the outer layer, which can improve the interaction<br />

with cells, together with the polarity of the micellar core,<br />

which controls the drug release rate, depending on the degree<br />

of hydrophobicity polymer core.<br />

Polymer micelles have been studied for the delivery of<br />

poorly soluble <strong>and</strong> toxic chemotherapeutic agents to treat<br />

cancer. These micelles can reach tumors by a passive<br />

mechanism (EPR effect) or they can be have targeting moieties<br />

conjugated to reach tumors by an active mechanism for<br />

enhanced delivery to a specific site [134].<br />

Some polymer micelle formulations of anticancer drugs<br />

have been reported to reach clinical trials.<br />

Kataoka’s group developed a micelle carrier (NK 911)<br />

[135] for anticancer therapy. This is the first c<strong>and</strong>idate of an<br />

antitumor drug on polymer micelles to have progressed to<br />

phase I clinical trials in Japan. NK911 is based on<br />

poly(ethylene oxide)-b-poly(aspartic acid) (PEO-b-PAsp)<br />

block copolymers conjugated with doxorubicin. PEO is believed<br />

to form the outer shell of the micelles <strong>and</strong> the doxoru-


22 Current <strong>Drug</strong> <strong>Delivery</strong>, 2012, Vol. 9, No. 4 Vilar et al.<br />

bicin conjugated poly(aspartic acid) chain is hydrophobic<br />

<strong>and</strong> forms the hydrophobic core of the micelles in aqueous<br />

media. The hydrophobic core enables NK911 to entrap a<br />

sufficient amount of doxorubicin. NK911 expresses higher in<br />

vivo antitumor activity than free doxorubicin because of the<br />

enhanced permeability <strong>and</strong> retention (EPR) effect. Parameters<br />

such as the maximum tolerated dose, dose limiting toxicities,<br />

<strong>and</strong> the recommended dose for phase II were determined.<br />

Another micelle formulation in clinical trials is SP1049C<br />

[136]. This is an anticancer agent that contains doxorubicin<br />

<strong>and</strong> two nonionic pluronic block copolymers. Doxorubicin is<br />

mixed with these systems, resulting in a spontaneous incorporation<br />

of the drug into micelles. In preclinical studies,<br />

SP1049C demonstrated increased efficacy compared to free<br />

doxorubicin. The toxicity profile, dose limiting toxicity,<br />

maximum tolerated dose <strong>and</strong> pharmacokinetic profile were<br />

determined in phase I.<br />

SP1049C was evaluated in phase II studies in patients<br />

with metastatic adenocarcinoma of the esophagus <strong>and</strong> esophageal<br />

junction. This agent showed a better response than<br />

when single drugs such as cisplatin, paclitaxel <strong>and</strong> docetaxel<br />

were used. In 2008, Supratek Pharma obtained FDA clearance<br />

for an investigative application for SP1049C for the<br />

treatment of metastatic adenocarcinoma. This formulation<br />

has been assigned to a pivotal phase III study protocol under<br />

the Special Protocol Assessment process.<br />

Finally, another biodegradable micelle, Genexol-PM, is<br />

in clinical trials [137]. Genexol is a micellar formulation<br />

characterized by a poly(ethylene glycol)-poly(D,L-lactide)<br />

copolymer that contains paclitaxel in the core. The copolymer<br />

residue increases the water-solubility of paclitaxel<br />

<strong>and</strong> allows the delivery of higher doses than free paclitaxel.<br />

The efficacy <strong>and</strong> safety of Genexol-PM <strong>and</strong> cisplatin were<br />

evaluated in phase II for the treatment of advanced nonsmall-cell<br />

lung cancer [138].<br />

2.3. <strong>Drug</strong> Carriers in which the <strong>Drug</strong> can be Embedded<br />

or Linked Covalently: Dendrimers<br />

The chemical structure of a dendrimer consists of a core<br />

or central structure that contains several branches, which in<br />

turn branch off, leading to a three-dimensional globular<br />

structure of concentric layers. Therefore, the dendrimeric<br />

structure is characterized by layers between each focal point<br />

(or cascade point) called generations, that is, a generation 4<br />

dendrimer is a system that has 4 focal points between the<br />

core <strong>and</strong> the surface. The nucleus is called generation zero<br />

because it does not have a focal point.<br />

The dendrimeric structure contains two key features that<br />

make them of interest as drug carriers. One is its interior<br />

hollows, which have precise chemical characteristics <strong>and</strong><br />

allow the encapsulation of drugs with hydrophobic properties.<br />

Another important feature of the dendrimer is that it<br />

contains specific surface functional groups that allow attachment<br />

of hydrophilic drugs by covalent bonds, electrostatic<br />

interactions or hydrogen bonds.<br />

The literature cites a variety of polymer configurations<br />

(lipid, polysaccharide, peptide, etc.). These dendrimers<br />

can have distinct functional groups, such as polyamines<br />

(dendrimer PPI), a mixture of amines <strong>and</strong> polyamides (dendrimer<br />

PAMAM) Fig. (1.21), <strong>and</strong> carboxylic acids. They<br />

may even hold more hydrophobic subunits such as poly(aryl<br />

ethers).<br />

Dendritic structures can be synthesized by two strategies,<br />

divergent or convergent synthesis. The former consists of<br />

synthesizing the dendrimer from the nucleus, the starting<br />

point on which the first layer will be joined. This results in a<br />

first-generation polymer that will be joined by the next layer<br />

<strong>and</strong> so the dendrimer will grow successively generation to<br />

generation. In the convergent synthesis, the segments of the<br />

dendrimer are synthesized separately, <strong>and</strong> once obtained,<br />

they are coupled to the core. The convergent synthesis starts<br />

building the dendrimer from the surface <strong>and</strong> ends up in the<br />

nucleus. Finally, a structure is obtained which is formed by a<br />

nucleus surrounded by a protective shell, since it forms a<br />

multivalent surface with a high number of reactive sites.<br />

Therefore, as mentioned above, the interior of a dendrimer is<br />

an ideal place to encapsulate guest molecules; alternatively,<br />

they can be conjugated or interacted with the multivalent<br />

surface of the dendrimer, which contains a high number of<br />

functional groups.<br />

Dendrimers are used for many treatments, including tumor<br />

[139], bacterial infection, antiviral, vaccine <strong>and</strong> gene<br />

therapy.<br />

Synthetic nanoscale dendrimers are multifunctional<br />

molecules capable of interacting specifically with tumor cells<br />

<strong>and</strong> they cause apoptosis. These controlled release drugs<br />

allow the administration of lower concentration doses <strong>and</strong> a<br />

therefore achieve a decrease in toxicity to healthy cells.<br />

One way of carrying dendrimers into malignant cells is<br />

through their conjugation with folic acid (FA) because the<br />

FA receptor is over-expressed in tumor cells. This is because<br />

FA is an essential ingredient for the replication of DNA <strong>and</strong><br />

therefore, is required in large amounts for the rapid cell division<br />

feature of cancer. The conjugation of dendrimers with<br />

FA also facilitates their internalization via receptor-mediated<br />

endocytosis.<br />

Quintana et al. [140] synthesized a generation 5 polyamidoamine<br />

dendrimer. The PAMAM dendrimer was conjugated<br />

with FA to obtain a specific focus, <strong>and</strong> with fluorescein<br />

isothiocynate (FITC), which is used as an imaging<br />

agent to track the localization of the dendrimer. And finally,<br />

it was also conjugated to methotrexate, a drug that inhibits<br />

FA metabolism. This drug specifically acts during DNA <strong>and</strong><br />

RNA synthesis; its cytotoxic activity occurs during the Sphase<br />

of the cell cycle. Once the dendrimer had been synthesized<br />

<strong>and</strong> conjugated, in vitro studies were performed with<br />

the KB cell line, which corresponds to human epidermoid<br />

carcinoma, a cell line that over-expresses folate receptors.<br />

These tests were performed with various dendrimer models.<br />

Amino groups that are on the surface of a structure of the<br />

generation 5 PAMAM, namely fluorescein <strong>and</strong> folic acid<br />

(G5-FITC-FA), <strong>and</strong> had not been conjugated by FA or FITC<br />

were capped with chemical groups such as carboxy or 2,3dihydroxypropyl<br />

or acetamide groups. Given the absence of<br />

repulsive forces from the charged amines, these new dendrimers<br />

showed a more relaxed structure than G5-FITC-FA.<br />

In contrast, G5-FITC-FA dendrimers bound poorly to KB


<strong>Polymers</strong> <strong>and</strong> <strong>Drug</strong> <strong>Delivery</strong> <strong>Systems</strong> Current <strong>Drug</strong> <strong>Delivery</strong>, 2012, Vol. 9, No. 4 23<br />

H 2N<br />

H 2N<br />

H<br />

N<br />

Fig. (1.21). PAMAM dendrimer.<br />

O<br />

H<br />

N<br />

cells, showing minimal uptake, thereby indicating the lack of<br />

specific interaction with the receptor. This finding is explained<br />

by the fact that some FA moieties are hidden inside<br />

the dendrimer, resulting in a reduced interaction with the<br />

receptor. In contrast, acetamide- G5-FITC-FA <strong>and</strong> 2,3dihydroxypropyl-<br />

G5-FITC-FA showed high interaction <strong>and</strong><br />

internalization in these cells. Therefore, these results demonstrated<br />

that the modification of the surface is crucial for improving<br />

the interaction of the dendrimer with the target cell.<br />

Majoros et al. [141] explained the synthesis, the characterization<br />

<strong>and</strong> in vitro <strong>and</strong> in vivo studies of PAMAM dendrimers<br />

conjugated with FA, FITC <strong>and</strong> methotrexate. A partial<br />

acetylation of the amino groups of the dendrimer surface<br />

was carried out to neutralize the structure <strong>and</strong> to prevent undesired<br />

reactions with molecules present in the body. The<br />

results of this study showed that dendrimers conjugated with<br />

FA inhibited the growth of KB cells carried out <strong>and</strong> showed<br />

a decrease in toxicity.<br />

Kono et al. [142] synthesized a polyether dendritic compound<br />

containing hydrazide groups on the surface which<br />

were conjugated with folate residues to carry methotrexate to<br />

tumor cells.<br />

O<br />

O<br />

N<br />

O<br />

N<br />

NH<br />

NH<br />

HN<br />

HN<br />

NH 2<br />

N<br />

O<br />

O<br />

NH 2<br />

H 2N<br />

O<br />

O NH<br />

N<br />

NH<br />

NH 2<br />

HN<br />

HN<br />

N<br />

O<br />

O<br />

N<br />

H<br />

O<br />

N N H<br />

O<br />

Furthermore, dendrimers have been used as carriers for<br />

other anticancer drugs, such as cisplatin [143], BH3 peptides<br />

[144] (which induce apoptosis), <strong>and</strong> 5-fluoro-uracil [145],<br />

the latter conjugated by an acid-labile linker to the surface of<br />

the dendrimer.<br />

A wide range of bacteria <strong>and</strong> yeast that cause human diseases<br />

are mediated by polyvalent interaction of carbohydrate<br />

<strong>and</strong> protein components with human cells.<br />

The main advantage of dendrimers as drugs is the number<br />

of potential sites for the attachment of functional surface<br />

groups that can grow exponentially generation to generation.<br />

The multiple surface groups on a single entity (multivalency<br />

or polyvalency) allow their association with pathogens. Dendrimer<br />

biocides that contain quaternary ammonium salts as<br />

functional end groups are an example. Quaternary ammonium<br />

compounds [146] are antimicrobial agents that disrupt<br />

bacterial membranes. PAMAM dendrimers have demonstrated<br />

significant antimicrobial activity because of their<br />

higher density of functional groups, which make the dendrimer<br />

more reactive to antimicrobial conjugation. The incorporation<br />

of antibacterial agents, such as sulfamethoxazole<br />

[147], silver salts [148], nadifloxacin <strong>and</strong> prulifloxacin<br />

NH 2<br />

NH 2


24 Current <strong>Drug</strong> <strong>Delivery</strong>, 2012, Vol. 9, No. 4 Vilar et al.<br />

[149], <strong>and</strong> niclosmide [150], has demonstrated significant<br />

antimicrobial activity.<br />

Moreover, lysine-based peptide dendrimers [151] have<br />

shown antimicrobial activity against Gram-positive (Staphylococcus<br />

aureus) <strong>and</strong> Gram-negative (Escherichia coli) bacteria<br />

as well as against fungal pathogens (C<strong>and</strong>ida albicans).<br />

Antiviral dendrimers attempt to mimic the anionic cell<br />

surface, so they are designed with surface anionic groups<br />

such as sulfonate residues or sialic acid, acidic carbohydrates<br />

that are present on the cell surface. Thus, the anionic dendrimer<br />

<strong>and</strong>/or the glycodendrimer compete with the cell surface<br />

by binding to the virus, leading to less likelihood of<br />

infection by the virus to the cell.<br />

PAMAM dendrimers [152] with a surface that had been<br />

modified covalently with naphthyl-sulfonate residues, inhibited<br />

HIV attachment <strong>and</strong> fusion by binding on the surface of<br />

the virus, thus preventing its binding to healthy cells. They<br />

also suppressed the early stages of viral replication.<br />

Furthermore, glycodendrimers, which are dendrimers that<br />

present multiple copies of carbohydrates on their surface, are<br />

widely used to address carbohydrate-protein interaction, conferring<br />

high selectivity to interact with the fusion protein of<br />

viruses. One example is the PAMAM dendrimer [153] conjugated<br />

with sialic acid. These sialic acids interact with glycoproteins<br />

localized on the surface of the influenza A virus,<br />

thus preventing viral adhesion to the sialic acid healthy cell<br />

surfaces. However, many viruses, such as influenza A virus<br />

<strong>and</strong> HIV, present a high degree of mutation, which leads to<br />

major changes in envelope protein glycosylation patterns<br />

when comparing different strains. This feature explains why<br />

it is so difficult to develop a vaccine against these viruses.<br />

Dendrimers are also used in vaccines. Haptens are low<br />

molecular weight molecules that induce a weak immune response<br />

unless their molecular weight is increased by polymerization<br />

or they are coupled to a high molecular weight carrier.<br />

Dendrimers are an excellent tool for vaccines because of<br />

their high capacity to combine a large number of haptens on<br />

the cell surface, thus inducing an immune response.<br />

MAP is the most widely used dendrimer for this purpose.<br />

This molecule has a multi-branched lysine core onto which<br />

haptens can be conjugated. MAPs have been developed as<br />

vaccines against cancer. Carbohydrate antigens, which are<br />

exposed on the surface of tumor cells, are potent targets for<br />

anticancer vaccines. Therefore, a multiple antigenic-linked<br />

dendrimer (MAG) that was synthesized carrying the Tn antigen<br />

linking to the carbohydrate tumor target, was developed.<br />

Hence, the dendrimer had the CD4 + T cell epitope to activate<br />

immune responses (lymphocytes T <strong>and</strong> B) specific for some<br />

types of carcinomas [154].<br />

In other experiments, Ota et al. [155] showed that MAP<br />

structures are internalized by dendritic cells, which are antigen-presenting<br />

cells. Their results showed that MAP structures<br />

are processed in the same way as an intracellular<br />

pathogen (cross presentation), leading to a peptide presented<br />

by a major histocompatibility complex (MHC) class I.<br />

MAP structures have also been used to transport protein<br />

antigens against microorganisms such as Plasmodium falciparum<br />

[156], which causes malaria in humans. To this end,<br />

epitopes of the major surface protein in this parasite were<br />

conjugated to the dendrimer. The results of experiments in<br />

mice showed an immune response.<br />

Finally, another interesting application of dendrimers is<br />

in gene therapy. Amino-terminated PAMAM or PPI are the<br />

dendrimers most commonly used as non-viral transfection<br />

agents of DNA to the nucleus. Kukowska-Latallo [157] synthesized<br />

several types of PAMAM dendrimers that differed<br />

in their core <strong>and</strong> number of generations. These were studied<br />

to improve transport <strong>and</strong> DNA transfection in a variety of<br />

mammalian cell lines. Various reporter genes were used to<br />

test the efficiency of transfection. In general, the results<br />

showed a highly efficient transfection in several cell lines<br />

with minimal cytotoxicity in eukaryotic cells.<br />

The capacity of dendrimers to transfect cells depends on<br />

their size, shape <strong>and</strong> number of primary amine groups on the<br />

polymer surface. Transfection is higher in dendrimers with<br />

an excess of primary amines on their surface than with DNA<br />

phosphate groups, thus giving the complex an overall positive<br />

charge <strong>and</strong> thereby support adherence of the complex to<br />

the negatively charged cell surface. Therefore, aminoterminated<br />

dendrimers carry DNA to the membrane <strong>and</strong> contribute<br />

to the transfection process by inducing the disruption<br />

of the membrane because, in theory, naked DNA, which is<br />

negatively charged, is repelled by the cell surface, which is<br />

also negatively charged.<br />

Navarro et al. [158] have studied PAMAM dendrimers as<br />

carriers for cancer gene therapy. PAMAM is used to protect<br />

DNA from degradation by nucleases <strong>and</strong> to deliver genetic<br />

material to tumor cells.<br />

Superfect [159] is a commercial transfection reagent consisting<br />

of activated dendrimer molecules that can carry larger<br />

amounts of genetic material than modified viruses for gene<br />

therapy. The structure of the dendrimer is formed by a central<br />

core <strong>and</strong> charged amino groups on the surface. These<br />

features allow the interaction of these charged groups with<br />

the negatively charged phosphate groups of nucleic acids <strong>and</strong><br />

enhance their adherence to the cell surface. Thus the dendrimers<br />

are taken up into the cell by non-specific endocytosis.<br />

Therefore, this new technology for gene transfer offers<br />

significant advantages, such as high gene transfer efficiencies<br />

<strong>and</strong> minimal toxicity, <strong>and</strong> it can be used with a broad<br />

range of cell types.<br />

3. CONCLUSION<br />

In the treatment of a pathophysiological process, it is<br />

desirable that the drug reaches its site of action at a particular<br />

concentration. This therapeutic dose range must be constant<br />

over a sufficiently long period of time to alter the process.<br />

The action of the drugs is limited by their degradation,<br />

interaction with other cells, <strong>and</strong> inability to penetrate tissues<br />

due to their chemical nature. Therefore, administration of<br />

relatively high doses is required to obtain the desired concentration<br />

at the site of action, thus leading to undesirable toxicological<br />

<strong>and</strong> immunological processes.<br />

For these reasons, research efforts are devoted to designing<br />

new formulations, being of particular interest polymeric


<strong>Polymers</strong> <strong>and</strong> <strong>Drug</strong> <strong>Delivery</strong> <strong>Systems</strong> Current <strong>Drug</strong> <strong>Delivery</strong>, 2012, Vol. 9, No. 4 25<br />

systems of drug carriers, to achieve a higher desired pharmacological<br />

response. These polymeric systems are suitable for<br />

site-specific <strong>and</strong> time-controlled delivery of drugs.<br />

The treatment of most diseases requires drug distribution<br />

at a specific site. This can be achieved by functionalizing the<br />

surface of polymeric particles with receptor recognition elements<br />

, such as antibodies <strong>and</strong> carbohydrates, or by linking<br />

the drug to the polymer through a specific linker that is degraded<br />

under certain conditions, such as in an acidic environment<br />

or by a specific enzyme. Other controlled-distribution<br />

systems use polymers, whose properties, such as swelling,<br />

change in response to an external stimulus such as pH.<br />

Time-controlled drug delivery can be achieved depending<br />

on the features of the polymer. Modulation of the chemical<br />

architecture of the polymer can regulate the release of the<br />

therapeutic agent <strong>and</strong> thus the profile of the final formulation,<br />

thus obtaining the following: an extended release formulation<br />

(swelling of polymeric matrix or degradation of<br />

polymeric capsules); a sustained-release formulation, in<br />

which the active substance is released with a kinetic profile<br />

of order 0 (osmotic pumps); a pulsatile-release formulation,<br />

in which drugs are released only when required by the body;<br />

<strong>and</strong> a delayed-release formulation, in which the active ingredient<br />

is released at a time other than at the time of administration<br />

(polymers sensitive to stimuli).<br />

From a polymer chemistry perspective, it is important to<br />

appreciate that the mechanisms of controlled-release require<br />

polymers with a variety of physico-chemical properties. Several<br />

types of polymers have been tested as potential drug<br />

delivery systems, including nano- <strong>and</strong> micro-particles, dendrimers,<br />

nano- <strong>and</strong> micro-spheres, capsosomes <strong>and</strong> micelles.<br />

In these systems, drugs can be encapsulated or conjugated<br />

into polymer matrices.<br />

Nano- <strong>and</strong> micro-particles are formulations in which the<br />

drug can be conjugated, trapped, or embedded within the<br />

polymer. Hydrogels such as PHEMA, <strong>and</strong> copolymers of<br />

PVA or PEG with acrylamides, are the most widely used.<br />

The three-dimensional structure of these hydrogels allows a<br />

controlled drug release.<br />

Nano- <strong>and</strong> micro-capsules are comprised of an outer shell<br />

that encapsulates the active agent. Usually, this kind of<br />

pharmaceutical formulation is used to transport drugs that<br />

are rapidly degraded in body fluids. Thus the drug is encapsulated<br />

inside the particle <strong>and</strong> it is released by the degradation<br />

of the polymer coating. Several polymers have been<br />

exploited for this kind of formulation, such as polyesters,<br />

polyanhydrides, polyamides, poly(orthoesters) <strong>and</strong> polysaccharides.<br />

Another formulation is capsosomes. These polymeric<br />

structures have a biodegradable outer layer <strong>and</strong> contain<br />

liposomes, which compartmentalize the interior of the capsule<br />

polymer. These systems show interesting features such<br />

as outer membrane permeability, which allows small molecules<br />

to cross the membrane by passive diffusion. Moroever,<br />

the division of the capsule into compartments by liposomes<br />

allows them to hold the same or distinct molecules.<br />

Micelles are composed of individual linear polymers that<br />

contain a hydrophobic <strong>and</strong> a hydrophilic segment, which<br />

have the capacity to form micellar structures in aqueous me-<br />

dia. The hydrophobic core of polymeric micelles facilitates<br />

the incorporation of hydrophobic drugs either through covalent<br />

or non-covalent interactions.<br />

Finally, there is another promising formulation for drug<br />

delivery, dendrimers. The chemical structure of the dendrimer<br />

consists of a core or central structure that contains<br />

several branches, which in turn branch off, leading to a<br />

three-dimensional globular structure with concentric layers.<br />

The dendrimeric structure allows the encapsulation of drugs<br />

with hydrophobic properties in its interior cavities while<br />

functional groups found on their surface can attach hydrophilic<br />

drugs by covalent bonds, electrostatic interactions <strong>and</strong><br />

hydrogen bonds.<br />

These polymeric systems have been used for various<br />

purposes including treatments for antineoplastic activity,<br />

bacterial infection <strong>and</strong> inflammatory process as well as for<br />

vaccines.<br />

Polymeric carriers of antineoplastic drugs can passively<br />

accumulate in cancerous tissues because of differences in the<br />

biochemical <strong>and</strong> physiological features of healthy <strong>and</strong> malignant<br />

tissues (EPR effect) while they can actively accumulate<br />

in the same tissues because they have been conjugated to<br />

targeting moieties. Furthermore, antineoplastic drugs are<br />

highly toxic <strong>and</strong> these delivery systems can prevent their<br />

degradation <strong>and</strong> interaction with other tissues.<br />

These formulations are also used against bacterial infections.<br />

Most intracellular infections are difficult to eradicate<br />

because bacteria inside phagosomes are protected from antibiotics.<br />

Thus, treatments for intracellular bacterial infection<br />

are hindered because most antibiotics have poor intracellular<br />

diffusion <strong>and</strong> some have a reduced activity in the acidic conditions<br />

of lysosomes. Therefore, the treatment of the diseases<br />

caused by these bacteria call for drug carriers such as dendrimers,<br />

which are endocytosed by macrophages, thus allowing<br />

drug release within them.<br />

Finally, another relevant application of drug delivery<br />

systems is vaccines. Vaccines are classified into two groups,<br />

namely proteins <strong>and</strong> nucleic acids. In both cases they require<br />

polymeric carriers because they are susceptible to degradation<br />

by peptidases or nucleases. The DNA vaccine is a newly<br />

developed system. The DNA in this vaccine system encodes<br />

a protein antigen of interest that induces activation of the<br />

immune system. This DNA can be encapsulated into polymeric<br />

carriers, thereby protecting it from degradation. It is then<br />

released into the phagosome, thus allowing it to reach the<br />

cell nucleus <strong>and</strong> express the foreign protein.<br />

In summary, the chemical nature of the many types of<br />

polymeric drug carriers available facilitates the distribution<br />

<strong>and</strong> interaction of drugs with their target tissue. These carriers<br />

also protect their drug cargo from degradation <strong>and</strong> prevent<br />

their side effects.<br />

CONFLICT OF INTEREST<br />

None declared.<br />

ACKNOWLEDGEMENT<br />

None declared.


26 Current <strong>Drug</strong> <strong>Delivery</strong>, 2012, Vol. 9, No. 4 Vilar et al.<br />

REFERENCES<br />

[1a] Seymour, L.; Miyamoto, Y.; Maeda, H.; Brerton, M.; Strohalm, J.;<br />

Ulbrich, K.; Duncan, R. Eur. J. Cancer., 1995, 31A, 766-770. (b)<br />

Seymour, L.W.; Duncan, R.; Strohalm, J.; Kopecek, J. J. Biomed.<br />

Mater. Res. 1987, 21, 1341-1358. (c) Gombotz, W.R.; Pettit, D.K.<br />

Bioconjug. Chem. 1995, 6, 332-351.<br />

[2a] Langer, R. Nature 1998, 392-395. (b) Brouwers, J. R. B. Pharm.<br />

World Sci. 1996, 18, 153-162.<br />

[3a] Wagenaar, B. W.; Müller, B. W. Biomaterials 1994, 15, 49-54. (b)<br />

Conforti, A.; Bertani, S.; Lussignoli, S.; Grigolini, L.; Terzi, M.;<br />

Lora, S.; Caliceti, P.; Marsilio, F.; Veronese, F. M. J. Pharm.<br />

Pharmacol. 1996, 48, 468-473. (c) Kalala, W.; Kinget, R.; Van den<br />

Mooter, G.; Samyn, C. Int. J. Pharm. 1996, 139, 187-195.<br />

[4] Schierholz, J. M.; Rump, A.; Pulverer, G. <strong>Drug</strong> Res. 1997, 47, 70-74.<br />

[5] Walter, K.A.; Tamargo, R.; Olivi, A.; Burger, P. C.; Brem, H.<br />

Neurosurgery. 1995, 37, 1129-1145.<br />

[6] Katayama, N.; Tanaka, R.; Ohno, Y.; Ueda, C.; Houjou, T.; Takada,<br />

K. Int. J. Pharm. 1995, 115, 87-93.<br />

[7] Maniar, M.; Domb, A.; Haffer, A.; Shah, J. J. Control. Release<br />

1994, 30, 233-239.<br />

[8] McGee, J.P.; Davis, S. S.; O’Hagan, D. T. J. Control. Release<br />

1994, 31, 55-60.<br />

[9] Wang, B., Siahaan, T. Soltero, R.A. In: <strong>Drug</strong> <strong>Delivery</strong>: principles<br />

<strong>and</strong> applications, Ed. Wiley, Inc., 2005.<br />

[10a] Sampath Kumar, K. P.; Bhowmik, D.; Chiranjib; Ch<strong>and</strong>ira, M.;<br />

Tripathi, K. K. J. Chem. Pharm. Res., 2010, 2, 349-360. b) Sáez,<br />

V.; Hernáez, E.; Angulo, L. S. Revista Iberoamericana de<br />

Polímeros, 2002, 3. c) Cho, X.; Wang, X.; Nie, S.; Chen, Z. G.;<br />

Shin, D. M. Clin. Cancer Res., 2008, 14, 1310-1316.<br />

[11] Michael E. A. In: Translation of Pharmaceutics: The Science of<br />

Dosage Form Design, 2e., 2002.<br />

[12a] http://pharmacycode.com/fda/Plendil_Er (Accessed November<br />

2011) (b) http://www.racgp.org.au/cmi/gwckapan.pdf (Accessed<br />

November 2011)<br />

[13] http://packageinserts.bms.com/pi/pi_sinemet_cr.pdf (Accessed<br />

November 2011)<br />

[14] San Juan, A.; Bala, M.; Hlawaty, H.; Portes, P.; Vranckx, R.;<br />

Feldman, L. J.; Letourneur, D. Biomacromolecules, 2009, 10,<br />

3074-80.<br />

[15] http://www.cordislabeling.com/pdf/5462872_4.pdf (Accesed December<br />

2011) http://www.cypherstent.com/Pages/index.aspx (Accesed<br />

December 2011)<br />

[16] http://www.ptca.org/articles/taxus_profileframe.html (Accesed<br />

December 2011)<br />

[17] Verma, R.K.; Mishra, B.; Garg, S. <strong>Drug</strong> Dev. Ind. Pharm. 2000,<br />

26, 695-708.<br />

[18] M. Anschütz, M.; Wonnemann, M.; Schug, B.; Toal, C.; Donath,<br />

F.; Pontius, A.; Pauli, K.; Brendel, E.; Blume, H. International J.<br />

Clin. Pharmacol. Thera., 2010, 48, 158-170.<br />

[19] Ravikumar Reddy J.; Veera, M.; Mohamed T. S.; Madhu C. S. J.<br />

Pharm. Sci. & Res. 2009, 1, 109-115.<br />

[20] Kost, J.; Horbett, T.A.; Ratner, B.D.; Singh, M. J. Biomed. Mater.<br />

Res. 1985, 19, 1117-1133.<br />

[21] Sadaphal, K.P.; Thakare, V.M.; G<strong>and</strong>hi, B.R.; Tekade, B.W. International<br />

Journal of <strong>Drug</strong> <strong>Delivery</strong>, 2011, 3, 348-356.<br />

[22a] Rihova, B.; Kopecek, J. J. Controlled Release, 1985, 2, 289-310. (b)<br />

Seymour, L. W., Flanagan, P. A.; Al-Shamkhani, A.; Subr. V.; Ulbrich,<br />

K.; Cassidy, J.; Duncan, R. Sel. Cancer Ther., 1991, 7, 59-73.<br />

[23] Wedge, S.R.; Duncan, R.; Kopeckova, P. Br. J. Cancer, 1991, 63,<br />

546-549<br />

[24] Clegg, J. A.; Hudecz, F.; Mezo, G.; Pimm, M. V.; Szerkerke, M.;<br />

Baldwin, R. W. Bioconjug. Chem., 1990, 1, 425-430.<br />

[25a] Allen, T.M. Nat. Rev. Cancer, 2002, 2, 750-763. (b) Brumlik, M.J.;<br />

Daniel, B.J.; Waehler, R.; Curiel, D.T.; Giles, F.J.; Curiel, T.J. Expert<br />

Opin. <strong>Drug</strong> Deliv., 2008, 5, 87-103. (c) Miller, K.; Erez, R.;<br />

Segal, E.; Shabat, D.; Satchi-Fainaro, R. Angew. Chem. Int. Ed.,<br />

2009, 48, 2949-2954.<br />

[26a] Qiu, Y.; Park, K. Adv. <strong>Drug</strong> Deliv. Rev., 2001, 53, 321-339<br />

(b)Gupta, P.; Vermani, K.; Garg, S. <strong>Drug</strong> Discov. Today, 2002, 7,<br />

569-579.<br />

[27] Brocchini, S.; Duncan, R. In: Polymer-drug conjugates: drug release<br />

from pendent linkers. In: Encyclopedia of controlled release.<br />

Mathiowitz E (Ed.). Wiley, NY, USA, 1999, 786-816.<br />

[28] Nori, A.; Kopecek, J. Adv. <strong>Drug</strong> Deliv. Rev. 2005, 57, 609-636.<br />

[29] Duncan, R. Nat. Rev. <strong>Drug</strong> Discov., 2003, 2, 347-360.<br />

[30] Ferruti, P.; Marchisio, M.A.; Duncan, R. Macromol. Rapid Comm.,<br />

2002, 23, 332-355.<br />

[31] Ulbrich, K.; Subr, V. Adv. <strong>Drug</strong> Deliv. Rev., 2004, 23, 1023-1050.<br />

[32] Duncan, R. In: Targeting <strong>and</strong> intracellular delivery of drugs. In:<br />

Encyclopedia of Molecular Cell Biology <strong>and</strong> Molecular Medicine,<br />

Meyers, R.A., Ed., Weinheim, Germany.Wiley-VCH Verlag,<br />

GmbH & Co. KgaA., 2005, 163-204.<br />

[33] Uhrich, K. E.; Cannizzarro, S.M.; Langer, R.S.; Shakesheff, K. M.<br />

Chem. Rev. 1999, 99, 3181-3198.<br />

[34] Panyam, J., Sahooo, S. K., Prabha, S., Bargar, T., Labhasetwar, V.<br />

Int. J. Pharm., 2003, 262, 1-11.<br />

[35] Tong, R.; Cheng, J. Polymer reviews, 2007, 47, 345-381.<br />

[36] Matsumura, Y.; Maeda, H. Cancer Res. 1986, 46, 6387-6392.<br />

[37a] Vicent, M. J.; Duncan, R. Trends Biotechnol. 2006, 24, 39-47.<br />

(b)Haag, R.; Kratz, F. Angew. Chem. Int. Ed. Engl., 2006, 45,<br />

1198-1215. (c) Duncan, R.; Vicent, M. J. Adv. <strong>Drug</strong> Deliv. Rev.,<br />

2010, 62, 272-282.<br />

[38] Lammers, T. Adv. <strong>Drug</strong> Deliv. Rev. 2010, 62, 203-230.<br />

[ 39] Thomson, A. H.; Vasey, P. A.; Murray, L. S.; Cassidy, J., Fraier,<br />

D., Frigerio, E., Twelves, C. J. Br. J. Cancer. 1999, 81, 99-107.<br />

[40a] Seymour, L. W.; Ferry, D. R.; Kerr, D. J.; Rea, D.; Whitlock, M.;<br />

Poyner, R.; Boivin, C.; Hesslewood, S.; Twelves, C.; Blackie, R.;<br />

Schatzlein, A.; Jodrell, D.; Bisset, D.; Calvert, H.; Lind, M.; Robbins,<br />

A.; Burtles, S.; Duncan, R.; Cassidy, J. Int. J. Oncol., 2009,<br />

34, 1629-1636. (b) Vasey, P.A.; Kaye, S.B.; Morrison, R. Clin.<br />

Cancer Res. 1999, 5, 83-94.<br />

[41a] Hopewel, J. W.; Duncan, R.; Wilding, D.; Chakrabarti, K. Hum.<br />

Exp. Toxicol., 2001, 20, 461-470. (b) Julyan, P. J.; Seymour, L. W.;<br />

Ferry, D. R.; Daryani, S.; Boivin, C. M.; Doran, J.; David, M.; Anderson,<br />

D.; Christodoulou, C.; Young, A. M.; Hesslewood, S.;<br />

Kerr, D. J. J. Control. Release, 1999, 57, 281-290.<br />

[42a] Campone, M.; Rademaker-Lakhai J. M.; Bennouna, J.; Howell, S.<br />

B.; Nowotnik, D. P.; Beijnen, J. H.; Schellens, J, H. Cancer Chemother.<br />

Pharmacol., 2007, 60, 523-33. (b) Rice, J. R.; Howell, S.<br />

B. <strong>Drug</strong>s Fut, 2004, 29, 561-566. (c) Kell<strong>and</strong>, L. Expert Opin. Investig.<br />

<strong>Drug</strong>s, 2007, 16, 1009-1021.<br />

[43] Nowotnik, D. P.; Cvitkovic, E. Adv. <strong>Drug</strong> Deliv. Rev. 2009, 61,<br />

1214-1219.<br />

[44] Shiose, Y.; Ochi, Y.; Kuga, H.; Yamashita, F.; Hashida, M. Biol.<br />

Pharm. Bull. 2007, 30, 2365-2370.<br />

[45] Soepenberg, O.; de Jone, M. J.; Sparreboom, A.; de Bruin, P.;<br />

Eskens, F. A.; de Heus, G.; W<strong>and</strong>ers, J.; Cheverton, P.; Ducharme,<br />

M. P.; Verweij, J. Clin. Cancer Res. 2005, 11, 703-711.<br />

[46a] Chipman, S.D.; Oldham, F.B.; Pezzoni, G.;, Singer, J.W. Int. J.<br />

Nanomed., 2006, 1, 375-383. (b) Shaffer, S.A.; Baker Lee, C.;<br />

Kumar, A.; Singer, J.W. Eur. J. Cancer, 2002, 38, 428.<br />

[47] Singer; J. W.; Shaffer, S.; Baker, B.; Bernareggi, A.; Stromatt, S.;<br />

Nienstedt, D.; Besman, M. Anticancer <strong>Drug</strong>s 2005, 16, 243-254.<br />

[48] Mita, M.; Mita, A.; Sarantopoulos, J.; Takimoto, C. H.; Rowinsky, E.<br />

K.: Romero, O.; Angiuli, P.; Allievi, C.; Eisenfeld, A.; Verschraegen,<br />

C. F. Cancer Chemother. Pharmacol., 2009, 64, 287-295.<br />

[49a] Singer, J. W.; Baker, B.; De Vries, P.; Kumar, A.; Shaffer, S.;<br />

Vawter, E.; Bolton, M.; Garzone, P. Adv.Exp. Med. Biol. 2003,<br />

519, 81-99. (b) Sabattini, P.; Brown, J.; Aghajanian, C.; Hensley,<br />

M. L.; Pezulli, S.; O’Flaherty, C.; Lovegren, M.; Funt, S.; Warner,<br />

M.; Mitchell, P.; Botton, M. G.; Spriggs, D.; Duggan, B. Proc. Am.<br />

Soc. Clin. Oncol., 2002, 21, (abst. 871). (c) Li, C.; Price, J. E.; Milas,<br />

L.; Hunter, N. R.; Ke, S.; Yu, D. F.; Charnsangavej, C.;<br />

Wallace, S. Clin. Cancer Res. 1999, 5, 891-897.<br />

[50a] Enzon Pharmaceuticals, Inc. www.enzon.com. (Accessed May<br />

2011) (b) Pastorino, F.; Loi, M.; Sapra, P.; Becherini, P.;Cilli, M.;<br />

Emionite, L.; Ribatti, D.; Greenberger, L. M.; Horak, I. D.; Ponzoni,<br />

M. Clin. Cancer Res. 2010, 16, 4809-4821.<br />

[51a] Davis, M. E. Adv. <strong>Drug</strong> Deliv. Rev. 2009, 61, 1189-92. (b) Schluep,<br />

T.; Hwang, J.; Cheng, J.J. Clin. Cancer Res. 2006, 12, 1606-1614<br />

[52] Homsi, J.; Simon, G. R.; Garrett, C. R.; Springett, G.; De Conti, R.;<br />

Chiappori, A. A.; Munster, P. N.; Burton, M. K.; Stromatt, S.; Allievi,<br />

C.; Angiuli, P.; Eisenfeld, A.; Sullivan, D. M.; Daud, A.I.<br />

Clin. Cancer Res., 2007, 13, 5855-5861. (b) Springett, G.M.;<br />

Takimoto, C.; McNamara, M. J. Clin. Oncol. 2004, 22, 3127.<br />

[53] Meerum Terwogt, J. M.; ten Bokkel Huinink, W. W.; Schellens, J.<br />

H.; Schot, M.; M<strong>and</strong>jes, I. A.; Zurlo, M. G.; Rochetti, M.; Rosing,<br />

H.; Koopman, F. J.; Beijnen, J. H. Anticancer <strong>Drug</strong>s, 2001, 12,<br />

315-323.<br />

[54a] Bisset, D.; Cassidy, J.; de Bono, J. S.; Muirhead, F.; Main, M.;<br />

Robson, L.; Fraier, D.; Magnè, M. L.; Pellizzoni, C.; Porro, M. G.;


<strong>Polymers</strong> <strong>and</strong> <strong>Drug</strong> <strong>Delivery</strong> <strong>Systems</strong> Current <strong>Drug</strong> <strong>Delivery</strong>, 2012, Vol. 9, No. 4 27<br />

Spinelli, R.; Speed, W.; Twelves, C. Br. J. Cancer, 2004, 91,50-55.<br />

(b) Schoemaker, N. E.; Van Kesteren, C.; Rosing, H.; Jansen, S.;<br />

Swart, M.; Lieverst, J.; Fraier, D.; Breda, M.; Pellizzoni, C.;<br />

Spinelli, R.; Grazia Porro, M.; Beijnen, J. H.; Schellens, J. H.; ten<br />

Bokkel Huinink, W. W. Br. J. Cancer, 2002, 87, 608-14. (c)<br />

Duncan, R. Adv. <strong>Drug</strong> Deliv. Rev. 2009, 61, 1131-1148.<br />

[55] Uhrich, K. E.; Cannizzaro, S. M.; Langer, R. S.; Shakesheff, K. M.<br />

Chem. Rev., 1999, 99, 3181-3198.<br />

[56] Gombotz, W. R.; Pettit, D. K. Bioconjug. Chem., 1995, 6, 332-351.<br />

[57] Russell, J. R.; Pishko, M. V.; Simonian, A. L.; Wild, J. R. Anal.<br />

Chem. 1999, 71, 4909-4912.<br />

[58] Syu, M.J,; Chang, Y.S. Biosens Bioelectron., 2009, 24, 2671-2677<br />

[59] Peppas, N. A.; Keys, K. B.; Torres-Lugo, M.; Lowman, A. M. J.<br />

Control. Release., 1999, 62, 81-87.<br />

[60] Varshosaz, J.; Koopaie, N. Iranian Polymer Journal, 2002, 11,<br />

123-131.<br />

[61] Nakamura, K.; Maitani, Y.; Lowman, A. M.; Takayama, K.; Peppas,<br />

N. A.; Nagai, T. J. Control. Release 1999, 61, 329-335.<br />

[62a] Govender, T.; Riley, T.; Ehtezazi, T.; Garnett, M. C.; Stolnik, S.;<br />

Illum, L.; Davis, S. S. Int. J. Pharm. 2000, 199, 95-110. (b)<br />

Panyam, J.; Williams, D.; Dash, A.; Leslie-Pelecky, D.; Labthasetwar,<br />

V., J. Pharm. Sci. 2004, 93, 1804-1814.<br />

[63] Liu, Z.; Cheung, R.; Wu, X. Y.; Ballinger, J. R.; Bendayan, R.;<br />

Rauth, A. M. J. Control. Release, 2001, 77, 213-224.<br />

[64a] Sáez, V.; Hernáez, E.; Angulo, L. S.; Katime, I. Revista<br />

Iberoamericana de Polímeros, 2004, 5, 87-101. (b) Yeo, Y,; Baek,;<br />

Park, K. Biotechnol. Bioprocess Eng., 2001, 6, 213-230.<br />

[65] de Jalón, E. G.; Blanco-Príeto, M. J.; Ygartua, P.; Santoyo, S. Int.<br />

J. Pharm. , 2001, 226, 181-184.<br />

[66] Wada, R.; Hyron S.H.; Ikada, Y. J. Pharm. Sci. 1990, 79, 919-924.<br />

[67] Iwata, M.; McGinity, J. W. Pharm. Res., 1993, 10, 1219-1227.<br />

[68] Nihant, N.; Stassen, S.; Gr<strong>and</strong>fils, C.; Jérome, R.; Teyssié, P.;<br />

Goffinet, G. Polymer International, 1994, 34, 289-299.<br />

[69] Naikwade, S.; Bajaj, A. Sci. Pharm., 2009, 77, 419-441.<br />

[70] Vaida, C.; Mela, P.; Kunna, K.; Sternberg, K.; Keul, H.; Möller, M.<br />

Macromol. Biosci., 2010, 10, 925-933.<br />

[71a] Bittner, B.; Ronneberger, B.; Zange, R.; Voll<strong>and</strong>, C.; Anderson,<br />

J.M.; Kissel, T. J. Microencapsul., 1998, 15, 495-514. (b) Uhrich,<br />

K.E.; Cannizzaro, S.M.; Langer, R.S.; Shakesheff, K.M. Chem.<br />

Rev., 1999, 99, 3181-3198.<br />

[72a] Xiao, R. Z. ; Zeng, Z. W.; Zhou, G. L. ; Wang, J. J. ; Li, F. Z. ;<br />

Wang, A. M. Int. J. Nanomedicine, 2010, 5, 1057-1065. (b) Parajó,<br />

Y.; d’Angelo, I,; Horváth, A.; Vantus, T.; György, K.; Welle, A.;<br />

Garcia-Fuentes, M.; Alonso, M. J. Eur. J. Pharm. Sci., 2010, 41,<br />

644-649.<br />

[73] Huo, D.; Deng, S.; Li, L.; Ji, J. Int. J. Pharm., 2005, 289,63-67.<br />

[74] Li, Y.; Lim, S.; Ooi, C. P. Pharm. Res. 2011 (DOI 10.1007/s11095-<br />

011-0600-9).<br />

[75] Ike, O.; Shimizu, Y.; Ikada, Y.; Watanabe, S.; Natsume, T.; Wada,<br />

R.; Hyon, S-H.; Hitomi, S. Biomaterials 1991, 12, 757-762.<br />

[76a] Ike, O.; Shimizu, Y.; Hitomi, S.; Wada, R.; Ikada, Y. Chest, 1991,<br />

99, 911-915. (b) Ike, O.; Hitomi, S.; Wada, R.; Watanabe, S. <strong>Drug</strong><br />

<strong>Delivery</strong> <strong>Systems</strong>, 1990, 5, 23-27.<br />

[77] Nagarwal, R. C.; Singh, P. N.; Kant, S.; Maiti, P.; P<strong>and</strong>it, J. K. J.<br />

Biomed. Nanotechnol., 2010, 6, 648-657.<br />

[78] Sharma, G.; Valenta, D. T.; Altman, Y.; Harvey, S.; Xie, H.; Mitragotri,<br />

S.; Smith, J. W. J. Control Release, 2010, 147, 408-412.<br />

[79] Italia, J. L.; Yahya, M. M.; Singh, D.; Ravi Kumar, M. N . Pharm.<br />

Res., 2009, 26, 1324-1331.<br />

[80] Espuelas, M.S.; Legr<strong>and</strong>, P.; Loiseau, P.M.; Bories, C.; Barratt, G.;<br />

Irache, J.M. J. <strong>Drug</strong> Target., 2002, 10, 593-599.<br />

[81] Nahar, M.; Jain, N. K. Pharm. Res., 2009, 26, 2588-2598.<br />

[82] P<strong>and</strong>ey, R.; Khuller, G.K. Chemotherapy., 2007, 53, 437-441.<br />

[83] Suarez, S.; O’Hara, P.; Kazantseva, C.; Newcomer, E.; Hopfer, R.;<br />

Mc Murray, D. N.; Hickey, A. J. J. Antimicrob. Chemother., 2001,<br />

48, 431-434.<br />

[84] Chen, G.; He, C. X.; Xu, D. H.; Liu, L.; Zhang, F. J.; Xu, J. Y.;<br />

Zheng, C. H.; Gao, J. Q.; Yan, M. Pharmazie, 2009, 64, 284-286.<br />

[85] Pillai, R.R.; Somayaji, S.N.; Rabinovich, M.; Hudson, M.C.; Gonsalves,<br />

K.E. Biomed. Mater., 2008, 3, 034114.<br />

[86] Jeong, Y.I.; Na, H.S.; Seo, D.H.; Kim, D.G.; Lee, H.C.; Jang,<br />

M.K.; Na, S.K.; Roh, S.H.; Kim, S.I.; Nah, J.W. Int. J. Pharm.,<br />

2008, 352, 317-323.<br />

[87] Esmaeili, F.; Hosseini-Nasr, M.; Rad-Malekshahi, M.; Samadi, N.;<br />

Atyabi, F.; Dinarv<strong>and</strong>, R. Nanomedicine., 2007, 3, 161-167.<br />

[88] <strong>Drug</strong>s information on line, <strong>Drug</strong>s.com. Web page:<br />

http://www.drugs.com/pro/nutropin-depot.html . (Accessed May<br />

2011)<br />

[89] Rafi, M.; Singh, S. M.; Kanchan, V.; Anish, C. K.; P<strong>and</strong>a, A. K. J.<br />

Microencapsul., 2010, 27, 552-560.<br />

[90] S<strong>and</strong>ers, L. M.; Mc Rae, G. I.; Vitale, K. M.; Kell, B. A. J. Control.<br />

Release, 1985, 2, 187-195.<br />

[91] Cui, F.D.; Tao, A. J.; Cun, D. M.; Zhang, L. Q.; Shi, K. J. Pharm.<br />

Sci., 2007, 96, 421-427.<br />

[92] Katti, D. S.; Lakshmi, S.; Langer, R.; Laurencin, C. T. Adv. <strong>Drug</strong><br />

Deliv. Rev. 2002, 54, 933-961.<br />

[93] Laurencin, C.; Domb, A.; Morris, C.; Brown, V.; Chasin, M.;<br />

McConnell, R.; Lange N.; Langer, R. J. Biomed. Mater. Res., 1990,<br />

24, 1463-1481.<br />

[94] Kumar N.; Langer, S. R.; Domb, A. J. Adv. <strong>Drug</strong> Deliv. Rev., 2002,<br />

54, 889-910.<br />

[95] Domb, A. J.; Rock, M.; Schwartz, J.; Perkin, C.; Yipchuck, G.;<br />

Broxup, B.; Villemure, J. G. Biomaterials 1994, 15, 681-688.<br />

[96] Kipper, M. J.; Shen, E.; Determan, A.; Narasimhan, B. Biomaterials<br />

2002, 23, 4405-4412.<br />

[97] Berckl<strong>and</strong>, C.; Kipper, M. J.; Narasimhan, B.; Kim, K. K.; Pack, D.<br />

W. J. Control. Release. 2004, 94, 129-41.<br />

[98] Petersen, L. K.; Sackett, C.K.;Narasimhan, B. J. Comb. Chem.,<br />

2010, 12, 51-56.<br />

[99] Manoharan, C.; Singh, J. J. Pharm. Sci., 2009, 98(11), 4237-4250.<br />

[100] Sun, L.; Zhou, S.; Wang, W.; Su, Q.; Li, X.; Weng, J. J. Mater Sci:<br />

Mater Med. 2009, 10, 2035-2042.<br />

[101] Determan, A. S.; Trewyn, B. G.; Lin, V. S.; Nilsen-Hamilton, M.<br />

N.; Narasimhan, B. J. Control. Release, 2004, 100, 97-109.<br />

[102] Carrillo-Conde, B.; Schiltz, E.; Yu, J.; Chris Minion, F.; Phillips,<br />

G. J.; Wannemuehler, M. J.; Narasimhan, B. Acta Biomater., 2010,<br />

6, 3110-3119.<br />

[103] Kipper, M. J.; Wilson, J. H.; Wannemuehler, M. J.; Narasimhan, B.<br />

J. Biomed. Mater. Res. A. 2006, 76, 798-810<br />

[104a] Marck, K. W.; Wildevuur, C. H.; Sederel, W. L.; Bantjes, A.; Feijen,<br />

J. J. Biomed. Mater. Res. 1997, 11, 405-22. (b) Martin E. C.; May, P.<br />

D.; McMahon, W. A. J. Biomed. Mater. Res. 1971, 5, 53-62.<br />

[105] Caponetti, G.; Hrkach, J. S.; Kriwet, B.; Poh, M.; Lotan, N.; Colombo,<br />

P.; Langer, R. J. Pharm. Sci.. 1999, 88, 136-141.<br />

[106] Hrkach, J. S.; Ou, J.; Lotan, N.; Langer, R. Hydrogels <strong>and</strong> Biodegradable<br />

<strong>Polymers</strong> for Bioapplications. 1996, 8, 93-102.<br />

[107] Vera, M.; Puiggalí, J.; Coudane, J. J. Microencapsul., 2006, 23,<br />

686-697.<br />

[108a] Heller, J.; Barr, J.; Ng, S. Y.; Abdellavoi, K. S.; Gurny, R. Adv.<br />

<strong>Drug</strong> Deliv. Rev., 2002, 54(7), 1015-1039. (b) Heller, J.; Barr, J.<br />

Biomacromolecules, 2004, 5, 1625-1632.<br />

[109] Heller, J. Annals of the New York Academy of Sciences. Macromolecules<br />

as <strong>Drug</strong> <strong>and</strong> as Carriers for Biologically Active Materials.<br />

1985, 446, 51-66.<br />

[110] Heller, J.; Barr, J. Polymeric <strong>Drug</strong> <strong>Delivery</strong> II, 2006, 3, 29-43.<br />

[111a] Bai, X. L.; Yang Y. Y.; Chung, T. S.; Ng, S.; Heller, J. J. Applied<br />

Polymer Science. 2001, 80, 1630-42. (b) Bai, X. L.; Yang Y. Y.;<br />

Chung, T. S.; Ng, S.; Heller, J. Processing <strong>and</strong> fabrication of advanced<br />

materials, VIII. 2001, 177-184.<br />

[112] Chia, H. H.; Yang, Y. Y.; Chung, T. S.; Ng, S.; Heller, J. J. Control.<br />

Release 2001, 75, 11-25.<br />

[113] Nguyen, D. N.; Raghavan, S. S.; Tashima, L. M.; Lin, E. C.; Fredette,<br />

S. J.; Langer, R. S.; Wang, C. Biomaterials 2008, 29, 2783-<br />

2793.<br />

[114] Wang, C.; Ge, Q.; Ting, D.; Nguyen, D.; Shen, H. R.; Chen, J.;<br />

Eisen, H. N.; Heller, J.; Langer, R.; Putman, D. Nat. Mater. 2004,<br />

3, 190-196.<br />

[115] Lin, Y. H.; Vasavada, R. C. J. Microencapsul. 2000, 17, 1-11.<br />

[116] Bouchemal, K.; Briançon, S.; Chaumonz, B.; Fessi, H.; Zydowicz,<br />

N. J. Microencapsulation 2003, 20, 637-651.<br />

[117a] Masotti, A.; Bordi, F.; Ortaggi, G.;Marino, F.; Palocci, C. In: Modern<br />

Research <strong>and</strong> Educational Topics in Microscopy; Mendez-Vilas A.,<br />

Díaz J., Eds.; 2007. (b) Masotti, A.; Bordi, F.; Ortaggi, G.; Marino,<br />

F.; Palocci, C. Nanotechnology 2008, 19, 055302/1-055302/6.<br />

[118] Özbas-Turan, S.; Aral, C.; Kabasakal, L.; Keyer-Uysal, M. Akbuga,<br />

J. J. Pharm. Pharm. Sci. 2003, 6, 27-32.<br />

[119] Grenha, A.; Al-Quadi, S.; Seijo, B.; Remunan-Lopez, C. J. <strong>Drug</strong><br />

Deliv. Sci. Tech., 2010, 20, 33-43.<br />

[120] Lee, E.; Kim, K.; Choi, M.; Lee, Y.; Park, J. W.; Kim, B. <strong>Drug</strong><br />

Deliv., 2010, 17, 573-580.


28 Current <strong>Drug</strong> <strong>Delivery</strong>, 2012, Vol. 9, No. 4 Vilar et al.<br />

[121] Lowman, A. M.; Morishita, M.; Kajita, M.; Nagai, T.; Peppas, N.<br />

A. J. Pharm. Sci., 1999, 88, 933-937.<br />

[122a] Ito, Y.; Casolaro, M.; Kono, K.; Yukio, I. J. Control. Release,<br />

1989, 10, 195-203. (b) Ishihara, K.; Kobayashi, M.; Shinohara, I.<br />

Polymer J., 1984, 16, 625-631.<br />

[123] Na, K.; Jung, J.; Lee, J.; Hyun, J. Langmuir 2010, 26, 11165-<br />

11169.<br />

[124] Curcio, M.; Puoci, F.; Spizzirri, U. G.; Iemma, F.; Cirillo, G.; Parisi,<br />

O. I.; Picci, N. AAPS. PharmSciTech., 2010, 11, 652-662.<br />

[125] Joung, Y. K.; Choi, J. H.; Park, K. M.; Park, K. D. Biomed. Mater.,<br />

2007, 2, 269-273 .<br />

[126a] Srivastava, R.; Brown, J. Q.; Zhu, H.; McShane, M. J. Macromol.<br />

Biosci., 2005, 5, 717-27.(b) Z.-l. Zhi, Z. I.; Haynie, D. T. Chem.<br />

Commun., 2006, 147-149.<br />

[127a] Petrov, A. I.; Volodkin, D. V.; Sukhorukov, G. B. Biotechnol.<br />

Prog. 2005, 21, 918-925. (b) Tiourina, O. P.; Antipov, A. A.; Sukhorukov,<br />

G. B.; Larionova, N. I.; Lvov, Y.; Mohwald, H. Macromol.<br />

Biosci., 2001, 1, 209-214.<br />

[128a] Caruso, F.; Trau, D.; Mohwald, H.; Renneberg, R. Langmuir, 2000,<br />

16, 1485-1488. (b)Wang, Y.; Caruso, F. Chem. Commun.(Camb),<br />

2004, 7, 1528-1529.<br />

[129a] Antipov, A.; Shchukin, D.; Fedutik, Y.; Zanaveskina, I.;<br />

Klechkovskaya, V.; Sukhoroukov, G.; Mohwald, H. Macromol.<br />

Rapid Commun., 2003, 24, 274-277. (b) Yu, A.; Gentle, I.; Lu, G.;<br />

Caruso, F. Chem. Commun. (Camb), 2006, 20, 2150-2152.<br />

[130a] Kreft, O.; Prevot, M.; Möhwald, H.; Sukhorukov, G. B. Angew.<br />

Chem. Int. Ed. Engl., 2007, 46, 5605-5608. (b) Mak, W. C.; Bai, J.;<br />

Chang, X. Y.; Trau, D. Langmuir, 2009, 25, 769-775.<br />

[131] Hosta, L.; Städler, B.; Yan, Y.; Nice, E. C.; Heath, J. K.; Albericio,<br />

F.; Caruso, F. Adv. Funct. Mater. 2010, 20, 59-66.<br />

[132] Kwon, G. S.; Naito, M.; Kataoka, K.; Yokoyama, M.; Sajurai, Y.;<br />

Okano, T. Colloids Surf. B., 1994, 2, 429-434.<br />

[133] Kwon, G.; Naito, M.; Yokohama, M.; Okano, T.; Sakurai, Y.;<br />

Kataoka, K. Pharm. Res., 1995, 12, 192-195.<br />

[134] Boddu, S. H. S.; Jwala, J.; Chowdhury, M. R.; Mitra, A. K. J. Ocul.<br />

Pharmacol. Ther., 2010, 26, 459-468.<br />

[135] Matsumura, Y.; Hamaguchi, T.; Ura, T.; Muro, K.; Yamada, Y.;<br />

Shimada, Y.; Shirao, K.; Okusaka, T.; Ueno, H.; Ikeda, M.; Watanabe,<br />

N. Br. J. Cancer 2004, 91, 1775-1781.<br />

[136] Danson, S.; Ferry,D.; Alakhov, V.; Margison, J.; Kerr, D.; Jowle,<br />

D.; Brampton, M.; Halbert, G.; Ranson, M. Br. J. Cancer 2004, 90,<br />

2085-2091.<br />

[137] Kim, T-Y.; Kim, D-W.; Chung, J-Y.; Shin, S. G.; Kim, S-C.; Heo,<br />

D. S.; Kim, N. K.; Bang, Y-J. Clin. Cancer Res. 2004, 10, 3708-<br />

3716.<br />

[138] Kim D-W.; Kim, S-Y.; Kim, H-K.; Kim, S-W.; Shin, S. W.; Kim,<br />

J. S.; Park, K.; Lee, M. Y.; Heo, D. S. Annals of Oncology., 2007,<br />

18, 2009-2014.<br />

[139] Baker, J. R. Jr. Hematololy Am Soc Hematol Educ Program., 2009,<br />

1, 708-719.<br />

[140] Quintana, A.; Raczka, E.; Piehler, L.; Lee, I.; Myc, A.; Mahoros, I.;<br />

Patri, A. K.; Thomas, T.; Mulé, J.; Baker, J. R. Pharm. Res., 2002,<br />

19, 1310-1316.<br />

Received: ??????????? Revised: ??????????? Accepted: ???????????<br />

[141] Majoros, I. J.; Williams, C. R.; Becker, A.; Baker, J. R. Jr. Wiley<br />

Interdiscip Rev. Nanomed. Nanobiotechnol., 2009, 1, 502-510.<br />

[142] Kono, K.; Liu, M.; Fréchet, J. M. Bioconjug. Chem., 1999, 10,<br />

1115-1121.<br />

[143] Malik, N.; Evagorou, E. G.; Duncan, R. Anticancer <strong>Drug</strong>s, 1999,<br />

10, 767-776.<br />

[144] Myc, A.; Patri, A. K.; Baker, J. R. Jr. Biomacromolecules, 2007, 8,<br />

2986-2989.<br />

[145] Zhuo, R. X.; Du, B.; Lu, Z. R. J. Control. Release, 1999, 57, 249-<br />

257.<br />

[146] Chen C. Z.; Cooper, S. L. Biomaterials, 2002, 23, 3359-3368.<br />

[147] Abeylath, S. C.; Turos, E.; Dickey, S.; Lim, D. V. Bioorg. Med.<br />

Chem., 2008, 16, 2412- 2418.<br />

[148] Balogh, L.; Swanson, D. R.; Tomalia, D. A.; Hagnauer, G. L.;<br />

McManus, A. T. Nano Lett., 2001, 1, 18-21.<br />

[149] Cheng, Y.; Qu, H.; Ma, M.; Xu, Z.; Xu, P.; Fang, Y.; Xu, T. Eur. J.<br />

Med. Chem., 2007, 42, 1032-1038.<br />

[150] Devarakonda, B.; Hill, R. A.; Liebenberg, W.; Brits, M; de Villiers,<br />

M. M. Int. J. Pharm., 2005, 304, 193-209.<br />

[151a] Klajnert, B.; Janiszewska, J.; Urbanczyk-Lipkowska, Z.;<br />

Bryszewska, M.; Shcharbin, D.; Labieniec, M. Int. J. Pharm., 2006,<br />

309, 208-217. (b) Klajnert, B.; Janiszewska, J.; Urbanczyk-<br />

Lipkowska, Z.; Bryszewska, M.; Ep<strong>and</strong>, R. M. Int. J. Pharm.,<br />

2006, 327, 145-152.<br />

[152a] Witvrouw, M.; Fikkert, V.; Pluymers, W.; Matthews, B.; Mardel,<br />

K.; Schols, D.; Raff, J.; Debyser, Z.; De Clercq, E.; Holan, G.;<br />

Pannecouque, C. Mol. Pharmacol., 2000, 58, 1100-1108. (b)<br />

Witvrouw, M.; Weigold, H.; Pannecouque, C.; Schols, D.; De<br />

Clercq, E.; Holan, G. J. Med. Chem., 2000, 43, 778-783.<br />

[153] L<strong>and</strong>ers, J. J.; Cao, Z.; Lee, I.; Piehler, L. T.;Myc P. P.; Myc, A.;<br />

Hamouda, T.; Galecki, A. T.; Baker, J. R. Jr. J. Infect. Dis., 2002,<br />

186, 1222-1230.<br />

[154a] Lo-Man, R.; Bay, S.; Vichier-Guerre, S.; Dériaud, E.; Cantacuzène,<br />

D.; Leclerc, C. Cancer Res., 1999, 59, 1520-1524. (b) Bay, S.; Lo-<br />

Man, R.; Osinaga, E.; Nakada, H.; Leclerc, C.; Cantacuzène, D. J.<br />

Pept. Res., 1997, 49, 620-625.<br />

[155] Ota, S.; Ono, T.; Morita, A.; Uenaka, A.; Harada, M.; Nakayama,<br />

E. Cancer Res., 2002, 62, 1471-1476.<br />

[156a] Moreno, C. A.; Rodriguez, R.; Oliveira, G. A.; Ferreira, V.;<br />

Nussenzweig, R. S.; Moya Castro, Z. R.; Calvo-Calle, J. M.;<br />

Nardin, E. Vaccine, 1999, 18, 89-99. (b) Joshi, M. B.; Gam, A. A.;<br />

Boykins, R. A.; Kumar, S.; Sacci, J.; Hoffman, S. L.; Nakhasi, H.<br />

L.; Kenney, R. T. Infect Immun., 2001, 61, 4884-4890.<br />

[157] Kukowska-Latallo, J. F.; Bielinska, A. U.; Johnson, J.; Spindler,<br />

R.; Tomalia, D. A.; Baker, J. R. Jr. Proc. Natl. Acad. Sci.U.S.A.,<br />

1996, 93, 4897-4902.<br />

[158] Navarro, D. G.; Tros de Ilarduya, A. M. An. R. Acad. Farm. 2008,<br />

74, 229-256.<br />

[159] Tang, M.X.; Redemann, C.T.;Szoka, F.C. Jr. Bioconjug. Chem.<br />

1996, 7, 703-714.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!