9

Carrier systems and biosensors for biomedical applications

F. Davis and S.P.J. Higson,    Cranfield University, UK

Abstract:

This chapter initially addresses why carrier systems are required and describes both natural and synthetic versions of these systems. Nanotechnology is becoming increasingly utilised in carrier systems so current and future applications of these materials are discussed. The section on biosensors begins with a history and descriptions of basic sensor formats. Glucose biosensors are described in more detail due to their market dominance. There follows a discussion of the latest developments is implantable glucose sensors and their potential for continuous glucose monitoring. The wider use of biosensors is then detailed, including their use for point-of-care applications.

Key words

drug delivery; liposome; hydrogel; biosensor; immunosensor; nanotechnology

9.1 Introduction

This chapter is intended as a sequel to our earlier chapter in Tissue Engineering using Ceramics and Polymers published in 2007 (Davis and Higson 2007), following up on much of our earlier writing and incorporating the advances of the last few years. This chapter addresses both carrier systems and biosensors which are often applied directly to tissues, either as skin patches, implanted or ingested by a variety of routes. There are a wide variety of these materials but since a full review would be impractical for a chapter of this size, we will mainly confine ourselves to carriers that are polymeric or ceramic in nature.

Sections 9.2 and 9.3 of this chapter will describe the use of carrier systems in biomedical applications. Initially a discussion is provided as to why carrier systems are required and this is followed by descriptions of different classes of material of natural and synthetic origin. This section will continue with a description of how nanotechnology is becoming utilised in carrier systems and will close with a discussion of current and future applications of these materials. Where this is especially relevant is when artificial scaffolds are used to aid tissue repair. There is the potential to utilise the scaffold also as a drug carrier system where controlled drug delivery of bioactive molecules from the scaffold would be an essential component of an engineering strategy for aiding the repair of defective tissue. This is somewhat outside the scope of this chapter but a number of reviews have been written on this subject (Habraken et al. 2007; Mouriño and Boccaccini 2010; Ekenseair et al. 2013).

Sections 9.49.6 of this chapter will be devoted to biosensors, beginning with a history and descriptions of basic sensor formats. Glucose biosensors will be described in more detail due to their dominance of the biosensor market. This will be followed by a description of the work progressing to the latest developments for implantable glucose sensors and their potential for continuous glucose monitoring. The wider use of biosensors will then be discussed, including the use of immunosensors to detect a wide range of physiological conditions and the use of biosensors for point-of-care applications. Possibly biosensors could be utilised in the vicinity of tissue engineering scaffolds to monitor the healing or any regeneration process or to provide an early warning system if anything should be beginning to go wrong.

9.2 Carrier systems

In many instances a simple one-off administration of a drug will suffice, e.g. taking an analgesic in response to a headache. However, this simple approach is often insufficient since often the condition may require continuous medication over a period of time, with many conditions requiring long-term drug therapy. Often a drug may have an optimum concentration within the body; too low and no benefit is derived, too high and unwanted side effects or toxicity can occur, endangering the health of the patient. However, when a drug is ingested or injected, the level within the patient’s blood tends to rise towards a plateau and then after some time falls again to zero unless more drug is given. In many instances a more efficient application method would be one that gives a stable level of treatment. Other problems can occur with drug stability: for a drug, for example, to be taken orally it must not be degraded by digestive processes. At the time of writing our original chapter, approximately 15% of the current world pharmaceutical market consisted of products which utilised a carrier system. Latest reports predict the US market for drug delivery systems is expected to reach $135 billion by 2015 (Taiyou Research 2011) with a 9% per year rise between 2010 and 2015. These writers predict that the largest share will be produced by oral drugs, with parenteral or injectable drug delivery systems also expected to grow at a rapid speed, especially in cancer applications. The fastest growth in the coming years is likely to be witnessed by implantable drug delivery systems.

To minimise drug degradation and optimise drug delivery, a wide variety of drug carrier and delivery systems are under investigation and numerous reviews have been published on this subject (Langer 1995, 1998; Kaparissides et al. 2006; Ganji and Vashedghani-Farahani 2009; Bose and Tarafder 2012; Dash and Konkimalla 2012, Fredenberg et al. 2012; Jain et al. 2012; Iwamoto 2013; Sultana et al. 2013).

A number of methods can be utilised to apply various carrier systems:

• Ingestion – the drug can be incorporated into a carrier system so that the composite material can be swallowed. This method is the most popular due to its simplicity and convenience. The carrier system must be capable of protecting the drug against the highly acidic medium of the stomach and/or enzymatic degradation throughout the digestive system. Sometimes the digestive process itself can be utilised to cause release of the drug. There are, however, a number of problems with this technique since it can lead to irritation of the bowel and also in many cases the drug must still penetrate the stomach/intestine wall.

• Inhalation – the drug can be dispersed within an aerosol, usually via a nebuliser and in this way can be directly inhaled. Especially suitable for treating respiratory diseases, this technique is widely used for the treatment of asthma. However, delivery via this method can still be adversely affected by the barrier between air and blood within the lung.

• Transdermal – the drug is incorporated within a patch, similar to the nicotine patches used to relieve ‘cravings’ of people attempting to stop smoking. This method avoids the problems of degradation of drugs by digestive process and can be used to provide local delivery, e.g. to a wound or skin conditions. Patches often need to be applied only once every several days and the method is non-invasive and painless. It also has the advantage that unlike oral and inhalation routes, patches can be easily and safely applied to unconscious patients. The skin barrier does lead to slow penetration rates and therefore in many instances only relatively low dosage levels can be attained. Other problems may include a lack of dosage flexibility.

• Injection – this technique has the disadvantage of being invasive and can be painful. Drug carrier systems can be utilised within this method, for example to prevent degradation of the drug within the bloodstream. As an alternative, a drug/carrier composite can be surgically implanted close to an affected site.

9.2.1 Classes of materials

Hydrophilic polymers

The pharmaceutical industry has shown great interest in the development of controlled release systems based on hydrophilic polymers. Hydrogels represent a common class of polymers used for drug delivery. In essence a hydrogel is a material based on polymers such as poly(vinyl alcohol) or poly(acrylic acid) which would normally be soluble in water. However, either during or after the polymer synthesis, a degree of crosslinking converts the linear polymer chains into a polymer network. This process renders the polymers insoluble; however, the high presence of hydrophilic groups within the network gives the structure a high affinity for water. Although the network does not dissolve, it is capable of adsorbing water with consequent swelling of the polymer matrix. The nature of the polymer and degree of crosslinking affects the swelling behaviour. A network with few crosslinks and a large number of hydrophilic groups will adsorb large amounts of water with a high degree of swelling. Less hydrophilic monomers, incorporation of hydrophobic co-monomers or a high degree of crosslinking all act to reduce water adsorption, usually leading to a firmer, more rigid gel. Due to the fact that they include a high water content, hydrogels often show high degrees of biocompatibility. The polymer network itself can be either bio-inert or be a biodegradable polymer network. Natural polymers can also be used with, for example, hydrogels based on chitosan, alginase or collagen having been utilised. The application of a wide variety of hydrophilic biodegradable polymers to delivery of proteins has been extensively reviewed (Gombotz et al. 1995; Ganji and Vashedghani-Farahani 2009). Frequently the drugs, especially if they are biologically derived (such as proteins) can be quite unstable. Besides enabling the controlled release of the active material, these hydrogels often act as a stabilizing media for these unstable agents. For example, a drug could be incorporated into a hydrogel either as it is synthesised or post-synthesis. This can then be applied to the patient via any of the techniques described above. Once in vivo, the drug is released by a number of means. These might include simple diffusion or alternatively the polymer may be eroded or dissolved, for example, by digestion. In an ideal situation this process will occur at a constant rate, leading to continual release of controlled amounts of the drug. If the polymer is utilised as a transdermal patch or is surgically implanted close to an affected site, the drug can be delivered where it is most needed.

A similar method utilises polymers which are actually degraded rather than just swollen. Examples include polymers which are based on poly(lactide) or poly(glycolide) (Fig. 9.1), and which are slowly hydrolysed in vivo to release an active agent, e.g. leuprolide acetate (Ogawa et al. 1988).

image
9.1 Structures of (a) poly(glycolide) and (b) poly(lactide).

A polymer must address several criteria before it is suitable for use as a carrier system. In the case of polymers that are effective in vivo, they must be both biocompatible and must biodegrade within a reasonable period of time. Irrespective of the manner of application, the polymer itself and any degradation products must be non-toxic and must create neither allergic nor inflammatory response. The method of release is often dependent on the nature of the drug itself; low molecular weight drugs are capable of diffusing out of the polymeric matrix and, if water soluble, will be rapidly released. However, larger molecules such as proteins will not diffuse as readily and often remain within the hydrogel matrix until the polymer itself is either degraded or enzymatic digestion releases them. Release rates depend on several factors including the quantity/dosing of drugs within the composite, the rate of degradation of the polymer, the water content (if it is a hydrogel-type material), and the presence and degree of crosslinking.

A property that makes hydrogels exceedingly suitable for drug encapsulation is that many of the materials used for these systems can be synthesised so as to be responsive to their environment. A change of pH, for example, can lead to protonation or deprotonation of active groups within the polymer. This can change its affinity for water, giving us a material whose swelling is pH responsive, thereby affecting many of its other physical properties such as permeability. This gives us the opportunity to design drug release agents which are selectively triggered by certain conditions, e.g. their responsiveness to pH means they can be designed to release active agents within a selected part of the digestive tract. Similar smart materials can be designed which respond to changes in physiological conditions and therefore only release the drug at times when it is needed. Hydrogels can be designed which respond to other stimuli as well as pH, such as ionic strength, temperature and electric field. They can even be designed to respond to the presence or absence of specific analytes.

As mentioned earlier, polymer-based delivery systems allow the application of active agents in many ways including ingestion, transdermal patches, suppositories, ocular and subcutaneous methods. A few examples are given here: nitroglycerin is often used in the treatment of angina; however, volatilisation of the active component can lead to loss of tablet activity (Markovich et al. 1997). This problem can be mitigated by the use of a acrylic-based hydrogel as a host for the nitroglycerin and incorporation of this composite in the construction of a transdermal patch.

There has been wide research into utilising these materials in the treatment of cancer for, for example, the delivery of ara-C for leukaemia. This drug gives rise to a number of side effects but these are mitigated when a constant infusion of the drug is introduced subcutaneously. As an alternative approach, crosslinked polyhydroxyethyl acrylate can be utilised as a host for this material. Discs of this composite display a steady controllable release of the active material (Teijon et al. 1997). A similar composite based upon polycaprolactone/polyethylene glycol (PCL/PEG) has been used as a matrix for the anticonvulsant drug clonazepam. Stable constant release properties were displayed for over 45 days (Cho et al. 1999).

Similar polymers are suitable for localised delivery of pharmaceuticals. In the treatment of brain cancer, one approach that has been successfully used is surgery to remove as much of the tumour as possible, followed by placing in the surgical site small wafers based on polyanhydrides. These contain the anti-cancer drug carmustine which is slowly released over a one month period to kill any remaining tumour cells (Brem et al. 1995). Alternatively thermosensitive hydrogels based on block copolymers of poly(ethylene oxide) and poly(lactide) have been made which are liquid at 45 °C and can be injected directly to the required site. Upon cooling to body temperature, the gel immediately sets, trapping any pharmaceutical compounds in the solution, so allowing them to be released slowly (Jeong et al. 1997).

Intelligent hydrogels

The materials mentioned so far are usually based on simple polymeric systems; however, more complex reactive systems can be designed. As mentioned earlier, hydrogels can be synthesised which respond to environmental conditions. These can be relatively simple changes such as temperature or pH, alternatively other chemical or biological species can be incorporated into these hydrogels to make them chemically or biologically responsive (Ganji and Vashedghani-Farahani 2009; Schmaljohann 2006).

A variety of polymers have been utilised to formulate responsive hydrogels. For example, a number of acrylamide-based polymers such as poly(N-isopropyl acrylamide) experience a transition from a open soluble form to a more compact insoluble form as temperature increases. Hydrogels based on these polymers tend to shrink as the temperature is increased above a critical value (Schmaljohann 2006) due to increased polymer–polymer hydrogen bonding. Other materials based on crosslinked polyacrylic acid/polyacrylamide swell as temperature increases.

Thermo-responsive polymers that exist as have been utilised as drug release agents. These polymers exist as a low viscosity fluid at room temperature, but when injected into the body, the increase in temperature causes them to gel. Examples of this include PLGA-PEG-PLGA (PLGA = poly(lactic-co-glycolic) acid) tri-block copolymers can be gelled along with the peptide calcitonin to allow steady zero-order delivery of the peptide over 100 hours (Ghahremankhani et al 2008). Other workers (W. C. Lee et al 2006) used copolymers of lactic acid and PEG to synthesise polymers which, on crosslinking with UV radiation, formed nanosized gel particles in suspension, which proved to be excellent controlled release agents for the hydrophobic anticancer drug camptothecin.

Hydrogels can also be synthesised that display a strong swelling change in response to pH. Such hydrogels are based on polymers that contain ionisable groups. These can be acidic polymers such as poly(acrylic acid), which becomes ionised and swells at higher pH as well as basic polymers such as poly(dimethylamino acrylate) which is protonated and swells at low pH. For example, the anti-cancer drug Adriamycin could be conjugated to a dimethyl maleic anhydride/vinyl pyrrolidinone copolymer and shown to release the drug gradually at neutral or slightly acidic pH (Kamada et al. 2004), with the resultant conjugate showing anticancer activity in mice with reduced toxic side effects than the free drug. Copolymers of pH and temperature-sensitive monomers have also been utilised to give hydrogels that are responsive to both stimuli, such as for example poly(N-isopropylacrylamide-co-butylmethacrylate-co-acrylic acid) which could be synthesised as beads containing entrapped calcitonin and release it in a controllable fashion (Serres et al. 1996).

One major field of research is hydrogels which will respond to the presence of certain biological species, enabling them to respond to certain physiological conditions. An example of this is a ‘smart’ porous membrane made from a polymethacrylic acid/PEG copolymer. This is used as a host for insulin but also contains encapsulated glucose oxidase. Glucose oxidase specifically catalyses the oxidation of glucose to gluconic acid, which causes a subsequent pH drop. This leads to shrinkage of the membrane and the controlled release of insulin in response to hyperglycaemia (Gander et al. 2001). A wide range of insulin-releasing polymers have been synthesised; a detailed review of these materials is outside the scope of this chapter but details of many of these and other bioresponsive polymers have been published (Miyata et al. 2002).

Polymers have also been synthesised which can actually target cancer cells. For example recently, catechol-based polymers were demonstrated to be taken up preferentially by cancer cells through cell surface receptor-mediated mechanisms (Su et al. 2011). These polymers could be used to carry the anti-cancer drug bortezomib to cancer cells and there release it. Using this method dramatically enhanced cellular uptake, proteasome inhibition and cytotoxicity toward breast carcinoma cells in comparison with nontargeting drug–polymer conjugates. Other work utilised temperature responsive hydrogel nanoparticles which had been conjugated within a ligand for targeting cancer cells, folic acid (Kim et al. 2006). Receptor mediated endocytosis meant these particles were preferentially adsorbed when incubated with liver cancer cells.

Natural polymers

Besides synthetic polymers, natural polymers have also been widely studied as drug carrier agents, especially as for many of them their biocompatibility and non-toxicity are well known since they form major components of our diets. Proteins are of great interest due to their wide variety of structures and their easy availability from sources such as, for example, egg-white, soybean and whey (Chen et al. 2006). Hydrogels can be easily generated from a wide variety of proteins and possess the desirable qualities of their synthetic analogues. Usually proteins can be gelated by methods such as heating, which unfolds the polypeptide chains within the protein structure. These then tend to aggregate, forming a three-dimensional structure crosslinked by hydrogen bonding and/or hydrophobic effects (Clark et al. 2001). Other methods for inducing gelation include crosslinking with calcium ions (Maltais et al. 2005). A wide variety of hydrogel morphologies and microstructures can be generated by variation of the preparation technique (Chen et al. 2006). Applications of these gels include controlled release of tocopherol (Chen et al. 2006) from a lactoglobulin-based emulsion under simulated gastric conditions. Other examples include the incorporation of drug compounds within albumin (Sokoloski and Royer 1984; Tomlinson and Burger 1985) or corn protein (Liu et al. 2005) and a review has been published on this topic (Chen et al. 2006).

Responsive hydrogels have also been made using natural polymers, for example hyaluronic acid and poly(N-isopropylacrylamide) copolymers have been synthesised, shown to form gels above 30–33 °C and demonstrated controlled release of a model protein bovine serum albumin (BSA) in rabbits for 60 hours (Ha et al. 2006). Chitosan/PEG copolymers also could be utilised as controlled release agents for BSA, showing steady release in vitro for 70 hours (Bhattarai et al. 2005). Starch/acrylic acid/acrylamide polymers have been shown to form hydrogels which were responsive to pH and salt concentrations (Sadeghi and Hosseinzadeh 2008) and the release of ibuprofen from these materials studied. Dextran-based hydrogels have also been conjugated to a model drug 2-methoxyestradiol and shown to release it in simulated colonic fluid as a result of swelling induced by pH combined with increased enzymatic degradation (Casadei et al. 2008). Bioresponsive polymers have also been synthesised. For example, triblock copolymer gels containing certain peptide sequences have been synthesised where the peptide sequence is digested by disease-associated proteases (Law et al. 2006). Model systems show that exposure of the polymer gels to the protease leads to dissolution and release of any encapsulates.

9.2.2 Micelles, vesicles and liposomes

Micelles are formed by a wide range of amphiphilic surfactant molecules in aqueous solution. A typical amphiphile contains a long alkyl chain and a polar headgroup. In solution the interactions between the alkyl chains and water are extremely unfavourable and this drives aggregation of the alkyl chains together to minimise this interaction. A structure then forms as shown schematically in Fig. 9.2, where a spherical aggregate spontaneously assembles, with the polar headgroups on the outside of the sphere and a hydrophobic interior, stabilised by van der Waals interactions between the chains.

image
9.2 Structures of (a) a micelle and (b) a vesicle.

What makes micelles so useful as carrier agents is that hydrophobic species can be incorporated within the core of the micelle. This enables the transport of hydrophobic active agents at concentrations much higher than would be possible for a simple aqueous solution of these compounds. The hydrophobic environment of the core moreover protects the guest against hydrolysis and enzymatic degradation.

As an alternative to the classical long chain surfactants shown in Fig. 9.2, amphiphilic block copolymers are also very adept at forming micellar structures. For example, a block copolymer containing a hydrophobic block, such as polystyrene, and a hydrophilic block, such as polyethylene oxide, can dissolve in water to give a micellar structure with a polystyrene core surrounded by a polyethylene oxide corona, with the whole structure being typically 5–50 nm in diameter. Block copolymers are especially suitable for this purpose since they are available in a wide variety of molecular weights. Since a wide variety of monomers and the composition ratios can easily be selected, this allows fine control of the size, composition and morphology of the micelles. Incorporation of many functional groups is easily attained; for example, crosslinking groups can be utilised to stabilise the micelle structure and the surface properties, selectivity and biocompatibility of the micelles can all be tailored.

In a similar manner, a variety of compounds can be utilised to form vesicles. These are somewhat more complex structures than micelles (Fig. 9.2) in which a spherical membrane is formed containing an aqueous core. Liposomes are types of vesicles which are widely used within drug transport and consist of spherical phospholipid bilayers with an aqueous core and are the most commonly used of these carrier systems (Langer 1998; Kaparissides et al. 2006). They are capable of transporting either hydrophilic agents within the core or hydrophobic agents which are incorporated into the membrane. The membrane itself can be tailored to prevent access of enzymes to the core, thereby preventing enzymatic degradation, but do, however, allow the diffusion out of the active agent. For example, channel-type proteins may be incorporated within the membrane and retain their activity, thereby allowing transport of ions or other small solutes such as drugs through the membrane.

The use of liposomes can lead to a large increase in drug carrying capacity compared to the use of polymeric systems; however, disadvantages can be encountered such as shorter shelf-life and rapid destruction of the vesicle within the body. The biocompatibility of these systems, however, can be improved by modifying their surface properties, such as by the attachment of PEG units (Park et al. 1997; Lasic et al. 1999; Moghimi and Szebeni 2003). Also, antibodies can be attached to the surfaces of these systems, allowing them to specifically target sites that require treatment. Tumours, for example, can be targeted by substituting liposomes with antibodies against the HER2 proto-oncogene, found in breast and other cancers (Kirpotin et al. 1997; Park et al. 1998). Similarly amylopectin can be attached to the surface of liposomes to enable targeting of lung tissue (Vyas et al. 2004).

More recently there has been much attention paid to using liposomes to deliver protein or peptide-based drugs since they can maintain the guest in an internal aqueous environment whilst protecting from the external environment. For example, vasorestrictive intestinal peptide could be encapsulated into PEG-conjugated liposomes and then delivered via aerosol inhalation, with the liposome protecting the peptide from aggregation and enzymatic degradation (Hajos et al. 2008). Similarly liposomes could be used to deliver therapeutic proteins against Chlamydia into mice and demonstrated reduced rates of infection (Hansen et al. 2008). Insulin can be incorporated into multivesicular liposomes coated with chitosan or carbopol. The resultant liposomes were 26–34 μm in diameter, contained 58–62% protein and when applied by nasal or ocular administration showed sustained protein released as assessed by reduction of blood glucose in diabetic rats (Jain et al. 2007). Other workers have studied chitosan containing liposomes and demonstrated controlled release of calcitonin in rats after oral application (Werle and Takeuchi 2009).

Liposomes have also been widely studied since they offer the potential to introduce drugs into the body via application to the eyes. Much of the work on ocular administration of drugs and other compounds has been recently reviewed (Mishra et al. 2011). Recently there has also been interest in catansomes, vesicles made from oppositely charged surfactants rather than phospholipids. In a recent example, perfluorinated sodium octanoate could be mixed with a hydrogenated surfactant such as dodecyl pyridinium chloride to form vesicles 100–200 nm diameter which could encapsulate a fluorescent marker (calcein) and release it upon application of fatty acids (Rosholm et al. 2012).

9.2.3 Nanotechnology

In recent times, nanotechnology has become an intensely studied field of research. The application of nanosized materials is expected to make significant advances in biomedical applications such as drug delivery and gene therapy (Moghimi et al. 2001; Moghimi and Szeleni 2003; Sahoo and Labhasetwar 2003).

Earlier in this chapter the use of biodegradable polymers as drug carriers was discussed. Nanoparticles of these types of materials with sizes in the range 10–1000 nm can be synthesised and incorporate drug molecules by way of entrapment or binding. Similarly, nanocapsules can be synthesised in which a polymer membrane forms a vesicle-like structure with the drug molecules confined within a central cavity. Because of their small size, these systems are capable of passing through small capillaries and be taken up by cells, allowing efficient drug accumulation at a target size (Desai et al. 1997). In these materials, size and surface properties determine their distribution in the body (Sahoo and Labhasetwar 2003). Localised application of these nanoparticles can be achieved by their tendency to accumulate in tumours due to enhanced permeation effects; this has been extensively reviewed elsewhere (Maeda 2001). Alternatively the nanoparticles can be delivered locally to the size of interest such as within a specific artery after balloon angioplasty (Guzman et al. 1996) to deliver long-term release (14 days) of dexamethasone. More recently nanoparticulate PLGA has been used to deliver zinc phthalocyanine into mouse tumours for use in photodynamic therapy (Fadel et al. 2010). Other workers used PLGA nanoparticles to deliver paxitaxel (Yang et al. 2009) or to simultaneously deliver vincristine and verapamil (Song et al. 2009) into cancer cells. Chitosan nanoparticles substituted with folic acid could be used to selectively deliver oligonucleotides into tumour cells (J. Wang et al. 2010) and similarly modified chitosan nanoparticles could selectively deliver paxitaxel with good cytotoxicity (Sahu et al. 2011). Crossing the barrier between blood and the brain or central nervous system is a major challenge for many drugs; however, modified chitosan/PLGA nanoparticles have been shown to cross the barrier (Z. H. Wang et al. 2010). These and other polymeric nanoparticles for drug delivery have been recently extensively reviewed (Parveen et al. 2012).

Dendrimers are branched polymers grown from a central core with very precisely controlled degrees of polymerisation. Figure 9.3 shows a schematic of a fourth generation dendrimer. This makes their size, composition and molecular substitution controllable to an exact degree and is an alternative form of polymer nanoparticle to those made by classical emulsion-type polymerisations. There has been great interest in utilising dendrimers as drug carrier agents because not only can the bulk of the dendrimer be synthesised exactly but a wide variety of surface groups can be utilised, enabling further surface functionalisation. Dendrimers, for example, can be made with hydrophilic surfaces and hydrophobic interiors or vice versa. Reviews on this subject (Svenson and Tomalia 2005; Gupta et al. 2006) and on the biomedical applications of these materials (Mintzer and Grinstaff 2011) extensively detail many of the most pertinent advances in this field, so only a few highlights will be given here.

image
9.3 Schematic structure of a fourth generation dendrimer.

Dendrimers have been utilized to carry a variety of small molecule pharmaceuticals. Commercial poly(amido amine) (PAMAM) dendrimers have been used to encapsulate the anticancer drug cisplatin, giving conjugates that exhibited slower release, higher accumulation in solid tumours, and lower toxicity compared to the free drug (Malik et al. 1999). PAMAM dendrimers which had been functionalised with PEG chains had their encapsulation behaviour for the anticancer drugs adriamycin and methotrexate studied. Up to 6.5 adriamycin molecules or 26 methotrexate molecules per dendrimer could be incorporated for one of the materials studied. The drug release from this dendrimer was slow at low ionic strength but fast in isotonic solution (Kojima et al. 2000). Similar dendrimers encapsulated the anticancer drug 5-fluorouracil, showing reasonable drug loading, and reduced release rate and haemolytic toxicity (Bhadra et al. 2003). Up to 78 molecules of the anti-inflammatory drug ibuprofen molecules were complexed by PAMAM dendrimers through electrostatic interactions between the dendrimer amines and the carboxyl group of the drug. The drug was successfully transported into lung epithelial carcinoma cells by the dendrimers (Kolhe et al. 2003). Other recent studies have also indicated that low generation PAMAM dendrimers cross cell membranes (El-Sayed et al. 2003).

One of the main applications recently of dendrimers is as carrier agents for anticancer drugs (Mintzer and Grinstaff 2011). Glycerol succinic acid dendrimers have been shown to encapsulate hydrophobic molecules (Morgan et al. 2006), including the anti-cancer drugs 10-hydroxycamptothecin and 7-butyl-10-aminocamptothecin. The resultant materials display cytotoxicity against breast, lung and colonorectal cancer cells. In mice the hepatotoxicity of methotrexate and 6-mercaptopurine, both Food and Drug Administration (FDA)-approved anticancer drugs, was shown to be reduced by binding to a melamine-based dendrimer (Neerman et al. 2004). Dendrimers could be substituted with glucosamine, which enabled them to cross the blood–brain barrier, and loaded with methotrexate. The resultant composites showing higher efficiency and enhanced permeability compared to the drug alone (Dhanikula et al. 2008).

Covalent binding of the drug species to the dendrimer by a linkage which can then by hydrolysed or otherwise decomposed has also proved an effective means of drug carriage. Doxorubicin could be attached by acid-sensitive cis-aconityl linkages to a commercial PAMAM dendrimer conjugated with various amounts of PEG chains (Zhu et al. 2010). These, when mixed with ovarian cancer cells, showed clear cellular uptake of the dendrimer along with release of the drug. Higher levels of PEG substitution led to higher levels of uptake. An asymmetric biodegradable PEG substituted polyester dendrimer containing 8–10 wt % doxorubicin attached by covalent hydrazone linkages could be prepared (C. C. Lee et al. 2006) and injected into cancerous mice. The dendrimer composite was nine times more effective at being uptaken by tumours than the pure drug and led to 100% tumour regression and survival of the mice over 60 days whereas no regression was seen for the control group treated with doxorubicin alone. Targeted dendrimers such as the recent example substituted with folic acid and containing methotrexate have been shown to display specificity and cytotoxocity towards cancer cells (Zhang et al. 2010). Other very recent work has compared doxorubicin encapsulated in PEGylated dendrimers and PEGylated liposomes and concluded that although the two systems display similar superior anti-tumour efficiency and enhanced accumulation in rat tumours to the native drug, the dendrimer has lower systemic toxicity (Kaminskas et al. 2012).

Ceramic nanoparticles have also been studied due to their inherent advantages of ease of synthesis in a wide variety of sizes, shapes and porosities by methods similar to sol–gel processes. They are available in very small (< 30 nm) sizes making them capable of crossing cell membranes, possess surfaces which are easily chemically modified and do not display swelling processes with changes in pH (Sahoo and Labhasetwar 2003). For example, silica nanoparticles can be constructed with an anti-cancer drug entrapped within their cores (Roy et al. 2003) which are stable in aqueous systems. Tumour cells take up these nanoparticles, which upon irradiation generate singlet oxygen which significantly damages the tumour cells. Similar particles have also been shown to act as non-viral DNA carriers (Bharali et al. 2005) and to be capable of crossing the blood–brain barrier.

Other metal oxides have been studied, especially iron oxides since they are paramagnetic and therefore potentially allow for targeting certain locations in the body by using a magnetic field (Parveen et al. 2012). Hyaluronic acid coated iron oxide nanoparticles were shown to be capable of transferring peptides into cells (Kumar et al. 2007). Other workers demonstrated delivery of doxorubicin into breast and prostate cancer cells using polymer-coated iron oxide nanoparticles (Jain et al. 2005) or delivery of genistein using chitosan-modified nanoparticles (Si et al. 2010). Similar particles coated with polyethyleneimine were shown to be uptaken by cancer cells and when injected into rat carotid arteries, they could then be magnetically targeted into the brain (Chertok et al. 2010).

A wide range of other nanoparticles are being studied as delivery agents such as gold and other metals, and carbon nanotubes. These are outside the scope of this chapter but have been extensively reviewed elsewhere (Parveen et al. 2012). One major issue with any nanoparticle-based delivery system is potential long-term toxicology issues, since the toxicity of many nanoparticles is still unknown. However, there have been a range of studies showing that many materials, which whilst benign in bulk, are highly toxic in a nanosized form, a classical example being asbestos.

9.3 Commercial systems

There are a wide variety of commercial drug carrier agents available; we are only going to mention a few of these as a complete review would form a chapter itself.

9.3.1 Polyanhydrides

As previously mentioned, polyanhydrides are widely used as slow release agents since they biodegrade reproducibly with no toxic by-products. Commercial products include materials such as Decapeptyl SR®, manufactured by Ipsen Ltd which contains the active ingredient triptorelin acetate encapsulated in poly(lactide-co-glycolide). This is utilised in the treatment of prostate cancer. Similar products include Lupron Depot® (Abbott Laboratories, active ingredient leuprolide acetate) and Sandostatin LAR® (Novartis, active ingredient octreotide acetate).

Nitroglycerin is a problematic drug due to its loss of tablet activity, often by volatilisation of the active component (Markovich et al. 1997). This can be avoided by incorporation of the nitroglycerin into an acrylic-based hydrogel which is then incorporated into a transdermal patch, as exemplified by products such as Deponit® (Schwarz Pharma), Minitran® (3 M Pharma) and Nitrodisc® (G.D. Searle Company).

9.3.2 Liposomes

A number of commercial drug delivery systems have been developed, especially in the field of cancer treatment. Daunozome® is a liposomal form of daunorubicin, a chemotherapy drug given to treat AIDS-related Kaposi’s sarcoma. Doxorubicin can also be formulated within PEG-substituted liposomes to give Doxil® which can be used to treat a variety of cancers. Other anti-cancer drugs which have been formulated in this way include cytarabine to give Depocyt®. Various other commercial liposome formulations have also been brought to market (Zhang et al. 2008).

9.4 Biosensors

9.4.1 History and format of biosensors

The purpose of this section is to introduce the concept of using biological molecules as the selective recognition elements within biosensors. Most sensors consist of three principal components, as described below and shown in Fig. 9.4:

• The first of these includes a receptor species, usually biological in origin such as an enzyme, antibody or DNA strand capable of recognising the analyte of interest with a high degree of selectivity.

• The second component that must be present is a transducer, enabling the translation of the binding event into a measurable physical change; possible events include the generation of electrons, protons, an electrochemically active chemical species such as hydrogen peroxide or simple physical changes such as a change in conductivity, optical absorbance or fluorescence.

• Thirdly there must be inclusion of a method of measuring the change detected at the transducer and converting this into useful information.

image
9.4 Schematic of a biosensor.

There are several advantages associated with using biological molecules as the active recognition entity within a sensor. Usually they display unsurpassed selectivities; for example glucose oxidase will interact with glucose and no other sugar, and in this way will act as a highly selective receptor. In the case of glucose oxidase, the electrochemically inactive substrate glucose is oxidised to form gluconolactone along with the concurrent generation of the electroactive species hydrogen peroxide. Enzymes also generally display rapid turnover rates and this is often essential to (a) avoid saturation and (b) to allow sufficient generation of the active species in order to be detectable.

Antibodies bind solely to their antigens and achieve specificity via a complex series of multiple non-covalent bonds. Since the principle of immunoassay was first published by Yalow and Berson in 1959, there has been an exponential growth in both the range of analytes to which the technique has been successfully applied and the number of novel assay designs. Development of enzyme-labelled immunoanalytical techniques, e.g. enzyme-linked immunosorbent assay (ELISA) has provided analytical tests without the safety risks associated with radiolabelling-based techniques.

The rapid measurement of analytes of clinical significance, e.g. towards various disease markers, would permit earlier intervention, which in a medical setting is frequently of utmost importance. There has been much research on the development of direct immunosensors that do not rely on the use of a detectable label. Such a system will lead to simpler assay formats and ideally lower times of detection. A reusable and rapid detection system would, moreover, allow for continuous real-time measurement, so helping to maintain optimal homeostatic conditions.

Unfortunately there are also some disadvantages related to the construction and use of biosensors. Often the biological species can either be extremely expensive or difficult to isolate in sufficient purity. Immobilisation of these species can lead to loss of activity and the presence of various chemical species in the test solution can also cause loss of activity, e.g. enzymes can be easily poisoned by heavy metals. In biological samples such as blood or saliva, there can also be solutes that are electrochemically active and interfere with determinations of the target species or various species may be present which bind to the surface so causing fouling and loss of sensor response.

Antibody–antigen binding is based on multiple non-covalent interactions, and therefore there are no newly formed molecules, protons or electrons that are easily detectable, which has limited the development of direct antibody affinity-type sensors. Also affinity binding constants typically range from 105 to 1011 mol. L− 1 meaning that the antibody-antigen binding event is often irreversible. As a consequence of this, many contemporary immunosensors are only of use for ‘single-shot’ analyses and must be disposable in nature.

9.4.2 Glucose biosensors

A series of extensive reviews on biosensors and their history have been published elsewhere (Hall 1990; Eggins 1996; Wang 2001) and therefore only a brief history will be given here. Easily the most intensively researched area has been towards the development of glucose biosensors (Wang 2001; Newman et al. 2004). The reason for this is the prevalence of diabetes, which has become a world-wide public health problem. Diabetes represents an increasing epidemic with, at the time of writing, 346 million sufferers worldwide and this is estimated to double between 2005 and 2030 (World Health Organisation, www.who.org). The world market for biosensors is predicted to reach $15–16bn by 2016 and in 2009 approximately 32% of this market was for blood glucose monitoring. Point-of-care applications account for approximately half of the market (Thusu 2010).

These factors have led to the development of a number of inexpensive disposable electrochemical biosensors for glucose, incorporating glucose oxidase bound immobilised at various electrodes. They are generally amperometric sensors, with electrodes polarised at a set potential; the oxidation or reduction of a chosen electroactive species at the surface will then lead to generation of a detectable current. The principal classes of glucose and other types of biosensors are described below.

First generation biosensors

The first electrochemical glucose biosensor was based on an oxygen electrode (Clark and Lyons 1962). A film of immobilised glucose oxidase was laid down upon the oxygen electrode, which was overlaid with a semipermeable dialysis membrane. Upon exposure to glucose, the enzymatically catalysed oxidation reaction occurs, causing a localised consumption of oxygen.

glucose+O2Glucoseoxidasegluconolactone+H2O2 [9.1]

image [9.1]

This then leads to a drop in the current generated at the oxygen electrode. This device was subject to fluctuations caused by variable oxygen levels. However, further work (Updike and Hicks 1967) utilised two oxygen electrodes, one of which was coated with glucose oxidase. Measurement of differential current between the two electrodes served to cancel out these effects. Alternatively the production of hydrogen peroxide can be electrochemically determined instead. The first commercial glucose analyser was the Model 23 YSI analyser, launched by the Yellow Spring Instrument Company in 1975 and based on the Clark electrode. This device was capable of measuring the glucose level in 25 ml of whole blood.

The results from these so-called first generation sensors can be affected by fluctuations in the ambient oxygen concentration or by the presence of electroactive species such as ascorbate, that are capable of being oxidised at + 650 mV – giving rise to an erroneous result. Concentration of interferents at the electrode surface can be minimised by application of a permselective coating to the sensor, thereby reducing interference from electroactive species. Polymeric materials have led the way, with materials such as the fluorinated ionomer Nafion (Turner and Sherwood 1994) and cellulose acetate (Maines et al. 1997) being two of the most commonly used. A beneficial side effect is that these materials can also confer a degree of biocompatibility. An alternative approach has been to electropolymerise suitable monomers to form protective coatings. 1,2-Diaminobenzene (Myler et al. 1997), for example, when deposited at the bioelectrode surface serves to both stabilise the electrode due to its inherent high biocompatibility whilst also imparting selective exclusion of interferents such as ascorbate.

Diabetes as a condition requires regular monitoring of blood glucose levels, so hospital analysis is impractical for a normal lifestyle. The obvious solution has been the development of inexpensive home detection methods where the physiological sample, usually blood, can be analysed by the patient. The problems of cleaning the sensor are moreover negated by using disposable sensor strips.

Second generation biosensors

The reaction of glucose oxidase with glucose gives rise to the formation of gluconolactone and the reduced form of the enzyme, which is then reoxidised by oxygen. Direct transfer of electrons from the electrode would circumnavigate this reaction; however, the active site of glucose oxidase is encased in a protein sheath which inhibits this transfer. To facilitate this transfer, a suitable chemical species can be utilised to ‘shuttle’ electrons back and forth between active site and electrode. This moiety, known as a mediator, must react readily with the enzyme to avoid competition by ambient oxygen and, in both its reduced and oxidised forms, be stable and preferably require as low an over-potential to be oxidised as is feasible. This method sidesteps problems associated with detecting species such as oxygen or peroxide and also lowers the potential required for measurement of the enzyme catalysed reaction, thereby reducing the inference by redox active species present within the sample to be studied. A typical reaction scheme (Fig. 9.5) where a ferrocene compound acts as mediator is shown (Cass et al. 1984). The first home glucose testing kit, the Exactech® glucose biosensor, is based on this chemistry. The actual pen-sized device is produced by Medisense® and utilises a disposable strip upon which a single drop of blood is placed. A wide range of glucose sensors (Wang 2001) based on this method have since become commercially available for home glucose testing. Further work has concentrated on reducing blood volumes and produced devices such as the Pelikan® Sun device which only required microlitre blood volumes (Newman et al. 2004). Pelikan® Sun has been discontinued but other manufacturers such as Tiniboy (www.tiniboy.com) have concentrated on minimising lancet size.

image
9.5 The oxidation of glucose at an electrode, mediated by a ferrocene derivative.

Third generation sensors

Second generation biosensors are still somewhat limited in that they require use of a suitable mediator. Attempts have been made to circumvent this by developing methods to electronically directly connect or ‘wire’ the enzyme to the electrode, thereby allowing simple electron transfer from the enzyme to the electrode without the requirement for a mediator. Although these types of device have not as yet been developed commercially, they provide a possible alternative to mediated electron transfer.

Typical approaches involve the use of a polymeric coating which both immobilises the glucose oxidase and allows electron transfer to the electrode. One example is the use of polyvinyl pyridine where the synthetic polymer has been modified with a large number of osmium-based electron transfer relays (Degani and Heller 1987; Ohara et al. 1994). The polymer is co-immobilised at an electrode with glucose oxidase. When the enzyme reacts with glucose and is converted to the reduced form, the polymer allows facile transfer of electrons to convert it back to the oxidised form. Similar materials based on polyvinyl imidazole has also been utilised (Mano et al. 2005). Glucose oxidase and other enzymes have also been immobilised on various electrodes by a variety of chemical and physical methods (Davis and Higson 2005).

Conducting polymers are especially suitable for the immobilisation of enzymes at electrode surfaces and have been reviewed in greater detail elsewhere (Gerard et al. 2002; Barisci et al. 1996). A variety of monomers such as pyrrole or aniline can be electropolymerised on an electrode surface and under correct conditions form stable conductive films. If during this electropolymerisation process, enzymes are present in the solution they can be entrapped within the film during the deposition process, alternatively they can be adsorbed onto, or be chemically grafted, to the film following deposition (Gerard et al. 2002; Barisci et al. 1996). The close association between the conductive polymer and the enzyme facilitates rapid electron transfer between the enzyme and an electrode surface.

One of the first methods involved the simple entrapment of enzymes such as glucose oxidase within polyaniline films (Cooper and Hall 1992). Our group has taken this process further, utilising both non-conductive and conductive polymers to fabricate arrays of conductive microelectrodes with entrapped biological molecules such as glucose oxidase (Barton et al. 2004).

9.5 Continuous monitoring

One problem with the current commercial biosensors is the invasive procedure, requiring frequent withdrawal of blood for testing, which can be both tedious and painful. Devices which could be implanted within the human body would negate this aspect of glucose testing. Implantable sensors suitable for in vivo glucose monitoring require the device to be extremely small and show long-term stability (with minimal drift, thereby removing the need for frequent calibration), display no oxygen dependency and also show high biocompatibility. The problem of biocompatibility has been the most elusive of these targets and as yet in vivo glucose sensors have only limited lifetimes. Effects on sensor performance include fouling by protein deposition on the surface or formation of fibrous tissue around the sensor (D’Orazio 2003), leading to loss of sensor performance. Rejection by the immune system or thrombus formation if used intravascularly can also degrade device performance and risk harm to the user.

Subcutaneously applied sensors have been developed (Bindra et al. 1991; Henry 1998) which could monitor glucose concentrations and also be changed by the user. However the effects of biofouling mean their use was limited to periods of 1–2 weeks. Again, fouling of the sensor and its effect on sensor performance is a major consideration. Attempts have been made to improve this using polymeric coatings. For example, a needle-based electrochemical probe coated with Nafion has been developed (Moussy et al. 1993) which is just 0.5 mm in diameter and can be inserted subcutaneously through an 18-gauge needle. Again in vivo measurements could be carried out over periods of 2 weeks.

A different approach has been to implant a microdialysis fibre into subcutaneous tissue and an iso-osmotic electrolyte solution pumped through the fibre. Glucose diffuses into the fibre and the electrolyte. If flow rates are kept constant the glucose concentration in the outflow from the fibre can be directly related to the glucose concentration in the interstitial fluid. Rapid changes in blood glucose concentration can be determined, although there is a time delay, typically about 8 minutes (Jansson et al. 1988) between changes in blood and interstitial fluid concentrations. The use of these probes combined with glucose biosensors to monitor the outflow have achieved 4–7 day’s continuous use for glucose monitoring in humans (Myerhoff et al. 1992; Hashiguchi et al. 1994).

A subcutaneously implantable device, the CGMS® (Continuous Glucose Monitoring System), was first commercialized by Minimed (www.minimed.com). These devices simply measured glucose and also allowed the input of data such as mealtimes, etc., all of which was then later downloaded to a PC. Later models that have come more recently to market include the Guardian® REAL-Time device (Medtronic Ltd). The glucose biosensor probe is inserted just beneath the skin, usually in the abdomen, and can be used to monitor glucose for up to 72 h, with a reading every 5 min. Results are wirelessly transmitted to a display device which then immediately shows the current reading along with other information such as graphs of recent levels as well as short-term prediction of glucose levels. Traditional blood sampling of glucose is used to calibrate the device.

9.6 Immunosensors for point-of-care testing

There are a wide number of physiological conditions such as stoke, various degenerative diseases and infectious diseases that produce biological markers, which can be viral, bacterial, proteins or other agents. Early detection of many of these species can be highly beneficial and allow for rapid treatment. This has been the driving force for immunosensing approaches and led to the development of the ELISA assay; however, this is a laboratory-based technique and a point-of-care method of detecting these materials would be of great interest.

Immunosensors depend on the strong specific binding that occurs between antibodies and their antigens; this is by a combination of such forces as hydrogen bonding and other interactions and again does not give rise to electrons or other reaction products. This has meant that there has been a great deal of research into AC impedance-based immunosensors because of the potential for impedance to be used in label-free assays. Much of the work on these systems has been reviewed extensively (Daniels and Pourmand 2007). Polymers have been widely used in the construction of many of these devices.

One of the earliest immunosensors used interdigitated electrodes coated with a crosslinked protein film containing antibodies to human IgG; when exposed to a solution of the antigen, changes in impedance were noted. This allowed detection of as little as 50 nm · mL− 1 of IgG; the use of control electrodes containing no antibody allowed subtraction of non-specific binding effects (Taylor et al. 1991). Other workers (Maupas et al. 1996) used polysiloxane to immobilise anti-α-fetoprotein antibodies on silicon to give immunosensors with a dynamic range of 10 to 150 ng · mL− 1.

Conducting polymers again have been used to construct immunosensors. Biotinylated polypyrrole films could be used to immobilise antibodies by making use of biotin-avidin binding protocols to give AC impedance immunosensors which had a detection limit of 10 pg · mL− 1 and a linear range of 10–80 ng · mL− 1 IgG (Ouerghi et al. 2004). A simple co-deposition protocol could be used to entrap antibodies into polypyrrole films (Grant et al. 2005) to give AC impedance immunosensors for BSA and digoxin. Polyaniline could also be used as a transducer layer and could be chemically modified after deposition to give a surface substituted with biotin groups. Standard avidin-biotin affinity protocols could be used to immobilize antibodies at these surfaces, allowing detection of antigens such as myelin basic protein (Tsekenis et al. 2008). Further work showed that affinity-based protocols gave much lower detection limits (2–3 orders of magnitude) than simple entrapment procedures (Barton et al. 2009). Within the same work it was shown that using a sonochemically fabricated microarray of polyaniline microelectrodes also gave much lower detection limits than simple planar polyaniline electrodes. Combination of the two methods demonstrated a synergistic effect with detection limits as low as 1 pg · mL− 1 being obtained for the cancer marker prostate specific antigen (Barton et al. 2008a) and the stroke marker protein neuron-specific enolase (Barton et al. 2008b), even in the presence of large excesses of potentially interfering proteins.

9.7 Future trends

This is a brief section just to highlight several potential future applications which, although not in the current marketplace, are the subjects of intensive research efforts.

9.7.1 Point-of-care tests

Although the biosensor market is dominated by glucose testing applications, there is great interest in developing simple tests for a wide variety of medical problems which can be easily applied by the patient and detect physical problems which are beginning to occur whilst still in an early stage, thereby enabling the patient to seek early medical help. An example of this was the Multisense® system, initially developed by Oxford Biosensors, which detected early markers of coronary heart disease. Heart disease is a major killer in the western world and the risk of developing this condition is known to be related to factors such as diet, smoking, weight and high blood cholesterol.

The Multisense® systems was based on a hand-held device with disposable electrochemical test strips. Each strip contains several sensors which individually sense for cholesterol (high density and low density lipoprotein) and triglycerides. Testing utilises a single drop of blood in the same manner as a glucose sensor, with the intent of fast and simple screening, diagnosis and monitoring of patients who are at risk of heart disease. Unfortunately Oxford Biosensors went into receivership in 2009 and no product of this type has as yet come onto the market.

9.7.2 Artificial pancreas

For many diabetes type 1 sufferers, life is a constant round of glucose testing and insulin injections. Externally worn insulin pumps have been developed, such as those by Medtronic (www.minimed.com). These have the advantage that they remove the need for injections by introducing insulin subcutaneously though a cannula. When coupled together with a glucose biosensor, there is then the potential for the device to inject insulin ‘on demand’. This was the first device of its type to receive FDA approval (Newman et al. 2004). This has led to the development of such products as the MiniMed Paradigm® Revel™ combined glucose sensor and insulin pump, which can automatically deliver insulin in response to glucose levels. At present, current commercial devices usually can only be used for a few days at a time without maintenance and injection, and sensing sites need to be changed regularly. An implanted device which would automatically respond to changes in glucose levels and display longer-term stabilities would in effect act as an artificial pancreas and improve the quality of life for many millions of people.

Recently the first clinical trial of an wirelessly controlled implantable microchip-based drug delivery device that releases the human parathyroid hormone fragment [hPTH(1–34)], used in the treatment of osteoporosis, showed that the device could deliver up to 20 adminstrations of the drug in vivo rather than the standard practice of daily injection (Farra et al. 2012).

9.7.3 Warfarin monitors

Warfarin is routinely prescribed as an anti-coagulation agent for patients with an increased tendency for thrombosis or as prophylaxis in those individuals who have already formed a blood clot (thrombus) which required treatment. However, warfarin when given in too high a dose inhibits natural clotting, leaving the patient subject to uncontrolled bleeding, both externally and as gastrointestinal bleeding. At present, patients placed on warfarin require their blood to be regularly tested to ensure the clotting ability of their blood falls within required ranges (Poller et al. 2003). This requires withdrawal of blood and testing within a laboratory environment. Since warfarin is one of the world’s most proscribed drugs, a simple home test which could monitor warfarin levels or alternatively monitor blood clotting ability would have a huge global market.

9.8 Conclusions

Drug delivery spans a wide variety of fields such as medicine, chemistry, biology and materials science. As fields such as genomics, proteomics and immunology progress, we will see these approaches being utilised to target drugs to specific sites with greater specificity. Further investigation into the transport of drugs across obstacles such as the blood–brain or air–blood barriers will enable development of simplified delivery systems for these compounds.

There are a wide number of potential techniques that would prove highly beneficial to medical practitioners and are currently being widely researched, some of which are listed below.

• Biodegradable nanoparticles and methods for producing them.

• More complex controlled release characteristics, for example in response to changing physiological conditions.

• Delivery of large quantities of drugs in controlled and highly localised manner, for example to destroy a tumour without harming surrounding healthy tissue.

• Biodegradable coatings for implants, for example, for the release of agents that minimise rejection and/or promote healthy cell growth.

• Functionalising the delivery system to enable selective targeting of sites via immunochemical response.

There are many future opportunities in this field, with one of the most exciting being the application of nanotechnology to drug delivery. The potential advantages associated with nanoparticles such as easy tailoring of properties such as size, often easy transition across cell membranes and potential site specificity ensures these approaches will be intensively studied. Problems to be overcome include the fact that these materials can be cytotoxic and often display poor stability in biochemical environments (Kaparissides et al. 2006). However, these are challenges to be overcome and it can be seen that with the advances taking place in chemistry, materials, biology and other sciences that sooner or later they are likely to be achieved.

As can be seen, biosensors are a rapidly expanding field of research. Historically the market was dominated by glucose sensing which is still the largest single sector. Many other applications such as cholesterol monitoring, monitors for various drug treatments and many other analytes are being intensively researched. Many of these applications are now making the journey from the laboratory to field and commercial applications.

AC impedance biosensing approaches are potentially leading methods for utilisation within point-of-care devices capable of detecting a range of clinical biomarkers. Electrochemical detection methods such as AC can be small and portable, utilise inexpensive single-shot screen-printed electrodes and give rapid results with relatively simple operating protocols due to their ability to be used in a label-free format. For usable devices it is more likely that simultaneous analysis of a clinical sample for a number of targets will be required. We predict that much future work will be devoted to developing small, low cost and low power devices that will be capable of multiple analyses whilst requiring minimal sample pre-treatment.

9.9 References

1. Barisci JN, Conn C, Wallace GG. Conducting polymer sensors. Trends Polym Sci. 1996;4:301–311.

2. Barton AC, Collyer SD, Davis F, et al. Sonochemically fabricated microelectrode arrays for biosensors offering widespread applicability Part I. Biosens Bioelec. 2004;20:328–337.

3. Barton AC, Davis F, Higson SPJ. Labeless immunosensor assay for prostate specific antigen with pg mL− 1 limits of detection based upon an AC impedance protocol. Anal Chem. 2008a;80:6198–6205.

4. Barton AC, Davis F, Higson SPJ. Labeless immunosensor assay for the stroke marker protein neuron specific enolase based upon an AC impedance protocol. Anal Chem. 2008b;80:9411–9416.

5. Barton AC, Collyer SD, Davis F, et al. Labeless AC impedimetric antibody based sensors with pg mL− 1 sensitivities for point-of-care biomedical applications. Biosens Bioelec. 2009;24:1090–1095.

6. Bhadra D, Bhadra S, Jain S, Jain NK. A PEGylated dendritic nanoparticulate carrier of fluorouracil. Int J Pharm. 2003;257:111–124.

7. Bharali DJ, Klejbor I, Stachowiak EK, et al. Organically modified silica nanoparticles: a nonviral vector for in vivo gene delivery and expression in the brain. Proc Natl Acad Sci USA. 2005;102:11539–11544.

8. Bhattarai N, Ramay HR, Gunn J, Matsen FA, Zhang M. PEG-grafted chitosan as an injectable thermosensitive hydrogel for sustained protein release. J Control Rel. 2005;103:609–624.

9. Bindra D, Zhang Y, Wilson G. Design and in vitro studies of a needle type glucose sensor for subcutaneous monitoring. Anal Chem. 1991;63:1692–1696.

10. Bose S, Tarafder S. Calcium phosphate ceramic systems in growth factor and drug delivery for bone tissue engineering: a review. Acta Biomat. 2012;8:1401–1421.

11. Brem H, Piantadosi S, Burger PC, et al. Placebo-controlled trial of safety and efficacy of intraoperative controlled delivery by biodegradable polymers of chemotherapy for recurrent gliomas. Lancet. 1995;345:1008–1012.

12. Casadei MA, Pitarresi G, Calabrese R, Paolicelli P, Giammona G. Biodegradable and pH-sensitive hydrogels for potential colon-specific drug delivery: characterization and in vitro release studies. Biomacromolecules. 2008;9:43–49.

13. Cass AEG, Davis G, Francis GD, et al. Ferrocene-mediated enzyme electrode for amperometric determination of glucose. Anal Chem. 1984;56:667–671.

14. Chen L, Remondetto GL, Subirade M. Food-based protein materials as nutraceutical delivery systems. Food Sci Tech. 2006;17:272–283.

15. Chertok B, David AE, Yang VC. Polyethyleneimine-modified iron oxide nanoparticles for brain tumor drug delivery using magnetic targeting and intracarotid administration. Biomaterials. 2010;31:6317–6324.

16. Cho CS, Han SY, Ha JH, Kim SH, Lim DY. Clonazepam release from bioerodible hydrogels based on semi-interpenetrating polymer networks composed of poly(epsilon-caprolactone) and poly(ethylene glycol) macromer. Int J Pharm. 1999;181:235–242.

17. Clark AH, Kavanagh GM, Ross-Murphy SB. Globular protein gelation – theory and experiment. Food Hydrocolloids. 2001;15:383–400.

18. Clark L, Lyons C. Electrode systems for continuous monitoring in cardiovascular surgery. Ann NY Acad Sci. 1962;102:29–45.

19. Cooper JC, Hall EAH. Electrochemical response of an enzyme-loaded polyaniline film. Biosens Bioelec. 1992;7:473–485.

20. Daniels JS, Pourmand N. Label-free impedance biosensors: opportunities and challenges. Electroanalysis. 2007;19:1239–1257.

21. Dash TK, Konkimalla VB. Poly-epsilon-caprolactone based formulations for drug delivery and tissue engineering: a review. J Cont Release. 2012;158:15–33.

22. Davis F, Higson SPJ. Structured thin films as components in biosensors. Biosens Bioelec. 2005;21:1–20.

23. Davis F, Higson SPJ. Carrier systems and biosensors for biomedical applications. Tissue Engineering using Ceramics and Polymers Cambridge: Woodhead Publishing Limited; 2007.

24. Degani Y, Heller A. Direct electrical communication between chemically modified enzymes and metal-electrodes 1 Electron-transfer from glucose-oxidase to metal-electrodes via electron relays, bound covalently to the enzyme. J Phys Chem. 1987;91:1285–1289.

25. Desai MP, Labhasetwar V, Walter E, Levy RJ, Amidon GL. The mechanism of uptake of biodegradable microparticles in Caco-2 cells is size dependent. Pharm Res. 1997;14:1568–1573.

26. Dhanikula RS, Argaw A, Bouchard JF, Hildgen P. Methotrexate loaded polyether-copolyester dendrimers for the treatment of gliomas: enhanced efficacy and intratumoral transport capability. Mol Pharmaceut. 2008;5:105–116.

27. D’Orazio P. Biosensors in clinical chemistry. Clin hem Acta. 2003;334:41–69.

28. Eggins BR. Biosensors Chichester: Wiley; 1996.

29. Ekenseair A, Kasper F, Mikos A. Perspectives on the interface of drug delivery and tissue engineering. Adv Drug Deliv Rev. 2013;65:89–92.

30. El-Sayed M, Rhodes CA, Ginski M, Ghandehari H. Transport mechanism(s) of poly(amidoamine) dendrimers across Caco-2 cell monolayers. Int J Pharm. 2003;265:151–157.

31. Fadel M, Kassab K, Fadeel DA. Zinc phthalocyanine-loaded PLGA biodegradable nanoparticles for photodynamic therapy in tumor bearing mice. Lasers Med Sci. 2010;25:283–292.

32. Farra R, Sheppard NF, McCabe L, et al. First-in-human testing of a wirelessly controlled drug delivery microchip. Sci Transl Med. 2012;4:122ra21.

33. Fredenberg S, Wahlgren M, Reslow M, Axelsson A. The mechanisms of drug release in poly(lactic-co-glycolic acid)-based drug delivery systems – a review. Int J Pharmaceut. 2012;415:34–52.

34. Gander B, Meinel L, Walter E, Merkle HP. Polymers as a platform for drug delivery: reviewing our current portfolio on poly (lactide-co-glycolide) (PLGA) Microspheres. Chimia. 2001;55:212–217.

35. Ganji F, Vashedghani-Farahani E. Hydrogels in controlled drug delivery systems. Iran Polym J. 2009;18:63–88.

36. Gerard M, Chaubey A, Malhotra BD. Application of conducting polymers to biosensors. Biosens Bioelec. 2002;17:345–359.

37. Ghahremankhani AA, Dorkoosh F, Dinarvand R. PLGA-PEG-PLGA tri-block copolymers as in situ gel-forming peptide delivery system: effect of formulation properties on peptide release. Pharm Develop Technol. 2008;13:49–55.

38. Gombotz WR, Pettit D. Biodegradable polymers for protein and peptide drug-delivery. Bioconjugate Chem. 1995;6:332–351.

39. Grant S, Davis F, Law KA, et al. A reagentless immunosensor for the detection of BSA at platinum electrodes by an AC impedance protocol. Anal Chim Acta. 2005;537:163–168.

40. Gupta U, Agashe HB, Asthana A, Jain NK. Dendrimers: novel polymeric nanoarchitectures for solubility enhancement. Biomacromolecules. 2006;7:649–658.

41. Guzman LA, Labhasetwar V, Song CX, et al. Local intraluminal infusion of biodegradable polymeric nanoparticles – a novel approach for prolonged drug delivery after balloon angioplasty. Circulation. 1996;94:1441–1448.

42. Ha DI, Lee SB, Chong MS, Lee YM, Kim SY, Park YH. Preparation of thermo-responsive and injectable hydrogels based on hyaluronic acid and poly(N-isopropylacrylamide) and their drug release behaviors. Macromol Res. 2006;14:87–93.

43. Habraken WJ, Wolke JG, Jansen JA. Ceramic composites as matrices and scaffolds for drug delivery in tissue engineering. Adv Drug Deliv Rev. 2007;59:234–248.

44. Hajos F, Stark B, Hensler S, et al. Inhalable liposomal formulation for vasoactive intestinal peptide. Int J Pharm. 2008;357:286–294.

45. Hall EAC. Biosensors Maidenhead: Open University Press; 1990.

46. Hansen J, Jensen KT, Follmann F, et al. Liposome delivery of Chlamydia muridarum major outer membrane protein primes a Th1 response that protects against genital chlamydial infection in a mouse model. J Infect Disease. 2008;198:758–767.

47. Hashiguchi Y, Sakakida M, Nishida K, Uemura T, Kajiwara K, Shichiri M. Development of a miniaturized glucose monitoring system by combining a needle-type glucose sensor with microdialysis sampling method. Diabetes Care. 1994;17:387–396.

48. Henry C. Getting under the skin. Anal Chem. 1998;70:594A–598A.

49. Iwamoto T. Clinical application of drug delivery systems in cancer chemotherapy: review of the efficacy and side effects of approved drugs. Bio Pharm Bull. 2013;36:715–718.

50. Jain TK, Morales MA, Sahoo SK, Leslie-Pelecky DL, Labhasetwar V. Iron-oxide nanoparticles for sustained delivery of anticancer agents. Mol Pharm. 2005;2:194–205.

51. Jain AK, Chalasani KB, Khar RK, et al. Muco-adhesive multivesicular liposomes as an effective carrier for transmucosal insulin delivery. J Drug Target. 2007;15:417–427.

52. Jain K, Kesharwani P, Gupta U, Jain NK. A review of glycosylated carriers for drug delivery. Biomaterials. 2012;33:4166–4186.

53. Jansson PA, Fowelin J, Smith U, Lonnroth R. Characterization by microdialysis of intercellular glucose level in subcutaneous tissue in humans. Am J Physiol. 1988;255:E218–E220.

54. Jeong B, Bae YH, Lee DS, Kim SW. Biodegradable block copolymers as injectable drug-delivery systems. Nature. 1997;388:860–862.

55. Kamada H, Tsutsumi Y, Yoshioka Y, et al. Design of a pH-sensitive polymeric carrier for drug release and its application in cancer therapy. Clin Cancer Res. 2004;10:2545–2550.

56. Kaminskas LM, McLeod V, Kelly BD, et al. A comparison of changes to doxorubicin pharmacokinetics, antitumor activity, and toxicity mediated by PEGylated dendrimer and PEGylated liposome drug delivery systems. Nanomed Nanotech Biol Med. 2012;8:103–111.

57. Kaparissides C, Alexandridou S, Kotti K, Chaitidou S. Recent Advances in Novel Drug Delivery Systems. J Nanotech Online 2006. doi 10.2240/azojono0111.

58. Kim J, Singh N, Lyon LA. Label-free biosensing with hydrogel microlenses. Angew Chem Int Ed. 2006;45:1446–1449.

59. Kirpotin D, Park JW, Hong K, et al. Sterically modified anti-HER2 immunoliposomews: design and targeting to human breast cancer cells in vitro. Biochemistry. 1997;36:66–75.

60. Kojima C, Kono K, Maruyama K, Takagishi T. Synthesis of polyamidoamine dendrimers having poly(ethylene glycol) grafts and their ability to encapsulate anticancer drugs. Bioconjug Chem. 2000;11:910–917.

61. Kolhe P, Misra E, Kannan RM, Kannan S, Lieh-Lai M. Drug complexation, in vitro release and cellular entry of dendrimers and hyperbranched polymers. Int J Pharm. 2003;259:143–160.

62. Kumar A, Sahoo B, Montpetit A, Behera S, Lockey RF, Mohapatra SS. Development of hyaluronic acid-Fe2O3 hybrid magnetic nanoparticles for targeted delivery of peptides. Nanomedicine. 2007;3:132–137.

63. Langer R. Biomaterials and Biomedical Engineering. Chem Eng Sci. 1995;50:4109–4121.

64. Langer R. Drug delivery and targeting. Nature. 1998;392(suppl):5–10.

65. Lasic DD, Vallner JJ, Working PK. Sterically stabilized liposomes in cancer therapy and gene delivery. Curr Opin Mol Ther. 1999;1:177–185.

66. Law B, Weissleder R, Tung CH. Peptide-based biomaterials for proteaseenhanced drug delivery. Biomacromolecules. 2006;7:1261–1265.

67. Lee WC, Li YC, Chu IM. Amphiphilic poly(d, l-lactic acid)/poly(ethylene glycol)/poly(d, l-lactic acid) nanogels for controlled release of hydrophobic drugs. Macromol Biosci. 2006a;6:846–854.

68. Lee CC, Gillies ER, Fox ME, et al. A single dose of doxorubicin-functionalized bow-tie dendrimer cures mice bearing C-26 colon carcinomas. Proc Nat Acad Sci USA. 2006b;103:16649–16654.

69. Liu X, Sun Q, Wang H, Zhang L, Wang J. Microspheres of corn protein zein for an ivermectin drug delivery system. Biomaterials. 2005;26:109–115.

70. Maeda H. The enhanced permeability and retention (EPR) effect in tumor vasculature: the key role of tumor-selective macromolecular drug targeting. Adv Enz Reg. 2001;41:189–207.

71. Maines A, Ashworth D, Vadgama P. Diffusion restricting outer membranes for greatly extended linearity measurements with glucose oxidase enzyme electrodes. Anal Chim Acta. 1997;323:223–231.

72. Malik N, Evagorou EG, Duncan R. Dendrimer–platinate: a novel approach to cancer chemotherapy. Anticancer Drugs. 1999;10:767–776.

73. Maltais A, Remondetto GE, Gonsalves R, Subirade M. Formation of soy protein isolate cold-set gels: protein and salt effects. J Food Sci. 2005;70:67–73.

74. Mano N, Mao F, Heller A. On the parameters affecting the characteristics of the “wired” glucose oxidase anode. J Electroanal Chem. 2005;574:347–357.

75. Markovich RJ, Taylor AK, Rosen J. Drug migration from the adhesive matrix to the polymer film laminate facestock in a transdermal nitroglycerin system. J Pharm Biomed Anal. 1997;16:651–660.

76. Maupas H, Saby C, Martelet C, et al. Impedance analysis of Si/SiO2 heterostructures grafted with antibodies: An approach for immunosensor development. Electroanal Chem. 1996;406:53–58.

77. Mintzer MA, Grinstaff MW. Biomedical applications of dendrimers: a tutorial. Chem Soc Rev. 2011;40:173–190.

78. Mishra GP, Bagui M, Tamboli V, Mitra AK. Recent applications of liposomes in ophthalmic drug delivery. J Drug Deliv. 2011; Article ID 863734.

79. Miyata T, Uragami T, Nakamae K. Biomolecule–sensitive hydrogels. Adv Drug Deliv Rev. 2002;54:79–98.

80. Moghimi SM, Szebeni J. Stealth liposomes and long circulating nanoparticles: critical issues in pharmacokinetics, opsonization and proteinbinding properties. Prog Lipid Res. 2003;42:463–478.

81. Moghimi SM, Hunter AC, Murray JC. Long-circulating and target-specific nanoparticles: theory to practice. Pharmol Rev. 2001;53:283–318.

82. Morgan MT, Nakanishi Y, Kroll DJ. Dendrimer-encapsulated camptothecins: Increased solubility, cellular uptake, and cellular retention affords enhanced anticancer activity in vitro. Cancer Res. 2006;66:11913–11921.

83. Mouriño V, Boccaccini AR. Bone tissue engineering therapeutics: controlled drug delivery in three-dimensional scaffolds. J Roy Soc Interface. 2010;7:209–227.

84. Moussy F, Harrison DJ, O’Brien DW, Rajotte RV. Performance of subcutaneously implanted needle-type glucose sensors employing a novel trilayer coating. Anal Chem. 1993;65:2072–2077.

85. Myerhoff C, Bischof F, Sternberg F, Zier H, Pfeiffer EF. On line continuous monitoring of subcutaneous tissue glucose in men by combining portable glucose sensor with microdialysis. Diabetologia. 1992;35:1087–1092.

86. Myler S, Eaton S, Higson SPJ. Poly(o-phenylenediamine) ultra-thin polymer-film composite membranes for enzyme electrodes. Anal Chim Acta. 1997;357:55–61.

87. Neerman MF, Chen H-T, Parrish AR, Simanek EE. Reduction of drug toxicity using dendrimers based on melamine. Mol Pharmaceut. 2004;1:390–393.

88. Newman JD, Tigwell LJ, Turner APF, Warner PJ. Biosensors: a clearer view. Biosensors 2004 – The 8th World Congress on Biosensors New York: Elsevier; 2004.

89. Ogawa Y, Yamamoto M, Okada H, Yashiki T, Shimamoto T. A new technique to efficiently entrap leuprolide acetate into microcapsules of polylactic acid or copoly(lactic/glycolic) acid. Chem Pharm Bull. 1988;36:1095–1103.

90. Ohara T, Rajagopalan R, Heller A. Wired enzyme electrodes for amperometric determination of glucose or lactate in the presence of interfering substances. Anal Chem. 1994;66:2451–2457.

91. Ouerghi C, Touhami A, Jaffrezic-Renault N, Martelet C, Ben Ouada H, Cosnier S. Electrodeposited biotinylated polypyrrole as an immobilization method for impedimetric immunosensors. IEEE Sens J. 2004;4:559–567.

92. Park JW, Hong K, Kirpotin D, Papahadjopoulos D, Benz CC. Immunoliposomes for Cancer Treatment. Adv Pharmacol. 1997;40:399–435.

93. Park JW, Kirpotin D, Hong K, et al. Anti-HER2 immunoliposomes for Targeted Drug Delivery. Med Chem Res. 1998;8:383–391.

94. Parveen S, Misra R, Sahoo SK. Nanoparticles: a boon to drug delivery, therapeutics, diagnostics and imaging. Nanomed Nanotech Biol Med. 2012;8:147–166.

95. Poller L, Keown M, Chauhan N, et al. Reliability of international normalised ratios from two point of care test systems: comparison with conventional methods. Brit Med J. 2003;327:30–32A.

96. Rosholm KR, Arouri A, Hansen PL, González-Pérez A, Mouritsen OG. Characterization of fluorinated catansomes: a promising vector in drug-delivery. Langmuir. 2012;28:2773–2781.

97. Roy I, Ohulchanskyy TY, Pudavar HE, et al. Ceramic-based nanoparticles entrapping water-insoluble photosensitizing anticancer drugs: a novel drug-carrier system for photodynamic therapy. J Am Chem Soc. 2003;125:7860–7865.

98. Sadeghi M, Hosseinzadeh H. Synthesis of starch-poly(sodium acrylateco-acrylamide) superabsorbent hydrogel with salt and pH-responsiveness properties as a drug delivery system. J Bioact Compat Polym. 2008;23:381–404.

99. Sahoo SK, Labhasetwar V. Nanotech approaches to drug delivery and imaging. Drug Discovery Today. 2003;8:1112–1120.

100. Sahu SK, Maiti S, Maiti TK, Ghosh SK, Pramanik P. Hydrophobically modified carboxymethyl chitosan nanoparticles targeted delivery of paclitaxel. J Drug Target. 2011;19:104–113.

101. Schmaljohann D. Thermo- and pH responsive polymers in drug delivery. Adv Drug Delivery Rev. 2006;58:1655–1670.

102. Serres A, Baudys M, Kim SW. Temperature and pH-sensitive polymers for human calcitonin delivery. Pharm Res. 1996;13:196–201.

103. Si HY, Li DP, Wang TM, et al. Improving the anti-tumor effect of genistein with a biocompatible superpara-magnetic drug delivery system. J Nanosci Nanotechnol. 2010;10:2325–2331.

104. Sokoloski TD, Royer GP. Drug entrapment within native albumin beads. In: Davis SS, Illum L, McVie JG, Tomlinson E, eds. Microspheres and Drug Therapy, Pharmaceutical, Immunological and Medical Aspects. Amsterdam: Elsevier; 2005:295–307.

105. Song XR, Cai Z, Zheng Y, He G, et al. Reversion of multidrug resistance by co-encapsulation of vincristine and verapamil in PLGA nanoparticles. Eur J Pharm Sci. 2009;37:300–305.

106. Su J, Chen F, Cryns VL, Messersmith PB. Catechol polymers for pH-responsive, targeted drug delivery to cancer cells. J Am Chem Soc. 2011;133:11850–11853.

107. Sultana S, Khan MR, Kumar M, Kumar S, Ali M. Nanoparticles-mediated drug delivery approaches for cancer targeting: a review. J Drug Target. 2013;21(107):125.

108. Svenson S, Tomalia DA. Commentary – dendrimers in biomedical applications – reflections on the field. Adv Drug Delivery Rev. 2005;57:2106–2129.

109. Taiyou Research. Market for Drug Delivery Systems in the US South Korea: Taiyou Research; 2011.

110. Taylor RF, Marenchic IG, Spencer RH. Antibody-based and receptor-based biosensors for detection and process-control. Anal Chim Acta. 1991;249:67–70.

111. Teijon JM, Trigo RM, Garcia O, Blanco MD. Cytarabine trapping in poly (2-hydroxyethyl methacrylate) hydrogels: drug delivery studies. Biomaterials. 1997;18:383–388.

112. Thusu R. http://www.sensorsmag.com/specialty-markets/medical/strong-growth-predicted-biosensors-market-7640; 2010; Accessed 14th Feb 2012.

113. Tomlinson E, Burger JJ. Incorporation of water soluble drugs in albumin microspheres. In: Widder J, Green R, eds. New York: Academic; 1985:27–43. Methods in Enzymology.

114. Tsekenis G, Garifallou G, Davis F, et al. Labeless Immunosensor assay for myelin basic protein based upon an AC impedance protocol. Anal Chem. 2008;20:2058–2062.

115. Turner RBF, Sherwood CS. Biocompatibility of perfluorosulfonic acid polymer membranes for biosensor applications. ACS Symp Ser. 1994;556:211–221.

116. Updike S, Hicks G. Electrode systems for continuous monitoring in cardiovascular surgery. Nature. 1967;214:986.

117. Vyas SP, Kannan ME, Jain S, Mishra V, Singh P. Design of liposomal aerosols for improved delivery of rifampicin to alveolar macrophages. Int J Pharmaceut. 2004;269:37–49.

118. Wang J. Glucose biosensors: 40 years of advances and challenges. Electroanalysis. 2001;13:983–988.

119. Wang J, Tao X, Zhang Y, Wei D, Ren Y. Reversion of multidrug resistance by tumor targeted delivery of antisense oligodeoxynucleotides in hydroxypropylchitosan nanoparticles. Biomaterials. 2010a;31:4426–4433.

120. Wang ZH, Wang ZY, Sun CS, Wang CY, Jiang TY, Wang SL. Trimethylated chitosan-conjugated PLGA nanoparticles for the delivery of drugs to the brain. Biomaterials. 2010b;31:908–915.

121. Werle M, Takeuchi H. Chitosan-aprotinin coated liposomes for oral peptide delivery: development, characterisation and in vivo evaluation. Int J Pharmaceut. 2009;370:26–32.

122. Yalow RS, Berson SA. Assay of plasma insulin in human subjects by immunological methods. Nature. 1959;184:1648–1649.

123. Yang H, Li K, Liu Y, Liu ZH, Miyoshi H. Poly(D, L-lactide-coglycolide) nanoparticles encapsulated fluorescent isothiocyanate and paclitaxol: preparation, release kinetics and anticancer effect. J Nanosci Nanotech. 2009;9:282–287.

124. Zhang L, Gu FX, Chan JM, Wang AZ, Langer RS, Farokhzad OC. Nanoparticles in medicine: therapeutic applications and developments. Clin Pharma Therapeut. 2008;83:761–769.

125. Zhang Y, Thomas TP, Desai A, et al. Targeted dendrimeric anticancer prodrug: a methotrexate-folic acid-poly(amidoamine) conjugate and a novel, rapid, “one pot” synthetic approach. Bioconjugate Chem. 2010;21:489–495.

126. Zhu S, Hong M, Zhang L, Tang G, Jiang Y, Pei Y. PEGylated PAMAM dendrimer-doxorubicin conjugates: in vitro evaluation and in vivo tumor accumulation. Pharm Res. 2010;27:161–174.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset