You are on page 1of 30

Dalton Transactions

An international journal of inorganic chemistry


www.rsc.org/dalton Volume 40 | Number 24 | 28 June 2011 | Pages 63016576

ISSN 1477-9226

COVER ARTICLE Tremel et al. Synthesis and bio-functionalization of magnetic nanoparticles for medical diagnosis and treatment

Dalton Transactions
Cite this: Dalton Trans., 2011, 40, 6315 www.rsc.org/dalton

Dynamic Article Links

PERSPECTIVE

Synthesis and bio-functionalization of magnetic nanoparticles for medical diagnosis and treatment
Downloaded on 20 June 2011 Published on 01 March 2011 on http://pubs.rsc.org | doi:10.1039/C0DT00689K

Thomas D. Schladt,a Kerstin Schneider,a Hansj rg Schildb and Wolfgang Tremel*a o


Received 21st June 2010, Accepted 13th January 2011 DOI: 10.1039/c0dt00689k The synthesis of multifunctional magnetic nanoparticles (NPs) is a highly active area of current research located at the interface between materials science, biotechnology and medicine. By virtue of their unique physical properties magnetic nanoparticles are emerging as a new class of diagnostic probes for multimodal tracking and as contrast agents for MRI. Furthermore, they show great potential as carriers for targeted drug and gene delivery, since reactive agents, such as drug molecules or large biomolecules (including genes and antibodies), can easily be attached to their surface. On the other hand, the fate of the nanoparticles inside the body is mainly determined by the interactions with its local environment. These interactions strongly depend upon the size of the magnetic NPs but also on the individual surface characteristics, like charge, morphology and surface chemistry. This review not only summarizes the most common synthetic approaches for the generation of magnetic NPs, it also focuses on different surface modication strategies that are used today to enhance the biocompatibility of these NPs. Finally, key considerations for the application of magnetic NPs in biomedicine, as well as various examples for the utilization in multimodal imaging and targeted gene delivery are presented.

Introduction
Today, the development of nanomaterials has moved beyond the discovery of totally new materials and compositions. Instead the focus has shifted to the investigation of more complex, composite systems, in which the recombination of known materials into structures of higher complexity opens new possibilities of functionality.15 While scientic diligence of designing new composite or hybrid materials is speeding up, industrial producers have begun merchandising the earliest discovered nanomaterials, and, in fact, are developing novel applications for them to t the desired needs. Besides other commercial applications, magnetic nanoparticles (NPs) are intensively being investigated for utilization in many different scientic and industrial elds, ranging from catalysis to mass data storage.6,7 One of the fastest moving and most exciting research areas is the interface between nanotechnology, biology and medicine. It has been stated by numerous experts, that the application of nanotechnology in medicine, which is often referred to as nanomedicine, offers many exciting possibilities for healthcare in the future, and may revolutionize the areas of targeted drug delivery, disease detection and tissue engineering.811 An often used catch phrase in this context is theragnostics, which, as this
a Institut f r Anorganische Chemie und Analytische Chemie, Johannes u Gutenberg-Universit t, Duesbergweg 10-14, D-55099 Mainz, Germany. a E-mail: tremel@uni-mainz.de; Fax: +49 6131 39-25605; Tel: +49 6131 39-25135 b Institut f r Institut f r Immunologie, Johannes Gutenberg-Universit t, u u a Obere Zahlbacher Str. 67, D-55131 Mainz, Germany

word construction implies, is the combination of both, therapy and diagnostics into one powerful tool. In fact, the concept of using magnetic nanoparticles to target tumor cells inside the human body, and applying them to treat cancer, was rst conceived in the late 1970s.12 The key idea was to attach common anticancer drugs to small magnetic spheres outside the body before administering them to the patient. After injection into the blood stream strong external magnetic elds should concentrate the drug-loaded particles inside the tumor tissue. The authors predicted that by this approach the drug payload would be reduced signicantly, and thereby the unwanted side effects associated with the systemic distribution of chemotherapeutic agents, including nausea, hair loss and a compromised immune system could be avoided. Although not yet fully in clinical use, nanomedicine has come a long way from these initial ideas, and it is proceeding with remarkable speed. Concerning diagnostics, the ability to accelerate the proton relaxation of water molecules in different tissues has been proposed to be one of the most promising features of magnetic NPs for nanomedicine, since it allows the development of novel contrast agents for magnetic resonance imaging (MRI). In fact, contrast agents based on superparamagnetic iron oxide NPs (SPIONs) have been in clinical use for almost two decades. Moreover, progress in the synthesis of magnetic NPs permits a precise tuning of their physical properties and, as a result, has led to tremendous improvements in their MRI performance. On the other hand, owing to their size, nanoparticles can penetrate cell walls13 and deliver biomolecules or drugs for therapeutic14 and diagnostic purposes.9 The use of Dalton Trans., 2011, 40, 63156343 | 6315

This journal is The Royal Society of Chemistry 2011

Downloaded on 20 June 2011 Published on 01 March 2011 on http://pubs.rsc.org | doi:10.1039/C0DT00689K

nanoparticulate (polymeric or inorganic) carriers allows the transport of hydrophobic low molecular weight drugs, to enhance their efciency. Additionally, a feature that proves to be benecial for the application of nanoparticulate systems is the fact that the blood vessels sustaining tumor tissue exhibit comparatively large fenestrations, and the lymph system of tumors is poorly operational. Therefore, macromolecules and nanoparticles leaking from these blood vessels can accumulate inside the tumor tissue, a phenomenon known as the EPR (enhanced permeability and retention) effect (see Fig. 1).15,16 However, one of the most prevailing issues associated with the application of these NP systems, is their behavior in vivo. For instance, the recognition and clearance by the reticuloendothelial system (RES) is a major obstacle since it reduces the circulation times of the NPs within the blood stream, and therefore, obstructs a site specic accumulation of the administered NP probes at designated areas (see Fig. 1).17 The enhanced uptake in the liver, spleen, and bone marrow is largely attributed to the macrophages residing in the tissues, which are responsible for clearing particulates and macromolecules circulating in the blood. When nanoparticles are administered intravenously, a variety of serum proteins bind to the surface of the nanoparticles, which are recognized by the scavenger receptor on the macrophage cell surface and internalized, leading to a signicant loss of nanoparticles from the circulation.18 The serum proteins binding on the nanoparticles are termed opsonins, and the macrophages contributing the major loss of injected dose are constituents of the RES or mononuclear phagocyte system (MPS). Minimizing protein binding is the key point for developing a long circulation nanoparticle formulation.

their surface properties, including morphology, charge and surface chemistry. As a consequence, efcient strategies for the synthesis of magnetic nanoparticles and their subsequent surface modication are necessary to meet the challenges of a later application in biomedicine. This review article summarizes some recent achievements in the rapidly evolving elds of nano-biotechnology and nanomedicine. It consists of three parts; in the rst section the properties and synthesis of magnetic NPs will be discussed briey before in the second part some of the most common surface modication strategies are presented. Finally, the last section addresses the application of magnetic NPs in biomedicine.

1. Magnetic properties
Magnetic nanoparticles are among the most investigated nanomaterial systems, owing to the fact that their magnetic properties dramatically depend upon their size and their morphology. Several issues are responsible for these unique properties: nite size effects, which result from the quantum connement of the electrons inside the material, and surface effects caused by symmetry breaking of the crystal structure at the boundaries of the particles.2023 Concerning nite-size effects, the magnetic properties of (ferro-) magnetic nanoparticles are dominated by two key features: (1) The single domain limit and (2) the superparamagnetic limit, which both lead to individual material-dependent length scales, i.e. the single domain size and the superparamagnetic size. Both features will be discussed briey. 1.1. Single-domain-limit A ferromagnetic bulk material usually consists of many separate areas, in which all magnetic moments of the constituent atoms are pinned in the same direction. The reason for such an arrangement arises from the fact that the magnetostatic energy (DE MS ) of the materials is lowered once a large domain is broken up into several smaller domains. However, the formation of new domain walls requires energy (E D ), and therefore, there is a size limit, below which the energy needed for the creation of a smaller domain exceeds the amount of energy gained from decreasing the magnetostatic energy. Consequently, this means that a magnetic nanoparticle with a diameter comparable to, or lower than the size of the smallest possible magnetic domain, could only consist of a single domain, which, in turn, results in a narrowing of the magnetic hysteresis curve compared to the bulk material (see Fig. 2).22,24,25 1.2. Superparamagnetic limit To understand the super-paramagnetic effect the magnetic anisotropy energy per particle (E(H)) which pins the magnetic moments of a single domain particle in a certain direction, has to be considered. E(H) is proportional to K a V (V is the particle volume), the energy barrier which stops the magnetic moment ipping from one direction to the opposite. K a V is usually much higher than the thermal energy kB T, however, with decreasing particle size K a V decreases to values equal to, or below kB T. As a result, the magnetic moments are able to overcome the energy barrier and freely ip in any direction, i.e. for kB T K a V the This journal is The Royal Society of Chemistry 2011

Fig. 1 Passive tumor targeting due to the EPR effect. Larger particles and agglomerates are rapidly attacked by phagocytes, whereas smaller particles can travel longer through the blood vessels to reach the target tissue. Once at the tumor site, the magnetic NPs accumulate inside the tumor tissue due to the EPR effect.

In addition to that, controlling the overall size of the NPs is crucial, since particles with a mean diameter below 5 nm are usually eliminated by renal excretion, whereas larger particles (>100 nm) are taken up easily by macrophages.19 Furthermore, the ultimate fate of nanoparticles within the body is determined by the interaction with the local environment inside biological systems, which depends not only on the particle size but also on
6316 | Dalton Trans., 2011, 40, 63156343

existence of canted surface spins.21 Additionally, Bdker et al. reported, that the magnetic anisotropy K a of iron NPs increases with decreasing particle size, due to a higher contribution of the surface anisotropy K aS .26 As an example, antiferromagnetic NiO and MnO nanoparticles exhibit increasing net magnetization values and higher magnetic blocking temperatures with decreasing particle size, the reason being the larger number of uncompensated surface spins.20,27,28
Downloaded on 20 June 2011 Published on 01 March 2011 on http://pubs.rsc.org | doi:10.1039/C0DT00689K

1.4. Magnetic NPs as contrast agents for MRI MR imaging is one of the most powerful non-invasive imaging techniques in clinical use today.29,30 Just as NMR spectroscopy, it is based on measuring the relaxation of protons in an external magnetic eld after they have been excited with a radio-frequency pulse. Basically, there are two different types of relaxation depending on the nature of the corresponding interactions: T 1 : Longitudinal relaxation. After excitation with a 90 radio frequency pulse the magnetization in the eld direction M z is zero and the perpendicular in-plane magnetization M xy is maximal. Over time, the magnetization returns into the eld direction and M xy decreases. This process is associated with a release of energy to the environment (the lattice), therefore, it is also called spinlattice (or longitudinal) relaxation. The time constant, at which the energy is depleted (and M z increases) T 1 , depends on different factors, including eld strength, or the nature of the observed tissue.30 T 2 : Transverse relaxation. The loss of the transverse magnetization M xy is also associated with a dephasing of the spins precessing in the xy-plane. Directly after excitation all spins are in phase and M xy is maximal. In principle, two components determine the dephasing process: On the one hand, an energy exchange between the spins induced by local magnetic eld interferences among the spins themselves (spinspin or transverse relaxation), leads to a faster and slower precession in the xy-plane. The time constant T 2 , with which M xy is depleted, is therefore independent of the applied magnetic eld. On the other hand, inhomogeneities of the external magnetic eld, arising from the magnetic coils of the MRI scanner itself or the body of the patient, lead to an additional dephasing and a faster decay of M xy with a time constant T 2 *.29,30 The contrast in MR images is mainly determined by three different features: the proton density, the T 1 time and the T 2 time of the regarding tissue. While the proton density is given by the chemical and physical nature of the tissue, the use of contrast agents can greatly enhance both T 1 and T 2 contrast. Basically it is desired to shorten the relaxation times because this leads to a greater difference in signal intensity and thus to a better contrast in the MR image. In T 1 -weighted images, the signal intensity relies on how fast M z is restored. After a given time period t following proton excitation, M z has only recovered to a value M z if no contrast agent is applied. On the other hand, in the presence of a paramagnetic substance the excess proton energy is released more efciently and therefore, the magnetization in eld direction reaches a higher value M z (M z >M z ). As a result, T 1 enhancement leads to a brighter (positive) contrast in the image. On the other hand, in T 2 -weighted images the signal intensity Dalton Trans., 2011, 40, 63156343 | 6317

Fig. 2

Illustration of the single domain limit.

system behaves like a paramagnet. Since the individual atomic moments add up to one (super) moment for every particle, this phenomenon is called superparamagnetism. On the other hand, when the temperature is lowered for a given superparamagnetic particle, the thermal energy of the particle decreases until kB T is lower than the energy barrier K a V (kB T K a V ), i.e. the particle undergoes a magnetic transition from a superparamagnetic to a blocked state, in which the magnetic moments cannot ip freely. The temperature at which kB T = K a V is therefore called the magnetic blocking temperature T B .22,25 Experimentally, T B can be determined by measuring the magnetization of a nanoparticle sample versus increasing temperature under an applied external magnetic eld. More precisely, in a so called zero-eld-cooled/eld-cooled (ZFC/FC) experiment, the sample is rst cooled down to a temperature well below the expected blocking temperature (usually 5 K). Then, an external eld (generally 100 Oe) is applied and the sample is slowly heated up while simultaneously measuring the magnetization. Since the particles were frozen from room temperature, their magnetic moments are distributed homogeneously, and therefore, the net magnetization is zero. However, as the temperature, increases the particles are able to move their moments in the direction of the external eld. As a result, the net magnetization rises. As mentioned before, above T B the thermal energy exceeds K a V , leading to a decline of the magnetization. As a result, the ZFC curve passes a maximum at T B . 1.3. Surface effects

Surface effects depend directly on the ratio of surface spins to bulk spins. Therefore, surface effects become more pronounced with increasing surface to volume ratio, i.e. decreasing particle size. The symmetry breaking at the particle boundary can lead to variations in the electronic band structure, lattice constant, and atom coordination and, as a direct result, change the magnetic properties.25 Furthermore, surface effects can lead to a decrease of the magnetization of small particles, for instance ferromagnetic oxide nanoparticles, with respect to the bulk value. This reduction has been associated with different mechanisms, such as the existence of a magnetic dead layer on the particle surface, or the This journal is The Royal Society of Chemistry 2011

is determined by the decrease of M xy . Here, the presence of a contrast agent leads to a faster dephasing of the spins that precess in the xy-plane. Consequently, M xy drops faster (M xy <M xy , at a given time t) resulting in a darkening of the corresponding MR image compared to images where the contrast agent is absent. In general, the effect of MRI contrast agents is based on inuencing the resonance properties of the tissue by changing the local magnetic eld inside the body. This change depends on interactions between the protons of the tissue (water, fat, proteins) and the unpaired electrons, i.e. the magnetic moment in the inner sphere (T 1 ) or the outer sphere (T 2 ) of the contrast agent. Consequently, most commercially available contrast agents consist of paramagnetic substances with a large number of unpaired electrons and high magnetic moment (e.g. solutions of metal ions, like Gd3+ , Mn2+ or Fe3+ ). On the other hand, superparamagnetic NPs have a much higher magnetic moment (compared to free ions 10010,000-fold), making them promising candidates for contrast enhancing purposes. In fact, iron oxide NPs are already in clinical use as T 2 contrast agents.9 Unlike X-ray contrast agents for computer tomography that rely on the direct absorption of X-ray radiation, the effect of (super)paramagnetic substances in MRI is oblique. For example, paramagnetic T 1 contrast agents withdraw the excess energy after proton excitation, and hence, lead to a faster recovery of M z (see above). However, this effect is strongly related to the ability of water molecules to enter the periphery of the agent. The impact of T 2 agents, on the other hand, arises from the creation of local magnetic elds, which results in increased eld-uctuations, and thus in a faster dephasing of the in-plane spins. As a consequence, T 2 agents need to have a much larger magnetic moment than T 1 agents. The measure, by which the performance of MRI contrast agents is evaluated, is called relaxivity (r1 = 1/T 1 and r2 = 1/T 2 , per mol of agent). Additionally, the ratio of T 2 to T 1 relaxivity (r2 /r1 -ratio) can also be used to estimate the contrast enhancing potential (e.g. T 1 agents require a low r2 /r1 ratio, whereas for T 2 agents it should be large). The degree of performance is strongly dependent on the size, size distribution, composition and crystallinity of the NPs.3133 As already stated in the previous sections the particle size and composition mainly determines the magnetization and, therefore, a narrow size distribution is necessary to efciently exploit the magnetic properties for MRI. Furthermore, it was shown by Weller and coworkers, that the surface characteristics of magnetic NPs (i.e. nature of surface ligands, degree of ligand loading) also play a crucial role for the MRI contrast enhancing abilities.34 The authors demonstrated that the transverse relaxivity, in particular r2 *, is very much higher for micellar coated compared to polymer-functionalized particles using the same size nanoparticles. Although superparamagnetic iron oxide NPs (SPIONs) show a strong T 2 effect, there are numerous reports in which they can also be applied as effective agents for the shortening of the T 1 relaxation time, either by applying special measurement techniques, by reducing the particle diameter, and thus the magnetic moment, or by labeling Fe3 O4 NPs with Gd-complexes.3538 Apart from iron oxide NPs, other nanomaterials are also objects of increasing scientic interest as MRI contrast agents. In this respect, gadolinium-based NPs have been extensively investigated as positive (T 1 ) contrast agents.3942 However, superparamagnetic MnO nanoparticles have
6318 | Dalton Trans., 2011, 40, 63156343

recently attracted considerable attention as positive (T 1 ) contrast agents.4348

2. Synthesis of magnetic nanoparticles


For more than 30 years, the preparation of nanocrystals, i.e. crystalline materials with dimensions in the range between 1 100 nm, has been intensively investigated, not only for their fundamental scientic interest, but also for their many possible technological applications. During that time, numerous strategies have been developed to synthesize them with different compositions and homogeneous size distributions.4952,25 As mentioned in the previous section, nanomaterials exhibit physical and chemical properties that are quite different from their bulk counterparts. However, since these properties are strongly dependent on the dimensions of the nanocrystals, the synthesis of uniformly-sized NPs (i.e. monodisperse with a standard deviation of the size distribution s 5%) is of key importance. The most favorable synthetic approaches usually comprise co-precipitation, hydrothermal synthesis, synthesis in (reverse) micelles, and thermal decomposition and/or reduction of appropriate metalorganic precursors. Simple metal nanoparticles such as iron,5355 cobalt56 and nickel,57 as well as metal oxides, like Fe3 O4 ,5860 g-Fe2 O3 ,53 manganese oxides,6164,52,65,28 cobalt oxides,6669 or NiO,70,27 mixed metal oxides with spinel-type structure, such as MFe2 O4 (M = Mn, Co, Zn),58,7173 or intermetallics such as FePt,7479 NiPt,80,81 and CoPt3 82,83 have been prepared by these methods. 2.1. General considerations A key issue for the preparation of monodisperse magnetic nanoparticles is to understand the basic formation mechanism of nanocrystals in solution. It is in fact surprising, that, although many different synthetic strategies are known today (see above), the comprehensive understanding of nanoparticle generation and the structure of the nuclei are still very limited. However, the main factors that determine this process, i.e. nucleation and growth, will be discussed briey.52,84 The rst research efforts associated with the formation of monodisperse colloidal particles were carried out in the 1940s. LaMer and coworkers revealed, that a short nucleation burst followed by slow controlled growth, without further nucleation, is essential to produce colloids with a narrow size distribution.85,86 Although this model was actually developed for sulfur hydrosols and oil aerosols, it can easily be transferred to nanoparticulate systems. According to the LaMer model, the whole nucleation and growth process can be divided into three phases (see Fig. 3). In the rst phase, the monomer concentration increases gradually until it reaches the point of supersaturation (S). If no seeds (like dust particles or small crystallites) are present, which would initiate heterogeneous nucleation, only homogeneous nucleation is possible. However, since the energy barrier for the initiation of a homogeneous nucleation event is considerably high, the monomer concentration can further increase (phase II). In phase II, the supersaturation nally reaches a critical value (Sc ), which means that the system contains enough energy to overcome the energy barrier (see Fig. 2). Therefore, homogeneous nucleation throughout the entire reaction solution can take place any time. Eventually, a large number of nuclei are formed simultaneously This journal is The Royal Society of Chemistry 2011

Downloaded on 20 June 2011 Published on 01 March 2011 on http://pubs.rsc.org | doi:10.1039/C0DT00689K

Downloaded on 20 June 2011 Published on 01 March 2011 on http://pubs.rsc.org | doi:10.1039/C0DT00689K

Fig. 3 LaMer plot illustrating the separation of nucleation and growth during the synthesis of monodisperse NPs. S and Sc are supersaturation and critical supersaturation, respectively.

in a process that can best be described as a nucleation burst. As a consequence, the monomer concentration drops drastically below a point where no further nucleation is possible. Finally, in phase III, all nuclei grow at the same time, and, since their growth histories are identical, the NPs will end up having an exceptionally narrow size distribution. However, it should be noted at this point, that this theory has been derived for microparticle systems and, although it is feasible to apply most of these considerations to nanoparticle models, caution must be taken and the statements must be judged critically, since the surface free energy is much higher in this size range, and therefore, cannot be considered to be constant.87 In fact, the high ratio of surface atoms to bulk atoms in small NPs leads to an extremely high driving force of these particles to minimize the surface energy, which, consequently, results in uncontrolled growth and particle agglomeration. The most common techniques to prevent these unwanted features use surfactant molecules to stabilize even the smallest NPs in solution. These molecules usually consist of a long hydrophobic hydrocarbon chain with a functional endgroup, that strongly attaches onto the surface of the regarding nanomaterial. Typical functional endgroups include phosphines, amines, thiols and carboxylates, and, as a result, a hydrophobic shell is formed around each particle, providing, not only, protection against oxidation, but also long-term stability in non-polar solvents.50,88,58,64,89,90 2.2. Co-precipitation

Fig. 4 Magnetic NPs synthesized by co-precipitation of metal salts in basic aqueous solutions. (a) Mn-ferrite NPs (with permission from ref. 91), (b) Co-ferrite NPs (with permission from ref. 92) (c) Fe3 O4 NPs (with permission from ref. 93).

One of the most preferred methods to produce magnetite (Fe3 O4 ), maghemite (g-Fe2 O3 ), and mixed ferrite (MFe2 O4 ) nanoparticles is the co-precipitation of iron(II)-, iron(III), and other metal (M(II)) salts (e.g. chlorides, sulfates, nitrates) in aqueous solution by addition of a base (see Fig. 4).88,9395 The products of coprecipitation reactions, especially those performed at or near room temperature, are usually poorly crystalline, or even completely amorphous, requiring a subsequent annealing step, which makes particle agglomeration unavoidable. On the other hand, conducting the co-precipitation reaction at higher temperatures (50 100 C) automatically leads to condensation of the precipitated metal hydroxides to form crystalline metal oxides.91,96 Depending on the metals salts used as starting materials, temperature, the This journal is The Royal Society of Chemistry 2011

Fe2+ /Fe3+ ratio and the pH, particles of different size, morphology, and composition can be produced in a highly reproducible fashion Since Fe3 O4 nanoparticles are subject to oxidation and dissolution in an aqueous environment, controlled oxidation to the more stable maghemite (g-Fe2 O3 ) is often conducted as a follow-up step during nanoparticle synthesis. A better dispersibility in water can be obtained by subsequent acidication of the particle surface with nitric acid or by addition of tartrate ions, making the NPs stable in acidic and basic media over long periods of time.92 The major advantages of this method are, once all parameters are set, the high reproducibility and high yields that can easily exceed multi kilogram amounts. Therefore, it is especially interesting for industrial scale applications.25 The major drawback, however, is a poor control of the size distribution. Since magnetization and blocking temperature strongly depend on nanoparticle size, a narrow size distribution is essential to fully exploit these properties in subsequent applications. Partial control over the size distribution can be achieved by addition of organic stabilizing agents, like polyvinyl pyrrolidone (PVP) or polyvinyl alcohol (PVA), which cover the surface of the nanoparticles during the synthesis. Other authors have reported that the use of various organic salts like sodium citrate or tartrate can control the dispersity of the product (see Fig. 4b).92 The effects of several organic anions, such as carboxylate and hydroxy carboxylate ions, on the formation of iron oxides or oxyhydroxides have been studied extensively.97 However, oleic acid is by far the most often used stabilizing agent for magnetic NPs.98,99 2.3. Solvothermal/Hydrothermal synthesis Solvothermal/hydrothermal strategies have been used to produce a large variety of nanomaterials.100107 The key idea behind these processes is to bring the reaction mixture to temperatures well above the boiling point of the respective solvent. This is achieved by placing the mixture inside a sealed high-pressure reaction vessel (autoclave). In some cases the solvent is heated Dalton Trans., 2011, 40, 63156343 | 6319

above its critical point, where it becomes a supercritical uid. Supercritical uids exhibit very high viscosities and are able to dissolve compounds that have a very low solubility under ambient conditions.104 However, in most cases it is not necessary to heat the solvent above its critical point, because the solubility and reactivity of metal salts and complexes is high enough, even under more moderate subcritical conditions. In any event, solvothermal processing allows many inorganic materials to be prepared at temperatures well below those required by traditional solid-state reactions. Similar to the co-precipitation process, the addition of stabilizing molecules can control the particle morphology and signicantly narrow the size distribution. The main advantage compared to other low-temperature procedures, like coprecipitation and solgel processing, are the very high crystallinity of the resulting nanomaterials. Detailed descriptions of the theory of solvothermal/hydrothermal processes can be found in recent review articles.25,105,108110 The solvothermal synthesis of magnetic metal oxide nanoparticles can be performed starting from different metal precursors. Fe3 O4 nanoparticles with tunable sizes were prepared under hydrothermal conditions using FeCl2 4H2 O and ammonia. Here, the size of the NPs could be controlled by variation of the reaction conditions.111 On the other hand, Lu et al.100 reported the solvothermal synthesis of nearly monodisperse g-Fe2 O3 nanoplates (see Fig. 5a) by reacting a solution of Fe(NO3 )3 9H2 O in absolute ethanol with PVP as stabilizing agent at 240 C. Furthermore, monodisperse NPs of MFe2 O2 (M = Ni, Co, Mn) and g-Fe2 O3 were obtained at the watertoluene interface under conventional and microwave-assisted hydrothermal conditions using metal nitrates and chlorides and oleic acid as capping agent (see Fig. 5d).103 Monodisperse magnetite nanoparticles with an average size of 39 nm were synthesized by co-precipitation of ferrous Fe2+ and ferric Fe3+ ions in water with tetramethylammonium hydroxide (N(CH3 )4 OH) at 70 C followed by hydrothermal treatment at

Downloaded on 20 June 2011 Published on 01 March 2011 on http://pubs.rsc.org | doi:10.1039/C0DT00689K

250 C.106 The use of hydrothermal techniques also allows the formation of unusual morphologies. Nanorings of a-Fe2 O3 , g-Fe2 O3 and Fe3 O4 were prepared by hydrothermal treatment of FeCl3 with certain additives, like Na2 HPO4 and Na2 SO4 , at 220 C. The preformed a-Fe2 O3 nanorings could easily be converted into gFe2 O3 and Fe3 O4 by reduction in a hydrogen/argon atmosphere.107 As a further example, larger nanoparticles of hausmannite (Mn3 O4 ) and nanorods of manganese oxidehydroxide (MnOOH) were synthesized using KMnO4 as precursor in a solution of water and ethanol (see Fig. 5b).91,92,96,101,112,99,97 Among the metal complexes cupferronates are the most frequently used compounds. Ghosh et al.27 reported the successful synthesis of 614 nm MnO and NiO nanoparticles by decomposing manganese cupferronate Mn(C6 H5 N2 O2 )2 and nickel cupferronate Ni(C6 H5 N2 O2 )2 in the presence of trioctylphosphine oxide (TOPO) as capping agent under solvothermal conditions. 2.4. Preparation within micelles Another strategy that is often applied for the synthesis of magnetic nanoparticles uses micelles as nanoreactors in (reverse) microemulsions. By denition, and in contrast to conventional emulsions, microemulsions are thermodynamically stable isotropic dispersions of two immiscible liquids, in which a mono-layered lm of surfactant molecules stabilizes each microdomain of both liquids.113115 A special kind of dispersion is the waterin-oil microemulsion, where the aqueous phase is dispersed as microdroplets (typically 150 nm in diameter) in a non-polar hydrocarbon phase (e.g. cyclohexane). The aqueous droplets themselves are surrounded by a stabilizing monolayer of surfactant molecules. Since the size and size distribution of the synthesized nanoparticles strongly depends on the diameter of the micelles, efcient control over the micelle radius is crucial to obtain high quality nanocrystals. It has been shown in the past, that the size of the reverse microdroplets can be adjusted by varying the molar ratio of water to surfactant:116 w 0 = [H2 O]/[S] With this equation, a straight proportional relationship between the effective micelle radius and the molar water-to-surfactant ratio w 0 is evident. With this understanding, a precise control of the micelle size is possible. In a stirred reverse microemulsion the micelles continuously collide, fuse, and break up again. Therefore, a reaction system containing two or more precursors in individual aqueous droplets will eventually lead to intimate mixing of the reactants and initiation of the reaction.117 The nanoparticle formation itself is usually conducted at room temperature, but it can also be accelerated by heating the organic solvent to reux.118 After completion, the product can be isolated by breaking up the micelles with different solvents such as methanol or acetone followed by ltration or centrifugation. Commonly used surfactant molecules are either ionic, like cetyltrimethylammonium bromide (CTAB) or sodium dodecylbenzenesulfonate (NaDBS), or non-ionic, like Igepal CO-520 or Triton-X 100. In some cases an additional cosurfactant is also applied. A vast variety of different nanoparticles have been synthesized using (reverse) micelles, ranging from metallic or intermetallic materials to quaternary metal oxides. Microemulsions, consisting of octane and hydrazine as reducing agent in water, with CTAB as surfactant and 1-hexanol as co-surfactant, were employed to create This journal is The Royal Society of Chemistry 2011

Fig. 5 Magnetic NPs synthesized by hydro-/solvothermal reactions. (a) g - Fe2 O3 nanoplates synthesized from iron nitrate precursors in ethanol (with permission from ref. 100), (b) Mn3 O4 NPs prepared from KMnO4 in ethanolwater (with permission from ref. 101), (c) ZnFe2 O4 NPs synthesized from zinc granules and ferric chloride in aqueous ammonia (from ref. 102) and nearly monodisperse NiFe2 O4 NPs prepared from the metal nitrates at a watertoluene interface with oleic acid as capping agent (with permission from ref. 103).

6320 | Dalton Trans., 2011, 40, 63156343

Downloaded on 20 June 2011 Published on 01 March 2011 on http://pubs.rsc.org | doi:10.1039/C0DT00689K

fcc-structured nickel nanoparticles.118 Additionally, the large scale synthesis of uniform magnetite (Fe3 O4 ) nanoparticles was achieved by precipitation from ethanolic FeCl2 /Fe(NO3 )3 solution in xylene with NaDBS under reux conditions.119 Ternary 3d metal oxides, MFe2 O4 (M = Mn, Co, Ni, Cu, Zn, Mg, or Cd, etc.), of the spinel structure are among the most important magnetic nanomaterials and have been widely used for various applications. A common approach for the synthesis of spinel ferrites comprises the combination of aqueous solutions of metal nitrates (e.g. Mn(NO3 )2 and Fe(NO3 )3 ) in toluene with NaDBS as surfactant,120 and, as mentioned before, the average size of the resulting nanoparticles can be controlled by adjusting the molar ratio of water to surfactant. Furthermore, even more complex metal oxide systems, like the quaternary CoCrFeO4 could be synthesized by a similar approach using aqueous solutions of CoCl2 , CrCl3 and Fe(NO3 )3 in toluene with NaDBS.121 A more exotic example was given by Xu et al., who prepared BaFe12 O19 nanocrystals in a microemulsion system consisting of CTAC, n-hexanol, water and cyclohexanol with (NH4 )2 CO3 and NH4 OH as precipitants.122 In contrast to that, a combination of reverse microemulsion and microwave assisted synthesis was recently used to produce manganese-based nanoscale metalorganic frameworks, which show promising properties for magnetic resonance imaging (MRI).123 2.5. Thermal decomposition and/or reduction

By far the best results, for the reproducible synthesis of highly crystalline magnetic nanoparticles with a narrow size distribution, are obtained when metalorganic precursors are decomposed or metal salts are reduced in high-boiling non-polar solvents and in the presence of surfactant molecules. The key issue for the production of monodisperse nanoparticles, the separation of nucleation and growth according to the LaMer concept (see above),85 can easily be achieved by controlling the basic experimental parameters like reaction temperature, heating time, heating rate, and concentrations of precursors and surfactants.50 The required short nucleation burst can basically be initiated in two ways: either by injecting a cold precursor solution into a hot mixture of solvent and surfactants, or by heating up a mixture of all components at the same time in a controlled fashion. The rst method is based on the simultaneous formation of many nuclei when the precursor is injected.124 In the second procedure, all reactants are mixed at room temperature and subsequently heated at a well dened heating rate, until the decomposition temperature of the precursor is reached. Fortunately, this decomposition does not necessarily lead to the formation of nuclei, because a precise separation of nucleation and growth would be difcult, if not impossible. In some cases, the degradation of the precursor molecules rather leads to the formation of molecular clusters that can be viewed as monomers for the generation of NPs. As the temperature rises, the concentration of these clusters increases until it reaches a critical point of supersaturation (see Fig. 2). As a result, a large number of nuclei are formed; however, since the temperature only increases slowly, no considerable growth occurs. Once the reaction mixture has reached a temperature that allows sufcient growth, all particles grow at the same rate leading to a homogeneous size distribution.125 In most cases particle growth can be stopped by rapidly reducing the temperature. For nanoparticles synthesized in this way the standard deviation of the size distribution (s) is This journal is The Royal Society of Chemistry 2011

usually around 10%, however, further size selection techniques, like repetitive precipitation, can lower s below 5%.50,126 Some of the most popular precursors and additives for the synthesis of magnetic nanoparticles will be discussed in the following. Among magnetic nanomaterials, metallic NPs show higher magnetization values than metal oxides. For the synthesis of metal NPs, zero-valent precursors, such as metal carbonyls are preferably used.53,127,128 Fatty acids, long chain amines and thiols are generally used as surfactants to stabilize the as-prepared particles from oxidation and agglomeration.73,129,130 For instance, Alivisatos et al.131 were able to prepare monodisperse e-Co nanoparticles by rapid pyrolysis of dicobalt octacarbonyl ([Co2 (CO)8 ]) in the presence of a surfactant mixture containing oleic acid, lauric acid, and trioctylphosphine. The size of the particles could be controlled from 317 nm by adjusting the precursor/surfactant ratio, the temperature and the injection time. In fact, iron NPs were prepared by a similar method.54 On the other hand, the reduction of metal salts with suitable reducing agents, like hydrides or alcohols, can also lead to the formation of metal nanoparticles. Sun and Murray were among the rst to synthesize metallic nanoparticles by a variety of methods, e.g. e-Co NPs by injecting superhydride (LiBEt3 H) into a hot solution of CoCl2 in octyl ether with oleic acid and trialkylphosphines as surfactants.132 The size of the particles could be controlled by varying the chain length of the alkyl group in the trialkylphosphines. However, one of the most cited approaches for the synthesis of intermetallic NPs, is the preparation of monodisperse FePt NPs and their self assembly into two- and three-dimensional superlattices (see Fig. 6ad).

Fig. 6 (a) and (b) TEM images of three-dimensional assemblies of 6 nm as-synthesized Fe50 Pt50 NPs. (c) HR-SEM image of a 180 nm thick layer of assembled 4 nm Fe52Pt48 nanocrystals. (d) HR-TEM image of 4 nm Fe52Pt48 NPs (with permission from ref. 74). (e) Two- and (f) three-dimensional superlattices of monodisperse MnO NPs prepared from Mn oleate.

Precursors with cationic metal centers are preferably used for the synthesis of magnetic metal oxide nanoparticles. Yet, the preparation of metal nanoparticles followed by mild oxidation is also applied in some cases. For instance, monodisperse maghemite nanoparticles have been synthesized by injecting iron pentacarbonyl (Fe(CO)5 ) into a hot solution of oleic acid, oleylamine and 1,2-hexadecandiol in octyl ether. The initially formed Fe nanoparticles were subsequently oxidized by addition Dalton Trans., 2011, 40, 63156343 | 6321

Downloaded on 20 June 2011 Published on 01 March 2011 on http://pubs.rsc.org | doi:10.1039/C0DT00689K

of trimethylamine oxide ((CH3 )3 NO) as mild oxidant.53 However, metal organic compounds, such as formates,63 acetates,62 and acetylacetonates61 have evolved as the most common precursors for the preparation of metal oxide particles. Inspired by the idea, that during the heating procedure a complex of the metal ion and the surfactant is formed, a new strategy was recently proposed.64,28,73 By directly using a preprepared metalsurfactant complex, a much better control over the reaction process was obtained. In this way, highly crystalline nanoparticles of different metal oxides with exceptionally narrow size distribution (s 5%) were synthesized. As an example, Fig. 6e and f show monodisperse manganese oxide (MnO) NPs synthesized by decomposition of a manganese oleate precursor in 1-octadecene. Due to their exceptionally narrow size distribution they self-assemble to form two-dimensional and even threedimensional superlattices.28

3.
3.1.

Surface modication
General considerations

The previous section demonstrated that there have been numerous ground breaking developments in the synthesis of magnetic nanoparticles over the past decades. However, the most important issue, which needs to be accomplished after their preparation, is to provide the nanoparticles with a sufcient degree of stability. It has already been pointed out, that the use of surfactant molecules is essential to prevent the particles from aggregation and rapid oxidation. Especially, pure metal NPs, like Fe, Co and Ni (and their alloys) are extremely sensitive to air, and in fact, their vulnerability towards oxidation becomes more pronounced with smaller particle sizes. Furthermore, the necessity of an efcient coverage of the nanoparticle surface becomes obvious when considering a potential use in biomedical applications. The high chemical reactivity of the NP surface could pose health risk to patients, since the NPs could already inuence reaction pathways on the cellular level.

Additionally, the colloidal stability of the NPs under physiological conditions is another important issue, which needs to be addressed if an application in medical or biological respect is anticipated. As described above, the synthesis of monodisperse and highly crystalline nanoparticles requires non-hydrolytic synthetic routes and the use of hydrophobic surfactants that efciently cover the particle surface. Therefore, as-prepared nanoparticles are insoluble in water and a direct application of these unmodied particles into aqueous solutions would inevitably lead to particle aggregation and precipitation. All the worse, if applied directly to a living organism, these nanoparticle aggregates would pose an immediate threat to health, since they would clog up small blood vessels and obstruct a free blood circulation. Furthermore, even if particle agglomeration could be prevented, unprotected inorganic nanoparticles are prone to opsonization, a defense process initiated by the innate immune system, in which certain serum proteins attach to the particle surface and initiate phagocytosis. Therefore, it is necessary to exchange the hydrophobic protection shell by hydrophilic ligands, that guarantee, not only a high colloidal stability in aqueous environment, but also prevent opsonization by serum proteins under physiological conditions.133 The latter is often referred to as the stealth effect.134136 This section will focus on some of the most common strategies for the protection and stabilization of magnetic nanoparticles (see Fig. 7). Interestingly, all known surface modication approaches result in a coreshell-like structure, with the magnetic NPs being the core and some kind of inert organic or inorganic material being the protecting/stabilizing shell. Apart from surface passivation, precious-metal or carbon coating (which will not be discussed here), surface modication with surfactant micelles, bi-functionalor polymeric ligands or silica are the most prevailing methods. 3.2. Amphiphilic micelles As explained in the previous section, as-prepared magnetic NPs are covered with a shell of hydrophobic surfactant molecules, such as oleic acid or oleylamine.137,138 The functional group of these

Fig. 7 Different strategies for the surface modication of magnetic NPs.

6322 | Dalton Trans., 2011, 40, 63156343

This journal is The Royal Society of Chemistry 2011

stabilizing agents are strongly attached to the particle surface by a coordinative bond, whereas the long alkyl chains point into the periphery, creating an inert layer, which offers, not only colloidal stability, but also protection from oxidation and aggregation. Consequently, one of the simplest approaches to enhance the solubility of hydrophobic NPs in aqueous solution, uses the molecular structure of the hydrophobic layer for the incorporation of additional surfactant molecules, like CTAB or oleic acid. This surfactant addition strategy was rst developed for the preparation of water-soluble quantum dots and was later also successfully applied for dispersing magnetic NPs in aqueous media.139141 During this procedure, the long alkyl chains of the CTAB molecules are integrated in between the C-chains of the oleate/oleylamine molecules due to attractive van-der-Waals interactions, and on the other hand, the cationic trimethylammonium moieties point to the outside, creating repulsive forces and stabilizing the particles in suspension. Unfortunately, NPs prepared in this way exhibit only moderate long-term stabilities and tend to agglomerate at higher salt concentrations.142 Many of the existing strategies for NP phase transfer, that were developed in the following years, use amphiphilic polymers as surface coating ligands.139,143 The structure of these polymers usually consists of individual lipophilic and hydrophilic polar components, such as poly(maleic anhydride-alt-1-octadecene)-PEG block copolymer, or amphiphilic PEG-phospholipids.136,144147 Here, the lipids act as the nonpolar constituent of larger amphiphiles. Their hydrophobic tails interact with the nonpolar surface ligand of the nanoparticles (see Fig. 8), which leads to a complete encapsulation of the core and its original coating. The hydrophilic endgroup of the amphiphile makes the material polar and fully dispersible in water. The nanoparticles are assumed to remain encapsulated by the original protecting ligand shell that is never broken up, and particle aggregation is suppressed by steric repulsion. Dai and coworkers recently developed a similar PEGmodied phospholipid surface coating for graphite-protected FeCo NPs with a high magnetic moment.148 Again, the non-polar hydrocarbon chains are bound to the hydrophobic graphite shell to form a stable double layer, while the PEG chains are exposed

Downloaded on 20 June 2011 Published on 01 March 2011 on http://pubs.rsc.org | doi:10.1039/C0DT00689K

to the outside. Another example, which uses a similar strategy for monodisperse MnO NPs, was demonstrated by Hyeon et al. in 2007 and Lee et al. in 2009.43,149 Related methods used a variety of amphiphilic polymers, including polystyrenepolyacrylic acid (PS-PAA) block copolymers, tetradecylphosphonate and PEG-2tetradecyl ether, to increase the hydrophilicity of various inorganic NPs.150,151 Additionally, there are many different amphiphilic polymers commercially available today, which offer different functional groups, like carboxylic acids, thiols, amines, and biotin, for the immobilization of various biomolecules, such as peptides, proteins, or oligonucleotides The presence of multiple surface bilayers, however, leads to much larger hydrodynamic sizes than that of the starting material hampering an efcient application in living organisms.152154 Furthermore, since the amphiphilic bilayer is only connected to the NP core by considerably weak van-der-Waals interactions, the surfactant molecules are in a dynamic equilibrium with their unbound counterparts, i.e. by shifting the equilibrium they can be removed by washing and diluting, which would result in particle agglomeration and precipitation. 3.3. Polymeric ligands One of the most established methods to stabilize magnetic NPs in water or buffered solutions is provided by the use of polymers. PEG, perhaps the most common representative biocompatible polymer, has received great attention due to its nonfouling property, that supports a resistance to protein adsorption and an ability to bypass the RES and natural barriers, such as the nasal mucosa.155158 PEGs have been extensively used as stabilizing materials for many NPs used on biomedical applications, particularly in long circulating in vivo imaging systems. In principle, chemical attachment to the NP surface can be achieved by using different functional groups. According to Pearsons concept of hard and soft acids and bases (HSAB), the nature of these groups depends on its afnity toward the metal constituent of the core material.159 For example, gold nanoparticles show a high afnity towards thiol groups, whereas manganese and iron oxide nanoparticles favor oxygen containing ligands. The use of polymers for this purpose is reasonable, since they have the advantage that numerous functional groups can be applied on a single macromolecule. In this context, multidentate polymers, that possess more than one anchoring group, offer much stronger attachment to the nanoparticle surface. Suitable polymers include poly(analine), poly(methylidene malonate), poly(pyrrole), poly(lactic acid), poly(glycolic acid), poly(ethylene imine) and their copolymers.160165 In addition, the possibility of cross-linking between the polymer chains leads to more rigid structures on the NP surface. For example, dextran coated iron oxide nanoparticles can be cross-linked using epichlorohydrin, and subsequently equipped with free amino groups to allow further modication.166 Other examples for cross-linked polymer functionalized iron oxide nanoparticles include the use of chitosan and poly(ethyleneglycol)-co-fumarate.167,168 A variety of methods, that use special chelating ligands, have been established in recent years.169 In fact, there are reports where particle binding of polymeric ligands on Fe3 O4 nanoparticles was achieved via carboxylic, phosphate, phosphonate and sulfate groups.151,170,34,171 Dopamine is another anchoring group which Dalton Trans., 2011, 40, 63156343 | 6323

Fig. 8 (a) Chemical structure of a PEG-phospholipid. (b) Coating procedure: the hydrophobic lipid tails are incorporated between the alkyl chains of the capping molecules on the NP surface, whereas the hydrophilic PEG chains point to the outside creating a hydrophilic shell.

This journal is The Royal Society of Chemistry 2011

Downloaded on 20 June 2011 Published on 01 March 2011 on http://pubs.rsc.org | doi:10.1039/C0DT00689K

shows a high afnity towards metal and metal oxide surface,172 in fact, it has evolved as a very popular binding ligand.173,81 Owing to the multidentate coordination of the chelating moieties, the bonding strength of these polymers is generally very high. Furthermore, the use of reactive ester polymers, such as poly(pentauoro phenylacrylate) (PPFA), permits the successive substitution of the pentauorophenol group by dopamine and other functional groups, including amines, (PEG), or uorescent dyes (see Fig. 9). Metal oxide nanoparticles functionalized with such polymers containing several dopamine moieties were used in different applications ranging from protein separation, cell targeting and multimodal imaging.174179,45

one nanoparticle at a time. Since the binding of the surface ligands is an equilibrium process with continuous attachment/detachment of the chelating groups, there is a non-neglectable possibility that one polymer chain simultaneously covers several particles. As a consequence, the presence of large agglomerates cannot be excluded, which impedes a suitable application for biomedical purposes. 3.4. Bi-functional ligands A key factor for in vivo tracking and targeting applications is a high stability of the bonding between functional molecules and the nanoparticle surface. An early release or uptake of the nanoparticle/molecule system due to other surface reactions, could be detrimental to the application or possibly to the patient. One of the most often used methods for surface modication of magnetic nanoparticles is the exchange of the hydrophobic stabilizing ligands by hydrophilic bifunctional molecules (see Fig. 10). As the name implies, bi-functional molecules are generally composed of two parts: an anchor group, which efciently binds to the surface of the nanoparticle and a hydrophilic part which allows stability in aqueous media. Compared to the amphiphilic micelle coating, the bi-functional ligand exchange strategy principally provides much better colloidal stability of the magnetic NPs under physiological conditions, due to stronger (mostly ionic) interactions between the bidentate (or, in some cases, multidentate) functional group and the metal-/metal oxide surface. Typical anchor groups for metal oxide nanoparticles include carboxylates,183,184 phosphonates,185,186 and catechols,173 whereas thiols187 are preferred for metallic and intermetallic NPs. However, the key requirement for an efcient attachment on the surface is a stronger binding strength of the anchor group, than that of the former hydrophobic stabilizing agent. Although initial attempts to modify hydrophobic nanoparticles were carried out with simple bi-functional molecules, like 2,5-dihydroxybenzoic acid,183 or tetramethylammonium hydroxide,188 PEG has evolved as the most often used hydrophilic linker, owing to its outstanding solubility, stability, and excellent biocompatibility.

Fig. 9 (a) Chemical structure of a typical multifunctional polymeric ligand. The co-polymer contains 3-hydroxytyramine (dopamine) as anchor group, a uorescent dye, PEG chains, and free amino groups for further functionalization. (b) Illustration of a magnetic NP modied with a multifunctional polymer.

A different approach to functionalize magnetic nanoparticles with polymers is enabled by atom transfer radical polymerisation (ATRP).180,181 With this method, polymers are grown on the surface of nanoparticles after previous treatment with an initiator.182 As an example, Lattuada and coworkers used this method to design water soluble Fe3 O4 nanoparticles. The NPs were coated with poly(lactic acid), followed by esterication through acylation, which enabled the addition of halide moieties to transform the nanoparticle surfaces into macro initiators.152 In that way, different ligands could be used to grow polymer coatings with, either cationic, anionic or neutral endgroups. A major drawback for the use of polymers for the surface modication of magnetic nanoparticles is the fact that the presence of multiple anchoring groups facilitates the bonding to more than
6324 | Dalton Trans., 2011, 40, 63156343

Fig. 10 Surface modication of magnetic NPs using bi-functional ligands. These ligands consist of an anchor group conjugated to a hydrophilic linker carrying a functional endgroup. A key requirement for a successful exchange of the hydrophobic capping agents is the complete dispersability of the NPs and the ligand molecules in the same solvent.

This journal is The Royal Society of Chemistry 2011

Among the above mentioned anchor groups, special focus has been addressed to catechols. Especially, dopamine is one of the most considered high afnity binding groups to stabilize metal oxide NPs in water and physiological environment.173,189191 It is a derivative of L-3,4-dihydroxyphenylalanine (L-DOPA), an amino acid, that occurs naturally in the adhesive protein of some marine mussels.172,192195 Indeed, the binding interaction of dopamine with the surface of iron oxide NPs has been thoroughly investigated in the past, and it was conrmed, that the strong attachment was due to an improved orbital overlap of the ve-membered ring and a reduced steric effect on the iron complex which can be obtained in a bidentate coordination.173,196,197 Since the dopamine anchor groups were conjugated with a PEG linker, the functionalized iron oxide NPs were readily soluble in aqueous solution and stable over long periods of time. An example, where two different anchor groups were used for the attachment of bi-functional ligands was provided by Hong et al. for intermetallic FePt NPs. Here, the surface of the magnetic particles was functionalized with a dopamine- and a thiol-containing PEG ligand.198 The use of two different binding groups is reasonable, giving the fact, that both molecules on their own would only ensure poor binding strength on FePt. Based on Pearsons hard and soft acids and bases (HSAB) concept, the reason for this behavior is simply the circumstance, that catechols, as well as other 1,2-diols, indeed show a high afnity toward the iron sites,199,200 but only a weak attraction toward platinum, whereas the opposite is the case for the thiol group. Consequently, the application of both ligands showed much better results compared to studies, where only one of them was used. The NPs were stable in various aqueous media, including water, ionic solutions, and cell culture medium.198 In a similar way, FePt NPs coated with a protective iron oxide shell were stabilized with a dopamine-PEG ligand.201,202 The resulting particles exhibited high cytoxicity due to the presence of the FePt core and strong T 2 MR contrast enhancement due to the presence of an iron oxide shell. Sun et al. also reported the modication of monodisperse iron/iron oxide core/shell NPs with a dopamine-PEG (DOPAPEG) ligand. The ligand was synthesized by a simple EDC/NHS coupling reaction of dopamine with a PEG-dicarboxylic acid. The modied Fe/Fe3 O4 core/shell NPs were stable in water and phosphate buffer solutions.55 The same group investigated the inuence of the PEG-chain length (HOOC-PEG(n) -COOH, M w (PEG) = 600, 3,000, 6,000, 20,000 g mol-1 ) on uptake in macrophage cells.191 The authors revealed, that macrophage uptake decreased with increasing PEG chain length (M w (PEG)). Additionally, the free carboxylic acid groups of the ligands permitted further conjugation of biomolecules. In a similar way, Wang et al. coupled chromones, a group of naturally occurring compounds that show antifungal, antiviral, antihypertensive, and anticancer properties, to PEGylated Fe3 O4 NPs. In this report the PEG chains were terminated with free amino groups, to which the drug molecules could easily be attached. Furthermore, the authors demonstrated that the drug could be released by variation of the pH value.203 However, the use of dopamine-PEG ligands can also be applied for more complex nanostructures. Cheng et al. reported a new approach for the targeted delivery and controlled release of the anticancer drug cisplatin using PEGylated porous and hollow Fe3 O4 NPs. The particles themselves were prepared by controlled oxidation of Fe NPs, followed by acid etching to create porous This journal is The Royal Society of Chemistry 2011

Downloaded on 20 June 2011 Published on 01 March 2011 on http://pubs.rsc.org | doi:10.1039/C0DT00689K

Fe3 O4 nanocrystals with spherical cavities in the middle. The presence of pores on the surface permitted the slow diffusion of cisplatin inside the hollow structures, as well as the slow release of the drug under physiological conditions. Once coupled to the humanized IgG1 monoclonal antibody Herceptin, the cisplatinloaded hollow NPs were able to target SK-BR-3 breast cancer cells.204 However, there have also been critical remarks concerning the application of dopamine as anchor group for metal oxide NPs. Especially, some authors reported on rather rapid degradation of dopamine due to a surface reaction, once it is attached to the surface of iron oxide NPs.197,205 Additionally, catechols are known to be liable to oxidation under aerobic conditions, forming dark and insoluble polymers.206 Apart from dopamine, other catechol anchor groups were also extensively investigated and found to be more reliable for biomedical applications.205,207 Careful selection of the different functionalities allowed close control of the hydrodynamic diameter and the interfacial chemistry of iron oxide nanoparticles. 3.5. Silica coating The surface modication strategies presented in the previous sections relied on the coordinative attachment of organic ligands carrying functional groups to the metal atoms on the particle surface. However, it has to be assumed that this protecting organic ligand shell around the nanoparticle is not densely packed, and therefore, permits diffusion of water molecules to the surface, which leads to dissolution of metal ions and, eventually, to the degradation of the nanoparticle itself. The advantage of using silica as a coating material is based on its exceptional stability, especially in aqueous media. From a technical point of view the easy regulation of the coating process, processability combined with chemical inertness, controlled porosity, and optical transparency are the most important factors.208,209 In addition, cations and positively charged molecules can be linked covalently to the polymeric silica layer at silica water interfaces under basic conditions, i.e. the silica layer provides steric and electrostatic protection for the cores, while acting simultaneously as a dispersing agent for electrostatic colloids. These advantages make silica an ideal, low-cost material to tailor the surface properties of various nanomaterials. Furthermore, silica coatings can equip magnetic NPs with many additional benets, such as high biocompatibility and the possibility of subsequent functionalization (see Fig. 11).210,211 Thus, the potential applicability increases signicantly, and, as an example, allows the use of these core@silica nanomaterials for diagnostic and therapeutic purposes (e.g. MRI).212,213 Over the years, various approaches have been exploited to fabricate silica coatings for colloidal nanoparticles, e.g. the classical Stober synthesis,214216 the formation of biosilica coatings,178,217 or the use of silane coupling agents.218 Although, these strategies have shown good results with various nanomaterials, the downside, however, is the limitation of these processes to aqueous solutions, and therefore, the need to start from water-soluble NPs. As already stated above, high-quality monodisperse NPs are protected by a hydrophobic surfactant layer, and thus, are only soluble in non-polar solvents. Hence, the above mentioned strategies are not applicable on these NPs. An alternative method, which Dalton Trans., 2011, 40, 63156343 | 6325

Fig. 11 Illustration of a multifunctional silica coated magnetic NP.

has become quite popular recently, is based on a microemulsion synthesis, where micelles or inverse micelles are used to conne and control the coating of silica on the core of the inner nanoparticle (see Fig. 12).219222 Microscopically they consist of small heterogeneous domains of water in a surrounding hydrophobic phase (oil) separated by a surfactant monolayer. The great advantage of this approach is the possibility to dissolve as-prepared hydrophobic nanocrystals in the oil phase before entrapping them inside the micelles (see Fig. 12). By optimizing the synthetic conditions it is even possible to capture only one NP per micelle. Furthermore, the thickness of the silica shell can easily be controlled. The silica coating procedure is then initiated by adding the desired silanes as precursors.223 As microemulsions are thermodynamically stable systems (as opposed to emulsions), both the content of the droplets and the surfactant molecules at the interfaces, are constantly and rapidly exchanged between different droplets, thereby facilitating chemical reactions involved in particle synthesis.98 An interesting feature of silica coating is the possibility to incorporate uorescent dyes into the SiO2 matrix. The advantage of this practice is a prolonged stability and higher efciency of the dye inside the shell compared to simple attachment on the outside of the nanocomposite. The reason for this is an extremely effective protection of the dye molecules towards oxidation and photobleaching, as well as a reduction of quenching effects.224230 Another advantage of silica coating of magnetic NPs is the opportunity to create a mesoporous shell. In fact, coreshell magnetic mesoporous silica microspheres with strong magnetic responsivity, orientated, accessible mesopores and high colloidal stability are highly valuable for biomedical applications.232236 Hyeon and coworkers designed magnetically separable high-performance

biocatalysts by cross-linking enzyme molecules on the surface of the Fe3 O4 @SiO2 nanoparticles.237 In a similar fashion, discrete and monodisperse mesoporous silica nanoparticles consisting of a single Fe3 O4 nanocrystal core and a mesoporous silica shell were prepared (see Fig. 13ae).141 The integrated capability of the coremesoporous shell structured NPs as MR and uorescence imaging agents, along with their potential use as a drug delivery vehicles, makes them promising candidates for future cancer diagnosis and therapy. However, the highly porous structure of the silica shell permits an easy diffusion of water molecules to the core particle, leading to a possible degradation through metal ion leaching. An improvement of the core-mesoporous-shell approach was reported by Deng et al., who incorporated a thin non-porous silica layer between the magnetic core and the mesoporous shell (Fe3 O4 @nSiO2 @mSiO2 ) to provide additional protection.231 The approach is shown in Fig. 13fh. The nonporous silica interface layer (n) is thought to protect the magnetite from leaching, whereas the outer mesoporous (m) silica shell not only offers a high surface area for the derivation of numerous functional groups, but also provides a large accessible pore volume for the adsorption and encapsulation of various agents, like drugs, biomacromolecules or even functional nanoparticles. Additionally, the authors demonstrated that the Fe3 O4 @nSiO2 @mSiO2 microspheres still showed superparamagnetic behavior. Unfortunately, most synthetic routes for the preparation of mesoporous silica coated NPs, are based on templating procedures in aqueous solution.238 In most approaches, CTAB is used as templating agent to form long tubular structures, around which the silica shell is polymerized. Subsequently, the template is removed either thermally or by chemical treatment yielding highly ordered mesoporous matrices. Since microemulsion techniques are not applicable in this case, Stober-type methods have to be used, which limits the size of the nanocomposites to diameters larger than 50 nm. This hampers potential in vivo applications of these particles.

Downloaded on 20 June 2011 Published on 01 March 2011 on http://pubs.rsc.org | doi:10.1039/C0DT00689K

4. Biomedical applications
The use of nanotechnology in biology and medicine has led to tremendous breakthroughs in the past decades, and yet, since research activities in this scientic area are proceeding with great speed, nano-biotechnology and nanomedicine are offering numerous exciting possibilities for the future. Already in the 1960s, articial liposomes were suggested as carriers of proteins and drugs for the treatment of various diseases,239 and since then

Fig. 12 Synthetic scheme showing the silica coating procedure in a reverse (or water-in-oil) microemulsion. The hydrophobic NPs and surfactant molecules are dispersed in a non-polar solvent. Upon addition of aqueous NH4 OH micelles are formed in which the NPs are entrapped. Subsequently, the condensation of tetraethoxysilane (TEOS) leads to the formation of SiO2 on the NP surface. After complete reaction, the SiO2 coated magnetic NPs are retrieved by addition of methanol.

6326 | Dalton Trans., 2011, 40, 63156343

This journal is The Royal Society of Chemistry 2011

Downloaded on 20 June 2011 Published on 01 March 2011 on http://pubs.rsc.org | doi:10.1039/C0DT00689K

Fig. 13 Encapsulation of magnetic NPs with mesoporous silica. (a) Reaction scheme illustrating the synthetic procedure to create mesoporous silica (mSiO2 ) coated Fe3 O4 NPs, or hollow mesoporous silica particles. (b) and (c) TEM images of hollow mSiO2 NPs. (d) and (e) TEM images of Fe3 O4 @mSiO2 NPs (with permission from ref. 141). (f) Reaction scheme depicting the experimental process to create Fe3 O4 @nSiO2 @mSiO2 microspheres. (g) TEM image showing the thin nSiO2 layer between Fe3 O4 and mSiO2 , and (h) the porosity in the mSiO2 shell (with permission from ref. 231).

nanotechnology has had a major impact on the development of drug delivery systems. The application of nanotechnology is, however, not only limited to drug delivery, various organic/inorganic nanostructures are currently being applied or show promising potential in many different biomedical areas including in vitro diagnostics, in vivo imaging, therapy of various diseases (like cancer, cardiovascular and neurological diseases), biomaterials, and tissue engineering.8,10,7,11,2,240,241 Up to now, there are over 40 FDA approved nanomedicine products on the market and even more are currently in clinical trials,8,10 most of them being used for drug delivery applications. In fact, early clinical results suggest that nanoparticle therapeutics can show enhanced efciency compared to conventional drugs, while at the same time, negative side effects can be reduced signicantly.242,243 This is due to a more targeted localization inside the desired area of the body (e.g. tumor tissue), either by passive or active cell targeting, followed by active cellular uptake of the nanoparticle probe. Moreover, the utilization of magnetic nanoparticles for simultaneous real-time in vivo monitoring of drug delivery is also an area of intense interest, owing to their MR enhancing properties. One of the greatest challenges associated with the use of magnetic NPs in vivo is the ability to circumvent or overcome the different biological barriers. As an example, the targeting efciency This journal is The Royal Society of Chemistry 2011

of nanoparticulate probes is often limited by the precocious recognition and elimination by the reticuloendothelial system (RES). In this respect, the overall NP size, morphology, charge and surface chemistry are crucial properties that determine, not only the circulation time inside the blood vessels, but also the nal distribution within the body.244246 In the following, the basic principles of nanoparticle design for biomedical applications, the fundamental interactions with biological systems, as well as the biodistribution and clearance of magnetic NPs from living organisms will be discussed briey. 4.1. Design considerations As pointed out in the previous sections, there have been remarkable advances in nanomaterials synthesis and functionalization, which today allow a precise engineering of the composition, size, size distribution, morphology, and surface chemistry of magnetic NPs. Both, magnetic properties and surface features can be tailored individually to meet the particular challenges of in vitro and in vivo biomedical applications. Before magnetic NPs can be used for biomedical applications, they require thorough consideration concerning the assembly of different components (see Fig. 14). Dalton Trans., 2011, 40, 63156343 | 6327

particles tend to non-specically stick to the surface of cells.253 How great the impact of this non-specic interaction actually is, regarding the blood circulation time, was demonstrated by Papisov et al.254 On the other hand, strong negative charges on the particle surface are also unfavorable, since they lead to increased liver uptake.244 Therefore, it has to be estimated that NPs with a near neutral surface charge exhibit prolonged blood circulation times. 4.2. Passive cell targeting
Downloaded on 20 June 2011 Published on 01 March 2011 on http://pubs.rsc.org | doi:10.1039/C0DT00689K

Fig. 14 Illustration of a multifunctional magnetic nanoparticle (adapted from ref. 9).

1. The NPs themselves can act as the active part of the nanocomposite probe, because they represent the platform, on which all other constituents will be assembled, and, depending on their physical properties (magnetic, optical), they are the key factor that ensures the detection of diseases inside the body. 2. The NPs need an inert and hydrophilic shell that protects them and offers both, stability in biological systems and long circulation times in the cardiovascular system. However, this attribute was already thoroughly discussed in the previous section. 3. Certain uorophores that permit a subsequent detection by optical methods may also be included. 4. Additionally, specic targeting ligands, which recognize certain kinds of cells inside the body, can be attached to the surface of the nanocomposite to ensure a selective delivery to the target tissue. 5. Moreover, responsive moieties can also be included that will react to external excitation, e.g. photosensitizers to produce reactive oxygen species (ROS) upon optical excitation, or gold NPs which produce heat upon NIR illumination. 6. Finally, certain drugs may also be attached to the composite to allow a specic treatment in a spatially conned area. In this respect special responsive linker molecules, which will cleave under an external stimulus (e.g. heat, pH, enzymatic cleavage), could guarantee a safe delivery of the payload to the desired region. Additionally, the overall size of the multifunctional magnetic NPs must also be considered. Their hydrodynamic size, which is dened as the overall size of the particle including the solvent shell, is strongly correlated to the ability of the nanoparticle to overcome the initial biological defense system and penetrate through the vascular barriers. On the one hand, the size of the NPs must be sufciently small to avoid rapid splenic ltration, on the other hand, they should be large enough to circumvent renal clearance.9,247 It has been shown, that particles with sizes larger than 200 nm are quickly attacked by phagocytotic cells of the spleen, whereas particles with an average diameter below 5.5 nm are rapidly removed through the kidneys (see Fig. 1).248,249 However, particles of the appropriate size range are subject to opsonization, resulting in recognition and elimination by Kupffer cells and other tissue macrophages, unless they are protected by an inert hydrophilic shell.9,17 For these NPs, immediate opsonization and/or phagocytosis is reduced. Furthermore, the surface charge of the magnetic NPs also plays a crucial role in cellular uptake and blood half-life.246,250252 In fact, it has been shown, that positively charged polymers and small
6328 | Dalton Trans., 2011, 40, 63156343

After multifunctional NPs are intravenously injected, they travel along the blood stream through the cardiovascular system. At this point, the surface chemistry can change signicantly by the adsorption of proteins, demonstrating the importance of an efcient shielding of the NPs (stealth effect).133 The role of proteinnanoparticle interactions has been described using the concept of the nanoparticleprotein corona.255 This dynamic layer of proteins (and/or other biomolecules) adsorbs to nanoparticle surfaces immediately upon contact with biouids. The composition of the protein corona at any given time will be determined by the concentrations of the over 3700 proteins in plasma256 and the kinetic on and off rates (or equilibrium binding constants) of each protein for the particular nanoparticle. This corona may not immediately reach equilibrium when exposed to a biological uid.257 Proteins with high concentrations and high association rate constants will initially occupy the nanoparticle surface, but may also dissociate quickly to be replaced by proteins of lower concentration, slower exchange, and higher afnity.258 These interactions may prove unfavorable possibly resulting in an inefcient uptake or even overload of SPIONs depending on the presence of certain proteins in the blood plasma that may act synergistically with the behavior of NPs. Thus, it can be speculated that NPs undergo two main processes that are important determinants of their cellular uptake: transient NP protein binding (in culture media in vitro or body uids in vivo) and NPprotein interaction with cell surface/membrane macromolecules. During blood circulation, there is a constant exchange of blood substances through the capillary microcirculation, which is regulated by concentration gradients. Since after injection the local NP concentration in the blood is very high, compared to the extravascular space, the NPs will slowly travel outside the blood vessels into the surrounding extracellular matrix (ECM), as long as the gradient persists. Once in the extravascular space, some NPs are taken up by the cells of the tissue and others are retained in the interstitial uid, eventually entering the lymphatic system.259261 However, these processes are comparatively slow. On the other hand, a faster diffusion of macromolecules and NPs into the ECM is observed in tumor tissue, a feature commonly referred to as enhanced permeability and retention (EPR) effect (see Fig. 1).262 This effect originates from the intrinsic vascular characteristics of tumor tissue and the lack of an efcient lymphatic recovery system in solid tumors. In contrast to the vasculature in healthy tissue, the blood vessels, that sustain a tumor, are leaky, with large fenestrations and without an effective clearance by the lymphatic system. This leads to a preferred accumulation of the NPs in the interstitial space inside the tumor, and eventually, the uptake of the nanocomposites into the tumor cells.15,263 This journal is The Royal Society of Chemistry 2011

Downloaded on 20 June 2011 Published on 01 March 2011 on http://pubs.rsc.org | doi:10.1039/C0DT00689K

Passive targeting can also be exploited through the clearance of NPs by the reticuloendothelial system (RES) (i.e. liver, spleen, lymph nodes and bone marrow). Indeed, the rst clinically approved T 2 contrast agents for MRI, based on magnetic iron oxide NPs, beneted from the rapid uptake of the NPs by Kupffer cells of healthy hepatic parenchyma. Thus, healthy tissue appeared darker than diseased tissue in the MR image, due to enhanced T 2 contrast, leading to an improved diagnostic differentiation.264266 However, a successful passive targeting of tumor tissue can only be achieved, if the blood circulation time of the NPs is reasonably long. This is another reason, why efcient surface functionalization strategies are of supreme importance. 4.3. Active cell targeting

There is no doubt about the fact, that the passive accumulation of magnetic NPs in tumor tissue due to the EPR effect is a great benet for the application of these NPs in the diagnosis and treatment of cancer. Nevertheless, a longer residence time and a more site-specic accumulation of the NP probes is desired in most cases. For this reason, active targeting strategies have been developed, in which magnetic NPs were further functionalized with specic targeting molecules that possess a high afnity toward unique molecular signatures found on malignant cells (see Fig. 15a and b).266 The key issue in this case is the fact that cancer cells overexpress certain kinds of receptors compared to healthy cells. Therefore, NPs carrying ligands which will specically bind to these receptors will preferably accumulate in tumor tissue and, eventually, will be taken up by the cancer cells. Several targeting ligands, including proteins, peptides, aptamers and small molecules, have been examined for this purpose.267273 The rst targeting agents, that were used for the specic delivery of magnetic NPs, comprised monoclonal antibodies (mAbs).274,275 Up to now, mAbs have evolved as one of the most popular targeting ligands, which is mainly due to their extremely high specicity.276,277 During the last decade, the development of HerceptinTM (trastuzumab), an FDA-approved mAb that binds to the HER2/neu receptor on the surface of certain cancer cells (overexpressed in breast, ovarian, stomach cancer), has led to various applications of magnetic NPs for selective targeting of cancer cells.278,279 However, a major drawback of mAbs is their comparatively large size and inherent immunogenicity, which impedes a sufcient circulation and diffusion through biological barriers.280 Another popular technique for the targeted delivery of magnetic NPs to tumor cells is the use of short peptides or small molecules conjugated to the NP surface. The advantage, in these cases, is the possibility to attach several (hundreds, or even thousands of) targeting ligands on each NP. As a result, more cell receptors can be addressed simultaneously, leading to the formation of more binding sites, and therefore, an increased binding afnity.269 The most investigated ligand in this respect is folic acid, since the receptors of this vitamin can be found on the surface of many human cancer cells, including breast, ovarian, lung, renal, and colon cancer.281 Indeed, folic acid modied magnetic NPs have been used in the past, not only to improve the MRI detectability of various tumors in vivo282284 but also to destroy cancer cells via hyperthermia treatment.285 A great advantage of short peptides and other small molecules is their high bonding strength once This journal is The Royal Society of Chemistry 2011

Fig. 15 Illustration of active tumor targeting using magnetic NPs with cell-specic ligands. (a) healthy cells only carry a certain number of cell receptors, which can be used for active targeting, in cancer cells (b) these receptors are overexpressed, therefore, more NPs can attach to the malignant cell. (c) Receptor-mediated endocythosis of the NPs: (d) formation of an endosome carrying the NPs, (e) internalization of the endosome, (f) endosomal acidication by proton pumps leads to elevated osmotic pressure and swelling, (g) nally, the endosome ruptures and releases the magnetic NPs.

they are conjugated to the NP surface. Compared to large peptides or proteins, there is a reduced risk of bond breaking and loss of functionality after administration of the NPs into the body. Another important issue associated with the targeted delivery of magnetic NPs to tumor cells is the internalization of the particles into the malignant cell. In principle, there are several possible mechanisms for the cellular uptake of NPs. For example, it has been shown, that small PEGylated NPs can easily diffuse through the cell membrane,282 on the other hand, permeation enhancers, such as TAT peptides, can signicantly increase the degree of NP uptake.286,287,221,288 The typical pathway for a receptor-mediated cellular uptake of NPs is shown in Fig. 15cg. After attachment on the cell surface, the NPs are internalized by formation of an endosome. Fig. 15 also illustrates another common challenge of targeted (drug) delivery that needs to be addressed for NP design, i.e. the controlled release of the NPs, and especially the loaded therapeutic agents, once the NPs have entered the interior of the cell. Various strategies have been developed to accomplish this task. The release mechanisms of most delivery systems depend on processes such as diffusion, dissolution, chemical and enzymatic reactions, or changes in various environmental factors, including temperature, pH, solvent effects, and ionic concentrations.289293 As an example, Dalton Trans., 2011, 40, 63156343 | 6329

Downloaded on 20 June 2011 Published on 01 March 2011 on http://pubs.rsc.org | doi:10.1039/C0DT00689K

an efcient release can be achieved by providing a positive surface charge to the NP probe. The incorporation of cationic polymers on NP surfaces has proved to be an appropriate method to induce osmotic swelling of the endosome caused by the proton sponge effect.294,295 As a result the endosomal membrane is disrupted and the NPs are released into the cytosol. The downside of this approach, however, is the possibility of uncontrolled release due to different release mechanisms depending on the location in and the character of the respective cell. For this reason various methods have been developed that rely on physical triggers, such as heat or radiation, to initiate payload release.296299 For these purposes specic stimuli responsive chemical linkers can be tailored and subsequently inserted between the NP and the drug. These strategies can involve bond breaking by radiative excitation, slow diffusion mediated processes from hollow or mesoporous NP structures, or by cleavage of the linker bond by pH reduction inside the lysosome.300308 4.4. Magnetic NPs as MRI contrast agents in biomedicine

Superparamagnetic iron oxide NPs (SPIONs) have been used as T 2 contrast agents for over two decades.265,274,309 Although enormous progress has been achieved in the technological development of MRI, especially sophisticated pulse sequences for image generation, there has hardly been any improvement in the development of the contrast agents which are in clinical use today. Typical T 2 contrast agents consist of polydisperse and aggregated iron oxide NPs, which show a comparably weak performance. However, the tremendous advances in the synthesis and functionalization of magnetic NPs discussed in the previous sections have led to a variety of new possibilities in the future design of MRI contrast enhancing probes.310312 The basic physical phenomena leading to an enhancement in MR contrast have already been presented. As mentioned previously, the nanoparticles must be monodisperse, highly crystalline, and water soluble to provide reproducible quality, high magnetization values and good biocompatibility under biological conditions. In this respect, magnetic NPs have been studied thoroughly over the past few years as MRI contrast agents for applications, such as cancer imaging, cell migration, gene expression, angiogenesis, apoptosis, cardiovascular disease imaging, or molecular imaging.166,313319 Specially designed NPs with tailored properties and cell specic surface ligands have a great potential to pinpoint biological targets, such as cancer, in the early stages of the disease. Concerning cancer imaging, the development of multifunctional magnetic NPs has led to many improvements in the detection, diagnosis and treatment of solid tumors. Weissleder et al. demonstrated the transgene expression in gliosarcoma cells in vivo using dextran coated iron oxide NPs.313 As an example for magnetic NPs being used as molecular sensors, Perez et al. developed biocompatible magnetic nanosensors that act as magnetic relaxation switches to detect molecular interactions (e.g. DNADNA, or proteinprotein interactions) by the reversible self-assembly of SPIONs.31 The changes in the magnetic relaxation which are associated with the particle self-assembly could then be detected by MRI. On the other hand, Hu et al. showed that PEGylated Fe3 O4 NPs conjugated to the monoclonal antibody rch 24, effectively targeted human colon carcinoma xenograft tumors implanted in nude mice.320
6330 | Dalton Trans., 2011, 40, 63156343

Another example, in which magnetic NPs were used to detect cancer markers by MRI in vivo, was given by Cheon and coworkers (see Fig. 16).33 The authors developed novel magnetic nanoprobes by systematically tuning the spin, size and composition of metal ferrite NPs, followed by conjugation of different cancer specic antibodies, such as Herceptin. These NPs showed enhanced MRI sensitivity for the detection of cancer and enabled the visualization of small implanted tumors in mice. Hultman et al., on the other hand, created immunotargeted SPIONs to investigate the MHC class II expression in renal medulla.321 For this purpose, monodisperse Fe3 O4 NPs were encapsulated in a phospholipid shell and further conjugated to RT1 anti-MHC Class II antibodies that are capable of targeting normal cells expressing specic target antigens. Enhanced binding of the RT1 functionalized SPIONs indicated a denitive specicity for the renal medulla and thus potential for disease detection. Recently, nanocomposite particles consisting of silica spheres decorated with Fe3 O4 NPs were synthesized by Lee et al.322 Additionally, a uorescent dye was incorporated into the silica domain to permit both MRI and optical detection. The use of mesoporous silica furthermore offered the possibility to load, and slowly release, an anti-cancer drug into the pores of the silica sphere. Although the use of manganese oxide NP as T 1 contrast agents is not as common yet, there are a number of reports dealing with the design of different manganese oxide NPs and their application for MR contrast enhancement. Hyeon and coworkers were the rst to prepare T 1 a contrast agent based on monodisperse MnO NPs and investigate its performance in vivo (see Fig. 17).43 The particles were coated with an amphiphilic phospholipid, functionalized with Herceptin and showed no severe toxicity even after several weeks. The authors showed that the Herceptin functionalized NPs accumulated in breast cancer tissue, whereas un-functionalized control NPs also illuminated the surrounding tissue. In a similar way, Shapiro et al. showed that un-functionalized Mn3 O4 and MnCO3 NPs can be internalized within phagocytotic cells and subsequently shuttled to endosomes and/or lysosomes.323 The following decomposition of the NPs inside the lysosome leads to the release of Mn2+ ions which act as a T 1 agent. It was also shown that a combination of SPIONs and MnO NPs can be used to track transplanted cells.316 While the SPIONs produced a negative contrast, the MnO NPs generated a positive contrast and by combination of both imaging techniques, simultaneous imaging with opposite contrast offers the possibility for MR double labeling of different cell populations. Additionally, Yang et al. developed silica coated Mn3 O4 NPs functionalized with folic acid and contained Rhodamine B isothiocyanate (RBITC) as uorescent dye for optical detection.47 These particles accumulated selectively in cancer cells overexpressing folic acid receptors and could be traced by MRI. Very recently, our group reported on the development of highly water soluble MnO NPs conjugated with protoporphyrin IX as multifunctional agents for MRI and photodynamic therapy.46 Protoporphyrin IX, as a photosensitizer initiates the production of reactive singlet oxygen when irradiated with visible light. We were able to show, that the NPs not only exhibited excellent T 1 contrast, but also induced apoptosis in Caki1 cells once illuminated with visible light. Apart from sole MRI contrast enhancement, magnetic NPs may also serve as probes for multimodal imaging. The combination of different materials enables the selective utilization of the This journal is The Royal Society of Chemistry 2011

Downloaded on 20 June 2011 Published on 01 March 2011 on http://pubs.rsc.org | doi:10.1039/C0DT00689K

Fig. 16 Specically engineered MFe2 O4 NPs as MRI contrast agents. (a) The magnetic properties can be tuned by varying the composition of the NPs. A higher magnetic moment results in an increase of the magnetic relaxivity and therefore, enhanced T 2 contrast effect. (b) In vivo MR detection of cancer cells using Herceptin conjugated MnFe2 O4 NPs (left) and bare cross-linked iron oxide NPs (CLIOs) (right). Over time, an increase in contrast enhancement due to preferable accumulation of Herceptin-conjugated MnFe2 O4 NPs in the tumor tissue is observed (with permission from ref. 33).

individual properties. For example, Choi et al. reported that the incorporation of 124 I into the biocompatible shell of manganese ferrite NPs, permitted the use of these hybrid NPs as dual imaging agents for MRI and positron emission tomography (PET) and, therefore, combining the benets of MRI together with the high sensitivity of PET (see Fig. 18).324 A similar approach was recently reported where the authors used MnO NPs coated with human serum albumin (HSA) and which were further conjugated with a 64 Cu containing complex.48 With this method it was possible

to visualize xenografted U87MG glioblastoma cells in vivo, both by MRI and PET. Another example for multimodal imaging was demonstrated by water soluble FePt NPs that were used for in vitro and in vivo imaging by MRI and computer tomography (CT).325 Different sized FePt NPs were functionalized with cysteamine to ensure hydrophilicity and the possibility for further biofunctionalization. Upon conjugation with the antibody AntiHer2, the NPs preferably accumulated in Her2-overexpressing cancer cells, as conrmed by MRI and CT. On the other hand,

This journal is The Royal Society of Chemistry 2011

Dalton Trans., 2011, 40, 63156343 | 6331

Downloaded on 20 June 2011 Published on 01 March 2011 on http://pubs.rsc.org | doi:10.1039/C0DT00689K

Fig. 17 MnO NPs as T1 contrast agents. (a) TEM images of monodisperse MnO NPs of different sizes. (b) T1-weighted images of aqueous MnO NP solutions. (c) Herceptin conjugated MnO NPs increase the contrast inside the tumor tissue, whereas bare MnO NPs (d) also led to an enhanced contrast in the surrounding tissue (with permission from ref. 43).

Fig. 18 Bimodal imaging using 124 I conjugated MnFe2 O4 NPs. (a) TEM images of monodisperse MnFe2 O4 NPs, (b) hydrodynamic size, (c) aqueous solutions of these NPs with different NaCl concentrations and varying pH. (d) Preparation scheme fot the conjugation of 124 I onto MnFe2 O4 NPs. (e) MRI and PET images of free 124 I, MnFe2 O4 NPs, and 124 I-conjugated MnFe2 O4 NPs (with permission from ref. 324).

we were able to combine the properties of MnO NPs and Au NPs for simultaneous MRI and optical detection by creating biocompatible Au@MnO nano-owers (see Fig. 19).45 4.5. Toxicity of magnetic NPs

The need to evaluate the possible risk of human exposure to different kinds of nanomaterials has become a central issue in modern materials and biomedical science. Especially, as the number of different nanomaterials increases with an unprecedented speed, there is the predominant opinion among both, proponents and sceptics, that the vast potential of nanotechnology requires immediate attention to safety issues. Particularly, in the eld of nano-biotechnology, direct exposure to the human body is denitely intended, and therefore, understanding the behavior and properties of magnetic NPs on the organism after
6332 | Dalton Trans., 2011, 40, 63156343

administration, is essential before considering clinical use. Since these NPs are intentionally designed to interact with cells, it is crucial to investigate the cellular responses upon NP exposure, to counteract any possible harmful effects, already during NP development. However, since most research on the toxicology of nanomaterials has focused on the effects of nanoparticles that enter the body accidentally, reliable methodologies to estimate the toxicological impact upon intended administration are still lacking, although there are already numerous studies on cytotoxic effects of different kinds of nanoparticles.326 Despite the fact that an in vivo application of any nanomaterial requires a thorough comprehension of the kinetic processes and the toxicological impact this material exerts inside a living organism, the most frequently used screening studies are based on in vitro cytotoxicity experiments, since they are simpler, faster and less expensive compared to in vivo studies.327330 This journal is The Royal Society of Chemistry 2011

Downloaded on 20 June 2011 Published on 01 March 2011 on http://pubs.rsc.org | doi:10.1039/C0DT00689K

Fig. 19 (a) Schematic illustration of Au@MnO nano-owers separately functionalized using a multifunctional polymeric ligand carrying a catechol anchor group and a green uorescent dye tagged to the PEG side groups. The Au core is functionalized with a red uorescent dye tagged thiolated oligonucleotide. (b) TEM image of polymer coated Au@MnO nanoparticles (with permission from ref. 45).

The main goal of most toxicity studies is to cover the main events in nanoparticle mediated cellular interactions. Particular emphasis is devoted to particle adhesion, cell entry and impact on cell viability and metabolic pathways to gain a comprehensive understanding of the main reasons of nanoparticle mediated cytotoxic effects on cells of the blood compartment. Contemporary research on nanotoxicity deals with two major questions that are largely unsolved to date. What are the uptake mechanisms and efcacies of particles as a function of size, shape and surface functionalization and which biochemical pathways are triggered that eventually cause cell toxicity in cells mirrored either by apoptosis or necrosis? An often examined parameter to assess possible harmful effects of nanomaterials inside living organisms is an inammatory response.331 For instance, recent studies revealed that TiO2 NPs, which are a common component in many commercially available products, such as cosmetics or sun screen lotions, can cause inammatory responses and generation of reactive oxygen species (ROS) leading to DNA damage.332 Moreover, it was also shown that in vitro incubation of single-walled carbon nanotubes (SWNTs) with keratinocytes and bronchial epithelial cells, led to lipid peroxidation, oxidative stress, mitochondrial dysfunction, and changes in cell morphology.333,334 Additionally, a size-dependent toxicity was reported for silver and gold NPs exposed to alveolar macrophages, connective tissue broblasts, epithelial cells, macrophages, and melanoma cells.335337 On the other hand, quantum dots can also initiate an inammatory response and the production of ROS.338 Concerning magnetic metal oxide NPs (above all, Fe3 O4 , gFe2 O3 and MnO), most in vitro cytotoxicity studies revealed rather negligible toxic effects. However, it was stated by several authors that naked iron oxide NPs show a higher toxic potential compared to those that were modied with an inert biocompatible coating.339 Furthermore, the toxicity depends on various factors such as surface-coating or its breakdown products, chemical composition of the cell-medium, the oxidation state of iron in and proteinNP interaction. A study investigating the effect of different surface coatings on cell behavior and morphology showed that dextran-coated Fe3 O4 NPs led to cell death and reduced proliferation quite similar to the interaction of cells with naked magnetite NPs.339 This was attributed to a collapse of the dextran shell exposing the cellular components to magnetite NPs. However, the cell behavior and morphology of cells treated This journal is The Royal Society of Chemistry 2011

with dextran-coated magnetite was different from the uncoated NPs, which showed more severe membrane disruptions.339,340 The membrane disruption after exposure to albumin functionalized Fe3 O4 NPs turned out to be related to the interaction between albumin and membrane fatty acids and phospholipids. However, in spite of the cytotoxicity of uncoated and dextran-coated Fe3 O4 NP the albumin-coated particles did not lead to cell death. Another cytotoxicity study on a mouse broblast cell line showed that the uncoated particles induce greater toxicity than the polyvinyl alcohol (PVA)-coated particles.341 However, the toxicity induced by uncoated particles was signicantly reduced compared to surface-saturated SPIONs which were prepared by pre-incubating the media with the NPs prior to exposure to cells. During this pre-incubation biomolecules from the cell medium adsorb onto the SPIONs which masks their reactive surface thereby preventing unfavorable cellnanoparticle or serum proteinnanoparticle interactions that in turn led to reduced cellular uptake and lower toxicity. Furthermore, the toxicity was governed by compositional changes in the media because serum proteins can bind to the negatively charged uncoated NPs. This resultant altered composition of the cell medium to which the cells are exposed results in the observed cytotoxicity.342 This effect may not occur in vivo because homeostasis maintained by the liver and kidneys efciently regulates any changes in pH, ionic strength and chemical composition of the blood plasma. Another important aspect is the stability of these SPIONs with time. Magnetite can undergo oxidation to form maghemite in the presence of air, light and moisture according to: Fe3 O4 + 2H+ g-Fe2 O3 + Fe2+ + H2 O.343 Magnetite (Fe3 O4 ) and maghemite (g-Fe2 O3 ) can show different cellular responses because of their ability to undergo oxidation/reduction reactions. In fact, magnetite has been shown to cause higher levels of oxidative DNA lesions (using the comet assay) in A549 human lung epithelial cell line in the absence of decreased cell viability as compared to maghemite owing to its potential to undergo oxidation.344 The toxicity can, however, be decreased by coating magnetite particles resulting in fewer oxidative sites that are less reactive.177 The importance of an efcient and biocompatible surface coating was further demonstrated by Gupta et al. The authors reported that PEGylated iron oxide NPs had almost no toxic effect on broblast cells, even at concentrations as high as 2 mg mL-1 , whereas uncoated NPs caused a 50% decrease in cell viability already at comparatively low concentrations (250 mg mL-1 ).117 The low cytotoxicity of PEGylated Fe3 O4 NPs was further conrmed by other authors using different kinds of PEG.345,153 Besides PEG, iron oxide NPs with other surface coatings, including various polymers or silica, also showed hardly any toxic effect.346348 The reason for the higher cytotoxicity of naked iron oxide NPs was attributed to both, cellular uptake of the NPs and the production of ROS, caused by Fe-ion leaching.349 For instance, Brunner et al. reported a signicant reduction in cell viability in human mesothelioma cells upon addition of only 3.75 ppm iron oxide.350 The authors proposed that this high toxicity was due to ironinduced free radical production (e.g. hydroxyl radicals) by Fenton or HaberWeiss reactions. Dalton Trans., 2011, 40, 63156343 | 6333

A further issue in this context is the dependence of the degree of NP internalization on the quality of the surface coating. For instance, De Cuyper and coworkers designed a cationic amphiphile to maximize internalization of SPION without inducing any cytotoxic effects on neural progenitor cells and human blood outgrowth endothelial cells.351 However, high doses of these NPs caused interference with the actin cytoskeleton resulting in decreased cell proliferation indicating the possibility that nontoxic doses could cause other forms of cellular stress. Similarly it was demonstrated that magnetoliposomes can affect the actin cytoskeleton architecture as well as the formation of focal adhesion complexes and impair the cell proliferation.352 Another case of SPION-mediated cellular stress involves disruption of a cytoskeleton protein, tubulin that has been shown to be associated with the uptake of transferrin-derivatised SPION.353 However, these results further substantiate the importance of effective surface coating techniques for magnetic NPs in biomedical applications. The majority of nanoparticle types studied so far are found to bind apolipoproteins.258 As apolipoproteins are known to be involved in lipoprotein complexes, which themselves have sizes on the nanoscale (e.g. 100 nm for chylomicron to ~10 nm for high density lipoproteins), there may be size-dependent interactions that drive the binding of apolipoproteins to nanoparticles. Apolipoprotein E is known to be involved in trafcking to the brain,354 and therefore, some NPs may be able to traverse the bloodbrain barrier. Thus, nanoparticles could be classied in terms of their biomolecule corona, which mediates their interaction with cellular machinery. This would represent a truly new paradigm in the eld of nanoscale toxicology, and in the design of nanocarriers for nanomedicine. In fact, a recent study on protein amyloid aggregation (relevant for Parkinsons and Alzheimers disease or type II diabetes) found the protein aggregates were signicantly reduced by magnetic Fe3 O4 NPs. These NPs not only inhibited lysozyme amyloid aggregation by blocking the nucleation process but also induced depolymerisation of lysozyme aggregates by interacting with and interrupting the adjoining protein sheets; lysozyme adsorption seemed to govern both these processes.355 Currently there are no clinical drugs to reverse or prevent the formation of aggregates and based on this study the Fe3 O4 NPs could potentially be used as novel therapeutic agents in the treatment of protein amyloid aggregation-associated human pathologies.356 On the other hand, the toxic properties of MnO NPs have hardly been investigated so far. However, most reports on the biomedical application of MnO NPs suggest that the toxicity of this material is comparatively low.316,47 For instance, Choi et al. investigated the toxic potential of Fe3 O4 and MnO NPs coated with a hydrophilic micellar phospholipid shell and revealed that, although the MnO NPs showed a slightly higher toxicity than the Fe3 O4 NPs, the overall toxic potential is acceptable.357 Furthermore, our group demonstrated recently that both, Au@MnO nanoowers and PEGylated MnO NPs carrying protoporphyrin IX as a photosensitizer, exhibited no cytoxic effect in CaKi-1 cells in concentrations as high as 140 mg mL-1 .45,46 Concerning free manganese ions, literature data suggest that, although minor concentrations of these ions are not considered to be harmful to the human body, thermodynamically stable Mn compounds which are not prone to Mn ion-leaching should be preferred as NP platforms for in vivo applications.358
6334 | Dalton Trans., 2011, 40, 63156343

Downloaded on 20 June 2011 Published on 01 March 2011 on http://pubs.rsc.org | doi:10.1039/C0DT00689K

Undoubtedly, a wealth of information has been collected on the toxicity of chemically pure, geometrically isotropic, and chemically monofunctional (mostly iron oxide) nanoparticles as a function of their size, shape and chemical nature. These studies must be extended in the future to elucidate of the impact of particles with intrinsic chemical and physical asymmetry, namely multifunctional or so-called Janus particles, which have been poorly assessed up to now. These particles possess new challenges for biological cells associated with high dipole moments, amphiphilicity, shape anisotropy, and chemical diversity/patchiness. Specically, the following scientic questions must be addressed: (i) How does multifunctionality and shape anisotropy impact the interaction of particles with biomembranes? (ii) What are the possible consequences for the uptake of heterogenic multifunctional nanoparticles? Do Janus particles exert curvature in cell membranes that inevitably leads to endocytosis?359 What exposure routes (basal or apical) lead to highest uptake rates? (iii) What kind of nanoparticle-mediated intracellular processing mechanisms can be identied?

4.6. Biodistribution and clearance An essential topic that is often neglected in literature is the long-term fate of magnetic NPs after they have fullled their duty inside the body. For this purpose it is not only important to understand the behavior and interaction of magnetic NPs with living organisms, it is also crucial to determine where nanoparticulate probes will eventually end up and how they are nally excreted from the body.326 As already stated in the previous sections, NPs can enhance the delivery of drugs to certain tissues. However, this may not only cause new side effects, but it may also lead to signicant alterations in the pharmacokinetic proles of the parent drug and the drug attached to the NPs. Therefore, the pharmacokinetics and biodistribution of NPs must be monitored to understand and predict their efcacy and side effects. The pharmacokinetic prole and the internalization of metal oxide NPs are determined by their chemical and physical properties, such as size, charge, and surface chemistry. After administration, larger particles with diameter > 200 nm are easily sequestered by the spleen and eventually removed by the cells of the phagocyte system, resulting in decreased blood circulation times. Small particles with diameters less than 5.5 nm are rapidly removed through extravasations and renal clearance. Magnetic NPs are usually within this suitable size range. Reduced liver metabolism and renal clearance of drugs encapsulated in the nanoparticles often result in prolonged blood circulation with an increased chance of accumulation in the target tissue. Opsonization is a major issue that induces MPS uptake of NPs, and therefore the surface characteristics and surface functionalization greatly determine their pharmacokinetic prole. In general, NPs with diameters of between 5 and 100 nm and a neutral and hydrophilic polymer extended surface exhibit prolonged blood circulation, moderate liver uptake and an increased level of drug delivery as determined by radiolabeling (PET/MR).360 This journal is The Royal Society of Chemistry 2011

4.7.

Magnetic NPs for immunotherapy

Downloaded on 20 June 2011 Published on 01 March 2011 on http://pubs.rsc.org | doi:10.1039/C0DT00689K

With the development of DNA technology, especially recombination technology, gene therapy, which transfers DNA or RNA into target cells, has become a novel approach for disease therapy.361 Finding effective transfection methods is a major objective in gene therapy research due to rapid degradation of DNA and RNA by enzymes and their poor diffusion across cell membranes.362 In addition, functionalized nanoparticles represent a promising tool for immunotherapy of tumors. Tumor cells express a wide variety of proteins that can be recognized by the immune system. They comprise tumor-specic proteins containing mutations or fusions and also tumor-associated proteins normally expressed in a developmentally or tissue-restricted fashion. These proteins represent ideal targets for therapeutic interventions as they allow the specic detection by cells or components of the immune system. Established cancer therapies employ a variety of manipulations to activate antitumor immunity. These include (i) passive immunization with monoclonal antibodies, (ii) active immunization by the application of adjuvants alone or in combination with tumor antigens, and (iii) the systemic or local delivery of cytokines. A variety of novel strategies have been developed based on fundamental advances in our understanding of the interactions between tumors and the immune system. Collectively, these strategies attempt to augment protective antitumor immunity and to disrupt the immune-regulatory circuits that are critical for maintaining tumor tolerance. Antibodies. Antibodies play an essential role in providing protective immunity to several pathogens, and the administration of tumor-targeting monoclonal antibodies has proven to be one of the most successful forms of immune therapy for cancer at the moment. The infusion of manufactured monoclonal antibodies can generate an immediate immune response while bypassing many of the limitations that impede endogenous immunity. Monoclonal antibodies have been approved for the treatment of several solid and non-solid tumors (see Table 1). The conjugation of antibodies to nanoparticles now generates a new product that combines the physico-chemical properties of nanoparticles like thermal, imaging, drug carrier, or magnetic characteristics with the ability of antibodies to specically recognize antigens on the surface of tumor cells.363 The possible applications of antibody-conjugated nanoparticles are numerous and can be divided in therapy and diagnosis. Therapeutic applications include targeted drug delivery, gene delivery, magnetic hyperthermia,

photodynamic therapy and the delivery of encapsulated molecules in vaccination strategies. An example for a therapeutic application is the use of AuFe3 O4 hetero-nanoparticles coupled with Herceptin and a platinum complex as target-specic nanocarriers for delivery of platinum into Her2-positive breast cancer cells.364 In diagnosis, the applications can be divided into those using in vivo and those using in vitro experimentation, and include contrast agents for magnetic resonance imaging (MRI), sensing, cell sorting, bioseparation, enzyme immobilization, immunoassays, transfection (gene delivery), purication, and so forth. Lee et al.33 showed that Herceptin-functionalized MFe2 O4 nanoparticles (M = Mn, Fe, Co or Ni) showed enhanced sensitivity for cancer cell detection and also made the in vivo imaging of small tumors possible. This suggests that by using appropriate cancer targeting molecules, the ultra-sensitive MR detection of various types of cancers should be possible. The improved sensitivity of assays using antibody-coated nanoparticles has also recently been shown by Thaxton et al. for the detection of the prostate specic antigen (PSA) in the serum of patients after radical prostatectomy).365,366 Here, sensitivity could be improved about 300-fold in comparison to the commercially available immunoassay used so far and allowed the prediction of disease relapse followed by adjuvant and salvage therapies at a much earlier time. Immune adjuvants. During the initial phase of infection, pathogen sensing by the immune system is based on recognition of a limited number of microbial molecular signatures, the pathogen associated molecular patterns (PAMPs) as summarized in Fig. 20.368,367 PAMPs are produced only by pathogens and not by host cells allowing the early distinction between self and microbial nonself. Many known PAMPs are typical nucleic acids or conserved components of cell wall structure from microorganisms. Their detection is mediated by a limited number of pattern recognition receptors (PRRs), which include membrane-bound receptors in the cell surface or in intracellular compartments, or soluble proteins secreted into the blood stream and tissue uids. Toll-like receptors (TLRs) represent a group of pattern recognition receptors with a great variety of different ligand specicities. Their discovery in 1997 by cloning and characterization of TLR4369 gave new impulses in the eld of innate immunity and in the understanding how innate and adaptive immunity are tightly interwoven. TLRs are type I transmembrane glycoproteins characterized by an extracellular leucine-rich repeat domain and a conserved intracellular domain, which is homologous to the cytosolic domain of the IL-1 receptor and therefore named Toll/IL-1 receptor (TIR) domain. Today, eleven different mammalian TLRs

Table 1 Monoclonal antibodies approved for tumor therapy in humans Product Rituximab (Rituxan) Trastuzumab (Herceptin R ) Gemtuzumab Ozogamicin (Mylotarg) Alemtuzumab (Campath) Cetuximab (Erbitux) Panimumab (Vectibix) Company Genentech Inc. Genentech Inc. Wyeth Averst Millennium/Ilex Partners LP ImClone Systems/Bristol-Myers Squibb/Merck KgaA Amgen Specicity Chimeric Ig anti-CD20 Humanized IgG anti-HER2 Humanized Ig anti CD33 Humanized Ig anti CD52 IgG1 , anti EGFR Human anti EGFR Disease Non-Hodgkin lymphomas breast cancer acute myeloid leukaemia chronic lymphocitic leukaemia colorectal tumor Metastatic colorectal carcinoma

This journal is The Royal Society of Chemistry 2011

Dalton Trans., 2011, 40, 63156343 | 6335

Downloaded on 20 June 2011 Published on 01 March 2011 on http://pubs.rsc.org | doi:10.1039/C0DT00689K

Fig. 20 Signalling through TLR3, TLR7, TLR8 and TLR9 in response to endosomal nucleic acids of viral origin. The Toll-like receptors (TLRs) that sense nucleic acids can operate in non-infected cells of many types to detect the production of infection in other cells. Following the recognition of viral double-stranded RNA (dsRNA), single-stranded RNA (ssRNA) or CpG-containing DNA by TLR3 or TLR7, TLR8 and TLR9 that are expressed in the endosome, signalling proceeds through TIR (Toll/interleukin-1 receptor (IL-1R))-domain-containing adaptor protein inducing interferon-b (TRIF) or myeloid differentiation primary-response gene 88 (MyD88), respectively. UNC93B is a multiple-transmembrane-spanning protein that is predominantly located in the endoplasmic reticulum (ER), but is known to associate with these endosomal TLRs and to be required for them to signal. TRIF, through the recruitment of tumour-necrosis factor receptor (TNFR)-associated factor 6 (TRAF6) and receptor-interacting protein 1 (RIP1), as well as TANK-binding kinase 1 (TBK1) and inducible IkB (inhibitor of nuclear factor-kB (NF-kB)) kinase (IKKi) activate interferon (IFN)-regulatory factor 3 (IRF3) and NF-kB. MyD88 recruits TRAF6 and IL-1R-associated kinase (IRAK) and activates IRF7 and NF-kB. TLR4 (not shown) also detects viruses, signalling in response to specic virally encoded proteins through MyD88, TRIF and/or TRAM (TRIF-related adaptor molecule). NF-kB, IRF7 and IRF3 translocate to the nucleus to induce the transcription of genes encoding cytokines such as TNF, IL-6 and type I IFNs (with permission from ref. 367).

have been identied, with their genes dispersed throughout the genome. They recognize a wide range of pathogen associated molecular patterns (PAMPs) derived from microbes and viruses such as ds-RNA, ss-DNA, and lipopolysaccharides (LPS) as well as intrinsic stress proteins. Upon ligand binding TLRs dimerize, thereby undergoing conformational changes required for the recruitment of the adaptor molecule MyD88 (myeloid differentiation primary-response protein 88). MyD88 consists of a C-terminal TIR domain, which interacts with the TIR domain of the receptor, and ultimately results in the activation of several kinases inducing phosphorylation, followed by ubiquitylation and subsequent degradation of IkB, thereby releasing NF-kB (nuclear factor-kB). NF-kB is consequently free to translocate into the nucleus and induce the expression of its target genes. Besides this MyD88-dependent signalling cascade, additional receptorproximal adaptor proteins have been described. They contribute
6336 | Dalton Trans., 2011, 40, 63156343

to a MyD88-independent activation of the NF-kB pathway and include: TIRAP (TIR-domain containing adaptor protein, also known as MyD88-adaptor-like protein, MAL),370,371 TRIF (TIRdomain-containing adaptor protein inducing IFN-b; also known as TIR-domain-containing molecule 1; TICAM1)372 and TRAM (TRIF-related adaptor molecule, also known as TIR-domaincontaining molecule 2; TICAM2)373377 In an orchestrated interplay pathogen recognition by TLRs links innate and adaptive immune responses by inducing the expression of diffusible chemotactic factors and cell surface adhesion molecules attracting innate immune cells, such as monocytes, neutrophils, basophils, eosinophils and NK cells as well as adaptive immune cells, and facilitating their migration to the inamed tissue.378 TLR triggered activation of DCs having captured microbial antigens does not only lead to the upregulation of co-stimulatory and MHC molecules but also to a switch in This journal is The Royal Society of Chemistry 2011

Downloaded on 20 June 2011 Published on 01 March 2011 on http://pubs.rsc.org | doi:10.1039/C0DT00689K

Fig. 21 Modication of manganese oxide nanoparticles with a multifunctional polymeric ligand and linkage to ssDNA. The ssDNA conjugated nanoparticled trigger the immune cascade by activating NF-kB (reproduced with permission from ref. 387).

chemokine receptor expression and to the secretion of cytokines and chemokines nally resulting in the generation of effector responses including T helper cell and CTL responses. Because of these features, TLR ligands represent a group of molecules ideally suited for the use as immune adjuvants. For example, synthetic oligodeoxynucleotides (ODNs) that contain unmethylated CpG motifs, which are present at a much higher frequency in the genomes of prokaryotes than of eukaryotes379,380 stimulate a powerful innate immune responses by interacting with TLR9.381384 Therefore, ODNs are currently evaluated for the immunotherapy of cancer, including the treatment of kidney, skin, breast, uterine and immune malignancies. Because cancer cells often express a variety of abnormal proteins that can serve as targets for an immune response (antigens) local administration of adjuvants can induce tumor-associated inammation and protective immunity. To further improve the use of TLR ligands as adjuvants immobilization on nanoparticles represents a very attractive approach and provides a tool to limit the systemic release of proinammatory cytokines associated with TLR ligand application in solution. This has been demonstrated for siRNA molecules complexed to polyethylenimine-based nanoparticles triggering DC activation in a TLR5-dependent manner,385 for CpG oligonucleotides386,387 encapsulated within liposomal nanoparticles.177 These different nanoparticles efciently activated the TLR9 signaling pathway (Fig. 21). This journal is The Royal Society of Chemistry 2011

However, optimal induction of adaptive immune responses does not only rely on efcient innate immune cell activation but requires the presentation of antigens at the same time. Nanoparticles represent an ideal tool to combine these two stimuli in newly developed vaccination protocols. This has been demonstrated for cationized gelatin nanoparticles carrying CpG oligonucleotides and the model antigen ovalbumin.388 A similar approach was taken by Uto and colleagues. Here CpG oligonucleotides and ovalbumin was immobilized using biodegradable glutamic acid nanoparticles.389 Potent induction of proinammatory cytokine release in a TRL9-dependent manner and activation of cytotoxic T cells was observed. In addition, nanoparticles have been used to activate the inammasome in an approach to induce potent immune responses. This was achieved by the incorporation of LPS on the surface of poly(lactic-co-glycolic acid) nanoparticles loaded recombinant West Nile envelope protein. Immunization of mice resulted in the protection against a murine model of West Nile encephalitis.390 However, most approaches so far immobilize TLR ligands and antigens in an unspecic manner by adsorption or encapsulation. This has the disadvantage that the exact composition of the nanoparticles is difcult to control especially when functionalization using several molecules is intended. A solution for this problem will be the development of nanoparticles carrying different functional groups on their surface and therefore allowing a controllable immobilization of TLR ligands for innate immune cell activation, antigens for adaptive immune cell activation and Dalton Trans., 2011, 40, 63156343 | 6337

molecules, like antibodies or lectins, that will allow the targeting of nanoparticles to certain cell subsets best suited for the induction of appropriate adaptive immune responses.

5.

Summary and Outlook

Functionalized nanoparticles are important platforms for multimodal imaging and targeted drug delivery. In this eld, materials scientists provide tailor-made tools for medical research, diagnosis and treatment. These tools are rationally designed to have dened functions. Still, the value of these tools can only be determined by the users in medical sciences that develop assays for applying these tools. The recent developments in the synthesis and functionalization of magnetic nanoparticles permit the use of those particles in diagnosis and therapy. The next generation of particles is expected to contain more potent uorophors, enhanced MRI contrast and various attachment sites for specialized drugs. Still, little is known about the impact of multifunctional particles that display intrinsic chemical and physical asymmetry which poses new challenges for cells associated with the amphiphilicity, dipole moments and chemical diversity/patchiness of the functionalized nanoparticles. Why is it important to study the impact of anisotropic multifunctional particles on biological cells extending the intricacy of the problem even further? Current nanotechnology projects that started during the few years focus more and more on the supramolecular weak binding of functionalized particles with the goal to form larger ensembles exhibiting novel functionalities. Thus one may anticipate new and so far untouched phenomena associated with the exposure of human tissue to the primary building blocks of these new materials. Therefore, in future work, several scientic questions need to be addressed: (i) How does multifunctionality, shape- and chemical anisotropy impact the interaction of particles with biomembranes? (ii) How do these particles enter cells? Do they exert curvature in cell membranes that inevitably leads to vesiculation due to longrange attraction as observed from viral endocytosis? (iii) What are the possible biochemical consequences for the uptake of multifunctional nanoparticles? Despite the challenges that have to met, multifunctional nanoparticles provide fascinating opportunities for tailoring properties that are not possible with other types of therapeutics. As more clinical data become available, the nanoparticle strategy may improve to such an extent that more sophisticated tools actually reach the clinic. Results from current trials are fuelling the enthusiasm of researchers.

Acknowledgements
We are grateful to Center for Complex Matter (COMATT) for support. K.S. is a recipient of a fellowship through funding of the Excellence Initiative (DFG/GSC 266). T. D. Schladt is recipient of a Carl-Zeiss Fellowship.

Notes and references


1 2 3 4 N. Nath and A. Chilkoti, Anal. Chem., 2002, 74, 504. M. Ferrari, Nat. Rev. Cancer, 2005, 5, 161. J. Cheon and J.-H. Lee, Acc. Chem. Res., 2008, 41, 1630. J. Gao, H. Gu and B. Xu, Acc. Chem. Res., 2009, 42, 1097.

5 R. Hao, R. Xing, Z. Xu, Y. Hou, S. Gao and S. Sun, Adv. Mater., 2010, 22, 2729. 6 Q. A. Pankhurst, J. Connolly, S. K. Jones and J. Dobson, J. Phys. D: Appl. Phys., 2003, 36, R167. 7 Q. A. Pankhurst, N. T. Thanh, S. K. Jones and J. Dobson, J. Phys. D: Appl. Phys., 2009, 42, 224001. 8 V. Wagner, A. Dullaart, A.-K. Bock and A. Zweck, Nat. Biotechnol., 2006, 24, 1211. 9 C. Sun, J. S. H. Lee and M. Zhang, Adv. Drug Delivery Rev., 2008, 60, 1252. 10 M. E. Davis, Z. Chen and D. M. Shin, Nat. Rev. Drug Discovery, 2008, 7, 771. 11 J. Shi, A. R. Votruba, O. C. Farokhzad and R. Langer, Nano Lett., 2010, 10, 3223. 12 A. Senyei, K. Widder and G. Czerlinski, J. Appl. Phys., 1978, 49, 3578. 13 A. Verma, O. Uzun, Y. Hu, Y. Hu, H.-S. Han, N. Watson, S. Chen, D. J. Irvine and F. Stellacci, Nat. Mater., 2008, 7, 588. 14 N. L. Rosi, D. A. Giljohann, C. S. Thaxton, A. K. R. Lytton-Jean, M. S. Han and C. A. Mirkin, Science, 2006, 312, 1027. 15 Y. Matsumura and H. Maeda, Cancer Res., 1986, 46, 6387. 16 D. Peer, J. M. Karp, S. Hong, O. C. Farokhzad, R. Margalit and R. Langer, Nat. Nanotechnol., 2007, 2, 751. 17 S. M. Moghimi, A. C. Hunter and J. C. Murray, Pharmacol. Rev., 2001, 53, 283. 18 M. Hashida, P. Opanasopit and M. Nishikawa, Crit. Rev. Ther. Drug Carrier Syst., 2002, 19, 191. 19 D. Venturoli and B. Rippe, Am. J. Physiol.: Renal Physiol., 2004, 288, F605. 20 R. H. Kodama, S. A. Makhlouf and A. E. Berkowitz, Phys. Rev. Lett., 1997, 79, 1393. 21 R. H. Kodama, J. Magn. Magn. Mater., 1999, 200, 359. 22 K. J. Klabunde, Nanoscale materials in chemistry, Wiley-Interscience, New York, NY, 2001. 23 X. Battle and A. Labarta, J. Phys. D: Appl. Phys., 2002, 35, R15. 24 D. L. Leslie-Pelecky and R. D. Rieke, Chem. Mater., 1996, 8, 1770. 25 A.-H. Lu, E. Salabas and F. Schuth, Angew. Chem., Int. Ed., 2007, 46, 1222. 26 F. Bodker, S. Morup and S. Linderoth, Phys. Rev. Lett., 1994, 72, 282. 27 M. Ghosh, K. Biswas, A. Sundaresan and C. N. R. Rao, J. Mater. Chem., 2006, 16, 106. 28 T. D. Schladt, T. Graf and W. Tremel, Chem. Mater., 2009, 21, 3183. 29 D. G. Mitchell and M. S. Cohen, MRI principles, Elsevier Saunders, Philadelphia, Pe., 2004. 30 M. A. Brown and R. C. Semelka, MRI, Wiley-Blackwell, Hoboken, 2010. 31 J. Perez, L. Josephson, T. OLoughlin, D. Hogemann and R. Weissleder, Nat. Biotech., 2002, 20, 816. 32 J.-F. Berret, N. Schonbeck, F. Gazeau, D. El Kharrat, O. Sandre, A. Vacher and M. Airiau, J. Am. Chem. Soc., 2006, 128, 1755. 33 J.-H. Lee, Y.-M. Huh, Y.-W. Jun, J.-w. Seo, J.-T. Jang, H.-T. Song, S. Kim, E.-J. Cho, H.-G. Yoon, J.-S. Suh and J. Cheon, Nat. Med., 2006, 13, 95. 34 U. I. Tromsdorf, N. C. Bigall, M. G. Kaul, O. T. Bruns, M. S. Nikolic, B. Mollwitz, R. A. Sperling, R. Reimer, H. Hohenberg, W. J. Parak, S. Forster, U. Beisiegel, G. Adam and H. Weller, Nano Lett., 2007, 7, 2422. 35 F. Eibofner, G. Steidle, R. Kehlbach, R. Bantleon and F. Schick, Magn. Reson. Med., 2010, NA. 36 U. I. Tromsdorf, O. T. Bruns, S. C. Salmen, U. Beisiegel and H. Weller, Nano Lett., 2009, 9, 4434. 37 T. \c{C}ukur, M. Yamada, W. R. Overall, P. Yang and D. G. Nishimura, Magn. Reson. Med., 2010, 63, 427. 38 K. H. Bae, Y. B. Kim, Y. Lee, J. Hwang, H. Park and T. G. Park, Bioconjugate Chem., 2010, 21, 505. 39 F. Evanics, P. R. Diamente, F. C. J. M. van Veggel, G. J. Stanisz and R. S. Prosser, Chem. Mater., 2006, 18, 2499. 40 J.-L. Bridot, A.-C. Faure, S. Laurent, C. Rivi` re, C. Billotey, B. Hiba, e M. Janier, V. Josserand, J.-L. Coll, L. V. Elst, R. Muller, S. Roux, P. Perriat and O. Tillement, J. Am. Chem. Soc., 2007, 129, 5076. 41 J. Y. Park, M. J. Baek, E. S. Choi, S. Woo, J. H. Kim, T. J. Kim, J. C. Jung, K. S. Chae, Y. Chang and G. H. Lee, ACS Nano, 2009, 3, 3663. 42 M. F. Warsi, R. W. Adams, S. B. Duckett and V. Chechik, Chem. Commun., 2010, 46, 451. 43 H. B. Na, J. H. Lee, an Kwangjin, Y. I. Park, M. Park, S. Lee, D. H. Nam, S. T. Kim, S. H. Kim, S. W. Kim, K. H. Lim, K. S.

Downloaded on 20 June 2011 Published on 01 March 2011 on http://pubs.rsc.org | doi:10.1039/C0DT00689K

6338 | Dalton Trans., 2011, 40, 63156343

This journal is The Royal Society of Chemistry 2011

44 45

46 47

48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 63 64 65 66 67 68 69 70 71 72 73 74 75 76 77 78 79 80 81 82

Kim, S. O. Kim and T. Hyeon, Angew. Chem., Int. Ed., 2007, 46, 5397. H. B. Na and T. Hyeon, J. Mater. Chem., 2009, 19, 6267. T. D. Schladt, M. I. Shukoor, K. Schneider, M. N. Tahir, F. Natalio, I. Ament, J. Becker, F. D. Jochum, S. Weber, O. Kohler, P. Theato, L. M. Schreiber, C. Sonnichsen, H. C. Schroder, W. E. Muller and W. Tremel, Angew. Chem. Int. Ed., 2010, 49, 3976. T. D. Schladt, K. Schneider, M. I. Shukoor, F. Natalio, H. Bauer, M. N. Tahir, S. Weber, L. M. Schreiber, H. C. Schroder, W. E. G. Muller and W. Tremel, J. Mater. Chem., 2010, 20, 8297. H. Yang, Y. Zhuang, H. Hu, X. Du, C. Zhang, X. Shi, H. Wu and S. Yang, Adv. Funct. Mater., 2010, 20, 1733. J. Huang, J. Xie, K. Chen, L. Bu, S. Lee, Z. Cheng, X. Li and X. Chen, Chem. Commun., 2010, 46, 6684. G. Schmid, Nanoparticles, Wiley-VCH, Weinheim, 2006. T. Hyeon, Chem. Commun., 2003, 927. R. C. OHandley, Modern magnetic materials, Wiley, New York, NY, 2000. J. Park, J. Joo, S. G. Kwon, Y. Jang and T. Hyeon, Angew. Chem., Int. Ed., 2007, 46, 4630. T. Hyeon, S. S. Lee, J. Park, Y. Chung and H. B. Na, J. Am. Chem. Soc., 2001, 123, 12798. D. Farrell, S. A. Majetich and J. P. Wilcoxon, J. Phys. Chem. B, 2003, 107, 11022. S. Peng, C. Wang, J. Xie and S. Sun, J. Am. Chem. Soc., 2006, 128, 10676. R. Grass, E. Athanassiou and W. Stark, Angew. Chem., Int. Ed., 2007, 46, 4909. D.-H. Chen and C.-H. Hsieh, J. Mater. Chem., 2002, 12, 2412. S. Sun, H. Zeng, D. B. Robinson, S. Raoux, P. M. Rice, S. X. Wang and G. Li, J. Am. Chem. Soc., 2004, 126, 273. N. Bao, L. Shen, Y. Wang, P. Padhan and A. Gupta, J. Am. Chem. Soc., 2007, 129, 12374. C. T. Yavuz, J. T. Mayo, W. W. Yu, A. Prakash, J. C. Falkner, S. Yean, L. Cong, H. J. Shipley, A. Kan, M. Tomson, D. Natelson and V. L. Colvin, Science, 2006, 314, 964. W. S. Seo, H. H. Jo, K. Lee, B. Kim, S. J. Oh and J. T. Park, Angew. Chem., Int. Ed., 2004, 43, 1115. M. Yin and S. OBrien, J. Am. Chem. Soc., 2003, 125, 10180. X. Zhong, R. Xie, L. Sun, I. Lieberwirth and W. Knoll, J. Phys. Chem. B, 2006, 110, 2. J. Park, K. An, Y. Hwang, J. G. Park, H. J. Noh, J. Y. Kim, J. H. Park, N. M. Hwang and T. Hyeon, Nat. Mater., 2004, 3, 891. Y. Chen, E. Johnson and X. Peng, J. Am. Chem. Soc., 2007, 129, 10937. M. Verelst, T. O. Ely, C. Amiens, E. Snoeck, P. Lecante, A. Mosset, M. Respaud, J. M. Broto and B. Chaudret, Chem. Mater., 1999, 11, 2702. T. He, D. Chen, X. Jiao, Y. Wang and Y. Duan, Chem. Mater., 2005, 17, 4023. M. Ghosh, E. V. Sampathkumaran and C. N. R. Rao, Chem. Mater., 2005, 17, 2348. W. S. Seo, J. H. Shim, S. J. Oh, E. K. Lee, N. H. Hur and J. T. Park, J. Am. Chem. Soc., 2005, 127, 6188. C. L. Carnes, J. Stipp, K. J. Klabunde and J. Bonevich, Langmuir, 2002, 18, 1352. H. Zeng, P. M. Rice, S. X. Wang and S. Sun, J. Am. Chem. Soc., 2004, 126, 11458. Q. Song and Z. Zhang, J. Am. Chem. Soc., 2004, 126, 6164. N. R. Jana, Y. Chen and X. Peng, Chem. Mater., 2004, 16, 3931. S. Sun, C. B. Murray, D. Weller, L. Folks and A. Moser, Science, 2000, 287, 1989. E. Shevchenko, D. Talapin, A. Kornowski, F. Wiekhorst, J. Kotzler, M. Haase, A. Rogach and H. Weller, Adv. Mater., 2002, 14, 287. L. C. Varanda and M. Jafelicci, J. Am. Chem. Soc., 2006, 128, 11062. S. Sun, Adv. Mater., 2006, 18, 393. M. Chen, J. Liu and S. Sun, J. Am. Chem. Soc., 2004, 126, 8394. C. Liu, X. Wu, T. Klemmer, N. Shukla, X. Yang, D. Weller, A. G. Roy, M. Tanase and D. Laughlin, J. Phys. Chem. B, 2004, 108, 6121. Y. Li, X. L. Zhang, R. Qiu, R. Qiao and Y. S. Kang, J. Phys. Chem. C, 2007, 111, 10747. H. M. Song, Y. J. Kim and J. H. Park, J. Phys. Chem. C, 2008, 112, 5397. E. V. Shevchenko, D. V. Talapin, A. L. Rogach, A. Kornowski, M. Haase and H. Weller, J. Am. Chem. Soc., 2002, 124, 11480.

83 E. V. Shevchenko, D. V. Talapin, H. Schnablegger, A. Kornowski, O. Festin, P. Svedlindh, M. Haase and H. Weller, J. Am. Chem. Soc., 2003, 125, 9090. 84 J. van Embden, J. E. Sader, M. Davidson and P. Mulvaney, J. Phys. Chem. C, 2009, 113, 16342. 85 V. K. LaMer and R. H. Dinegar, J. Am. Chem. Soc., 1950, 72, 4847. 86 T. Sugimoto, Monodispersed particles, Elsevier, Amsterdam, New York, 2001. 87 A. Alivisatos, J. Phys. Chem., 1996, 100, 13226. 88 J. Cheon, N.-J. Kang, S.-M. Lee, J.-H. Lee, J.-H. Yoon and S. J. Oh, J. Am. Chem. Soc., 2004, 126, 1950. 89 J. Rockenberger, E. C. Scher and A. P. Alivisatos, J. Am. Chem. Soc., 1999, 121, 11595. 90 Y.-W. Jun, J.-S. Choi and J. Cheon, Chem. Commun., 2007, 1203. 91 Z. Zhang, Z. L. Wang, B. C. Chakoumakos and J. S. Yin, J. Am. Chem. Soc., 1998, 120, 1800. 92 S. Neveu, A. Bee, M. Robineau and D. Talbot, J. Colloid Interface Sci., 2002, 255, 293. 93 Y. S. Kang, S. Risbud, J. F. Rabolt and P. Stroeve, Chem. Mater., 1996, 8, 2209. 94 P. C. Kuo and T. S. Tsai, J. Appl. Phys., 1989, 65, 4349. 95 C.-Y. Hong, I. J. Jang, H. E. Horng, C. J. Hsu, Y. D. Yao and H. C. Yang, J. Appl. Phys., 1997, 81, 4275. 96 T. Fried, G. Shemer and G. Markovich, Adv. Mater., 2001, 13, 1158. 97 X. Lu, M. Niu, R. Qiao and M. Gao, J. Phys. Chem. B, 2008, 112, 14390. 98 B. L. Cushing, V. L. Kolesnichenko and C. J. OConnor, Chem. Rev., 2004, 104, 3893. 99 A. L. Willis, N. J. Turro and S. OBrien, Chem. Mater., 2005, 17, 5970. 100 J. Lu, X. Jiao, D. Chen and W. Li, J. Phys. Chem. C, 2009, 113, 4012. 101 W. Zhang, Z. Yang, Y. Liu, S. Tang, X. Han and M. Chen, J. Cryst. Growth, 2004, 263, 394. 102 C. Rath, K. Sahu, S. Anand, S. Date, N. Mishra and R. Das, J. Magn. Magn. Mater., 1999, 202, 77. 103 B. Baruwati, M. N. Nadagouda and R. S. Varma, J. Phys. Chem. C, 2008, 112, 18399. 104 F. Cansell, B. Chevalier, A. Demourgues, J. Etourneau, C. Even, V. Pessey, S. Petit, A. Tressaud and F. Weill, J. Mater. Chem., 1999, 9, 67. 105 U. K. Gautam, M. Ghosh, M. Rajamathi and R. Seshadri, Pure Appl. Chem., 2002, 74, 1643. 106 T. Daou, G. Pourroy, S. B gin-Colin, J. Gren` che, C. Ulhaq-Bouillet, e e P. Legar , P. Bernhardt, C. Leuvrey and G. Rogez, Chem. Mater., e 2006, 18, 4399. 107 C.-J. Jia, L.-D. Sun, F. Luo, X.-D. Han, L. J. Heyderman, Z.-G. Yan, C.-H. Yan, K. Zheng, Z. Zhang, M. Takano, N. Hayashi, M. Eltschka, M. Kl ui, U. Rudiger, T. Kasama, L. Cervera-Gontard, R. E. Dunina Borkowski, G. Tzvetkov and J. Raabe, J. Am. Chem. Soc., 2008, 130, 16968. 108 M. Rajamathi and R. Seshadri, Curr. Opin. Solid State Mater. Sci., 2002, 6, 337. 109 G. Demazeau, J. Mater. Chem., 1999, 9, 15. 110 M. A. Willard, L. K. Kurihara, E. E. Carpenter, S. Calvin and V. G. Harris, Int. Mater. Rev., 2004, 49, 125. 111 S. Ge, X. Shi, K. Sun, C. Li, C. Uher, J. R. Baker, M. M. Banaszak Holl and B. G. Orr, J. Phys. Chem. C, 2009, 113, 13593. 112 J. Lee, T. Isobe and M. Senna, Colloids Surf., A, 1996, 109, 121. 113 T. P. Hoar and J. H. Schulman, Nature, 1943, 152, 102. 114 G. Gillberg, H. Lehtinen and S. Friberg, J. Colloid Interface Sci., 1970, 33, 40. 115 D. Langevin, Annu. Rev. Phys. Chem., 1992, 43, 341. 116 B. K. Paul and S. P. Moulik, Curr. Sci., 2001, 80, 990. 117 A. K. Gupta and M. Gupta, Biomaterials, 2005, 26, 3995. 118 D.-H. Chen and S.-H. Wu, Chem. Mater., 2000, 12, 1354. 119 Y. Lee, J. Lee, C. Bae, J.-G. Park, H.-J. Noh, J.-H. Park and T. Hyeon, Adv. Funct. Mater., 2005, 15, 503. 120 C. Liu, B. Zou, A. J. Rondinone and Z. Zhang, J. Phys. Chem. B, 2000, 104, 1141. 121 C. R. Vestal and Z. Zhang, Chem. Mater., 2002, 14, 3817. 122 P. Xu, X. Han and M. Wang, J. Phys. Chem. C, 2007, 111, 5866. 123 K. M. Taylor, W. J. Rieter and W. Lin, J. Am. Chem. Soc., 2008, 130, 14358. 124 C. B. Murray, D. J. Norris and M. Bawendi, J. Am. Chem. Soc., 1993, 115, 8706. 125 V. F. Puntes, K. M. Krishnan and A. Alivisatos, Science, 2001, 291, 2115.

Downloaded on 20 June 2011 Published on 01 March 2011 on http://pubs.rsc.org | doi:10.1039/C0DT00689K

This journal is The Royal Society of Chemistry 2011

Dalton Trans., 2011, 40, 63156343 | 6339

126 S. Stoeva, K. J. Klabunde, C. M. Sorensen and I. Dragieva, J. Am. Chem. Soc., 2002, 124, 2305. 127 D. P. Dinega and M. Bawendi, Angew. Chem., Int. Ed., 1999, 38, 1788. 128 F. X. Redl, C. T. Black, G. C. Papaefthymiou, R. L. Sandstrom, M. Yin, H. Zeng, C. B. Murray and S. P. OBrien, J. Am. Chem. Soc., 2004, 126, 14583. 129 C. B. Murray, C. R. Kagan and M. Bawendi, Annu. Rev. Mater. Sci., 2000, 30, 545. 130 A. C. S. Samia, K. Hyzer, J. A. Schlueter, C.-J. Qin, J. S. Jiang, S. D. Bader and X.-M. Lin, J. Am. Chem. Soc., 2005, 127, 4126. 131 V. F. Puntes, K. M. Krishnan and P. Alivisatos, Appl. Phys. Lett., 2001, 78, 2187. 132 S. Sun and C. B. Murray, J. Appl. Phys., 1999, 85, 4325. 133 T. Cedervall, I. Lynch, S. Lindman, T. Berggard, E. Thulin, H. Nilsson, K. A. Dawson and S. Linse, Proc. Natl. Acad. Sci. U. S. A., 2007, 104, 2050. 134 M. Bruchez, JR. , M. Moronne, P. Gin, S. Weiss and A. P. Alivisatos, Science, 1998, 281, 2013. 135 W. Chan and S. Nie, Science, 1998, 281, 2016. 136 X. Michalet, F. F. Pinaud, L. A. Bentolila, J. M. Tsay, S. Doose, J. J. Li, G. Sundaresan, A. M. Wu, S. S. Gambhir and S. Weiss, Science, 2005, 307, 538. 137 J. Qin, S. Laurent, Y. Jo, A. Roch, M. Mikhaylova, Z. Bhujwalla, R. Muller and M. Muhammed, Adv. Mater., 2007, 19, 1874. 138 R. de Palma, S. Peeters, M. J. van Bael, H. Van Den Rul, K. Bonroy, W. Laureyn, J. Mullens, G. Borghs and G. Maes, Chem. Mater., 2007, 19, 1821. 139 L. Shen, P. E. Laibinis and T. A. Hatton, Langmuir, 1999, 15, 447. 140 M. H. Sousa, F. A. Tourinho, J. Depeyrot, G. J. da Silva and M. C. F. L. Lara, J. Phys. Chem. B, 2001, 105, 1168. 141 J. Kim, H. S. Kim, N. Lee, T. Kim, H. Kim, T. Yu, I. C. Song, W. K. Moon and T. Hyeon, Angew. Chem., Int. Ed., 2008, 47, 8438. 142 Y.-W. Jun, J.-H. Lee and J. Cheon, Angew. Chem., Int. Ed., 2008, 47, 5122. 143 A. Wooding, M. Kilner and D. B. Lambrick, J. Colloid Interface Sci., 1991, 144, 236. 144 B. Dubertret, P. Skourides, D. J. Norris, V. Noireaux, A. H. Brivanlou and A. Libchaber, Science, 2002, 298, 1759. 145 N. Nitin, L. LaConte, O. Zurkiya, X. Hu and G. Bao, JBIC, J. Biol. Inorg. Chem., 2004, 9, 706. 146 I. L. Medintz, H. Uyeda, E. R. Goldman and H. Mattoussi, Nat. Mater., 2005, 4, 435. 147 J. Qin, Y. Jo and M. Muhammed, Angew. Chem., Int. Ed., 2009, 48, 7845. 148 W. S. Seo, J. H. Lee, X. Sun, Y. Suzuki, D. Mann, Z. Liu, M. Terashima, P. C. Yang, M. V. McConnell, D. G. Nishimura and H. Dai, Nat. Mater., 2006, 5, 971. 149 J. Shin, R. Anisur, M. Ko, G. Im, J. Lee and I. Lee, Angew. Chem., Int. Ed., 2009, 48, 321. 150 D. B. Robinson, H. H. Persson, H. Zeng, G. Li, N. Pourmand, S. Sun and S. X. Wang, Langmuir, 2005, 21, 3096. 151 S.-W. Kim, S. Kim, J. B. Tracy, A. Jasanoff and M. G. Bawendi, J. Am. Chem. Soc., 2005, 127, 4556. 152 M. Lattuada and T. A. Hatton, Langmuir, 2007, 23, 2158. 153 W. W. Yu, E. Chang, C. M. Sayes, R. Drezek and V. L. Colvin, Nanotechnology, 2006, 17, 4483. 154 W. W. Yu, E. Chang, J. C. Falkner, J. Zhang, A. M. Al-Somali, C. M. Sayes, J. Johns, R. Drezek and V. L. Colvin, J. Am. Chem. Soc., 2007, 129, 2871. 155 J. M. Harris, Poly(ethylene glycol) chemistry, Plenum Press, New York, 1992. 156 R. Gref, Y. Minamitake, M. T. Peracchia, V. Trubetskoy, V. Torchilin and R. Langer, Science, 1994, 263, 1600. 157 D. Bazile, C. Prudhomme, M. T. Bassoullet, M. Marlard, G. Spenlehauer and M. Veillard, J. Pharm. Sci., 1995, 84, 493. 158 K. Knop, R. Hoogenboom, D. Fischer and U. Schubert, Angew. Chem., Int. Ed., 2010, 49, 6288. 159 R. G. Pearson, J. Am. Chem. Soc., 1963, 85, 3533. 160 M. Wan and J. Li, J. Polym. Sci., Part A: Polym. Chem., 1998, 36, 2799. 161 M. D. Butterworth, S. A. Bell, S. P. Armes and A. W. Simpson, J. Colloid Interface Sci., 1996, 183, 91. 162 P. Tartaj, M. P. Morales, T. Gonz lez-Carreno, S. Veintemillasa Verdaguer and C. J. Serna, J. Magn. Magn. Mater., 2005, 290291, 28.

163 G. Barratt, Cell. Mol. Life Sci., 2003, 60, 21. 164 T. Xia, M. Kovochich, M. Liong, H. Meng, S. Kabehie, S. George, J. I. Zink and A. E. Nel, ACS Nano, 2009, 3, 3273. 165 D. R. Radu, C.-Y. Lai, K. Jeftinija, E. W. Rowe, S. Jeftinija and V. S. Y. Lin, J. Am. Chem. Soc., 2004, 126, 13216. 166 L. Josephson, C. H. Tung, A. Moore and R. Weissleder, Bioconjugate Chem., 1999, 10, 186. 167 D.-L. Zhao, X.-X. Wang, X.-W. Zeng, Q.-S. Xia and J.-T. Tang, J. Alloys Compd., 2009, 477, 739. 168 M. Mahmoudi, A. Simchi, M. Imani and U. O. H feli, J. Phys. Chem. a C, 2009, 113, 8124. 169 H. Uyeda, I. L. Medintz, J. K. Jaiswal, S. M. Simon and H. Mattoussi, J. Am. Chem. Soc., 2005, 127, 3870. 170 M. S. Nikolic, M. Krack, V. Aleksandrovic, A. Kornowski, S. Forster and H. Weller, Angew. Chem., 2006, 118, 6727. 171 N. A. Frey, S. Peng, K. Cheng and S. Sun, Chem. Soc. Rev., 2009, 38, 2532. 172 J. H. Waite and M. L. Tanzer, Science, 1981, 212, 1038. 173 C. Xu, K. Xu, H. Gu, R. Zheng, H. Liu, X. Zhang, Z. Guo and B. Xu, J. Am. Chem. Soc., 2004, 126, 9938. 174 M. Shukoor, F. Natalio, V. Ksenofontov, M. Tahir, M. Eberhardt, P. Theato, H. Schroder, W. Muller and W. Tremel, Small, 2007, 3, 1374. 175 M. I. Shukoor, F. Natalio, M. N. Tahir, V. Ksenofontov, H. A. Therese, P. Theato, H. C. Schroder, W. E. G. Muller and W. Tremel, Chem. Commun., 2007, 4677. 176 H. C. Schroder, F. Natalio, M. Wiens, M. N. Tahir, M. I. Shukoor, W. Tremel, S. I. Belikov, A. Krasko and W. E. G. Muller, Mol. Immunol., 2008, 45, 945. 177 M. Shukoor, F. Natalio, N. Metz, N. Glube, M. Tahir, H. Therese, V. Ksenofontov, P. Theato, P. Langguth, J. P. Boissel, H. C. Schroder, W. E. G. Muller and W. Tremel, Angew. Chem., Int. Ed., 2008, 47, 4748. 178 M. I. Shukoor, F. Natalio, H. A. Therese, M. N. Tahir, V. Ksenofontov, M. Panthofer, M. Eberhardt, P. Theato, H. C. Schroder, W. E. G. Muller and W. Tremel, Chem. Mater., 2008, 20, 3567. 179 M. I. Shukoor, F. Natalio, M. N. Tahir, M. Divekar, N. Metz, H. A. Therese, P. Theato, V. Ksenofontov, H. C. Schroder, W. E. G. Muller and W. Tremel, J. Magn. Magn. Mater., 2008, 320, 2339. 180 C. R. Vestal and Z. Zhang, J. Am. Chem. Soc., 2002, 124, 14312. 181 Y. Wang, X. Teng, J.-S. Wang and H. Yang, Nano Lett., 2003, 3, 789. 182 S. Edmondson, V. L. Osborne and W. T. S. Huck, Chem. Soc. Rev., 2004, 33, 14. 183 A. Bourlinos, A. Bakandritsos, V. Georgakilas and D. Petridis, Chem. Mater., 2002, 14, 3226. 184 D. Q. Vo, E.-J. Kim and S. Kim, J. Colloid Interface Sci., 2009, 337, 75. 185 E. S. Gawalt, M. J. Avaltroni, M. P. Danahy, B. M. Silverman, E. L. Hanson, K. S. Midwood, J. E. Schwarzbauer and J. Schwartz, Langmuir, 2003, 19, 71477147. 186 C. A. Traina and J. Schwartz, Langmuir, 2007, 23, 9158. 187 M. G. Warner, S. M. Reed and J. E. Hutchison, Chem. Mater., 2000, 12, 3316. 188 V. Salgueirino-Maceira, L. M. Liz-Marz n and M. Farle, Langmuir, a 2004, 20, 6946. 189 J. Xie, C. Xu, Z. Xu, Y. Hou, K. L. Young, S. X. Wang, N. Pourmand and S. Sun, Chem. Mater., 2006, 18, 5401. 190 H. Gu, Z. Yang, J. Gao, C. Chang and B. Xu, J. Am. Chem. Soc., 2005, 127, 34. 191 J. Xie, C. Xu, N. Kohler, Y. Hou and S. Sun, Adv. Mater., 2007, 19, 3163. 192 E. Vaccaro and J. H. Waite, Biomacromolecules, 2001, 2, 906. 193 N. Holten-Andersen, G. E. Fantner, S. Hohlbauch, J. H. Waite and F. W. Zok, Nat. Mater., 2007, 6, 669. 194 N. Holten-Andersen, T. E. Mates, M. S. Toprak, G. D. Stucky, F. W. Zok and J. H. Waite, Langmuir, 2009, 25, 3323. 195 M. J. Harrington, A. Masic, N. Holten-Andersen, J. H. Waite and P. Fratzl, Science, 2010, 328, 216. 196 C. Xu, K. Xu, H. Gu, X. Zhong, Z. Guo, R. Zheng, X. Zhang and B. Xu, J. Am. Chem. Soc., 2004, 126, 3392. 197 M. D. Shultz, J. Reveles, S. N. Khanna and E. E. Carpenter, J. Am. Chem. Soc., 2007, 129, 2482. 198 R. Hong, N. O. Fischer, T. Emrick and V. M. Rotello, Chem. Mater., 2005, 17, 4617. 199 L. X. Chen, T. Liu, M. C. Thurnauer, R. Csencsits and T. Rajh, J. Phys. Chem. B, 2002, 106, 8539.

Downloaded on 20 June 2011 Published on 01 March 2011 on http://pubs.rsc.org | doi:10.1039/C0DT00689K

6340 | Dalton Trans., 2011, 40, 63156343

This journal is The Royal Society of Chemistry 2011

200 A. K. Boal, K. Das, M. Gray and V. M. Rotello, Chem. Mater., 2002, 14, 2628. 201 K. Somaskandan, T. Veres, M. Niewczas and B. Simard, New J. Chem., 2008, 32, 201. 202 J. Gao, G. Liang, J. S. Cheung, Y. Pan, Y. Kuang, F. Zhao, B. Zhang, X. Zhang, E. X. Wu and B. Xu, J. Am. Chem. Soc., 2008, 130, 11828. 203 B. Wang, C. Xu, J. Xie, Z. Yang and S. Sun, J. Am. Chem. Soc., 2008, 130, 14436. 204 K. Cheng, S. Peng, C. Xu and S. Sun, J. Am. Chem. Soc., 2009, 131, 10637. 205 E. Amstad, T. Gillich, I. Bilecka, M. Textor and E. Reimhult, Nano Lett., 2009, 9, 4042. 206 S. Zurcher, D. W ckerlin, Y. Bethuel, B. Malisova, M. Textor, S. Tosatti a and K. Gademann, J. Am. Chem. Soc., 2006, 128, 1064. 207 E. Amstad, S. Zurcher, A. Mashaghi, J. Y. Wong, M. Textor and E. Reimhult, Small, 2009, 5, 1334. 208 H. S. Nalwa, Handbook of surfaces and interfaces of materials, Academic Press, San Diego, Calif., 2001. 209 A. Guerrero-Martnez, J. P rez-Juste and L. M. Liz-Marz n, Adv. e a Mater., 2010, 22, 1182. 210 D. Knopp, D. Tang and R. Niessner, Anal. Chim. Acta, 2009, 647, 14. 211 I. K. Herrmann, R. N. Grass, D. Mazunin and W. J. Stark, Chem. Mater., 2009, 21, 3275. 212 Y. Piao, A. Burns, J. Kim, U. Wiesner and T. Hyeon, Adv. Funct. Mater., 2008, 18, 3745. 213 S. T. Selvan, T. T. Y. Tan, D. K. Yi and N. R. Jana, Langmuir, 2010, 26, 11631. 214 W. Stober, A. Fink and E. Bohn, J. Colloid Interface Sci., 1968, 26, 62. 215 M. Ohmori and E. Matijevic, J. Colloid Interface Sci., 1993, 160, 288. 216 L. M. Liz-Marz n and A. P. Philipse, J. Colloid Interface Sci., 1995, a 176, 459. 217 H. C. Schroder, X. Wang, W. Tremel, H. Ushijima and W. E. G. Muller, Nat. Prod. Rep., 2008, 25, 455. 218 E. P. Plueddemann, Silane coupling agents, Plenum Press, New York, 1991. 219 S.-Y. Chang, L. Liu and S. A. Asher, J. Am. Chem. Soc., 1994, 116, 6745. 220 S. Santra, R. Tapec, N. Theodoropoulou, J. Dobson, A. Hebard and W. Tan, Langmuir, 2001, 17, 2900. 221 S. Santra, H. Yang, D. Dutta, J. T. Stanley, P. H. Holloway, W. Tan, B. M. Moudgil and R. A. Mericle, Chem. Commun., 2004, 2810. 222 D. K. Yi, S. Selvan, S. S. Lee, G. C. Papaefthymiou, D. Kundaliya and J. Y. Ying, J. Am. Chem. Soc., 2005, 127, 4990. 223 M. Li, H. Schnablegger and S. Mann, Nature, 1999, 402, 393. 224 A. van Blaaderen and A. Vrij, Langmuir, 1992, 8, 2921. 225 A. Imhof, M. Megens, J. Engelberts, D. de Lang, R. Sprik and W. Vos, J. Phys. Chem. B, 1999, 103, 1408. 226 Z. Li and E. Ruckenstein, Nano Lett., 2004, 4, 1463. 227 A. Burns, H. Ow and U. Wiesner, Chem. Soc. Rev., 2006, 35, 1028. 228 L. Wang, K. Wang, S. Santra, X. Zhao, L. R. Hilliard, J. E. Smith, Y. Wu and W. Tan, Anal. Chem., 2006, 78, 646. 229 J. E. Fuller, G. T. Zugates, L. S. Ferreira, H. S. Ow, N. N. Nguyen, U. B. Wiesner and R. S. Langer, Biomaterials, 2008, 29, 1526. 230 A. A. Burns, J. Vider, H. Ow, E. Herz, O. Penate-Medina, M. Baumgart, S. M. Larson, U. Wiesner and M. Bradbury, Nano Lett., 2009, 9, 442. 231 Y. Deng, D. Qi, C. Deng, X. Zhang and D. Zhao, J. Am. Chem. Soc., 2008, 130, 28. 232 T. Sen, A. Sebastianelli and I. J. Bruce, J. Am. Chem. Soc., 2006, 128, 7130. 233 Q. He, Z. Zhang, Y. Gao, J. Shi and Y. Li, Small, 2009, 5, 2722. 234 K. K. Coti, M. E. Belowich, M. Liong, M. W. Ambrogio, Y. A. Lau, H. A. Khatib, J. I. Zink, N. M. Khashab and J. F. Stoddart, Nanoscale, 2009, 1, 16. 235 Y. Zhu, T. Ikoma, N. Hanagata and S. Kaskel, Small, 2010, 6, 471. 236 T. Suteewong, H. Sai, J. Lee, M. Bradbury, T. Hyeon, S. M. Gruner and U. Wiesner, J. Mater. Chem., 2010, 20, 7807. 237 J. Lee, Y. Lee, J. K. Youn, H. B. Na, T. Yu, H. Kim, S.-M. Lee, Y.-M. Koo, J. H. Kwak, H. G. Park, H. N. Chang, M. Hwang, J.-G. Park, J. Kim and T. Hyeon, Small, 2008, 4, 143. 238 C. Gao and S. Che, Adv. Funct. Mater., 2010. 239 A. D. Bangham and R. W. Horne, J. Mol. Biol., 1964, 8, 660668, IN2-IN10.

240 S. A. Wickline, A. M. Neubauer, P. M. Winter, S. D. Caruthers and G. M. Lanza, J. Magn. Reson. Imaging, 2007, 25, 667. 241 C. Corot, K. G. Petry, R. Trivedi, A. Saleh, C. Jonkmanns, J.-F. Le Bas, E. Blezer, M. Rausch, B. Brochet, P. Foster-Gareau, D. Bal riaux, e S. Gaillard and V. Dousset, Invest. Radiol., 2004, 39, 619. 242 A. S. Lubbe, C. Bergemann, H. Riess, F. Schriever, P. Reichardt, K. Possinger, M. Matthias, B. Dorken, F. Herrmann, R. Gurtler, P. Hohenberger, N. Haas, R. Sohr, B. Sander, A.-J. Lemke, D. Ohlendorf, W. Huhnt and D. Huhn, Cancer Res., 1996, 56, 4686. 243 A. S. Lubbe, C. Alexiou and C. Bergemann, J. Surg. Res., 2001, 95, 200. 244 C. Chouly, D. Pouliquen, I. Lucet, J. Jeune and P. Jallet, J. Microencap.: Micro Nano Carr., 1996, 13, 245. 245 H. S. Choi, B. I. Ipe, P. Misra, J. H. Lee, M. G. Bawendi and J. V. Frangioni, Nano Lett., 2009, 9, 2354. 246 A. Verma and F. Stellacci, Small, 2010, 6, 12. 247 E. K. Larsen, T. Nielsen, T. Wittenborn, H. Birkedal, T. Vorup-Jensen, M. H. Jakobsen, L. Astergaard, M. R. Horsman, F. Besenbacher, K. A. Howard and J. Kjems, ACS Nano, 2009, 3, 1947. 248 A. Tanimoto and S. Kuribayashi, Eur. J. Radiol., 2006, 58, 200. 249 H. Soo Choi, W. Liu, P. Misra, E. Tanaka, J. P. Zimmer, B. Itty Ipe, M. G. Bawendi and J. V. Frangioni, Nat. Biotechnol., 2007, 25, 1165. 250 M. Massignani, C. LoPresti, A. Blanazs, J. Madsen, S. P. Armes, A. L. Lewis and G. Battaglia, Small, 2009, 5, 2424. 251 Y. Zhang, M. Yang, J.-H. Park, J. Singelyn, H. Ma, M. J. Sailor, E. Ruoslahti, M. Ozkan and C. Ozkan, Small, 2009, 5, 1990. 252 R. R. Arvizo, O. R. Miranda, M. A. Thompson, C. M. Pabelick, R. Bhattacharya, J. Robertson, V. M. Rotello, Y. Prakash and P. Mukherjee, Nano Lett., 2010, 10, 2543. 253 T. Fujita, M. Nishikawa, Y. Ohtsubo, J. Ohno, Y. Takakura, H. Sezaki and M. Hashida, J. Drug Targeting, 1994, 2, 157. 254 M. I. Papisov, A. Bogdanov, JR. , B. Schaffer, N. Nossiff, T. Shen, R. Weissleder and T. J. Brady, J. Magn. Magn. Mater., 1993, 122, 383. 255 I. Lynch, A. Salvati and K. A. Dawson, Nat. Nanotechnol., 2009, 4, 546. 256 B. Muthusamy, G. Hanumanthu, S. Suresh, B. Rekha, D. Srinivas, L. Karthik, B. M. Vrushabendra, S. Sharma, G. Mishra, P. Chatterjee, K. S. Mangala, H. N. Shivashankar, K. N. Chandrika, N. Desphande, M. Suresh, N. Kannabiran, V. Niranjan, A. Nalli, T. S. Prasad, K. S. Arun, R. Reddy, S. Chandran, T. Jadhav, D. Julie, M. Mahesh, S. L. John, K. Palvankar, D. Sudhir, P. Bala, N. S. Rashmi, G. Vishnupriya, K. Dhar, S. Reshma, R. Chaerkady, T. K. Gandhi, H. C. Harsha, S. S. Mohan, S. Desphande, M. Sarker and A. Pandey, Proteomics, 2005, 5, 3531. 257 C. Rocker, M. Potzl, F. Zhang, W. J. Parak and G. U. Nienhaus, Nat. Nanotechnol., 2009, 4, 577. 258 T. Cedervall, I. Lynch, M. Foy, T. Berggard, S. C. Donnelly, G. Cagney, S. Linse and K. A. Dawson, Angew. Chem., Int. Ed., 2007, 46, 5754. 259 B. Ballou, B. Lagerholm, L. A. Ernst, M. P. Bruchez and A. S. Waggoner, Bioconjugate Chem., 2004, 15, 79. 260 S. H. Bhang, N. Won, T.-J. Lee, H. Jin, J. Nam, J. Park, H. Chung, H.-S. Park, Y.-E. Sung, S. K. Hahn, B.-S. Kim and S. Kim, ACS Nano, 2009, 3, 1389. 261 T. Pons, E. Pic, N. Lequeux, E. Cassette, L. Bezdetnaya, F. Guillemin, F. Marchal and B. Dubertret, ACS Nano, 2010, 4, 2531. 262 H. Maeda, J. Wu, T. Sawa, Y. Matsumura and K. Hori, J. Controlled Release, 2000, 65, 271. 263 H. Maeda, Adv. Enzyme Regul., 2001, 41, 189. 264 O. Cl ment, N. Siauve, M. Lewin, E. de Kerviler, C.-A. Cu nod and e e G. Frija, Biomed. Pharmacother., 1998, 52, 51. 265 D. D. Stark, R. Weissleder, G. Elizondo, P. F. Hahn, S. Saini, L. E. Todd, J. Wittenberg and J. T. Ferrucci, Radiology, 1988, 168, 297. 266 C. Corot, P. Robert, J.-M. Id e and M. Port, Adv. Drug Delivery Rev., e 2006, 58, 1471. 267 X. Gao, Y. Cui, R. M. Levenson, L. W. K. Chung and S. Nie, Nat. Biotechnol., 2004, 22, 969. 268 W. Cai, D.-W. Shin, K. Chen, O. Gheysens, Q. Cao, S. X. Wang, S. S. Gambhir and X. Chen, Nano Lett., 2006, 6, 669. 269 R. Weissleder, K. Kelly, E. Y. Sun, T. Shtatland and L. Josephson, Nat. Biotechnol., 2005, 23, 1418. 270 A. M. Smith, H. Duan, A. M. Mohs and S. Nie, Adv. Drug Delivery Rev., 2008, 60, 1226. 271 M. Liong, J. Lu, M. Kovochich, T. Xia, S. G. Ruehm, A. E. Nel, F. Tamanoi and J. I. Zink, ACS Nano, 2008, 2, 889.

Downloaded on 20 June 2011 Published on 01 March 2011 on http://pubs.rsc.org | doi:10.1039/C0DT00689K

This journal is The Royal Society of Chemistry 2011

Dalton Trans., 2011, 40, 63156343 | 6341

272 H. S. Choi, W. Liu, F. Liu, K. Nasr, P. Misra, M. G. Bawendi and J. V. Frangioni, Nat. Nanotechnol., 2009, 5, 42. 273 J. Rosenholm, C. Sahlgren and M. Linden, J. Mater. Chem., 2010, 20, 2707. 274 S. Cerdan, H. R. Lotscher, B. Kunnecke and J. Seelig, Magn. Reson. Med., 1989, 12, 151. 275 L. X. Tiefenauer, G. Kuehne and R. Y. Andres, Bioconjugate Chem., 1993, 4, 347. 276 F. X. Gu, R. Karnik, A. Z. Wang, F. Alexis, E. Levy-Nissenbaum, S. Hong, R. S. Langer and O. C. Farokhzad, Nano Today, 2007, 2, 14. 277 J. Lee, Y. Choi, K. Kim, S. Hong, H.-Y. Park, T. Lee, G. J. Cheon and R. Song, Bioconjugate Chem., 2010, 21, 940. 278 D. Artemov, N. Mori, R. Ravi and Z. M. Bhujwalla, Cancer Res., 2003, 63, 2723. 279 Y.-M. Huh, Y.-W. Jun, H.-T. Song, S. Kim, J.-S. Choi, J.-H. Lee, S. Yoon, K.-S. Kim, J.-S. Shin, J.-S. Suh and J. Cheon, J. Am. Chem. Soc., 2005, 127, 12387. 280 S. Nie, Y. Xing, G. J. Kim and J. W. Simons, Annu. Rev. Biomed. Eng., 2007, 9, 257. 281 J. Ross, P. Chaudhuri and M. Ratnam, Cancer, 1994, 73, 2432. 282 Y. Zhang, N. Kohler and M. Zhang, Biomaterials, 2002, 23, 1553. 283 N. Kohler, G. E. Fryxell and M. Zhang, J. Am. Chem. Soc., 2004, 126, 7206. 284 S. Wang, X. Shi, M. Van Antwerp, Z. Cao, S. Swanson, X. Bi and J. Baker, Adv. Funct. Mater., 2007, 17, 3043. 285 F. Sonvico, S. Mornet, S. Vasseur, C. Dubernet, D. Jaillard, J. Degrouard, J. Hoebeke, E. Duguet, P. Colombo and P. Couvreur, Bioconjugate Chem., 2005, 16, 1181. 286 P. Wunderbaldinger, L. Josephson and R. Weissleder, Bioconjugate Chem., 2002, 13, 264. 287 M. Lewin, N. Carlesso, C. H. Tung, X. W. Tang, D. Cory, D. T. Scadden and R. Weissleder, Nat. Biotechnol., 2000, 18, 410. 288 J. S. Wadia and S. F. Dowdy, Adv. Drug Delivery Rev., 2005, 57, 579. 289 J. Panyam and V. Labhasetwar, Adv. Drug Delivery Rev., 2003, 55, 329. 290 C. K. Kim, P. Ghosh, C. Pagliuca, Z.-J. Zhu, S. Menichetti and V. M. Rotello, J. Am. Chem. Soc., 2009, 131, 1360. 291 D. A. LaVan, T. McGuire and R. Langer, Nat. Biotechnol., 2003, 21, 1184. 292 S. Y. Kim, I. G. Shin, Y. M. Lee, C. S. Cho and Y. K. Sung, J. Controlled Release, 1998, 51, 13. 293 A. Chilkoti, M. R. Dreher, D. E. Meyer and D. Raucher, Adv. Drug Delivery Rev., 2002, 54, 613. 294 O. Boussif, F. Lezoualch, M. A. Zanta, M. D. Mergny, D. Scherman, B. Demeneix and J. P. Behr, Proc. Natl. Acad. Sci. U. S. A., 1995, 92, 7297. 295 A. Asati, S. Santra, C. Kaittanis and J. M. Perez, ACS Nano, 2010, 4, 5321. 296 M. A. Kay, J. C. Glorioso and L. Naldini, Nat. Med., 2001, 7, 33. 297 V. Sokolova and M. Epple, Angew. Chem., Int. Ed., 2008, 47, 1382. 298 K. A. Whitehead, R. Langer and D. G. Anderson, Nat. Rev. Drug Discovery, 2009, 8, 129. 299 A. K. Shalek, J. T. Robinson, E. S. Karp, J. S. Lee, D.-R. Ahn, M.-H. Yoon, A. Sutton, M. Jorgolli, R. S. Gertner, T. S. Gujral, G. MacBeath, E. G. Yang and H. Park, Proc. Natl. Acad. Sci. U. S. A., 2010, 107, 1870. 300 M. S. Yavuz, Y. Cheng, J. Chen, C. M. Cobley, Q. Zhang, M. Rycenga, J. Xie, C. Kim, K. H. Song, A. G. Schwartz, L. V. Wang and Y. Xia, Nat. Mater., 2009, 8, 935. 301 S. Nappini, F. B. Bombelli, M. Bonini, B. Norden and P. Baglioni, Soft Matter, 2010, 6, 154. 302 A. N. Zelikin, ACS Nano, 2010, 4, 2494. 303 R. Weissleder, C.-H. Tung, U. Mahmood and A. Bogdanov, Nat. Biotechnol., 1999, 17, 375. 304 L. M. Bareford and P. W. Swaan, Adv. Drug Delivery Rev., 2007, 59, 748. 305 C.-H. Tung, U. Mahmood, S. Bredow and R. Weissleder, Cancer Res., 2000, 60, 4953. 306 J. You, G. Zhang and C. Li, ACS Nano, 2010, 4, 1033. 307 C. R. Gordijo, A. J. Shuhendler and X. Y. Wu, Adv. Funct. Mater., 2010, 20, 1404. 308 B. P. Timko, T. Dvir and D. S. Kohane, Adv. Mater., 2010, 22, 4925. 309 R. Weissleder, A. Bogdanov, E. A. Neuwelt and M. Papisov, Adv. Drug Delivery Rev., 1995, 16, 321.

310 L. Frullano and T. Meade, JBIC, J. Biol. Inorg. Chem., 2007, 12, 939. 311 Y.-W. Jun, J.-H. Lee and J. Cheon, Angew. Chem., Int. Ed., 2008, 47, 5122. 312 H. B. Na, I. C. Song and T. Hyeon, Adv. Mater., 2009, 21, 2133. 313 R. Weissleder, A. Moore, U. Mahmood, R. Bhorade, H. Benveniste, E. Chiocca and J. P. Basilion, Nat. Med., 2000, 6, 351. 314 J. W. M. Bulte, T. Douglas, B. Witwer, S.-C. Zhang, E. Strable, B. K. Lewis, H. Zywicke, B. Miller, P. van Gelderen, B. M. Moskowitz, I. D. Duncan and J. A. Frank, Nat. Biotechnol., 2001, 19, 1141. 315 R. Weissleder, Science, 2006, 312, 1168. 316 A. A. Gilad, P. Walczak, M. T. McMahon, H. Bin Na, J. H. Lee, K. An, T. Hyeon, P. C. M. van Zijl and J. W. M. Bulte, Magn. Reson. Med., 2008, 60, 1. 317 H. W. Kang, L. Josephson, A. Petrovsky, R. Weissleder and A. Bogdanov, Bioconjugate Chem., 2002, 13, 122. 318 M. Zhao, D. A. Beauregard, L. Loizou, B. Davletov and K. M. Brindle, Nat. Med., 2001, 7, 1241. 319 Y. W. Jun, Y. M. Huh, J. S. Choi, J. H. Lee, H. T. Song, S. Kim, S. Yoon, K. S. Kim, J. S. Shin, J. S. Suh and J. Cheon, J. Am. Chem. Soc., 2005, 127, 5732. 320 F. Hu, L. Wei, Z. Zhou, Y. Ran, Z. Li and M. Gao, Adv. Mater., 2006, 18, 2553. 321 K. L. Hultman, A. J. Raffo, A. L. Grzenda, P. E. Harris, T. R. Brown and S. O TM Brien, ACS Nano, 2008, 2, 477. a 322 J. E. Lee, N. Lee, H. Kim, J. Kim, S. H. Choi, J. H. Kim, T. Kim, C. in Song, S. P. Park, W. K. Moon and T. Hyeon, J. Am. Chem. Soc., 2010, 132, 552. 323 E. M. Shapiro and A. P. Koretsky, Magn. Reson. Med., 2008, 60, 265. 324 J.-S. Choi, J. Park, H. Nah, S. Woo, J. Oh, K. Kim, G. Cheon, Y. Chang, J. Yoo and J. Cheon, Angew. Chem., Int. Ed., 2008, 47, 6259. 325 S.-W. Chou, Y.-H. Shau, P.-C. Wu, Y.-S. Yang, D.-B. Shieh and C.-C. Chen, J. Am. Chem. Soc., 2010, 132, 13270. 326 M. Garnett and P. Kallinteri, Occup. Med., 2006, 56, 307. 327 V. E. Kagan, H. Bayir and A. A. Shvedova, Nanomed.: Nanotechnol., Biol. Med., 2005, 1, 313. 328 H. C. Fischer and W. C. W. Chan, Curr. Opin. Biotechnol., 2007, 18, 565. 329 I. Linkov, F. K. Satterstrom and L. M. Corey, Nanomed.: Nanotechnol., Biol. Med., 2008, 4, 167. 330 N. Lewinski, V. Colvin and R. Drezek, Small, 2008, 4, 26. 331 A. E. Nel, L. M dler, D. Velegol, T. Xia, E. M. Hoek, P. Somasuna daran, F. Klaessig, V. Castranova and M. Thompson, Nat. Mater., 2009, 8, 543. 332 B. C. Schanen, A. S. Karakoti, S. Seal, D. R. Drake, W. L. Warren and W. T. Self, ACS Nano, 2009, 3, 2523. 333 A. Shvedova, V. Castranova, E. Kisin, D. Schwegler-Berry, A. Murray, V. Gandelsman, A. Maynard and P. Baron, J. Toxicol. Environ. Health, Part A, 2003, 66, 1909. 334 C.-W. Lam, J. T. James, R. McCluskey and R. L. Hunter, Toxicol. Sci., 2003, 77, 126. 335 C. Carlson, S. M. Hussain, A. M. Schrand, L. K. Braydich-Stolle, K. L. Hess, R. L. Jones and J. J. Schlager, J. Phys. Chem. B, 2008, 112, 13608. 336 K. B. Male, B. Lachance, S. Hrapovic, G. Sunahara and J. H. T. Luong, Anal. Chem., 2008, 80, 5487. 337 Y. Pan, S. Neuss, A. Leifert, M. Fischler, F. Wen, U. Simon, G. Schmid, W. Brandau and W. Jahnen-Dechent, Small, 2007, 3, 1941. 338 A. Nel, T. Xia, L. M dler and N. Li, Science, 2006, 311, 622. a 339 C. C. Berry, S. Wells, S. Charles and A. S. G. Curtis, Biomaterials, 2003, 24, 4551. 340 C. C. Berry, S. Wells, S. Charles, G. Aitchison and A. S. G. Curtis, Biomaterials, 2004, 25, 5405. 341 M. Mahmoudi, A. Simchi, M. Imani, M. A. Shokrgozar, A. S. Milani, U. O. H feli and P. Stroeve, Colloids Surf., B, 2010, 75, 300. a 342 M. Mahmoudi, A. Simchi, M. Imani, A. S. Milani and P. Stroeve, Nanotechnology, 2009, 20, 225104. 343 S. Laurent, D. Forge, M. Port, A. Roch, C. Robic, L. Vander Elst and R. N. Muller, Chem. Rev., 2008, 108, 2064. 344 H. L. Karlsson, J. Gustafsson, P. Cronholm and L. Moller, Toxicol. Lett., 2009, 188, 112. 345 F. Y. Cheng, C. H. Su, Y. S. Yang, C. S. Yeh, C. Y. Tsai, C. L. Wu, M. T. Wu and D. B. Shieh, Biomaterials, 2005, 26, 729. 346 A. Petri-Fink, M. Chastellain, L. Juillerat-Jeanneret, A. Ferrari and H. Hofmann, Biomaterials, 2005, 26, 2685.

Downloaded on 20 June 2011 Published on 01 March 2011 on http://pubs.rsc.org | doi:10.1039/C0DT00689K

6342 | Dalton Trans., 2011, 40, 63156343

This journal is The Royal Society of Chemistry 2011

347 F. Cengelli, D. Maysinger, F. Tschudi-Monnet, X. Montet, C. Corot, A. Petri-Fink, H. Hofmann and L. Juillerat-Jeanneret, J. Pharmacol. Exp. Ther., 2006, 318, 108. 348 H. Tan, J. M. Xue, B. Shuter, X. Li and J. Wang, Adv. Funct. Mater., 2010, 20, 722. 349 F. Hu, K. G. Neoh, L. Cen and E. T. Kang, Biomacromolecules, 2006, 7, 809. 350 T. J. Brunner, P. Wick, P. Manser, P. Spohn, R. N. Grass, L. K. Limbach, A. Bruinink and W. J. Stark, Environ. Sci. Technol., 2006, 40, 4374. 351 S. J. H. Soenen, N. Nuytten, S. F. de Meyer, S. C. de Smedt and M. de Cuyper, Small, 2010, 6, 832. 352 S. J. H. Soenen, E. Illyes, D. Vercauteren, K. Braeckmans, Z. Majer, S. C. De Smedt and M. De Cuyper, Biomaterials, 2009, 30, 6803. 353 C. C. Berry, S. Charles, S. Wells, M. J. Dalby and A. S. G. Curtis, Int. J. Pharm., 2004, 269, 211. 354 H. R. Kim, K. Andrieux, S. Gil, M. Taverna, H. Chacun, D. Desmaele, F. Taran, D. Georgin and P. Couvreur, Biomacromolecules, 2007, 8, 793. 355 A. Bellova, E. Bystrenova, M. Koneracka, P. Kopcansky, F. Valle, N. Tomasovicova, M. Timko, J. Bagelova, F. Biscarini and Z. Gazova, Nanotechnology, 2010, 21, 65103. 356 M. N. Vieira, J. D. Figueroa-Villar, M. N. Meirelles, S. T. Ferreira and F. G. De Felice, Cell Biochem. Biophys., 2006, 44, 549. 357 J. Choi, S. Lee, H. Na, an Kwangjin, T. Hyeon and T. Seo, Bioprocess Biosyst. Eng., 2009, 33, 21. 358 B. Misselwitz, A. Muhler and H. J. Weinmann, Invest. Radiol., 1995, 30, 611. 359 B. J. Reynwar, G. Illya, V. A. Harmandaris, M. M. Muller, K. Kremer and M. Deserno, Nature, 2007, 447, 461. 360 C. Glaus, R. Rossin, M. J. Welch and G. Bao, Bioconjugate Chem., 2010, 21, 715. 361 J. Yang, C.-H. Lee, H.-J. Ko, J.-S. Suh, H. G. Yoon, K. Lee, Y.-M. Huh and S. Haam, Angew. Chem., Int. Ed., 2007, 46, 8836. 362 D. T. Denhardt, Ann. N. Y. Acad. Sci., 1992, 660, 70. 363 S. C. McBain, H. H. Yiu and J. Dobson, Int. J. Nanomed., 2008, 3, 169. 364 C. Xu, B. Wang and S. Sun, J. Am. Chem. Soc., 2009, 131, 4216. 365 C. S. Thaxton, R. Elghanian, A. D. Thomas, S. I. Stoeva, J.-S. Lee, N. D. Smith, A. J. Schaeffer, H. Klocker, W. Horninger, G. Bartsch and C. A. Mirkin, Proc. Natl. Acad. Sci. U. S. A., 2009, 106, 18437. 366 J. M. Nam, C. S. Thaxton and C. A. Mirkin, Science, 2003, 301, 1884. 367 B. Beutler, C. Eidenschenk, K. Crozat, J.-L. Imler, O. Takeuchi, J. A. Hoffmann and S. Akira, Nat. Rev. Immunol., 2007, 7, 753. 368 C. A. Janeway, Jr. and R. Medzhitov, Annu. Rev. Immunol., 2002, 20, 197. 369 R. Medzhitov, P. Preston-Hurlburt and C. A. Janeway, Jr., Nature, 1997, 388, 394. 370 T. Horng, G. M. Barton and R. Medzhitov, Nat. Immunol., 2001, 2, 835.

371 K. A. Fitzgerald, E. M. Palsson-McDermott, A. G. Bowie, C. A. Jefferies, A. S. Mansell, G. Brady, E. Brint, A. Dunne, P. Gray, M. T. Harte, D. McMurray, D. E. Smith, J. E. Sims, T. A. Bird and L. A. J. ONeill, Nature, 2001, 413, 78. 372 M. Yamamoto, S. Sato, H. Hemmi, H. Sanjo, S. Uematsu, T. Kaisho, K. Hoshino, O. Takeuchi, M. Kobayashi, T. Fujita, K. Takeda and S. Akira, Nature, 2002, 420, 324. 373 L.-H. Bin, L.-G. Xu and H.-B. Shu, J. Biol. Chem., 2003, 278, 24526. 374 H. Oshiumi, M. Sasai, K. Shida, T. Fujita, M. Matsumoto and T. Seya, J. Biol. Chem., 2003, 278, 49751. 375 M. Yamamoto, S. Sato, H. Hemmi, S. Uematsu, K. Hoshino, T. Kaisho, O. Takeuchi, K. Takeda and S. Akira, Nat. Immunol., 2003, 4, 1144. 376 M. Yamamoto, S. Sato, H. Hemmi, K. Hoshino, T. Kaisho, H. Sanjo, O. Takeuchi, M. Sugiyama, M. Okabe, K. Takeda and S. Akira, Science, 2003, 301, 640. 377 K. A. Fitzgerald, D. C. Rowe, B. J. Barnes, D. R. Caffrey, A. Visintin, E. Latz, B. Monks, P. M. Pitha and D. T. Golenbock, J. Exp. Med., 2003, 198, 1043. 378 Q. Huang, D. Liu, P. Majewski, L. C. Schulte, J. M. Korn, R. A. Young, E. S. Lander and N. Hacohen, Science, 2001, 294, 870. 379 A. Razin and J. Friedman, Prog. Nucleic Acid Res. Mol. Biol., 1981, 25, 33. 380 L. R. Cardon, C. Burge, D. A. Clayton and S. Karlin, Proc. Natl. Acad. Sci. U. S. A., 1994, 91, 3799. 381 S. Yamamoto, T. Yamamoto, T. Katoaka, E. Kuramoto, O. Yano and T. Tokunaga, Proc. Jpn. Soc. Immunol., 1989, 48, 272. 382 S. Yamamoto, T. Yamamoto, T. Kataoka, E. Kuramoto, O. Yano and T. Tokunaga, J. Immunol., 1992, 148, 4072. 383 A. M. Krieg, A.-K. Yi, S. Matson, T. J. Waldschmidt, G. A. Bishop, R. Teasdale, G. A. Koretzky and D. M. Klinman, Nature, 1995, 374, 546. 384 D. M. Klinman, A.-K. Yi, S. L. Beaucage, J. Conover and A. M. Krieg, Proc. Natl. Acad. Sci. U. S. A., 1996, 93, 2879. 385 J. R. Cubillos-Ruiz, X. Engle, U. K. Scarlett, D. Martinez, A. Barber, R. Elgueta, L. Wang, Y. Nesbeth, Y. Durant, A. T. Gewirtz, C. L. Sentman, R. Kedl and J. R. Conejo-Garcia, J. Clin. Invest., 2009, 119, 2231. 386 G. Chikh, S. D. de Jong, L. Sekirov, S. G. Raney, M. Kazem, K. D. Wilson, P. R. Cullis, J. P. Dutz and Y. K. Tam, Int. Immunol., 2009, 21, 757. 387 M. I. Shukoor, F. Natalio, M. N. Tahir, M. Wiens, M. Tarantola, H. A. Therese, M. Barz, S. Weber, M. Terekhov, H. C. Schroder, W. E. G. Muller, A. Janshoff, P. Theato, R. Zentel, L. M. Schreiber and W. Tremel, Adv. Funct. Mater., 2009, 19, 3717. 388 C. Bourquin, D. Anz, K. Zwiorek, A.-L. Lanz, S. Fuchs, S. Weigel, C. Wurzenberger, P. von der Borch, M. Golic, S. Moder, G. Winter, C. Coester and S. Endres, J. Immunol., 2008, 181, 2990. 389 T. Uto, T. Akagi, T. Hamasaki, M. Akashi and M. Baba, Immunol. Lett., 2009, 125, 46. 390 S. L. Demento, S. C. Eisenbarth, H. G. Foellmer, C. Platt, M. J. Caplan, W. M. Saltzman, I. Mellman, M. Ledizet, E. Fikrig, R. A. Flavell and T. M. Fahmy, Vaccine, 2009, 27, 3013.

Downloaded on 20 June 2011 Published on 01 March 2011 on http://pubs.rsc.org | doi:10.1039/C0DT00689K

This journal is The Royal Society of Chemistry 2011

Dalton Trans., 2011, 40, 63156343 | 6343

You might also like