3

Nanomedicine

J. Cancino-Bernardi
V.S. Marangoni
V. Zucolotto    Group of Nanomedicine and Nanotoxicology, Institute of Physics of São Carlos, University of São Paulo (Universidade de São Paulo—USP), São Carlos, São Paulo, Brazil

Abstract

The possibility of applying nanomaterials to the diagnosis, treatment, and prevention of diseases makes nanomedicine one of the most attractive areas of nanotechnology. The application of nanomaterials in molecular imaging, drug delivery, and therapeutic interventions promises to have a positive impact in this area owing to the unique properties of nanoparticles, which will allow them to overcome cellular and physiological barriers. Nanoparticles modified with antibodies, transferases, lectins, and avidins confer greater specificity to the delivery system by increasing the release, absorption, biodistribution, and subsequent elimination of the active ingredient, improving the efficiency of diagnosis and treatment—as is the case with photodynamic therapy, which is directly linked to advances in controlled drug delivery. This chapter describes the importance of nanomedicine, the scientific advances in this field, and a brief discussion of the toxicity and regulations of nanomaterials in medicine.

Keywords

nanomedicine
nanoparticles
nanomaterials
diagnosis
therapy
drug delivery
cancer
biofunctionalization
nanotoxicology
regulation

3.1. Nanomedicine

Nanomedicine is among the numerous opportunities and advances promoted by nanotechnology. The definitions of nanomedicine accepted today were established by the National Institutes of Health of the United States and the European Science Foundation, which define nanomedicine as the “science that uses nanomaterials to the development of diagnosis, treatment and prevention of specific medical application.” [1,2] The advancements of this technology have promoted innovations in different medical fields, including controlled drug delivery, biomarkers, molecular imaging, and biosensing [35]. Nanomedicine can make significant contributions to the health industry, including the areas of medicine, pharmaceutical sciences, and dentistry. These advances are a result of partnerships between companies and research centers, including universities, and reflect the interdisciplinarity of this area of research [6].
Many pharmaceutical companies have supported and encouraged research in drug delivery. The greatest advantage of using nanomedicine compared with traditional medicine is the use of analytical tools and treatments, such as nanoparticles. In this context, nanoparticles can overcome cellular and physiological barriers and allow for imaging and diagnostic applications at the cellular and molecular levels because of their nanoscale dimension [710].
After the first reports of the potential use of nanomedicine, several research centers and government agencies have contributed to the development of this area with the allocation of new investment funds for research and development or the creation of new research centers at universities. Part of the investment in nanotechnology research is devoted to nanomedicine, which includes the development of complex drug delivery systems and advanced therapeutic and imaging techniques. Many products and treatments using nanomaterials are already available on the global market, including hormone therapy, bone implants, appetite controllers, and nanodrugs for the treatment of cancer. Other products are in clinical phases of development [11].

3.2. Nanomaterials Applied to Diagnosis and Therapy

3.2.1. Use of Nanomaterials in Medicine

The rapid advancement of nanomedicine has allowed for the development of nanomaterials with novel properties, structures, and specific functions, and their application in diagnostics and therapeutics. Several nanomaterials are used to promote these changes. These nanoparticles can be metallic [12,13], magnetic [14], polymers [15], dendrimers [16,17], nanoparticles with ceramic materials [18,19], nanoparticles with silica [20], quantum dots, [21] nanotubes [22], carbon nanofibers [23], or graphene [24], all amenable to modifications [25]. Each of these nanomaterials can be modified with polymers, nanocomplexes, other nanoparticles, dendrimers, or biomolecules such as DNA, proteins, antibodies, and antigens, and we can produce a multitude of modifications to achieve higher selectivity [2628]. For the development of nanomaterials with high specificity regardless of the application, several factors should be taken into account, including their size, shape, modifications, composition, stability, biodegradation, and biodispersion. These parameters can help identify new or improved functions of nanomaterials and consequently their capabilities. Examples of such materials are found in Fig. 3.1 [29].
image
Figure 3.1 Examples of nanoparticles and UV-VIS spectra of aqueous suspensions of (A) gold nanohexapods, (C) gold nanorods, and (E) nanocages. Transmission electron microscopy (B, D, F) of the corresponding gold structures. Reprinted with permission from Y. Wang et al., Comparison study of gold nanohexapods, nanorods, and nanocages for photothermal cancer treatment, ACS Nano 7 (3) (2013) 2068–2077, Copyright (2014), American Chemical Society [29].
Molecular imaging, drug delivery, and photodynamic and photothermal therapies using nanomaterials promise to have a major impact on diagnosis and therapy [5,30]. Several products are already available on the market, primarily in the United States and Asia, with an emphasis on the field of oncology, and the results of preclinical and clinical studies using nanostructured carrier systems, nanomaterials as contrast agents, and magnetic nanoparticles have shown improved diagnosis and treatment in terms of drug delivery, image quality, and electromagnetic functions [3134].
The development of novel nanomaterials for use in medicine has proved viable in clinical and preclinical proof-of-concept experiments. Dendrimers have been applied in diagnostics and therapeutics for the development of a new cardiac test. This test can identify any damage to the heart muscle and significantly decrease the waiting time for blood test results [35,36]. Clinical trials with magnetic and metallic nanoparticles have yielded excellent results in the treatment of cancer via application of hyperthermia, in which the tumor is subjected to an electromagnetic field. For this application, nanoparticles are conjugated to specific antibodies or proteins to mark cancerous cells or tissues [37]. External laser irradiation in these tissues increases the temperature of the tissue to approximately 55°C and destroys tumor cells [38]. The application of nanomedicine is already a reality, and examples of this application are presented later.

3.2.2. Medical Applications of Magnetite and Core–Shells

The use of iron oxide nanoparticles in biomedical applications has increased because of their high biocompatibility, biodegradability, and superparamagnetic characteristics when applied in a magnetic field. Magnetite (Fe3O4) and its oxidized and stable form γ-Fe2O3 are by far the most commonly used magnetic nanoparticle contrast agents in magnetic resonance imaging, therapies to detect metastases, and clinical trials in nanomedicine [39,40]. Nanocrystals of magnetic iron oxide consist of magnetic metals such as iron, nickel, and cobalt. These nanomaterials are used in diagnostics and therapeutics for controlled drug delivery or imaging thanks to their nanometric scale, biocompatibility, and ability to be manipulated by an external magnetic field [41].
Many studies have focused on the encapsulation of new iron oxide nanoparticles with increasingly specific physicochemical characteristics to mark human erythrocytes by internalization of these nanoparticles through the membrane pores of these cells, without affecting cell viability [42]. The effectiveness of magnetic nanoparticles in the identification tumor metastasis is superior to that of other noninvasive methods [43,44]. Magnetic nanoparticles are versatile in the diagnostic imaging of inflammatory processes, including multiple arterioscleroses and rheumatoid arthritis in macrophages [41,45]. Magnetic nanoparticles and fluorophores are combined in preoperative treatments to diagnose glioma based on the response of these nanoparticles to fluorescence and magnetic resonance [46].
A new type of magnetic nanoparticle used in diagnostic imaging is radioluminescent core–shell nanoparticles. Their significant advantage is that a single nanoparticle allows for responses in terms of magnetic resonance, fluorescence, and radioluminescence. That is, they have magnetic and optical properties. Furthermore, magnetic luminescent nanoparticles can be used for controlled drug delivery [47]. Superparamagnetic nanoparticles with a size of 5–10 nm are used as intracellular markers for infectious or viral diseases. These nanoparticles are functionalized with antibodies, peptide ligands, or enzymes to provide greater specificity in the target cells [3,48,49]. Researchers intend to use core–shell magnetic nanoparticles coated with gold as marker probes coupled with drug delivery using an HIV antibody. These nanoparticles would trace and treat the viruses that were not eliminated with the available anti-AIDS cocktails, thereby promoting a revolution in the treatment of AIDS [49]. There are numerous applications of magnetic nanoparticles, either magnetite or core–shell, to mark tumors or lesions in tissues for improved diagnostic imaging.

3.2.3. Controlled Drug Delivery

The area that has developed the most thanks to nanomedicine is controlled drug delivery, which is the method or process of delivery of a pharmaceutical compound to achieve a therapeutic effect on an organism [5052]. New drug delivery technologies aim to increase the release, adsorption, biodistribution, and subsequent elimination of the active ingredients, improving efficiency and the benefits to the patient. In some cases, gene-based drugs can be delivered using normal release routes, including the oral, topical, transmucosal, and inhalation routes [9,46,51,52]. In such cases, the active ingredient may be susceptible to enzymatic degradation or not be efficiently absorbed owing to its physicochemical properties [53]. Advances in nanomedicine can overcome these drawbacks. Together with pharmaceutical advances, new nanomedical technologies promise to improve the activity and delivery of medications [4].
Current advancements in drug delivery include the development of carrier systems, in which the active ingredient is released only to the target area, for example, cancerous tissues [5]. The challenge today is to develop timed-release formulations in which the active ingredient is released over a period in preset doses [50]. For the drug to reach the target site efficiently, the delivery system should avoid the defense mechanisms of the body so that the system only releases the active ingredient to the target site [54]. The greatest challenge will be to prevent physiological mechanisms from limiting the activity of delivery systems, to block drug dispersion, and to prevent immune defense responses from causing physical or biochemical changes in these systems [55].
Drug delivery systems were initially developed using polymers and later using lipid vesicles. The interest in protecting vitamins from oxidation using polymers in the 1980s marked the beginning of the development of encapsulation systems. Soon afterward, polymers such as poly(lactic-co-glycolic acid) (PLGA) and polylactic acid (PLA) were used in drug delivery systems owing to their biodegradability and biocompatibility [56]. However, their disadvantages were associated with their low solubility and permeability. To address these disadvantages, the development of drug carriers using polypeptide vesicles increased rapidly [51]. However, disadvantages were also associated with the use of peptide vesicles, including low bioavailability due to low stability and high degradation in biological systems, low permeability in biological membranes, and short half-life in the circulatory system. Studies in nanomedicine indicate that several nanoparticles could penetrate epithelial tissue. Therefore, nanoparticles have received more attention than vesicular systems owing to their therapeutic potential and high stability in biological fluids [50,52]. In addition, drug delivery to specific organs began to become a reality. The combination of nanoparticles with antibodies, transferases, lectins, and avidins led to the development of delivery systems with greater specificity (Fig. 3.2).
image
Figure 3.2 The efficacy of nanoparticles in drug delivery is highly dependent on their size and shape.
The size of the nanoparticles affects their dispersion into and out of the vascular system, and the shape of nanoparticles affects their incorporation and penetration into the vessel walls. Reprinted with permission from O.C. Farokhzad, R. Langer, Impact of nanotechnology on drug delivery, ACS Nano 3 (1) (2009) 16–20, Copyright (2014) American Chemical Society [50].
The advantage of using nanoparticles in drug delivery systems is a result of two basic properties. The first is their small size, which allows them to penetrate into cells and small capillaries, causing the accumulation of active ingredient into the delivery systems. The second characteristic is the possibility of using biodegradable materials in the preparation of nanoparticles. The primary benefits of using nanotechnology in drug delivery are the improved delivery of the active ingredient, longer shelf life of the drugs, low cost of new formulations compared with the discovery of new molecules, and particularly the reduction of the doses indicated, which reduces the costs of the drugs.
The research on new biomarkers and drug delivery systems has focused on cardiovascular disease, cancer, and immunological diseases. Nanomaterials such as dendrimers, metal and silica nanoparticles, polymer nanoparticles, quantum dots, and carbon nanotubes (CNTs) have exceptional potential for direct delivery of drugs, including nitric oxide (NO) in patients with endocrine disorders [46]. In this context, the advantage of NO delivery systems using dendrimers includes the drug release control for time and site specificity and improved biocompatibility compared with conventional pharmacological applications. The primary advantage of NO release using dendrimer nanoparticles compared with other release systems is their increased capacity to store NO owing the pore structure of dendrimer molecules [57]. Furthermore, it is possible to change the external characteristics of dendrimer molecules, increasing their solubility or altering a specific function. Another example is the use of CNTs in the preferred delivery of acetylcholine in the brain for the treatment of diseases such as Alzheimer’s, without damaging the mitochondria or nonspecific organelles [58].
The development of drug carriers is highly dependent on the physicochemical properties of the nanoparticles but not on the active ingredient, as was once thought. Therefore, the physicochemical properties of these carriers can be optimized to target a specific tissue or to be nonspecifically absorbed by the target cells. However, to achieve this goal, nanomedicine should be used for the correct development of these carriers by determining the best techniques for encapsulation, the solvents used, the solubility of the active ingredient and solvent, and the types of polymers and stabilizers; in addition, it allows us to control other variables, including nanoparticle size, encapsulation efficiency, temperature, pH, ionic strength, and, most importantly, the residue of the polymer or nanoparticle associated with the drug in the patient’s body. The elimination of the polymer or nanoparticle from the patient’s body is a major cause for concern because of the possible toxicity of these materials [59].
One challenge in developing drug delivery systems is the use of theranostic nanomaterials, which combine diagnostics with therapeutics [60]. An ideal theranostic system will combine diagnosis and treatment in a single step using preventive monitoring before disease symptoms manifest [33,46,60]. This system includes the detection and removal of fat plaques in arteries for the prevention of cardiovascular disease. The development of drug-delivery systems with multiple active ingredients is promising because it allows for the synergistically combined delivery of several drugs to biological targets, for example, an antiangiogenic compound and a DNA intercalating agent, and this approach can increase the therapeutic efficacy.

3.2.4. Nanoparticles in Photothermal and Photodynamic Therapy

The use of the plasmon absorption band of nanoparticles for phototherapy has been promising for the treatment and diagnosis of cancer, and the technique is minimally invasive. However, for photothermal treatments in vivo, it is necessary to use nanoparticles with radiation in the near-infrared (NI) region (wavelengths between 650 and 900 nm) because the absorption coefficients of hemoglobin and water are smaller in this region [20,61]. Therefore, the plasmon resonance band of the nanoparticles in the NI region has been used for therapeutic purposes, that is, the excitation of electrons from the surface of nanoparticles causes localized heating (due to the effect of localized surface plasmon resonance—LEPR) and death of cancer cells. The nanoparticles most studied for this application include rod-shaped gold nanoparticles, known as nanorods, as well as multifunctional liposomes and polymer plasmonic nanoparticles [6163].
Standard photodynamic therapy has been applied to cancerous tumors, in which a photosensitive drug is administered to the patient, accumulates near the tumor tissue, and induces cell death after light activation. The photosensitive agents used for therapy since 1990 are porphyrin compounds, and for this reason, the photosensitive agents are classified based on the presence or absence of porphyrin [64]. Although porphyrins have a rapid adsorption and low toxicity, the disadvantages include low selectivity to smaller tumors and high hydrophobicity. Therefore, the development of highly selective photosensitive agents is the primary research topic in this area.
One of the most significant problems in photodynamic therapy is the controlled delivery of the photosensitive agent. Therefore, the development of this type of treatment is directly correlated with advances in systems for controlled drug delivery. Other limitations of photodynamic therapy are the low selectivity to tumors, low solubility, and high aggregation of these agents under physiological conditions [65]. With the advancement of research in drug delivery systems in nanomedicine, new strategies have been developed to increase the solubility of hydrophobic photosensitive agents and selectivity by modifying the surface area of nanoparticles loaded with the agent. The development of highly selective photoactive nanoparticles will ensure that this therapy targets cancer cells and preserves healthy cells [66].
Photosensitive agent formulations using nanoparticles are already available. One of the most promising formulations is the encapsulation of temoporfin as liposome (Foscan) [67], which overcome the limitations of existing commercial agents by causing a more efficient destruction of tumors and reducing damage and toxicity to healthy tissue compared with commercial agents. The encapsulation of porphyrin-based photosensitive agents in dendrimers is also possible. In this case, porphyrins and phthalocyanines are stabilized externally by dendrimers, which prevents aggregation [54]. CNTs are another possibility for delivery of photosensitive agents because they can be modified to increase specificity and adsorption in the body. However, a challenge that needs to be overcome is that CNTs absorb light in the NI region, which may cause the death of healthy cells owing to excessive local heating [68,69].
To improve the selectivity and specificity of the photosensitive agent in cancer cells, a combination of nanostructures with monoclonal antibodies and antibodies against transferrin receptors that are expressed on the surface of many solid tumors can be used [70,71]. Nanostructures can also be combined with other ligands, including vitamins, glycoproteins, peptides, aptamers, DNA, and growth factors. Another approach is the combination of different treatment modalities. For example, a single type of nanoparticle can be used in photodynamic therapy and radiotherapy. In this case, luminescent nanoparticles combined with photosensitive agents can be exposed to radiation such as X-rays, which causes the nanoparticles to emit luminescence, which in turn activates the photosensitive agents [72]; this is a major advantage for in vivo applications because the nanoparticles are luminescent and therefore no external light source is required.

3.2.5. Nanoparticles for Upconversion—Imaging of Cancer Cells

Nanoparticles for upconversion have emerged as a new class of agents for bioimaging. Upconversion is the absorption of two or more lower-energy photons and the consequent release of a higher-energy photon. This technique has been important in the development of more efficient systems in medicine [73]. Typically, these nanoparticles are excited at a wavelength in the NI region (900–1000 nm) and emit wavelengths with higher energy in the ultraviolet–visible region. This process is different from multiphoton absorption in organic molecules and quantum dots, and in general their efficiency is much higher, which allows for the use of low-cost continuous-wave diode lasers in contrast to the ultrashort pulse lasers used in the excitation of nonlinear multiphotons [73,74].
Nanoparticles used in upconversion usually consist of nanostructured systems doped with lanthanides, that is, rare-earth trivalent ions such as Yb3+, Tm3+, Er3+, and Ho3+, dispersed in nanostructures in a suitable dielectric medium [75,76]. These lanthanide dopants act as optically active centers and emit light when excited with the appropriate wavelength. These optical properties can be adjusted by selecting the lanthanide dopants to produce systems with different emission wavelengths, ranging from visible to NI or ultraviolet. Furthermore, the selection of the nanostructured host system, that is, the nanomaterials that will be doped with lanthanides, is essential for the production of highly efficient systems. Characteristics such as transparency in the spectral region of interest and chemical stability are critical in the selection of this system [73].
Nanoparticles of sodium yttrium fluoride doped with ytterbium and erbium (NaYF4: Yb,Er) are efficient in converting infrared radiation into visible radiation and have shown promise in the detection of biological interactions [77]. The modification of dopants of NaYF4 nanoparticles to Tm3+ and Yb3+ allows for the upconversion from NI to visible radiation, resulting in high penetration of these particles into biological tissues and acquisition of images with high optical contrast, which is supported by in vitro and in vivo experiments [78]. Several other host nanostructures have been explored in the development of these systems, including ZrO2 [79] and LaF3 [80].
Furthermore, the possibility to combine different materials into a single structure has made it possible to obtain systems with multiple and synergistic properties. For example, multifunctional nanoparticles can be manufactured by adsorption of Fe3O4 nanoparticles on the surface of upconverting nanostructures followed by the formation of a thin layer of gold on the surface [81]. Therefore, these materials have optical and magnetic properties and can be exploited as contrast agents in magnetic resonance imaging. Alternatively, the introduction of Gd3+ to these structures allows it to be used as a T1 contrast agent in magnetic resonance imaging [82]. The presence of gold nanoparticles on the surface of upconverting nanoparticles enables the modulation of these systems, increasing their versatility [83].
Therefore, these nanoparticles have many advantages for tumor imaging applications, including stable emission, high resistance to photodegradation, the ability to be detected in biological tissues using light in the NI region, and the possibility of biofunctionalization of their surface. Despite their great potential, rapid heating can overheat the tissue around the tumor, and lanthanides can increase the toxicity of these materials; therefore, new strategies for synthesis and functionalization are required to minimize these adverse effects [34,73,74].

3.3. Synthesis of Nanomaterials for Application in Nanomedicine

The successful use of nanomaterials in medicine is closely correlated with their properties. Characteristics such as stability, size variation, morphology, surface charge, and toxicity should be chosen carefully to achieve the desired results. In this context, increasingly effective methods have been developed to produce nanomaterials with highly controlled physicochemical parameters. The following topics represent the main methods used in the production of different types of nanomaterials explored in medicine.

3.3.1. Gold Nanoparticles

Among the metal nanoparticles, nanoparticles of noble metals such as gold have been extensively studied in recent years. These particles have a plasmon resonance band owing to collective oscillation of free electrons on their surface, whose wavelength can be changed by controlling some parameters, including size, shape, and surface coating [84].
Gold nanoparticles can be manufactured using the precipitation method, in which a reducing agent is added to a gold salt solution in the presence of a stabilizer. In the most common route of synthesis, nanoparticles are synthesized in an aqueous medium in the presence of citric acid or sodium citrate, which acts as a reducing agent and stabilizer because of its negative charge [85]. Other stabilizing molecules have been explored, including polymers. This functionalization allows us to obtain organic and inorganic hybrid nanomaterials, whose parameters, including size, shape, and surface functional groups, can be adjusted by the appropriate choice of the surface polymer [86].
Sulfur establishes a strong bond with atomic gold [87] and has been extensively explored for the preparation and stabilization of gold nanoparticles. In one method of synthesis, gold nanoparticles stabilized with sodium 3-mercaptopropionate molecules are obtained by the simultaneous addition of a citrate salt and sodium 3-mercaptopropionate to an aqueous solution of chloroauric acid (HAuCl4) under boiling [88]. The particle size can be controlled by modifying the ratio between the concentration of stabilizer and the concentration of gold ions in solution [88].

3.3.2. Magnetic Nanoparticles

Magnetic nanoparticles have been extensively studied for medical applications and have proven to be highly effective as contrast agents in nuclear magnetic resonance imaging and treatment by hyperthermia [89]. These particles can be formed by different types of cubic ferrites with the general expression M2+Fe2O4, where M2+ is a metal such as Mg, Co, Zn, or Fe [as in Fe3O4 or maghemite (γ-Fe2O3)] [90]. Fe3O4 can be synthesized using a simple method known as coprecipitation, which involves a controlled alkalization reaction in which a precipitator is added to an aqueous solution of iron salt at a pH between 9 and 10 at room temperature [89]. Fe3O4 nanoparticles synthesized from iron (II) and iron (III) salts follow the stoichiometry of the overall reaction
Fe2+ + 2Fe3+ + 8OH → Fe3O4 + 4H2O
Using only iron (II) salt, the reaction occurs by different mechanisms and the diameter of the obtained nanoparticles varies between 30 and 50 nm:
Fe2+ + 2OH → Fe(OH)2
3Fe(OH)2 + ½ O2 → Fe(OH)2 + 2FeOOH + H2O
Fe(OH)2 + 2FeOOH → Fe3O4 + 2H2O
This method is quite simple, but it has limitations on the control of the shape and size of the obtained particles. Methods that use organic solvents and high temperatures have been the most efficient to date. In one such method, nanoparticles (MFe2O4) are obtained from metal acetylacetonate (M = Fe, Co, Mn) in 1,2-hexadecanodiol in the presence of oleic acid and olamine at high temperatures [90]. After synthesis, these particles can be washed by magnetic separation and functionalized with polymeric molecules for use in an aqueous medium.

3.3.3. Core–Shell-Type Structures

Nanoparticles that incorporate multiple components and functionalities are crucial for the development of new systems for medical applications. Among them, core–shell nanostructures with different components have become increasingly important owing to the possibility of easily adjusting their new properties. For example, Oldenburg et al. [91] showed that the absorption band of gold nanoparticles can be shifted from visible to infrared by a coating of silica particles (core) with a gold layer, forming core–shell-type structures. Therefore, the absorption wavelength can be adjusted by the relative thickness of the gold layer around the silica nanoparticles [91].
Another prominent candidate for application in medicine is nanoparticles of Fe3O4@Au. The advantage of this system compared with iron oxide nanoparticles is the possibility of developing systems that combine optical and magnetic properties in a single process. Furthermore, the coating of iron oxide nanoparticles protects the Fe3O4 core against oxidation without drastically reducing their magnetic properties [92].
Several strategies have been developed to form such structures. One example is the reduction of gold ions in the presence of Fe3O4 nanoparticles, 1,2-hexadecanodiol, oleic acid, and oleylamine in phenyl ether under magnetic stirring, argon atmosphere, and a temperature of 190°C [93].
Another synthesis strategy is the initial coating of iron oxide nanoparticles with silica using tetraethylorthosilicate (TEOS) and ammonium hydroxide in an alcohol solution. These particles are then functionalized with 3-aminopropyltriethoxysilane (APTES). This system is added to a suspension of gold colloids in the presence of tetrakis(hydroxymethyl)phosphonium chloride (THPC), which interacts with amino groups from the APTES. The formation of a gold nanoparticle-THPC suspension involves the reduction of chloroauric acid (HAuCl4) with THPC, leading to the formation of small particles (approximately 2 nm). Finally, the nanoshell is prepared via reduction of HAuCl4 with formaldehyde in a solution containing potassium carbonate (K2CO3), and the magnetic nanoparticles are coated with silica prepared beforehand [94,95].

3.3.4. Biofunctionalization of Nanomaterials

The possibility of incorporating various types of molecules on the surface of these nanomaterials is essential for the development of specific agents for disease diagnosis and treatment and drug delivery systems. These particles can be functionalized with biomolecules such as proteins [96], peptides [97], aptamers [98], and specific antibodies [99] that specifically recognize the target cells or tissues, providing additional recognition properties to the nanomaterials. Furthermore, fluorescent or antitumor molecules can be added to these nanostructures, allowing for their detection by fluorescence spectroscopy or the attenuation of the side effects caused by these drugs, respectively. The organization of these nanocomplexes is directed by ionic, covalent, and noncovalent interactions and by hydrogen bonds between the molecules to be incorporated on the surface of the nanoparticles [100] (Fig. 3.3).
image
Figure 3.3 Nanoparticle with various ligands that confer multifunctionality using a single platform.
For example, gold nanoparticles conjugated to oligonucleotides are of great interest because the complementarity of DNA base pairs increases the specificity of these complexes and allows their use in biosensors for the detection of specific DNA sequences associated with abnormalities or diseases [101], intracellular gene regulation [102], and supramolecular organization of nanostructures [103]. Gold nanoparticles can be functionalized with oligonucleotides by modifying the oligonucleotides with thiol groups, which form highly stable bonds with gold [104].

3.4. Nanotoxicology

Despite the rapid progress and acceptance of nanomedical technologies in diagnostics and therapeutics, their potential effect on human health due to prolonged exposure to these compounds has not been established. It is important to realize that regulatory issues in nanomedicine involve the approval of therapeutic processes, which are complex and expensive but necessary, regardless of the nanomaterial or application [105]. In the past 5 years, research in nanotoxicity has focused on cell culture systems and animal testing for the establishment of regulatory processes. However, data from these studies are still inaccurate, which warrants the development of new methods to elucidate these effects [106,107]. In in vitro systems, the interactions between nanostructures and biological components are complex because of the diversity of proteins and substances in the cellular media, and even more complex in in vivo systems that contain biological fluids, organelles, and biomolecules [108,109].
To properly assess the potential risks of nanoparticles to health, their life cycle should be measured in all development phases, including production, storage, distribution, application, and elimination [110]. The effect on humans or the environment may vary in different stages of the life cycle. In addition, metabolic and immune responses induced by these materials are still not fully understood. Toxicological studies suggest that nanomaterials can cause adverse health effects; however, the fundamental cause–effect relationships are still unclear. Many nanomaterials can interact with cellular components, interrupting or changing specific functions or even causing the overproduction of reactive oxygen and nitrogen species [106].
Before nanomedicine is applied in diagnostics and therapeutics, we need to evaluate critically how the human body will react to different nanomaterials because they are foreign bodies. Reactions against these nanomaterials may not be identical to those produced against viruses or bacteria, as in the former case there are no specific enzymes or antibodies against these agents [111]. Moreover, although these materials represent substances unknown to the human body, they are immediately sensed when they reach the bloodstream (the first access route for drug delivery systems) and quickly interact with blood components through physical, chemical, and biological processes [112]. The interactions of nanomaterials with nuclei of mitochondrial cells have been considered the main source of cytotoxicity. Nanomaterials, such as silver nanoparticles coated with fullerenes, micelles, copolymers, and CNTs, can interact with mitochondrial cells and cause apoptosis and DNA damage [113,114]. Owing to the complexity of the mechanisms involved in the interactions between nanomaterials and biological samples, the biophysical and biochemical aspects are difficult to investigate, especially in vivo and in real time.
The use of nanoparticles in nanomedicine has advantages and disadvantages. For example, the benefits of the oral, transdermal, intravenous, or respiratory administration of nanoparticles are their limited invasiveness and large surface area. The disadvantages are associated with possible interactions with mucosal membranes, local irritation, and liver toxicity. Therefore, regardless of the application of nanoparticles in medicine, their toxicity should be assessed before clinical trials. The scientific community is continually developing new analytical methods to overcome the limited data on the toxicity of nanomaterials [115117]. However, many basic standard toxicology tests are not applied to nanotoxicology because of the high reactivity of nanomaterials with chemicals used in these assays. Despite the limitations in the development of appropriate methods to assess the specific interactions between nanomaterials and biological systems, many studies using in vitro and in vivo analytical protocols are available [118121].

3.4.1. In Vitro Assays

One type of test that is most commonly applied to evaluate the intrinsic toxicity of nanomaterials is assays with different in vitro cell lines [108]. One of the most studied nanomaterials to assess the degree of toxicity is CNTs [122]. CNTs have gained prominence because of their capacity to be used as carriers combined with the possibility of making modifications on their surface, which can lead to the development of novel nanomaterials with therapeutic applications, for example, in cancer [123]. However, CNTs can internalize into human T-cells and kidney cells and cause time- and dose-dependent cell death [124126]. Single-wall CNTs (SWCNTs) are more toxic to alveolar macrophages compared with double-wall CNTs. The toxicity of CNTs also depends on other factors, including their concentration, shape, degree of surface functionalization, and adsorption of metals during the production process. In other cases, CNTs are not toxic, such as when they are administered in lung cells at a dose between 1.5 and 800 mg mL−1 [127].
Metallic and magnetic nanoparticles are biocompatible because they are made of inert material. However, nanoparticles made of inert material can produce morphological changes, loss of function, inflammation, and even irreversible cell damage. One study on the toxicity of gold nanoparticles in human sperm indicated changes in motility, morphology, and sperm fragmentation [128]. However, the morphology of primary endothelial cells from rat brains (rBMEC) did not change after 24 h of exposure to gold nanoparticles [129].
PLGA polymeric nanoparticles were not toxic to the A549 cell line, even at high concentrations [130]. Silica nanoparticles administered to A549 cells in a time- and dose-dependent manner reduced cell viability and increased the levels of oxidative stress, indicating a cytotoxic effect [131]. Silver nanoparticles at concentrations higher than 50 μg mL−1 applied to fibroblast cells (NIH3T3) caused apoptosis associated with ROS production [132]. The administration of silver nanoparticles at concentrations close to 2 μg mL−1 reduced the viability of HEK cells and induced the production of cytokines responsible for immune responses [133]. In contrast, silver nanoparticles at a concentration of 6.3 μg mL−1 showed no evidence of causing cell death or damage to A431 human carcinoma cells [134].
In summary, toxicological studies of nanoparticles in in vitro systems are still inconsistent regarding the adverse and toxic effects of nanoparticles, indicating that more research is needed to address this issue [135]. In addition, further systematic in vitro studies are necessary to regulate the medical use of nanoparticles.

3.4.2. In Vivo Studies

The in vivo characteristics of these nanoparticles are a major issue that needs to be addressed [106,119,136]. It is still not known whether all nanoparticles have undesirable effects on the environment and living organisms in general. At present, four routes of entry of nanoparticles into the body have been described: inhalation, ingestion, absorption through the skin, and injection during medical procedures. Once inside the body, nanoparticles are highly mobile and can remain anywhere in the body [137]. So far, neither the nanoparticles nor the products and materials that contain them are subject to regulation.
In rats, the accumulation and potential adverse effects of CNTs in several organs, including the liver, lung, and spleen, are reported [110]. Various studies have evaluated the toxicity of carbon-based nanomaterials, including multiwall CNTs, carbon nanofibers, and carbon nanoparticles, on the basis of the radius and surface chemistry of these materials. The nanotoxicity was tested in vivo using a tumor cell line culture. The results indicated that these materials are toxic in a dose-dependent manner. Furthermore, cytotoxicity was enhanced when the surface of these nanomaterials was functionalized after acid treatment [138]. Another study inoculated CNTs into primary skins and subjected this material to conjunctival irritation and sensitivity tests. The negative metagenesis results suggested that SWCNTs are not carcinogenic. In addition, the lethal dose for hamsters was 2000 mg kg−1 body weight, a dose higher than will ever be used in nanomedicine [139].
Changes in gene expression were observed in pregnant mice exposed to titanium oxide (TiO2) nanoparticles. Exposure of mouse fetuses to TiO2 nanoparticles during the prenatal period affected the expression of genes related to the development and function of the central nervous system [140]. Considering the potential toxicity of nanoparticles and nanomaterials in general and the lack of legislation governing such effects, new analytical methods and additional systematic studies are necessary to assess the nanotoxicity of these materials. It is expected that, as the field of nanomedicine progresses, new concepts in nanotoxicity will guide the creation of laws to regulate the use of nanomaterials.

References

[1] Oberdorster G. Safety assessment for nanotechnology and nanomedicine: concepts of nanotoxicology. J. Intern. Med. 2010;267(1):89105.

[2] Foundation EES. Nanomedicine, an ESF–European Medical Research Councils (EMRC) Forward Look Report. France: ESF; 2004.

[3] Blanco E, Hsiao A, Mann AP, Landry MG, Meric-Bernstam F, Ferrari M. Nanomedicine in cancer therapy: innovative trends and prospects. Cancer Sci. 2011;102(7):12471252.

[4] Wang J, Byrne JD, Napier ME, DeSimone JM. More effective nanomedicines through particle design. Small. 2011;7(14):19191931.

[5] Boulaiz H, Alvarez PJ, Ramirez A, Marchal JA, Prados J, Rodriguez-Serrano F, et al. Nanomedicine: application areas and development prospects. Int. J. Mol. Sci. 2011;12(5):33033321.

[6] Caruso F, Hyeon T, Rotello VM. Nanomedicine. 2012;41(7):25372538.

[7] Labhasetwar V, Zborowski M, Abramson AR, Basilion JP. Nanoparticles for imaging, diagnosis, and therapeutics. Mol. Pharm. 2009;6(5):12611262.

[8] Estanqueiro M, Amaral MH, Conceicao J, Lobo JM Sousa. Nanotechnological carriers for cancer chemotherapy: the state of the art. Colloid Surf. B. 2015;126:631648.

[9] Alivisatos AP. Less is more in medicine—sophisticated forms of nanotechnology will find some of their first real-world applications in biomedical research, disease diagnosis and, possibly, therapy. Sci. Am. 2001;285(3):6673.

[10] Bertrand N, Wu J, Xu XY, Kamaly N, Farokhzad OC. Cancer nanotechnology: the impact of passive and active targeting in the era of modern cancer biology. Adv. Drug Deliv. Rev. 2014;66:225.

[11] Nanotecnhologies TPoE. 2016 [cited 2016 03022016], Available from: http://www.nanotechproject.org/inventories/medicine/apps/

[12] Arvizo RR, Bhattacharyya S, Kudgus RA, Giri K, Bhattacharya R, Mukherjee P. Intrinsic therapeutic applications of noble metal nanoparticles: past, present and future. Chem. Soc. Rev. 2012;41(7):29432970.

[13] Dreaden EC, Alkilany AM, Huang X, Murphy CJ, El-Sayed MA. The golden age: gold nanoparticles for biomedicine. Chem. Soc. Rev. 2012;41(7):27402779.

[14] Piao Y, Burns A, Kim J, Wiesner U, Hyeon T. Designed fabrication of silica-based nanostructured particle systems for nanomedicine applications. Adv. Funct. Mater. 2008;18(23):37453758.

[15] Zhang L, Chan JM, Gu FX, Rhee J-W, Wang AZ, Radovic-Moreno AF, et al. Self-assembled lipid-polymer hybrid nanoparticles: a robust drug delivery platform. ACS Nano. 2008;2(8):16961702.

[16] Yang F, Jin C, Subedi S, Lee CL, Wang Q, Jiang Y, et al. Emerging inorganic nanomaterials for pancreatic cancer diagnosis and treatment. Cancer Treat. Rev. 2012;38(6):566579.

[17] Ciolkowski M, Petersen JF, Ficker M, Janaszewska A, Christensen JB, Klajnert B, et al. Surface modification of PAMAM dendrimer improves its biocompatibility. Nanomedicine. 2012;8(6):815817.

[18] Rosenholm JM, Sahlgren C, Linden M. Multifunctional mesoporous silica nanoparticles for combined therapeutic, diagnostic and targeted action in cancer treatment. Curr. Drug Targets. 2011;12(8):11661186.

[19] Vallet-Regi M, Ruiz-Hernandez E. Bioceramics: from bone regeneration to cancer nanomedicine. Adv. Mater. 2011;23(44):51775218.

[20] Croissant J, Zink JI. Nanovalve-controlled cargo release activated by plasmonic heating. J. Am. Chem. Soc. 2012;134(18):76287631.

[21] Lee KH, Galloway JF, Park J, Dvoracek CM, Dallas M, Konstantopoulos K, et al. Quantitative molecular profiling of biomarkers for pancreatic cancer with functionalized quantum dots. Nanomedicine. 2012;8(7):10431051.

[22] Mehdipoor E, Adeli M, Bavadi M, Sasanpour P, Rashidian B. A possible anticancer drug delivery system based on carbon nanotube-dendrimer hybrid nanomaterials. J. Mater. Chem. 2011;21(39):1545615463.

[23] Andrews RJ, Nanotechnology, Neurosurgery. J. Nanosci. Nanotechnol. 2009;9(8):50085013.

[24] Liu Z, Robinson JT, Tabakman SM, Yang K, Dai H. Carbon materials for drug delivery & cancer therapy. Mater. Today. 2011;14(7-8):316323.

[25] Swierczewska M, Liu G, Lee S, Chen X. High-sensitivity nanosensors for biomarker detection. Chem. Soc. Rev. 2012;41(7):26412655.

[26] Perfezou M, Turner A, Merkoci A. Cancer detection using nanoparticle-based sensors. Chem. Soc. Rev. 2012;41(7):26062622.

[27] Palivan CG, Fischer-Onaca O, Delcea M, Itel F, Meier W. Protein-polymer nanoreactors for medical applications. Chem. Soc. Rev. 2012;41(7):28002823.

[28] Siqueira Jr JR, Abouzar MH, Poghossian A, Zucolotto V, Oliveira Jr ON, Schoening MJ. Penicillin biosensor based on a capacitive field-effect structure functionalized with a dendrimer/carbon nanotube multilayer. Biosens. Bioelectron. 2009;25(2):497501.

[29] Wang Y, Black KCL, Luehmann H, Li W, Zhang Y, Cai X, et al. Comparison study of gold nanohexapods, nanorods, and nanocages for photothermal cancer treatment. ACS Nano. 2013;7(3):20682077.

[30] Kranz C, Eaton DC, Mizaikoff B. Analytical challenges in nanomedicine. Anal. Bioanal. Chem. 2011;399(7):23092311.

[31] Vivero-Escoto JL, Huxford-Phillips RC, Lin W. Silica-based nanoprobes for biomedical imaging and theranostic applications. Chem. Soc. Rev. 2012;41(7):26732685.

[32] Taylor A, Wilson KM, Murray P, Fernig DG, Levy R. Long-term tracking of cells using inorganic nanoparticles as contrast agents: are we there yet? Chem. Soc. Rev. 2012;41(7):27072717.

[33] Sumer B, Gao J. Theranostic nanomedicine for cancer. Nanomedicine. 2008;3(2):137140.

[34] R. Tong, D.S. Kohane, New strategies in cancer nanomedicine, in: P.A. Insel (Ed.), Annual Review of Pharmacology and Toxicology, vol. 56, Annual Reviews, Palo Alto, 2016, pp. 41–57.

[35] Bai SH, Thomas C, Rawat A, Ahsan F. Recent progress in dendrimer-based nanocarriers. Crit. Rev. Ther. Drug Carr. Syst. 2006;23(6):437495.

[36] Buxton DB, Lee SC, Wickline SA, Ferrari M. National Heart, Lung, and Blood Institute Nanotechnology Working Group.Recommendations of the National Heart, Lung, and Blood Institute Nanotechnology Working Group. Circulation. 2003;108(22):27372742.

[37] Nie S. Understanding and overcoming major barriers in cancer nanomedicine. Nanomedicine. 2010;5(4):523528.

[38] Lee N, Hyeon T. Designed synthesis of uniformly sized iron oxide nanoparticles for efficient magnetic resonance imaging contrast agents. Chem. Soc. Rev. 2012;41(7):25752589.

[39] Li Y-F, Chen C. Fate and toxicity of metallic and metal-containing nanoparticles for biomedical applications. Small. 2011;7(21):29652980.

[40] Huang S-H, Juang R-S. Biochemical and biomedical applications of multifunctional magnetic nanoparticles: a review. J. Nanoparticle Res. 2011;13(10):44114430.

[41] McCarthy JR, Weissleder R. Multifunctional magnetic nanoparticles for targeted imaging and therapy. Adv. Drug Deliv. Rev. 2008;60(11):12411251.

[42] Antonelli A, Sfara C, Manuali E, Bruce IJ, Magnani M. Encapsulation of superparamagnetic nanoparticles into red blood cells as new carriers of MRI contrast agents. Nanomedicine. 2011;6(2):211223.

[43] Bourne MW, Margerun L, Hylton N, Campion B, Lai JJ, Derugin N, et al. Evaluation of the effects of intravascular MR contrast media (gadolinium dendrimer) on 3D time of flight magnetic resonance angiography of the body. J. Magn. Reson. Imag. 1996;6(2):305310.

[44] Brown G, Richards CJ, Newcombe RG, Dallimore NS, Radcliffe AG, Carey DP, et al. Rectal carcinoma: thin-section MR imaging for staging in 28 patients. Radiology. 1999;211(1):215222.

[45] Sun C, Lee JSH, Zhang M. Magnetic nanoparticles in MR imaging and drug delivery. Adv. Drug Deliv. Rev. 2008;60(11):12521265.

[46] Shubayev VI, Pisanic II TR, Jin S. Magnetic nanoparticles for theragnostics. Adv. Drug Deliv. Rev. 2009;61(6):467477.

[47] Chen H, Colvin DC, Qi B, Moore T, He J, Mefford OT, et al. Magnetic and optical properties of multifunctional core-shell radioluminescence nanoparticles. J. Mater. Chem. 2012;22(25):1280212809.

[48] Dykman L, Khlebtsov N. Gold nanoparticles in biomedical applications: recent advances and perspectives. Chem. Soc. Rev. 2012;41(6):22562282.

[49] Could P. Nanoparticles probes biosystems. Mater. Today. 2004;7:3643.

[50] Farokhzad OC, Langer R. Impact of nanotechnology on drug delivery. ACS Nano. 2009;3(1):1620.

[51] Sahoo SK, Labhasetwar V. Nanotech approaches to delivery and imaging drug. Drug Discov. Today. 2003;8(24):11121120.

[52] Sahoo SK, Parveen S, Panda JJ. The present and future of nanotechnology in human health care. Nanomedicine. 2007;3(1):2031.

[53] Guo S, Wang E. Functional micro/nanostructures: simple synthesis and application in sensors, fuel cells, and gene delivery. Acc. Chem. Res. 2011;44(7):491500.

[54] Kateb B, Chiu K, Black KL, Yamamoto V, Khalsa B, Ljubimova JY, et al. Nanoplatforms for constructing new approaches to cancer treatment, imaging, and drug delivery: what should be the policy? Neuroimage. 2011;54:S106S124.

[55] Veiseh O, Gunn JW, Zhang M. Design and fabrication of magnetic nanoparticles for targeted drug delivery and imaging. Adv. Drug Deliv. Rev. 2010;62(3):284304.

[56] Leserman LD, Barbet J, Kourilsky F, Weinstein JN. Targeting to cells of fluorescent liposomes covalently coupled with monoclonal antibody or protein-A. Nature. 1980;288(5791):602624.

[57] Seabra AB, Duran N. Nitric oxide-releasing vehicles for biomedical applications. J. Mater. Chem. 2010;20(9):16241637.

[58] Yang Z, Zhang Y, Yang Y, Sun L, Han D, Li H, et al. Pharmacological and toxicological target organelles and safe use of single-walled carbon nanotubes as drug carriers in treating Alzheimer disease. Nanomedicine. 2010;6(3):427441.

[59] Menjoge AR, Kannan RM, Tomalia DA. Dendrimer-based drug and imaging conjugates: design considerations for nanomedical applications. Drug Discov. Today. 2010;15(5-6):171185.

[60] Chen X, Gambhlr SS, Cheon J. Theranostic nanomedicine. Acc. Chem. Res. 2011;44(10):841.

[61] Song JB, Huang P, Duan HW, Chen XY. Plasmonic vesicles of amphiphilic nanocrystals: optically active multifunctional platform for cancer diagnosis and therapy. Acc. Chem. Res. 2015;48(9):25062515.

[62] Frazier N, Ghandehari H. Hyperthermia approaches for enhanced delivery of nanomedicines to solid tumors. Biotechnol. Bioeng. 2015;112(10):19671983.

[63] Liu Y, Ashton JR, Moding EJ, Yuan HK, Register JK, Fales AM, et al. A plasmonic gold nanostar theranostic probe for in vivo tumor imaging and photothermal therapy. Theranostics. 2015;5(9):946960.

[64] Dougherty TJ, Gomer CJ, Henderson BW, Jori G, Kessel D, Korbelik M, et al. Photodynamic therapy. J. Natl. Cancer Inst. 1998;90(12):889905.

[65] Paszko E, Ehrhardt C, Senge MO, Kelleher DP, Reynolds JV. Nanodrug applications in photodynamic therapy. Photodiagnosis Photodyn. Ther. 2011;8(1):1429.

[66] Perni S, Prokopovich P, Pratten J, Parkin IP, Wilson M. Nanoparticles: their potential use in antibacterial photodynamic therapy. Photochem. Photobiol. Sci. 2011;10(5):712720.

[67] Compagnin C, Bau L, Mognato M, Celotti L, Miotto G, Arduini M, et al. The cellular uptake of meta-tetra (hydroxyphenyl)chlorin entrapped in organically modified silica nanoparticles is mediated by serum proteins. Nanotechnology. 2009;20(34):345101345113.

[68] Zhu Z, Tang ZW, Phillips JA, Yang RH, Wang H, Tan WH. Regulation of singlet oxygen generation using single-walled carbon nanotubes. J. Am. Chem. Soc. 2008;130(33):1085610857.

[69] Erbas S, Gorgulu A, Kocakusakogullari M, Akkaya EU. Non-covalent functionalized SWNTs as delivery agents for novel Bodipy-based potential PDT sensitizers. Chem. Commun. 2009;(33):49564958.

[70] Zhang L, Gu FX, Chan JM, Wang AZ, Langer RS, Farokhzad OC. Nanoparticles in medicine: therapeutic applications and developments. Clin. Pharmacol. Ther. 2008;83(5):761769.

[71] Torchilin VP. Micellar nanocarriers: pharmaceutical perspectives. Pharm. Res. 2007;24(1):116.

[72] Zhao B, Yin J-J, Bilski PJ, Chignell CF, Roberts JE, He Y-Y. Enhanced photodynamic efficacy towards melanoma cells by encapsulation of Pc4 in silica nanoparticles. Toxicol. Appl. Pharmacol. 2009;241(2):163172.

[73] Chen GY, Qju HL, Prasad PN, Chen XY. Upconversion nanoparticles: design, nanochemistry, and applications in theranostics. Chem. Rev. 2014;114(10):51615214.

[74] Wu S, Butt HJ. Near-infrared-sensitive materials based on upconverting nanoparticles. Adv. Mater. 2016;28(6):12081226.

[75] Passuello T, Piccinelli F, Pedroni M, Polizzi S, Mangiarini F, Vetrone F, et al. NIR-to-visible and NIR-to-NIR upconversion in lanthanide doped nanocrystalline GdOF with trigonal structure. Opt. Mater. 2011;33(10):15001505.

[76] Tian L, Xu Z, Zhao S, Cui Y, Liang Z, Zhang J, et al. The upconversion luminescence of Er3+/Yb3+/Nd3+ triply-doped beta-NaYF4 nanocrystals under 808-nm excitation. Materials. 2014;7(11):72897303.

[77] Yi GS, Lu HC, Zhao SY, Yue G, Yang WJ, Chen DP, et al. Synthesis, characterization, and biological application of size-controlled nanocrystalline NaYF4 : Yb,Er infrared-to-visible up-conversion phosphors. Nano Lett. 2004;4(11):21912196.

[78] Nyk M, Kumar R, Ohulchanskyy TY, Bergey EJ, Prasad PN. High contrast in vitro and in vivo photoluminescence bioimaging using near infrared to near infrared up-conversion in TM3+ and Yb3+ doped fluoride nanophosphors. Nano Lett. 2008;8(11):38343838.

[79] Patra A, Friend CS, Kapoor R, Prasad PN. Upconversion in Er3+: ZrO2 nanocrystals. J. Phys. Chem. B. 2002;106(8):19091912.

[80] Sivakumar R, van Veggel F, Raudsepp M. Bright white light through up-conversion of a single NIR source from sol-gel-derived thin film made with Ln(3+)-doped LaF3 nanoparticles. J. Am. Chem. Soc. 2005;127(36):1246412465.

[81] Cheng L, Yang K, Li Y, Chen J, Wang C, Shao M, et al. Facile preparation of multifunctional upconversion nanoprobes for multimodal imaging and dual-targeted photothermal therapy. Angew. Chem. Int. Ed. 2011;50(32):73857390.

[82] Zhou J, Yu M, Sun Y, Zhang X, Zhu X, Wu Z, et al. Fluorine-18-labeled Gd3+/Yb3+/Er3+ co-doped NaYF4 nanophosphors for multimodality PET/MR/UCL imaging. Biomaterials. 2011;32(4):11481156.

[83] Zhang H, Li Y, Ivanov IA, Qu Y, Huang Y, Duan X. Plasmonic modulation of the upconversion fluorescence in NaYF4:Yb/Tm hexaplate nanocrystals using gold nanoparticles or nanoshells. Angew. Chem. Int. Ed. 2010;49(16):28652868.

[84] Jain PK, El-Sayed IH, El-Sayed MA. Au nanoparticles target cancer. Nano Today. 2007;2(1):1829.

[85] Turkevich J, Stevenson PC, Hillier J. A study of the nucleation and growth process in the synthesis of colloidal gold. Discuss. Faraday Soc. 1951;(11):55.

[86] Tuncel D, Demir HV. Conjugated polymer nanoparticles. Nanoscale. 2010;2(4):484494.

[87] Daniel MC, Astruc D. Gold nanoparticles: assembly, supramolecular chemistry, quantum-size-related properties, and applications toward biology, catalysis, and nanotechnology. Chem. Rev. 2004;104(1):293346.

[88] Yonezawa T, Kunitake T. Practical preparation of anionic mercapto ligand-stabilized gold nanoparticles and their immobilization. Colloid Surf. A. 1999;149(1-3):193199.

[89] Laurent S, Forge D, Port M, Roch A, Robic C, Elst LV, et al. Magnetic iron oxide nanoparticles: synthesis, stabilization, vectorization, physicochemical characterizations, and biological applications. Chem. Rev. 2008;108(6):20642110.

[90] Sun SH, Zeng H, Robinson DB, Raoux S, Rice PM, Wang SX, et al. Monodisperse MFe2O4 (M = Fe, Co, Mn) nanoparticles. J. Am. Chem. Soc. 2004;126(1):273279.

[91] Oldenburg SJ, Averitt RD, Westcott SL, Halas NJ. Nanoengineering of optical resonances. Chem. Phys. Lett. 1998;288(2-4):243247.

[92] Ban ZH, Barnakov YA, Golub VO, O’Connor CJ. The synthesis of core-shell iron@gold nanoparticles and their characterization. J. Mater. Chem. 2005;15(43):46604662.

[93] Wang LY, Luo J, Maye MM, Fan Q, Qiang RD, Engelhard MH, et al. Iron oxide-gold core-shell nanoparticles and thin film assembly. J. Mater. Chem. 2005;15(18):18211832.

[94] Ji XJ, Shao RP, Elliott AM, Stafford RJ, Esparza-Coss E, Bankson JA, et al. Bifunctional gold nanoshells with a superparamagnetic iron oxide-silica core suitable for both MR imaging and photothermal therapy. J. Phys. Chem. C. 2007;111(17):62456251.

[95] Melancon MP, Lu W, Li C. Gold-based magneto/optical nanostructures: challenges for in vivo applications in cancer diagnostics and therapy. MRS Bull. 2009;34(6):415421.

[96] Nghiem THL, La TH, Vu XH, Chu VH, Nguyen TH, Le QH, et al. Synthesis, capping and binding of colloidal gold nanoparticles to proteins. Adv. Nat. Sci. 2010;1(2):025009.

[97] Oyelere AK, Chen PC, Huang XH, El-Sayed IH, El-Sayed MA. Peptide-conjugated gold nanorods for nuclear targeting. Bioconjug. Chem. 2007;18(5):14901497.

[98] Medley CD, Bamrungsap S, Tan WH, Smith JE. Aptamer-conjugated nanoparticles for cancer cell detection. Anal. Chem. 2011;83(3):727734.

[99] Yang LL, Mao H, Wang YA, Cao ZH, Peng XH, Wang XX, et al. Single chain epidermal growth factor receptor antibody conjugated nanoparticles for in vivo tumor targeting and imaging. Small. 2009;5(2):235243.

[100] Katz E, Willner I. Integrated nanoparticle-biomolecule hybrid systems: synthesis, properties, and applications. Angew. Chem. Int. Ed. 2004;43(45):60426108.

[101] Zhong GX, Liu AL, Chen XH, Wang K, Lian ZX, Liu QC, et al. Electrochemical biosensor based on nanoporous gold electrode for detection of PML/RAR alpha fusion gene. Biosens. Bioelectron. 2011;26(9):38123817.

[102] Rosi NL, Giljohann DA, Thaxton CS, Lytton-Jean AKR, Han MS, Mirkin CA. Oligonucleotide-modified gold nanoparticles for intracellular gene regulation. Science. 2006;312(5776):10271030.

[103] Pilo-Pais M, Goldberg S, Samano E, LaBean TH, Finkelstein G. Connecting the nanodots: programmable nanofabrication of fused metal shapes on DNA templates. Nano Lett. 2011;11(8):34893492.

[104] Mirkin CA, Letsinger RL, Mucic RC, Storhoff JJ. A DNA-based method for rationally assembling nanoparticles into macroscopic materials. Nature. 1996;382(6592):607609.

[105] Plata DL, Gschwend PM, Reddy CM. Industrially synthesized single-walled carbon nanotubes: compositional data for users, environmental risk assessments, and source apportionment. Nanotechnology. 2008;19(18):185706185720.

[106] Fischer HC, Chan WCW. Nanotoxicity: the growing need for in vivo study. Curr. Opin. Biotechnol. 2007;18(6):565571.

[107] Oberdorster G, Oberdorster E, Oberdorster J. Nanotoxicology: an emerging discipline evolving from studies of ultrafine particles. Environ. Health Perspect. 2005;113(7):823839.

[108] Donaldson K, Poland CA. Nanotoxicity: challenging the myth of nano-specific toxicity. Curr. Opin. Biotechnol. 2013;24(4):724734.

[109] Docter D, Strieth S, Westmeier D, Hayden O, Gao MY, Knauer SK, et al. No king without a crown—impact of the nanomaterial-protein corona on nanobiomedicine. Nanomedicine. 2015;10(3):503519.

[110] Singh S, Nalwa HS. Nanotechnology and health safety—toxicity and risk assessments of nanostructured materials on human health. J. Nanosci. Nanotechnol. 2007;7(9):30483070.

[111] Gatti AMM. Nanopathology: The Health Impact of Nanoparticles. Singapore: Pan Stanford Publishing; 2008.

[112] Witzmann FA. Effect of carbon nanotube exposure on keratinocyte protein expression. In: Monteiro-Riviere NA, Tran CL, eds. Nanotoxicology: Characterization, Dosing, Health, Effects. New York: Informa Healthcare; 2007.

[113] Stone V, Johnston H, Schins RPF. Development of in vitro systems for nanotoxicology: methodological considerations. Crit. Rev. Toxicol. 2009;39(7):613626.

[114] Johnston HJ, Hutchison G, Christensen FM, Peters S, Hankin S, Stone V. A review of the in vivo and in vitro toxicity of silver and gold particulates: particle attributes and biological mechanisms responsible for the observed toxicity. Crit. Rev. Toxicol. 2010;40(4):328346.

[115] Cancino J, Nobre TM, Oliveira ON, Machado SAS, Zucolotto V. A new strategy to investigate the toxicity of nanomaterials using Langmuir monolayers as membrane models. Nanotoxicology. 2013;7(1):6170.

[116] Chang Y, Yang S-T, Liu J-H, Dong E, Wang Y, Cao A, et al. In vitro toxicity evaluation of graphene oxide on A549 cells. Toxicol. Lett. 2011;200(3):201210.

[117] Arora S, Rajwade JM, Paknikar KM. Nanotoxicology and in vitro studies: the need of the hour. Toxicol. Appl. Pharmacol. 2012;258(2):151165.

[118] Donaldson K, Schinwald A, Murphy F, Cho WS, Duffin R, Tran L, et al. The biologically effective dose in inhalation nanotoxicology. Acc. Chem. Res. 2013;46(3):723732.

[119] Oberdoerster G. Nanotoxicology: in vitro-in vivo dosimetry. Environ. Health Perspect. 2012;120(1):A13A20.

[120] Cao Y, Jacobsen NR, Danielsen PH, Lenz AG, Stoeger T, Loft S, et al. Vascular effects of multiwalled carbon nanotubes in dyslipidemic ApoE(/) mice and cultured endothelial cells. Toxicol. Sci. 2014;138(1):104116.

[121] Seabra AB, Paula AJ, de Lima R, Alves OL, Duran N. Nanotoxicity of graphene and graphene oxide. Chem. Res. Toxicol. 2014;27(2):159168.

[122] Johnston HJ, Hutchison GR, Christensen FM, Peters S, Hankin S, Aschberger K, et al. A critical review of the biological mechanisms underlying the in vivo and in vitro toxicity of carbon nanotubes: the contribution of physico-chemical characteristics. Nanotoxicology. 2010;4(2):207246.

[123] Firme III CP, Bandaru PR. Toxicity issues in the application of carbon nanotubes to biological systems. Nanomedicine. 2010;6(2):245256.

[124] Monteiro-Riviere NA, Nemanich RJ, Inman AO, Wang YYY, Riviere JE. Multi-walled carbon nanotube interactions with human epidermal keratinocytes. Toxicol. Lett. 2005;155(3):377384.

[125] Bottini M, Bruckner S, Nika K, Bottini N, Bellucci S, Magrini A, et al. Multi-walled carbon nanotubes induce T lymphocyte apoptosis. Toxicol. Lett. 2006;160(2):121126.

[126] Baktur R, Patel H, Kwon S. Effect of exposure conditions on SWCNT-induced inflammatory response in human alveolar epithelial cells. Toxicol. In Vitro. 2011;25(5):11531160.

[127] Davoren M, Herzog E, Casey A, Cottineau B, Chambers G, Byrne HJ, et al. In vitro toxicity evaluation of single walled carbon nanotubes on human A549 lung cells. Toxicol. In Vitro. 2007;21(3):438448.

[128] Wiwanitkit V, Sereemaspun A, Rojanathanes R. Effect of gold nanoparticles on spermatozoa: the first world report. Fertil. Steril. 2009;91(1):E7E8.

[129] Trickler WJ, Lantz SM, Murdock RC, Schrand AM, Robinson BL, Newport GD, et al. Brain microvessel endothelial cells responses to gold nanoparticles: In vitro pro-inflammatory mediators and permeability. Nanotoxicology. 2011;5(4):479492.

[130] Tahara K, Sakai T, Yamamoto H, Takeuchi H, Hirashima N, Kawashima Y. Improved cellular uptake of chitosan-modi fled PLGA nanospheres by A549 cells. Int. J. Pharm. 2009;382(1-2):198204.

[131] Lin W, Huang Y-w, Zhou X-D, Ma Y. In vitro toxicity of silica nanoparticles in human lung cancer cells. Toxicol. Appl. Pharmacol. 2006;217(3):252259.

[132] Hsin Y-H, Chena C-F, Huang S, Shih T-S, Lai P-S, Chueh PJ. The apoptotic effect of nanosilver is mediated by a ROS- and JNK-dependent mechanism involving the mitochondrial pathway in NIH3T3 cells. Toxicol. Lett. 2008;179(3):130139.

[133] Samberg ME, Oldenburg SJ, Monteiro-Riviere NA. Evaluation of silver nanoparticle toxicity in skin in vivo and keratinocytes in vitro. Environ. Health Perspect. 2010;118(3):407413.

[134] Arora S, Jain J, Rajwade JM, Paknikar KM. Cellular responses induced by silver nanoparticles: in vitro studies. Toxicol. Lett. 2008;179(2):93100.

[135] Zhao J, Castranova V. Toxicology of nanomaterials used in nanomedicine. J. Toxicol. Environ. Health B. 2011;14(8):593632.

[136] Gerber A, Bundschuh M, Klingelhofer D, Groneberg DA. Gold nanoparticles: recent aspects for human toxicology. J. Occup. Med. Toxicol. 2013;8:6.

[137] Berger M. Nano-Society: Pushing the Boundaries of Technology. Cambridge, UK: Royal Society of Chemistry; 2009.

[138] Magrez A, Kasas S, Salicio V, Pasquier N, Seo JW, Celio M, et al. Cellular toxicity of carbon-based nanomaterials. Nano Lett. 2006;6(6):11211125.

[139] Miyawaki J, Yudasaka M, Azami T, Kubo Y, Iijima S. Toxicity of single-walled carbon nanohorns. ACS Nano. 2008;2(2):213226.

[140] Shimizu M, Tainaka H, Oba T, Mizuo K, Umezawa M, Takeda K. Maternal exposure to nanoparticulate titanium dioxide during the prenatal period alters gene expression related to brain development in the mouse. Part. Fibre Toxicol. 2009;6(20):18.

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

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