a concise review of nanomaterials for drug delivery and...

14
Send orders for print-reprints and e-prints to [email protected] Current Nanoscience, 2020, 16, 1-14 1 REVIEW ARTICLE 1573-4137/20 $65.00+.00 © 2020 Bentham Science Publishers A Concise Review of Nanomaterials for Drug Delivery and Release Alfonso Toro-Córdova a , Beatriz Sanz a and Gerardo F. Goya a,b,* a Institute of Nanoscience of Aragon, University of Zaragoza, Zaragoza, Spain; b Condensed Matter Physics Department, Faculty of Sciences, University of Zaragoza, Zaragoza, Spain A R T I C L E H I S T O R Y Received: October 05, 2018 Revised: November 16, 2018 Accepted: July 16, 2019 DOI: 10.2174/1573413715666190724150816 Abstract: This review provides an updated vision about the recent developments in the field of drug vectorization using functional nanoparticles and other nanovectors. From a large number of these nanotechnology-based drug delivery systems that emerge nearly every week, only a tiny fraction reaches a pre-clinical or clinical phase study. In this report, we intend to provide contextual informa- tion about those nanocarriers and release methods that have shown the best outcomes at in vitro and in vivo experiments, highlighting those with proven therapeutic efficiency in humans. From silica- based porous nanoparticles to liposomes or polymeric nanoparticles, each one of these nanosystems has its advantages and drawbacks. We describe and discuss briefly those approaches that, in our cri- terion, have provided significant advancements over existing therapies at the in vivo level. This work also provides a general view of those commercially available nanovectors and their specific area of therapeutic action. Keywords: Drug release, drug delivery system, nanoparticles, liposome, therapeutic drugs, targeting, in vivo drug delivery. 1. INTRODUCTION The idea of controlled drug release, as opposed to imme- diate release, refers to the control of the temporal and spatial presentation of a therapeutic molecule in the body. Sus- tained-release dosages aim to maintain the drug availability within specific therapeutic windows of concentration and time. The use of micro- and nanomaterials as drug carriers to improve the drug release is a rather old concept reported already in the ‘80s when micellar solutions were developed for controlled release formulations [1]. Each controlled- release system has to be designed not only with specific at- tention on the chemistry of the drug to be administrated but also considering the specificities of the targeted organ (e.g., crossing physiological barriers, protect from quick elimina- tion, etc.). The advent of nanotechnology to the field of medicine made possible novel surface-modified nanocarriers for drug targeting and controlled release. This review aims to provide a succinct revision of those recently developed drug release methodologies that, in our opinion, have shown the best performance in animal models or clinical applications. We start with a description of some types of carriers to be used in drug vectorization, followed by few in vitro success- ful experiments with those nanosystems. Finally, we ap- praise the state-of-the-art regarding those commercial prod- ucts of nanocarrier-based drug delivery systems (DDS) al- ready approved by the regulatory agencies. *Address correspondence to this author at the Institute of Nanoscience of Aragon(INA), Department of Condensed Matter Physics, Faculty of Sci- ences, University of Zaragoza, 50018, Zaragoza, Spain; Tel: ++34-876-555- 362; Fax: +34-976-762-776; E-mail: [email protected] 2. TYPES OF CARRIERS 2.1. Liposomes Lipidic carriers are suitable for in vivo applications due to their similitude with biological structures that veil the im- mune system and result in very long circulation times after intravenous administration. The low immune response against liposomes is partially due to a reduced phagocytosis, a mechanism demonstrated at the in vitro level using blood THP-1 monocyte cells in human plasma [2]. Lipid-based nanosystems can be divided into non-lamellar and lamellar types. The latter are usually obtained from the direct assem- bly of amphiphilic building blocks, while the former include liposomes, cubosomes and spongosomes. Liposomes have been reported about 50 years ago [3], but it was only two decades later that their ability as DDS for therapeutic appli- cations was reported [4]. Liposomes can be described as lipid vesicles composed of an aqueous core surrounded by lipid bi-or multilayers, with average sizes around 50-200 nm. The amphipathic character of phospholipids makes possible to encapsulate either hydrophilic molecules at the aqueous core or hydrophobic drugs within the lipid bilayers of the liposomes [5, 6]. 2.2. Extracellular Vesicles An interesting strategy for drug delivery recently devel- oped is the use of extracellular vesicles (EVs) of different intracellular origin [7]. The highly efficient cell uptake mechanisms of EVs allow boosting the delivery doses for not only synthetic drugs, but also RNAs or proteins. EVs- based strategies have the advantage of an easy overcome of

Upload: others

Post on 12-Aug-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: A Concise Review of Nanomaterials for Drug Delivery and ...unizar.es/gfgoya/index_archivos/PDF/CN_DrugRelease2019.pdf · 2 Current Nanoscience, 2020, Vol. 16, No. 00 Toro-Córdova

Send orders for print-reprints and e-prints to [email protected]

Current Nanoscience, 2020, 16, 1-14 1

REVIEW ARTICLE

1573-4137/20 $65.00+.00 © 2020 Bentham Science Publishers

A Concise Review of Nanomaterials for Drug Delivery and Release

Alfonso Toro-Córdovaa, Beatriz Sanza and Gerardo F. Goyaa,b,*

aInstitute of Nanoscience of Aragon, University of Zaragoza, Zaragoza, Spain; bCondensed Matter Physics Department, Faculty of Sciences, University of Zaragoza, Zaragoza, Spain

A R T I C L E H I S T O R Y

Received: October 05, 2018 Revised: November 16, 2018 Accepted: July 16, 2019 DOI: 10.2174/1573413715666190724150816

Abstract: This review provides an updated vision about the recent developments in the field of drug vectorization using functional nanoparticles and other nanovectors. From a large number of these nanotechnology-based drug delivery systems that emerge nearly every week, only a tiny fraction reaches a pre-clinical or clinical phase study. In this report, we intend to provide contextual informa-tion about those nanocarriers and release methods that have shown the best outcomes at in vitro and in vivo experiments, highlighting those with proven therapeutic efficiency in humans. From silica-based porous nanoparticles to liposomes or polymeric nanoparticles, each one of these nanosystems has its advantages and drawbacks. We describe and discuss briefly those approaches that, in our cri-terion, have provided significant advancements over existing therapies at the in vivo level. This work also provides a general view of those commercially available nanovectors and their specific area of therapeutic action.

Keywords: Drug release, drug delivery system, nanoparticles, liposome, therapeutic drugs, targeting, in vivo drug delivery.

1. INTRODUCTION

The idea of controlled drug release, as opposed to imme-diate release, refers to the control of the temporal and spatial presentation of a therapeutic molecule in the body. Sus-tained-release dosages aim to maintain the drug availability within specific therapeutic windows of concentration and time. The use of micro- and nanomaterials as drug carriers to improve the drug release is a rather old concept reported already in the ‘80s when micellar solutions were developed for controlled release formulations [1]. Each controlled-release system has to be designed not only with specific at-tention on the chemistry of the drug to be administrated but also considering the specificities of the targeted organ (e.g., crossing physiological barriers, protect from quick elimina-tion, etc.). The advent of nanotechnology to the field of medicine made possible novel surface-modified nanocarriers for drug targeting and controlled release. This review aims to provide a succinct revision of those recently developed drug release methodologies that, in our opinion, have shown the best performance in animal models or clinical applications. We start with a description of some types of carriers to be used in drug vectorization, followed by few in vitro success-ful experiments with those nanosystems. Finally, we ap-praise the state-of-the-art regarding those commercial prod-ucts of nanocarrier-based drug delivery systems (DDS) al-ready approved by the regulatory agencies.

*Address correspondence to this author at the Institute of Nanoscience of Aragon(INA), Department of Condensed Matter Physics, Faculty of Sci-ences, University of Zaragoza, 50018, Zaragoza, Spain; Tel: ++34-876-555-362; Fax: +34-976-762-776; E-mail: [email protected]

2. TYPES OF CARRIERS

2.1. Liposomes

Lipidic carriers are suitable for in vivo applications due to their similitude with biological structures that veil the im-mune system and result in very long circulation times after intravenous administration. The low immune response against liposomes is partially due to a reduced phagocytosis, a mechanism demonstrated at the in vitro level using blood THP-1 monocyte cells in human plasma [2]. Lipid-based nanosystems can be divided into non-lamellar and lamellar types. The latter are usually obtained from the direct assem-bly of amphiphilic building blocks, while the former include liposomes, cubosomes and spongosomes. Liposomes have been reported about 50 years ago [3], but it was only two decades later that their ability as DDS for therapeutic appli-cations was reported [4]. Liposomes can be described as lipid vesicles composed of an aqueous core surrounded by lipid bi-or multilayers, with average sizes around 50-200 nm. The amphipathic character of phospholipids makes possible to encapsulate either hydrophilic molecules at the aqueous core or hydrophobic drugs within the lipid bilayers of the liposomes [5, 6].

2.2. Extracellular Vesicles

An interesting strategy for drug delivery recently devel-oped is the use of extracellular vesicles (EVs) of different intracellular origin [7]. The highly efficient cell uptake mechanisms of EVs allow boosting the delivery doses for not only synthetic drugs, but also RNAs or proteins. EVs-based strategies have the advantage of an easy overcome of

Page 2: A Concise Review of Nanomaterials for Drug Delivery and ...unizar.es/gfgoya/index_archivos/PDF/CN_DrugRelease2019.pdf · 2 Current Nanoscience, 2020, Vol. 16, No. 00 Toro-Córdova

2 Current Nanoscience, 2020, Vol. 16, No. 00 Toro-Córdova et al.

biological barriers [8]. Additionally, it has been reported that curcumin-loaded EVs increase drug’s solubility, stability and protection against lipopolysaccharide (LPS)-induced micro-glial activation that causes brain inflammation [9]. The dif-ferent biogenesis of EVs determines the final biological function, and can be described with three main classes: exosomes, micro-vesicles and apoptotic bodies [10]. In gen-eral, all types of EVs are composed of a surrounding lipid bilayer, and their sizes can be found to vary widely in the literature from ≈30 nm to about 2 microns in diameter. Among these, exosomes are perhaps the most commonly used because their ability to degrade themselves after fusion with lysosomes at the intracellular space. It has been re-ported that exosomal encapsulation of doxorubicin and pacli-taxel are effective for targeting brain tumors [11]. Exosomes are released into the extracellular environment through an endo-lysosomal pathway, after fusion of multi vesicular en-dosomes with the cell surface (Fig. 1). It has been demon-strated that at both in vitro and in vivo levels, exosomes are mediators of intercellular communication through transfer of messenger molecules [12].

2.3. Nanoparticles

The use of functionalized nanoparticles (NPs) as carriers for controlled vectorization and release of drugs is based on the wide spectrum of physical properties of the NPs core,

offering different ways of on-site drug release by application of a remote physical stimulus such a light, magnetic fields, electromagnetic radiation or ultrasonic waves [13-17]. Each one of these strategies has its pros and cons, and the choice among them depends on the particularities of the specific clinical situation to be solved [18]. 2.3.1. Metallic Nanoparticles

The developments in synthetic chemistry of the last dec-ades have provided many new nanostructured materials with new physicochemical properties. These properties can origi-nate either from the larger surface/volume ratio of nanome-ter-sized materials, or from subtler size-effects of physical nature (e.g., plasmonic resonance, superparamagnetism, etc.). Some successful applications reported in biomedicine have been connected with the use of semiconductor nanoc-rystals [19], gold nanoparticles and silver nanoshells [20, 21]. Gold nanoparticles possess unique photodynamic prop-erties related to the absorption of near-infrared radiation, a mechanism useful for both diagnosis and therapy. These applications have been discussed in detail elsewhere, notably the comprehensive works of Dykman et al. [22] and Panchapakesan et al. [23]. Several noble metals (Pt, Pd, Ag, Au, etc.) are being explored as new, cutting-edge nanostruc-tured biocatalysts due to their abundant exposed active sites and highly accessible surfaces [24]. Moreover, the interac-tion of metallic Au particles with membranes, when heated

Fig. (1). Schematic view of the pathway formation of extracellular vesicles, including exosomes, microvesicles, and apoptotic vesicles. In many cases, extracellular vesicles have signalling functions involved in different biological processes like antigen presentation, activation of cell surfaces through major histocompatibility complex (MHC) molecules, or transporting miRNAs.

Page 3: A Concise Review of Nanomaterials for Drug Delivery and ...unizar.es/gfgoya/index_archivos/PDF/CN_DrugRelease2019.pdf · 2 Current Nanoscience, 2020, Vol. 16, No. 00 Toro-Córdova

Nanomaterials for Drug Release Current Nanoscience, 2020, Vol. 16, No. 00 3

with light, has been recently studied [25]. More rare materi-als like Rhodium (Rh) nanoparticles, synthesized by chemi-cal reduction on polymeric micelle templates, have shown better electro catalytic oxidation of methanol than the bulk counterparts [26], a property that could be adapted to modify cellular bio-catalytic pathways, provided that Rh-toxicity can be kept at low values. 2.3.2. Magnetic Nanoparticles

Magnetic nanoparticles (MNPs) have been projected for remotely controlled release at least since 1969 [27]. The pos-sibility of remote magnetic actuation on MNPs by dc mag-netic fields, and their coupling with ac magnetic fields to release heat have triggered a plethora of nanodevices in bio-medicine, many of them with great success [28, 29]. Al-though these applications are well-known and standardized, it is remarkable that the physical mechanisms that govern drug release under alternate magnetic fields are still a matter of controversy, i.e., whether a mechanical rather than a ther-mal effect is actually acting locally when radiofrequency (RF) magnetic fields are applied. This is partially related to the fact that measuring local temperatures (i.e., at the nanometer scale) at the particle surroundings has proven extremely difficult [30, 31]. In any case, the improvement of the release rates when an electromagnetic field (EMF) of a given frequency is applied has been clearly demonstrated experimentally [32]. The control of drug release with EMF operates in different types of carriers and reservoirs; the most commonly used being surface-coated NPs. However, other thermoresponsive carriers and reservoirs have been reported to be effective, such as thermoresponsive ferrogels [33], thermomagnetic polymer films [34] and polymer films with embedded NPs for synergic uses of drug release and thermal tissue ablation. Regarding MNPs, the selection of magnetic materials of the magnetic cores has been based primarily on biocompatibility requirements, making the iron-oxide based MNPs the most extensively reported so far [35, 36]. A second, more labile criterion for choosing the mag-netic core relies on those intrinsic magnetic parameters like saturation magnetization, magnetic anisotropy, Curie tem-perature, etc. that may improve the magnetic response.

In a recent work using MNPs in a liposomal formulation, Guo et al. [37, 38] were able to deliver anti-neoplastic drugs to the local tumor environment in an animal model. Moreo-ver, they used two concurrent physical stimuli (Infrared light and radiofrequency) to trigger the release bursts of the cyto-toxic drugs within the liposomes, achieving promising thera-peutic results in vivo.

Reviews on the different types of metallic and magnetic materials used to produce nanoparticles for drug delivery can be found elsewhere [18]. Here, we will mention only that for the purposes of drug delivery/release nanosystems translated into clinics, manufacturing these systems under good manu-facturing practice (GMP) standards is technically achievable. On the other side, as nicely described in Ref. [39], the steps to clinical implementation can be difficult and lengthy (see also Section 5.2 below) and it is along these stages that roughly 95% of the designed nano-devices fail to comply the stringent clinical requirements. A prospective analysis of the experimental works of the last years indicated that most of the successful applications in vivo are based on the use of

multi-stimulus DDS like the previously described. The emerging picture is that any single activation method by it-self would be less efficient than the combination of two or more of them, as in the successful case of MNPs and liposomes. These emerging strategies have not yet been fully explored.

3. RELEASE MECHANISMS

3.1. Solid Drugs and Solid Matrices

When a solid dosage form is required for the specific therapeutic situation, it is essential to warrant that drug dis-solution (release from the solid matrix) occurs in an appro-priate manner. Indeed, these drug release kinetic profiles are mandatory from regulatory authorities. The quantitative analysis of the values obtained in dissolution / release tests is usually numerically simulated using models that express the dissolution rate as a function of the chemical parameters for each dosage form. The models are often empirically derived from previous experiments that constitute a databank, al-though in simple cases theoretical models are available [40]. Drug release rates from solid dosage forms have been de-scribed by kinetic models based on the time-dependence of the dissolved amount of drug Q(t). These models are usually referred as zero-order, first-order, etc., reflecting the func-tional dependence of Q with time. The quantitative interpre-tation of the values obtained in the dissolution assay is facili-tated by the usage of a generic equation that mathematically translates the dissolution curve of each pharmaceutical dos-age form. In some cases, that equation can be deduced by a theoretical analysis of the process, as for example in zero-order kinetics.

3.2. Solid Drugs in Water-Soluble Polymers

Many therapeutic drugs have a too-high toxicity profile for a direct administration in vivo. For this reason, many molecules intended for clinical applications require a modi-fied formulation in order to minimize toxicity, as well as to improve efficacy of release kinetics. One of the proposed ways to achieve these goals has been the use of polymeric microspheres and nanospheres that could control the in vivo drug release profiles. Solid dispersions in water-soluble car-riers have been successful in this sense, especially by modi-fying solid-to-liquid release rates of highly hydrophobic drugs [41].

The concept of a ‘solid dispersion’ covers any dosage of one or more active ingredients dispersed in a biologically inert solid matrix [42, 43]. The mechanisms behind the ob-served improved release kinetics, when compared to solid-solid systems, are not yet completely understood, but there is some consensus regarding the key roles of the crystallo-graphic structure and dissolution properties on the final re-lease rates [44]. Some models of carrier-controlled kinetics have been proposed to calculate the release rates in terms of the concentration profile at the polymer layer adjacent to the solid surface [45].

A reported alternative to control the release kinetics is the use of cyclodextrins in polymeric nanospheres and hydrogels [46]. Cyclodextrins are cyclic oligosaccharides that have the ability to form non-covalent complexes, changing the phys-

Page 4: A Concise Review of Nanomaterials for Drug Delivery and ...unizar.es/gfgoya/index_archivos/PDF/CN_DrugRelease2019.pdf · 2 Current Nanoscience, 2020, Vol. 16, No. 00 Toro-Córdova

4 Current Nanoscience, 2020, Vol. 16, No. 00 Toro-Córdova et al.

icochemical properties of the carriers. However, only a few reports on the actual efficacy have been published [47], while none in vivo tests have been reported to the best of our knowledge.

3.3. Coated Tablets

Perhaps the best-known method used to improve high-density formulations is to coat tablets with functional films. Specifically, tablet coatings can be used to achieve drug sta-bilization and to modify/delay drug release when oral deliv-ery forms are required [48]. A group of successful systems developed last decade for colon-specific delivery, which reported good specificity in vivo included pressure-controlled colon delivery capsules (PCDCs) and colonic DDS based on pectin and galactomannan coating. The latter DDS was reported by Lee et al. [49] and consisted of a tablet coated with two specific polysaccharides (pectin and galac-tomannan). The solubility of this coating formulation was found to depend on the pH, being insoluble at pH ≥ 7 but dissolving fast at those pH ≤ 7 typical of intestinal fluids, although the authors did not discuss any model for this pH dependence.

3.4. Remote Radiofrequency-Triggered Release

Non-invasive strategies for remotely triggering drug re-lease have been proposed mainly for liposomal-based nanovectors [50]. These strategies can be based on different triggering stimuli, including (but not restricted to) enzymatic [51], temperature [52], light [53], magnetic fields [54] and ultrasound [55]. Of special clinical interest is the control of drug release profiles on demand by a remote alternating magnetic field of low frequency (i.e., at the low part of the RF spectrum, 100-800 kHz), because of the deep penetration of these waves without noticeable interaction with biological tissues. In order to load the MNPs with the therapeutic drug to be delivered, several coating/functional approaches are possible. Iron oxide MNPs attached to porous silica have been reported as an efficient way to store chemical payloads and release them upon triggering using an RF field [56]. In the same work, an increase of desorption rates under applied RF magnetic fields has been reported. Remotely triggered DDS for cancer applications relies deeply on the recent im-provements for the synthesis of MNPs and MNPs-based polymeric materials like ferrogels [57] or magnetoliposomes (MLPs) [58, 59]. The uniqueness of remote triggering by low-frequency magnetic fields relates with large penetration depth of these waves into the human body without interac-tion with the healthy organs or tissues. The price to pay, on the other hand, is the need of delivering a ‘heating agent’ i.e., the magnetic nanoparticles to the target site to be heated.

3.5. Remote Ultrasound-Triggered Release

Ultrasound is another non-invasive, remote stimulus that can be focused on targeted sites for triggering drug release. Since ultrasonic waves have the additional effect of increas-ing the permeability of blood-tissue barriers and cell mem-branes, they can be used synergistically to control release kinetics [60]. So far, the liposomal formulations are the nanocarriers with best response to ultrasound stimulation. Acoustically active liposomes with an average size of 800

nm have already been used as contrast agents for ultrasound image enhancement [61, 62], while magneto-liposomes offer the additional possibility of magnetic resonance imaging protocols [63]. In the case of acoustic imaging, the main mechanism seems to be associated with air vesicles encapsu-lated within liposomes, which determine their acoustic activ-ity [64]. Moreover, these liposomes have the potential to respond to ultrasound stress by releasing their contents.

4. TESTING DRUG DELIVERY/RELEASE IN VITRO

The main sequence of in vitro events involving targeting, internalization and drug release is common to almost any DDS, since target sites for most therapies are located at the intracellular space. As an example, cationic liposomes pos-sess the ability to form complexes (lipoplexes) with nucleic acids through electrostatic interaction, which determines their transfection efficiency. Improved internalization of DDS is usually achieved through surface functionalization with cell penetrating peptides or membrane permeable ligands.

Once at the intracellular space, different DDS are de-signed to react to the cytoplasm conditions to stimulate drug release. In such strategies, the nanocarrier is usually de-signed to be sensitive to cell temperature (T); pH changes; redox potential and/or elevated reactive oxygen species (ROS) levels (Fig. 2). The different physical stimuli used to remotely activate release mechanisms include radiofre-quency (RF), near-infrared light (NIR) and ultrasound (US), already discussed in the previous section.

In vitro release tests are recognized as a significant piece of pharmaceutical information since in vitro dosage form can influence the release kinetics [65]. Under certain conditions, it can be a good estimator for assessment of bioequivalence. Most of the nanosystems described in the previous section have transited different processes towards final clinical or biomedical applications. While many of them have been tested in vitro for a variety of cell lines, only a relatively small number of these proof-of-concept nanovectors have been tested in animal models (see below). In the following subsections we review the most representative approaches tested so far in primary cell types and some that have gone all the way through the in vitro tests to the clinical approval.

Polymer-based carriers are among the systems exten-sively tested in vitro [66]. The presence of primary, secon-dary, and/or tertiary amine groups provide effective, linkable surfaces with negatively charged molecules of biological interest (e.g., nucleic acids). Several cationic polymers have been tested, and Olden et al. [67] reported that a subset of sunflower-shaped polymer NPs could mediate gene transfec-tion on both cultured and primary cells with efficiencies up to 50% in the Jurkat human T cell line. Additionally, these NPs were reported to show minimal concomitant toxicity (> 90% viability), opening possible ways to optimize primary T-cell transfection conditions including activation time, cell density, DNA dose, culture media, and cytokine treatment.

Another approach that has consensus regarding its effi-cacy in vitro is the use of ferritin-based nanocarriers. Ferritin is an iron-storage protein consisting of 24 subunits that self-assemble to form spherical nanocages of around 12 nm in

Page 5: A Concise Review of Nanomaterials for Drug Delivery and ...unizar.es/gfgoya/index_archivos/PDF/CN_DrugRelease2019.pdf · 2 Current Nanoscience, 2020, Vol. 16, No. 00 Toro-Córdova

Nanomaterials for Drug Release Current Nanoscience, 2020, Vol. 16, No. 00 5

diameter with an interior cavity of 8 nm [68]. Chemical groups (N- and C- terminal sites) present at the outer surface provide the possibility to bind different compounds. Zhang and co-workers presented the use of apoferritin as a delivery agent, with reports of low cytotoxicity and low immune re-sponse [69]. Translocation of apoferritin cage allowed the encapsulation of doxorubicin (Dox) inside the cavity form-ing a stable complex (H-Dox). Due to the chemical and size properties, H-Dox is able to pass the blood–brain barrier in vivo. Zhang et al. showed its potential using cerebellar or-ganotypic cultures and suggested the possibility of using a double chambered cage containing different targets. This idea was further developed by Kin et al. [70] who described the formation of double-chambered ferritin cage NPs in which the first N-terminal chamber is loaded with peptide and the second C-terminal chamber to be loaded with a tu-mor-targeting pro-apoptotic peptide and the green fluores-cent protein (GFP), respectively.

Related to the above approach, micellar nanocarriers have been also used at the in vitro level to test their multi-functional potential. For example, Jing et al. [71] presented a micellar nano-platform capable of simultaneously showing high cell penetration and nuclear targeting through pH-triggered surface charge reversal. The results reported on the mice bearing 4T1 breast tumor showed enhanced cellular internalization and low side effects of encapsulated drugs, while keeping therapeutic efficacy. It is interesting to note that within the tumor tissue (acidic pH) the system exhibited negative to positive charge reversal, facilitating the cell in-

ternalization and subsequent nuclear targeting. The antican-cer drug model used 10-hydroxycamptothecin conjugated to methoxy polyethylene glycol in order to improve the cyto-toxicity.

Liposomes are perhaps the most successful formulation for clinical drug delivery which is discussed in detail in the next Section regarding their in vivo application. Here, we only mentioned that surface-modified liposomes having ar-ginine-rich cell-penetrating peptides and transferrin [72] have been reported to improve targeting onto A2780 ovarian carcinoma cells via the over-expressed transferrin receptors. In the same way, Patil et al. [73] reported liposomes contain-ing a dual chemotherapeutic load of lipophilic mitomycin C and doxorubicin.

A lipid-polymer hybrid liposomal nano-platform based on hyaluronic acid-magnetic nanoparticle-liposomes have shown also good results as a vehicle for docetaxel by im-proving the cellular uptake in human breast cancer cells [74]. In addition, they proved the double effect of chemotherapy and thermotherapy under near-infrared laser irradiation (808 nm) [75].

As mentioned in Section 1, inorganic metallic NPs have also been exploited to transport and release therapeutic drugs by chemical or physical stimuli. The most extensively inves-tigated material is Au-NPs. Functionalized gold nanoparti-cles are known to have good biocompatibility and versatile shape and size, providing a tunable response to physical stimuli. Recently, Hernández Montoro et al. [76] tested a

Fig. (2). Schematic illustration of drug carrier internalization into the intracellular space and its activation by different external or internal stimuli.

Page 6: A Concise Review of Nanomaterials for Drug Delivery and ...unizar.es/gfgoya/index_archivos/PDF/CN_DrugRelease2019.pdf · 2 Current Nanoscience, 2020, Vol. 16, No. 00 Toro-Córdova

6 Current Nanoscience, 2020, Vol. 16, No. 00 Toro-Córdova et al.

drug release system based on Au-nanostars coated with a mesoporous silica in HeLa cells. The silica shell was loaded with doxorubicin (Dox) and coated with octadecyl-trimethoxy-silane and paraffin to retard the release of Dox. Low cytotoxicity confirmed the efficacy of the coating to prevent Dox leaking, and using an 808 nm laser irradiation, release profile was accelerated. Spherical Au-NPs have been also tested in vitro, using different coatings, with different degrees of success [77].

The development of drug-vehicles made of materials with dual effect has also witnessed some successful in vitro experimentation. Rodrigues et al. [78] used graphene-based magnetic nanoparticles functionalized with pluronic F-127, demonstrating efficacy on concurrent hyperthermia and pH stimuli-responsive drug delivery on Hep-G2 cells. This nanosystem showed the potential ability to transport thera-peutic doses of doxorubicin, and to respond to both pH changes in tumor environment and to external magnetic fields. In addition, albumin-based MNPs with low toxicity and immunogenicity, have been tested in vitro by Nostari et al. [79], who applied albumin-coated MNPs attached with curcumin (CUR) in vitro, aiming to decrease tumor growth through proliferation suppression. Although this compound presents promising properties, it is not commonly used due to the short biological half-life and low solubility, which results in poor absorption and thus low bioavailability through the oral route.

5. TARGETED NANOCARRIERS FOR IN VIVO DRUG DELIVERY

Current nano-based DDS is still far from the concept of “magic bullet” created by Paul Ehrlich in the early 1900s. Most of them rely on passive drug accumulation in desired tissues, taking advantage of physiological conditions like the enhanced permeation and retention(EPR) effect, which has been described since approximately 30 years [80]. However, only a marginal increase in drug concentration has been reached with passively guided strategies, and this often rep-resents only a very small fraction of the total dose adminis-tered. For this reason, active targeting is necessary to achieve specific tissue/organ accumulation. Active drug transport is based on molecular recognition processes to deliver drugs to specific pathological sites. Most active guided nanosystems employ targeting ligands at their surface, including organic molecules, antibodies, aptamers, proteins or peptides [81].

Drug delivery to the central nervous system (CNS) is of-ten limited by the very low permeability of the blood-brain barrier (BBB), which is composed of a continuous and charge-polarized endothelial cell layer. Therefore, most of the drugs must travel across a trans-cellular hurdle employ-ing a set of transporters before targeting brain tissues. There are good recent reviews on the use of diverse nanocarriers for treating neurodegenerative and other CNS diseases [82, 83]; therefore, we only mention a few examples here. For instance, a PEGylated liposomal system functionalized with transferrin (Tf) and penetratin (Pen) for targeting BBB endo-thelial cells was recently tested for transfection both in vitro and in vivo [84]. In the latter, this PenTf-liposome was used to assess gene delivery of β-galactosidase-peptide through monitoring β-galactosidase expression in mice tissues. The

results showed that PenTf-liposomes were able to cross BBB and accumulate in brain (12% of the injected dose), and had better transfection capacity than plain liposomes.

Recently, Shahin et al. [85] synthesized a nanocarrier based on PEI-coated mesoporous silica nanoparticles (MSNs) conjugated with hyaluronic acid (HA) for targeting CD44 expressing-recurrent ovarian cancer stem cells. A novel small interfering RNA against TWIST protein (siT-WIST) was further conjugated with the MSNs-HA particles aiming to suppress TWIST and reduce chemoresistance when administered in combination with cisplatin. In vivo results showed 60% tumor growth suppression in the cisplatin group compared to control, MSN-siTWIST reduced it an additional 20%, but the MSN-HA-siTWIST + cisplatin group exhibited the best efficacy, reducing almost a 90% of the tumor vol-ume. Future work could show the potential applications of this system in drug delivery.

Another option for active drug delivery of anti-neoplastic drugs is targeting the tumor vasculature. RGD peptide can bind to αvβ5 integrin, which is often overexpressed in some types of tumors. Song et al. [86] developed a novel nanos-tructured lipid carrier (NLC) composed of a mixture of solid and liquid state lipids conjugated with RGD and encapsulat-ing anticancer drug temozolomide (TMZ) for glioblastoma multiforme treatment. RGD-TMZ/NLCs efficacy was tested on U87 MG-bearing mice, treatments were given intrave-nously and repeated once every three days over a period of 21 days. At the end of the study, RGD-TMZ/NLCs showed a 4-fold inhibition of tumor growth compared with the free drug treatment group. Furthermore, tumor growth inhibition was significantly increased compared with non-RGD guided NLCs, confirming active targeting to tumor vasculature via RGD peptide.

5.1. Stimuli-responsive Control of Drug Release

Controlling drug release rates using external stimuli on nanocarriers at target sites is a persistent challenge. This non-invasive strategy has still many issues that need to be solved before it can be safely translated into clinics. The research of new components for nanocarriers has allowed the development of systems with dual or multi stimuli-triggered drug release mechanisms through both internal (pH, tem-perature, redox potential, ROS generation, enzymes, hy-poxia) and external (ultrasound, light, radiofrequencies, magnetic fields) stimulus. There are many tests currently running on these novel carriers, both in vitro and in vivo [87-89].

The liposomal formulations are perhaps the more suc-cessful DDS currently employed in clinics, due to their phys-icochemical compatibility, pharmacokinetic behaviour and delivery efficacy [90]. Thus, liposomes have been used as the starting point to build many stimuli-responsive DDS. For example, pegylated liposomes loaded with cisplatin and MNPs (i.e., magnetoliposomes) have been tested in vitro and in vivo regarding their transition temperature (Tm), encapsu-lation efficiency and drug release profiles [91]. This new type of DDS showed the capacity of inhibiting tumor growth, and when combined with magnetic hyperthermia (MHT) treatments a complete obliteration of the tumors was reported [91].

Page 7: A Concise Review of Nanomaterials for Drug Delivery and ...unizar.es/gfgoya/index_archivos/PDF/CN_DrugRelease2019.pdf · 2 Current Nanoscience, 2020, Vol. 16, No. 00 Toro-Córdova

Nanomaterials for Drug Release Current Nanoscience, 2020, Vol. 16, No. 00 7

Another promising formulation of magneto-liposomes encapsulating both iron oxide magnetic nanoparticles and a commercially available photosensitizer (Foscan, m-THPC) was tested for a combination of MHT and photodynamic therapy (PDT) for tumor ablation. The formulation yielded satisfactory ratios of both agents and showed good stability in vitro (less than 20% of m-THPC released during the first seven days). In vivo, combination of MHT and PDT with the m-THPC-loaded MLPs showed better efficacy than individ-ual treatments and was able to completely eradicate tumours in A431 xenografts [92].

After the first applications of hyperthermia, surviving cancer cells often develop thermo-resistance associated with increased expression of heat shock proteins (HSPs) reducing subsequent heating efficacy and tumor relapses. 17-AGG, an HSP90 inhibitor derived from geldanamycin, has shown ca-pacity for killing tumor cells by inhibiting HSP90, however, its poor solubility in water limits its use. In order to over-come this limitation, Yang et al. [93] encapsulated 17-AGG inside thermosensitive MLPs and further modified them for targeting CD90+ liver cancer stem cells (CD90@17-AGG/TMs). To prove the antitumor efficacy of the formula-tion, CD90@17-AGG/TMs were intratumorally injected in CD90+LCSCs-bearing mice and 24h after administration, tumors were exposed to an alternating current magnetic field (ACMF, f=200 kHz, I=20 A) for 1 h, every other day for a week. Compared to the other experimental groups, CD90@17-AGG/TMs + ACFM showed a superior reduction of tumor size (83.44 ± 5.78% inhibition rate) and increased the number of apoptotic cells in tumor tissue.

Dual stimuli-responsive polymeric micelles composed of methoxy poly(ethylene glycol)-co-poly (lysine) (mPEG-PLys) and a mixture of fatty acids including linoleic acid, α-linoleic acid and araquidonic acid as building blocks were developed by Gao et al. Chlorin e6 (Ce6) was loaded within the micelle core. Ce6 is a photosensitizer, which, upon laser irradiation, can produce highly reactive oxygen species (ROS). These ROS not only exert a cytotoxic effect on can-cer cells but also act like a trigger mechanism, causing lipid peroxidation of the unsaturated fatty acids and destabilizing the vehicle for a rapid release of Ce6 and a maximization of the ROS effects. This hypothesis was tested in vivo in tumor-bearing mice after intravenous administration of either free Ce6 or its micellar form (RMAA). The authors found that after 24h post-administration, Ce6 concentration in tumor tissue was much higher in RMAA group; also RMAA exhibited the best antitumor outcome reducing tumor volume and im-proving survival time [94].

Shi et al. demonstrated another example of ROS-mediated drug release [95]. These authors succeeded in formulating mesoporous TiO2 nanoparticles with entrapped doxorubicin and coupling DNA fragments at the surface to block doxoru-bicin leakage. The nanosystem exhibited very good stability under physiological conditions. However, under ultrasound stimulation TiO2 can produce ROS that causes dsDNA shear-ing and therefore increasing doxorubicin release. This was first demonstrated in vitro, where nanoparticles without ultra-sound stimulation released only 11% of encapsulated doxoru-bicin after 48 h while 83% was released when ultrasound was applied. These results encouraged authors to analyze the pos-

sible antitumor efficacy in vivo. NBMTNs were intravenously injected in MCF-7/ADR tumor bearing mice and ultrasound irradiation (1W/cm2, 120S) was performed for 24h on tumor site. Results showed that NBMTNs + US causes a 5-fold higher inhibition of tumor growth compared with control group and 4-fold inhibition compared with free doxorubicin or NBMTNs without ultrasound.

A novel system composed of redox/enzyme self-assembling polymeric NPs has shown promising results [96]. This complex system (CSCD) was obtained by conjugating chondroitin sulphate (CS) with deoxicholic acid (DOCA) through a redox-sensitive disulphide bond. A therapeutic drug, docetaxel (DTX), was further encapsulated within these nanoparticles. Hyal-1 is a lysosomal enzyme overex-pressed in various types of cancer cells, associated with growth, metastasis and angiogenesis of tumors. Since HA can be degraded by Hyal-1 and the structural similarities between CS and HA, it is believed that Hyal-1 can also de-grade CS. Therefore, controlled release in these nanoparti-cles can be achieved by Hyal-1 induced CS degradation and glutation (GSH) induced cleavage of disulphide bonds. This hypothesis was tested both in vitro in B16F10 cells (murine melanoma) and in vivo using B16F10 tumor bearing mice. Compared with commercial docetaxel (Taxotere®), CSCD showed a 4-fold higher area under the receiver-operating curve in tumor. Furthermore, CSCD showed increased tumor growth inhibition and reduced lung metastasis.

Theranostic nanosystems appear to be the next step to-wards personalized medicine. Formulations that can be ap-plied for both diagnostic and therapeutic purpose have at-tracted much attention in the last years. As an example, the recently reported nanosystem composed of liposomes with co-encapsulated perfluorocarbon (PFC), hollow gold nano-spheres (HAuNS), and doxorubicin, was intended for ultra-sound imaging, photothermal therapy (PTT) and tempera-ture-controlled drug release purposes, all triggered by NIR light irradiation. This and other types of new, multipurpose nanosystems have started to provide quite promising results both in vitro and in vivo [37].

5.2. Translation into Clinics

Nanotechnology came into the field of DDS about two decades ago, opening new ways to improve the spatial and temporal control of the therapeutic release profiles. This re-sulted in a large number of new concepts and nanodevices with proven experimental success in vitro or in vivo. Fig. (3) resumes the main key-points involved from the initial devel-opment of any DDS to the final clinical approval and com-mercialization. Although many new developments subjected to clinical trials use clinically approved drugs, a full charac-terization phase for each DDS is required since small changes in composition or physicochemical parameters like particle size, zeta potential or drug-release patterns could cause a com-pletely different behavior. After the formulation phase, pre-clinical studies both in vitro and in vivo are necessary to prove the bioavailability, efficacy and safety properties of the nano-carrier before their use in patients. Furthermore, clinical stud-ies from phase-I to phase-III must guarantee the bioequiva-lence, safety profile and better/similar efficacy compared with the standard drug before finally being approved in clinics.

Page 8: A Concise Review of Nanomaterials for Drug Delivery and ...unizar.es/gfgoya/index_archivos/PDF/CN_DrugRelease2019.pdf · 2 Current Nanoscience, 2020, Vol. 16, No. 00 Toro-Córdova

8 Current Nanoscience, 2020, Vol. 16, No. 00 Toro-Córdova et al.

Fig. (3). Key-points along the efficacy/safety studies for new nanomaterial-based drug delivery systems, before they can be approved for clinical uses.

Table 1. Nanotechnology-based drug delivery systems approved for clinical use.

Category Name Drug Disease

Nanocrystals ZYPAdhera Olanzapine Aggression, bipolar disorders, manic episodes, schizophrenia

KadcylaTM Trastuzumab / emtansine Metastatic breast cancer

Adcetris® Brentuximabvedotin Hodgkin’s Lymphoma, anaplastic large cell lymphoma

Mylotarg® Gentuzumab / Ozogamicin Acute myeloid leukemia Antibody-drug conjugate

ZevalinTM 90Y Non-Hodgkin’s lymphoma

Doxil® / Caelyx® Doxorubicin Kaposi’s Sarcoma, ovarian and breast cancer, Multiple mye-

loma

Myocet® Doxorubicin Combination therapy for metastatic breast cancer

LipoDox ® Doxorubicin Kaposi’ sarcoma, breast and ovarian cancer

Daunoxome ® Daunorubicin Kaposi’s sarcoma

DepoCytTM Cytarabine Lymphomatous meningitis

Marqibo ® Vincristine Acute lymphoblastic leukemia

Liposomes

Onco TCS® Vinciristine Relapsed aggressive non-Hodgkin’s lymphoma

Table 1 contd….

Page 9: A Concise Review of Nanomaterials for Drug Delivery and ...unizar.es/gfgoya/index_archivos/PDF/CN_DrugRelease2019.pdf · 2 Current Nanoscience, 2020, Vol. 16, No. 00 Toro-Córdova

Nanomaterials for Drug Release Current Nanoscience, 2020, Vol. 16, No. 00 9

Category Name Drug Disease

Onivyde® Irinotecan Pancreatic cancer

Mepact Mifamurtide Osteosarcoma

Abelcet® Amphotericin B Severe fungal infections

Ambisome® Amphotericin B Fungal infections

Amphotec® Amphotericin B Aspergillosis, visceral leishmaniasis

DoceAqualipTM Docetaxel Breas, gastric, head and neck and prostate cancers

DepoDur® Morphine Postoperative pain

Epaxal® Hepatitis A vaccine Hepatitis A

Inflexal® V Influenza virus vaccine Influenza virus infections

Visudyne® Verteporfin Age-related macular degeneration, choroidal neovasculari-

zation

Zinostatinstimalamer Neocarzinostatin Liver cancer

Lupron Depot® Leuprolide Prostate, breast, ovarian and endometrial cancers. Endo-

metriosis, infertility, benign prostatic hyperplasia Polymeric Nanoparticles

Abraxane® Paclitaxel

Metastatic breast cancer, non-small cell lung cancer, pan-creatic, ovarian, fallopian tubes and primary peritoneal

cancers

Polymeric Micelles Genexol®PM Paclitaxel Metastatic or recurrent breast cancer, Advanced or metas-

tatic non-small cell lung cancer.

Oncaspar® L-asparaginase Acute lymphoblastic leukemia, Non-Hodgkin’s lymphoma

Adagen® Adenosine deaminase Severe combined immunodeficiency (SCID)

Pegasys® IFN-α-2a Hepatitis B and C Polymer-Drug conjugate

CImzia® Fab’ fragment against TNFα Active rheumatoid arthritis, axial spondyloarthritis, anky-losing spondylitis, no-radiographic axial spondyloarthritis

and psoriatic arthritis

Inorganic Nanoparticles NanoTherm® Iron Oxide Nanoparticles Glioblastoma

In the last few years, a large number of experimental

nanocarriers have been developed and tested in preclinical stages [97]. Unfortunately, despite the substantial research showing promising results both in vitro and in vivo with this nanomaterials for controlled drug delivery and release, there are only few of them that have been successfully translated to clinical studies and even less have been approved for clinical use (Table 1). However, owing to the increasing in-formation in molecular biology of different diseases and the positive results in preclinical stages, the number of nanoma-terials in clinical stages is expected to increase in the next years [98-100].

CONCLUSION

Many long-standing glitches that hindered therapeutic applications of DDS have been overcome by the incorpora-tion of nanotechnology and nanomaterials into the main original conception. Although solid matrix formulations are still irreplaceable for some therapeutic goals, nanosystems with active control of release kinetics are now a real possibil-

ity. The review of the current literature shows that those DDSs currently available have improved the pharmacologi-cal outcomes of several therapeutic strategies, making a real impact on the clinical activity. Among the DDS formulations with evident success, anticancer-drug delivery systems stand out. Moreover, the basic templates already developed for some clinically-approved DDS can be adapted for new drugs still to be synthesized from the (also) fast-growing field of chemical engineering and drug screening. On the other hand, the success regarding DDS having remote wireless control of drug release is yet to be proven at clinical stages. These would be the next generation DDS that could resemble the long-standing conception of a ‘magic bullet’.

LIST OF ABBREVIATIONS

BBB = blood-brain barrier Ce6 = chlorin e6 CNS = central nervous system CS = chondroitin sulphate

Page 10: A Concise Review of Nanomaterials for Drug Delivery and ...unizar.es/gfgoya/index_archivos/PDF/CN_DrugRelease2019.pdf · 2 Current Nanoscience, 2020, Vol. 16, No. 00 Toro-Córdova

10 Current Nanoscience, 2020, Vol. 16, No. 00 Toro-Córdova et al.

CSCD = chondroitin sulphate-ss-deoxycholic acid CUR = curcumin DDS = drug delivery systems DNA = deoxyribonucleic acid DOCA = deoxcholic acid Dox = Doxorubicin DTX = docexatel NPs = Nanoparticles EMF = electromagnetic field m-THPC = meta-tetra(hydroxyphenyl)chlorin EPR = enhanced permeation and retention EVs = Extracellular vesicles GFP = green fluorescent protein GHS = glutation GMP = Good Manufacturing Practices HA = hyaluronic acid HAuNS = hollow gold nanospheres HSP = heat shock protein LPS = Lipopolysacharide MHC = major histocompatibility complex mi-RNA = mitochondrial ribonucleic acid MHT = Magnetic Hyperthermia MLP = magneto-liposome MNPs = Magnetic nanoparticles MSNs = mesoporous silica particles NIR = near-infrared light NLC = nanostructured lipid carrier PCDCs = presusure-controlled colon delivery capsules Pen = penetratin PDT = photodynamic therapy PFC = perfluorocarbon PMAA = chlorin e6 micellar form PPT = photothermal therapy RGD = arginylglycylaspartic acid Rh = Rhodium RNAs = ribonucleic acids RF = radiofrequency ROS = reactive oxygen species T = temperature Tm = transmission temperature Tf = transferrin TMZ = temozolomide US = ultrasound.

CONSENT FOR PUBLICATION

Not applicable.

FUNDING  

The authors acknowledge financial support from the Spanish Ministerio de Ciencia, Innovación y Universidades (projects RTC-2017-6620-1 and MAT2016-78201-P) and the Aragon Regional Government (DGA, Project No. E26). Alfonso Toro-Córdova is obliged with Mexican CONACyT for a post-doctoral fellowship (CVU 257448) at INA, Zaragoza.  

CONFLICT OF INTEREST

The authors declare no conflict of interest, financial or otherwise.

ACKNOWLEDGEMENTS

All authors contributed equally to the design, discussion and writing of the manuscript.

REFERENCES [1] Nyström, C.; Bisrat, M. Coulter Counter measurements of solubil-

ity and dissolution rate of sparingly soluble compounds using mi-cellar solutions. J. Pharm. Pharmacol., 1986, 38(6), 420-425.

[http://dx.doi.org/10.1111/j.2042-7158.1986.tb04604.x] [PMID: 2873218]

[2] Bazile, D.; Prud’homme, C.; Bassoullet, M.T.; Marlard, M.; Spen-lehauer, G.; Veillard, M. Stealth Me.PEG-PLA nanoparticles avoid uptake by the mononuclear phagocytes system. J. Pharm. Sci., 1995, 84(4), 493-498.

[http://dx.doi.org/10.1002/jps.2600840420] [PMID: 7629743] [3] Bangham, A.D.; Horne, R.W. Negative staining of phospholipids

and their structural modification by surface-active agents as ob-served in the electron microscope. J. Mol. Biol., 1964, 8(5), 660-668.

[http://dx.doi.org/10.1016/S0022-2836(64)80115-7] [PMID: 14187392]

[4] Mezei, M.; Gulasekharam, V. Liposomes--a selective drug delivery system for the topical route of administration. Lotion dosage form. Life Sci., 1980, 26(18), 1473-1477.

[http://dx.doi.org/10.1016/0024-3205(80)90268-4] [PMID: 6893068]

[5] Sarangi, B.; Jana, U.; Mohanta, G.P.; Manna, P.K. Drug release kinetics study of lovastatin loaded solid lipid nanoparticles for oral delivery. Curr. Nanosci., 2018, 14(4), 319-328.

[http://dx.doi.org/10.2174/1573413714666180201163804] [6] Xing, H.; Hwang, K.; Lu, Y. Recent developments of liposomes as

nanocarriers for theranostic applications. Theranostics, 2016, 6(9), 1336-1352.

[http://dx.doi.org/10.7150/thno.15464] [PMID: 27375783] [7] Soussan, E.; Cassel, S.; Blanzat, M.; Rico-Lattes, I. Drug delivery

by soft matter: matrix and vesicular carriers. Angew. Chem. Int. Ed. Engl., 2009, 48(2), 274-288.

[http://dx.doi.org/10.1002/anie.200802453] [PMID: 19072808] [8] Armstrong, J.P.K.; Stevens, M.M. Strategic design of extracellular

vesicle drug delivery systems. Adv. Drug Deliv. Rev., 2018, 130, 12-16.

[http://dx.doi.org/10.1016/j.addr.2018.06.017] [PMID: 29959959] [9] Sun, D.; Zhuang, X.; Xiang, X.; Liu, Y.; Zhang, S.; Liu, C.;

Barnes, S.; Grizzle, W.; Miller, D.; Zhang, H-G. A novel nanopar-ticle drug delivery system: the anti-inflammatory activity of cur-cumin is enhanced when encapsulated in exosomes. Mol. Ther., 2010, 18(9), 1606-1614.

[http://dx.doi.org/10.1038/mt.2010.105] [PMID: 20571541] [10] EL Andaloussi, S.; Mäger, I.; Breakefield, X.O.; Wood, M.J.A.

Extracellular vesicles: biology and emerging therapeutic opportuni-ties. Nat. Rev. Drug Discov., 2013, 12(5), 347-357.

Page 11: A Concise Review of Nanomaterials for Drug Delivery and ...unizar.es/gfgoya/index_archivos/PDF/CN_DrugRelease2019.pdf · 2 Current Nanoscience, 2020, Vol. 16, No. 00 Toro-Córdova

Nanomaterials for Drug Release Current Nanoscience, 2020, Vol. 16, No. 00 11

[http://dx.doi.org/10.1038/nrd3978] [PMID: 23584393] [11] Yang, T.; Martin, P.; Fogarty, B.; Brown, A.; Schurman, K.;

Phipps, R.; Yin, V.P.; Lockman, P.; Bai, S. Exosome delivered anticancer drugs across the blood-brain barrier for brain cancer therapy in Danio rerio. Pharm. Res., 2015, 32(6), 2003-2014.

[http://dx.doi.org/10.1007/s11095-014-1593-y] [PMID: 25609010] [12] Stoorvogel, W.; Kleijmeer, M.J.; Geuze, H.J.; Raposo, G. The

biogenesis and functions of exosomes. Traffic, 2002, 3(5), 321-330. [http://dx.doi.org/10.1034/j.1600-0854.2002.30502.x] [PMID:

11967126] [13] El-Boubbou, K. Magnetic iron oxide nanoparticles as drug carriers:

preparation, conjugation and delivery. Nanomedicine (Lond.), 2018, 13(8), 929-952.

[http://dx.doi.org/10.2217/nnm-2017-0320] [PMID: 29546817] [14] Giannaccini, M.; Calatayud, M.P.; Poggetti, A.; Corbianco, S.;

Novelli, M.; Paoli, M.; Battistini, P.; Castagna, M.; Dente, L.; Parchi, P.; Lisanti, M.; Cavallini, G.; Junquera, C.; Goya, G.F.; Raffa, V. Magnetic nanoparticles for efficient delivery of growth factors: Stimulation of peripheral nerve regeneration. Adv. Healthc. Mater., 2017, 6(7)

[http://dx.doi.org/10.1002/adhm.201601429] [PMID: 28156059] [15] Cao, Y.; Chen, Y.; Yu, T.; Guo, Y.; Liu, F.; Yao, Y.; Li, P.; Wang,

D.; Wang, Z.; Chen, Y.; Ran, H. Drug release from phase-changeable nanodroplets triggered by low-intensity focused ultra-sound. Theranostics, 2018, 8(5), 1327-1339.

[http://dx.doi.org/10.7150/thno.21492] [PMID: 29507623] [16] Schroeder, A.; Avnir, Y.; Weisman, S.; Najajreh, Y.; Gabizon, A.;

Talmon, Y.; Kost, J.; Barenholz, Y. Controlling liposomal drug re-lease with low frequency ultrasound: mechanism and feasibility. Langmuir, 2007, 23(7), 4019-4025.

[http://dx.doi.org/10.1021/la0631668] [PMID: 17319706] [17] Schwerdt, J.I.; Goya, G.F.; Calatayud, M.P.; Hereñú, C.B.; Reg-

giani, P.C.; Goya, R.G. Magnetic field-assisted gene delivery: achievements and therapeutic potential. Curr. Gene Ther., 2012, 12(2), 116-126.

[http://dx.doi.org/10.2174/156652312800099616] [PMID: 22348552]

[18] McNamara, K.; Tofail, S.A.M. Nanoparticles in biomedical appli-cations. Adv. Phys. X, 2017, 2(1), 54-88.

[19] Noh, Y.W.; Lim, Y.T.; Chung, B.H. Noninvasive imaging of den-dritic cell migration into lymph nodes using near-infrared fluores-cent semiconductor nanocrystals. FASEB J., 2008, 22(11), 3908-3918.

[http://dx.doi.org/10.1096/fj.08-112896] [PMID: 18682573] [20] Huang, L.; Wan, J.; Wei, X.; Liu, Y.; Huang, J.; Sun, X.; Zhang,

R.; Gurav, D.D.; Vedarethinam, V.; Li, Y.; Chen, R.; Qian, K. Plasmonic silver nanoshells for drug and metabolite detection. Nat. Commun., 2017, 8(1), 220.

[http://dx.doi.org/10.1038/s41467-017-00220-4] [PMID: 28790311]

[21] España-Sánchez, B.L.; Ávila-Orta, C.A.; Padilla-Vaca, L.F.; Bar-riga-Castro, E.D.; Soriano-Corral, F.; González-Morones, P.; Ramírez-Wong, D.G.; Luna-Bárcenas, G. Early stages of antibacte-rial damage of metallic nanoparticles by TEM and STEM-HAADF. Curr. Nanosci., 2017, 14(1), 54-61.

[http://dx.doi.org/10.2174/2468187307666170906150731] [PMID: 29399015]

[22] Dykman, L.A.; Khlebtsov, N.G. Multifunctional gold-based nano-composites for theranostics. Biomaterials, 2016, 108, 13-34.

[http://dx.doi.org/10.1016/j.biomaterials.2016.08.040] [PMID: 27614818]

[23] Panchapakesan, B.; Book-Newell, B.; Sethu, P.; Rao, M.; Iru-dayaraj, J. Gold nanoprobes for theranostics. Nanomedicine (Lond.), 2011, 6(10), 1787-1811.

[http://dx.doi.org/10.2217/nnm.11.155] [PMID: 22122586] [24] Koyani, R.; Perez-Robles, J.; Cadena-Nava, R.D.; Vazquez-Duhalt,

R. Biomaterial-based nanoreactors, an alternative for enzyme de-livery. Nanotechnol. Rev., 2017, 6(5), 405-419.

[http://dx.doi.org/10.1515/ntrev-2016-0071] [25] Torchi, A.; Simonelli, F.; Ferrando, R.; Rossi, G. Local enhance-

ment of lipid membrane permeability induced by irradiated gold nanoparticles. ACS Nano, 2017, 11(12), 12553-12561.

[http://dx.doi.org/10.1021/acsnano.7b06690] [PMID: 29161019] [26] Jiang, B.; Li, C.; Dag, Ö.; Abe, H.; Takei, T.; Imai, T.; Hossain,

M.S.A.; Islam, M.T.; Wood, K.; Henzie, J.; Yamauchi, Y.

Mesoporous metallic rhodium nanoparticles. Nat. Commun., 2017, 8, 15581.

[http://dx.doi.org/10.1038/ncomms15581] [PMID: 28524873] [27] Figge, F.H.; Flower, G., Jr Method of administering therapeutic

agents; Inc, A., Ed.; AMP Inc.: USA, 1969. [28] Thanh, N.T. Clinical Applications of Magnetic Nanoparticles:

From Fabrication to Clinical Applications. In: CRC Press; , 2018. [http://dx.doi.org/10.1201/9781315168258] [29] Dobson, J.; El Haj, A.J.; Bin, H.; Markides, H.; Henstock, J.R.

Applications of magnetic nanoparticles in tissue engineering and regenerative medicine.Nanomagnetic Actuation in Biomedicine, Donson, J.; Rinaldi, C; CRC Press, 2018, pp. 205-228.

[http://dx.doi.org/10.1201/9781315356525] [30] Chiu-Lam, A.; Rinaldi, C. Nanoscale thermal phenomena in the

vicinity of magnetic nanoparticles in alternating magnetic fields. Adv. Funct. Mater., 2016, 26(22), 3933-3941.

[http://dx.doi.org/10.1002/adfm.201505256] [PMID: 29225561] [31] Alphandery, E.; Haidar, D.A.; Seksek, O.; Thoreau, M.; Traut-

mann, A.; Bercovici, N.; Gazeau, F.; Guyot, F.; Chebbi, I. A Fluo-rescent nanoprobe for the detection of in situ temperature changes during hyperthermia treatment of tumors. Biophys. J., 2018, 114(3), 361a.

[http://dx.doi.org/10.1016/j.bpj.2017.11.2006] [32] Hayashi, K.; Ono, K.; Suzuki, H.; Sawada, M.; Moriya, M.; Saka-

moto, W.; Yogo, T. High-frequency, magnetic-field-responsive drug release from magnetic nanoparticle/organic hybrid based on hyperthermic effect. ACS Appl. Mater. Interfaces, 2010, 2(7), 1903-1911.

[http://dx.doi.org/10.1021/am100237p] [PMID: 20568697] [33] Hernández, R.; Sacristán, J.; Asín, L.; Torres, T.E.; Ibarra, M.R.;

Goya, G.F.; Mijangos, C. Magnetic hydrogels derived from poly-saccharides with improved specific power absorption: potential de-vices for remotely triggered drug delivery. J. Phys. Chem. B, 2010, 114(37), 12002-12007.

[http://dx.doi.org/10.1021/jp105556e] [PMID: 20806925] [34] Criado, M.; Sanz, B.; Goya, G.F.; Mijangos, C.; Hernández, R.

Magnetically responsive biopolymeric multilayer films for local hyperthermia. J. Mater. Chem. B Mater. Biol. Med., 2017, 5(43), 8570-8578.

[http://dx.doi.org/10.1039/C7TB02361H] [35] Valdiglesias, V.; Fernández-Bertólez, N.; Kiliç, G.; Costa, C.;

Costa, S.; Fraga, S.; Bessa, M.J.; Pásaro, E.; Teixeira, J.P.; Laffon, B. Are iron oxide nanoparticles safe? Current knowledge and future perspectives. J. Trace Elem. Med. Biol., 2016, 38, 53-63.

[http://dx.doi.org/10.1016/j.jtemb.2016.03.017] [PMID: 27056797] [36] Patil, U.S.; Adireddy, S.; Jaiswal, A.; Mandava, S.; Lee, B.R.;

Chrisey, D.B. In vitro/in vivo toxicity evaluation and quantification of iron oxide nanoparticles. Int. J. Mol. Sci., 2015, 16(10), 24417-24450.

[http://dx.doi.org/10.3390/ijms161024417] [PMID: 26501258] [37] Li, W.; Hou, W.; Guo, X.; Luo, L.; Li, Q.; Zhu, C.; Yang, J.; Zhu,

J.; Du, Y.; You, J. Temperature-controlled, phase-transition ultra-sound imaging-guided photothermal-chemotherapy triggered by NIR light. Theranostics, 2018, 8(11), 3059-3073.

[http://dx.doi.org/10.7150/thno.23885] [PMID: 29896302] [38] Guo, Y.; Zhang, Y.; Ma, J.; Li, Q.; Li, Y.; Zhou, X.; Zhao, D.;

Song, H.; Chen, Q.; Zhu, X. Light/magnetic hyperthermia triggered drug released from multi-functional thermo-sensitive magnetol-iposomes for precise cancer synergetic theranostics. J. Control. Re-lease, 2018, 272, 145-158.

[http://dx.doi.org/10.1016/j.jconrel.2017.04.028] [PMID: 28442407]

[39] Tietze, R.; Zaloga, J.; Unterweger, H.; Lyer, S.; Friedrich, R.P.; Janko, C.; Pöttler, M.; Dürr, S.; Alexiou, C. Magnetic nanoparticle-based drug delivery for cancer therapy. Biochem. Biophys. Res. Commun., 2015, 468(3), 463-470.

[http://dx.doi.org/10.1016/j.bbrc.2015.08.022] [PMID: 26271592] [40] Paixão, P.; Gouveia, L.F.; Silva, N.; Morais, J.A.G. Evaluation of

dissolution profile similarity - Comparison between the f2, the mul-tivariate statistical distance and the f2 bootstrapping methods. Eur. J. Pharm. Biopharm., 2017, 112, 67-74.

[http://dx.doi.org/10.1016/j.ejpb.2016.10.026] [PMID: 27865857] [41] Bibby, D.C.; Davies, N.M.; Tucker, I.G. Mechanisms by which

cyclodextrins modify drug release from polymeric drug delivery systems. Int. J. Pharm., 2000, 197(1-2), 1-11.

Page 12: A Concise Review of Nanomaterials for Drug Delivery and ...unizar.es/gfgoya/index_archivos/PDF/CN_DrugRelease2019.pdf · 2 Current Nanoscience, 2020, Vol. 16, No. 00 Toro-Córdova

12 Current Nanoscience, 2020, Vol. 16, No. 00 Toro-Córdova et al.

[http://dx.doi.org/10.1016/S0378-5173(00)00335-5] [PMID: 10704788]

[42] Chiou, W.L.; Riegelman, S. Pharmaceutical applications of solid dispersion systems. J. Pharm. Sci., 1971, 60(9), 1281-1302.

[http://dx.doi.org/10.1002/jps.2600600902] [PMID: 4935981] [43] Juarez, J.M.; Cussa, J.; Gomez Costa, M.B.; Anunziata, O.A.

Nanostructured ketorolac-tromethamine/MCF: Synthesis, charac-terization and application in drug release system. Curr. Nanosci., 2018, 14(5), 432-439.

[http://dx.doi.org/10.2174/1573413714666180222134742] [44] Corrigan, O.I. Mechanisms of dissolution of fast release solid dis-

persions. Drug Dev. Ind. Pharm., 1985, 11(2-3), 697-724. [http://dx.doi.org/10.3109/03639048509056896] [45] Craig, D.Q. The mechanisms of drug release from solid dispersions

in water-soluble polymers. Int. J. Pharm., 2002, 231(2), 131-144. [http://dx.doi.org/10.1016/S0378-5173(01)00891-2] [PMID:

11755266] [46] Quaglia, F.; Varricchio, G.; Miro, A.; La Rotonda, M.I.; Larobina,

D.; Mensitieri, G. Modulation of drug release from hydrogels by using cyclodextrins: the case of nicardipine/beta-cyclodextrin sys-tem in crosslinked polyethylenglycol. J. Control. Release, 2001, 71(3), 329-337.

[http://dx.doi.org/10.1016/S0168-3659(01)00242-5] [PMID: 11295225]

[47] Aloisio, C.; Gomes de Oliveira, A.; Longhi, M. Characterization, inclusion mode, phase-solubility and in vitro release studies of in-clusion binary complexes with cyclodextrins and meglumine using sulfamerazine as model drug. Drug Dev. Ind. Pharm., 2014, 40(7), 919-928.

[http://dx.doi.org/10.3109/03639045.2013.790408] [PMID: 23627444]

[48] Sastry, S.V.; Nyshadham, J.R.; Fix, J.A. Recent technological advances in oral drug delivery - a review. Pharm. Sci. Technol. To-day, 2000, 3(4), 138-145.

[http://dx.doi.org/10.1016/S1461-5347(00)00247-9] [PMID: 10754543]

[49] Lee, S.S.; Lim, C.B.; Pai, C.M.; Lee, S.; Park, I.; Seo, M.G.; Park, H. Composition and pharmaceutical dosage form for colonic drug delivery using polysaccharides. US6413494B1., 2002.

[50] Wang, Y.; Kohane, D.S. External triggering and triggered targeting strategies for drug delivery. Nat. Rev. Mater., 2017, 2, 17020.

[http://dx.doi.org/10.1038/natrevmats.2017.20] [51] Meers, P. Enzyme-activated targeting of liposomes. Adv. Drug

Deliv. Rev., 2001, 53(3), 265-272. [http://dx.doi.org/10.1016/S0169-409X(01)00205-8] [PMID:

11744171] [52] Burke, C.; Dreher, M.R.; Negussie, A.H.; Mikhail, A.S.; Yar-

molenko, P.; Patel, A.; Skilskyj, B.; Wood, B.J.; Haemmerich, D. Drug release kinetics of temperature sensitive liposomes measured at high-temporal resolution with a millifluidic device. Int. J. Hyper-thermia, 2018, 34(6), 786-794.

[http://dx.doi.org/10.1080/02656736.2017.1412504] [PMID: 29284329]

[53] Vankayala, R.; Hwang, K.C. Near-infrared-light-activatable nano-material-mediated phototheranostic nanomedicines: An emerging paradigm for cancer treatment. Adv. Mater., 2018, 30(23)e1706320

[http://dx.doi.org/10.1002/adma.201706320] [PMID: 29577458] [54] Hoare, T.; Timko, B.P.; Santamaria, J.; Goya, G.F.; Irusta, S.; Lau,

S.; Stefanescu, C.F.; Lin, D.; Langer, R.; Kohane, D.S. Magneti-cally triggered nanocomposite membranes: a versatile platform for triggered drug release. Nano Lett., 2011, 11(3), 1395-1400.

[http://dx.doi.org/10.1021/nl200494t] [PMID: 21344911] [55] Salkho, N.M.; Turki, R.Z.; Guessoum, O.; Martins, A.M.; Vitor,

R.F.; Husseini, G.A. Liposomes as a promising ultrasound-triggered drug delivery system in cancer treatment. Curr. Mol. Med., 2017, 17(10), 668-688.

[http://dx.doi.org/10.2174/1566524018666180416100142] [PMID: 29663885]

[56] Soltys, M.; Kovatcik, P.; Lhotka, M.; Ulbrich, P.; Zadrazil, A.; Stepanek, F. Radiofrequency controlled release from mesoporous silica nano-carriers. Microporous Mesoporous Mater., 2016, 229, 14-21.

[http://dx.doi.org/10.1016/j.micromeso.2016.04.009] [57] Cezar, C.A.; Kennedy, S.M.; Mehta, M.; Weaver, J.C.; Gu, L.;

Vandenburgh, H.; Mooney, D.J. Biphasic ferrogels for triggered

drug and cell delivery. Adv. Healthc. Mater., 2014, 3(11), 1869-1876.

[http://dx.doi.org/10.1002/adhm.201400095] [PMID: 24862232] [58] Pradhan, P.; Giri, J.; Rieken, F.; Koch, C.; Mykhaylyk, O.; Döblin-

ger, M.; Banerjee, R.; Bahadur, D.; Plank, C. Targeted temperature sensitive magnetic liposomes for thermo-chemotherapy. J. Control. Release, 2010, 142(1), 108-121.

[http://dx.doi.org/10.1016/j.jconrel.2009.10.002] [PMID: 19819275]

[59] Nappini, S.; Bombelli, F.B.; Bonini, M.; Norden, B.; Baglioni, P. Magnetoliposomes for controlled drug release in the presence of low-frequency magnetic field. Soft Matter, 2010, 6(1), 154-162.

[http://dx.doi.org/10.1039/B915651H] [60] Budhwani, K.I.; Dettmann, M.A.; Saleh, M.N.; Thomas, V. Nano

and microbubble systems for on-demand cancer drug delivery. Curr. Nanosci., 2018, 14(1), 33-41.

[61] Huang, S.L.; Hamilton, A.J.; Nagaraj, A.; Tiukinhoy, S.D.; Kle-german, M.E.; McPherson, D.D.; Macdonald, R.C. Improving ul-trasound reflectivity and stability of echogenic liposomal disper-sions for use as targeted ultrasound contrast agents. J. Pharm. Sci., 2001, 90(12), 1917-1926.

[http://dx.doi.org/10.1002/jps.1142] [PMID: 11745750] [62] Centelles, M.N.; Wright, M.; So, P-W.; Amrahli, M.; Xu, X.Y.;

Stebbing, J.; Miller, A.D.; Gedroyc, W.; Thanou, M. Image-guided thermosensitive liposomes for focused ultrasound drug delivery: Using NIRF-labelled lipids and topotecan to visualise the effects of hyperthermia in tumours. J. Control. Release, 2018, 280, 87-98.

[http://dx.doi.org/10.1016/j.jconrel.2018.04.047] [PMID: 29723616]

[63] Fite, B.Z.; Kheirolomoom, A.; Foiret, J.L.; Seo, J.W.; Mahakian, L.M.; Ingham, E.S.; Tam, S.M.; Borowsky, A.D.; Curry, F.E.; Fer-rara, K.W. Dynamic contrast enhanced MRI detects changes in vascular transport rate constants following treatment with ther-mally-sensitive liposomal doxorubicin. J. Control. Release, 2017, 256, 203-213.

[http://dx.doi.org/10.1016/j.jconrel.2017.04.007] [PMID: 28395970]

[64] Klibanov, A.L. Ligand-carrying gas-filled microbubbles: ultra-sound contrast agents for targeted molecular imaging. Bioconjug. Chem., 2005, 16(1), 9-17.

[http://dx.doi.org/10.1021/bc049898y] [PMID: 15656569] [65] Lin, Z.; Zhou, D.; Hoag, S.; Qiu, Y. Influence of drug properties

and formulation on in vitro drug release and biowaiver regulation of oral extended release dosage forms. AAPS J., 2016, 18(2), 333-345.

[http://dx.doi.org/10.1208/s12248-015-9861-2] [PMID: 26769249] [66] Mukherjee, S.; Shunmugam, R. Polymer based nano-assemblies:

Very efficient carrier in the field of cancer chemotherapy. J. Nanomed. Res., 2017, 5(6), 00137.

[67] Olden, B.R.; Cheng, Y.; Yu, J.L.; Pun, S.H. Cationic polymers for non-viral gene delivery to human T cells. J. Control. Release, 2018, 282, 140-147.

[http://dx.doi.org/10.1016/j.jconrel.2018.02.043] [PMID: 29518467]

[68] Khoshnejad, M.; Parhiz, H.; Shuvaev, V.V.; Dmochowski, I.J.; Muzykantov, V.R. Ferritin-based drug delivery systems: Hybrid nanocarriers for vascular immunotargeting. J. Control. Release, 2018, 282, 13-24.

[http://dx.doi.org/10.1016/j.jconrel.2018.02.042] [PMID: 29522833]

[69] Zhang, L.; Li, L.; Di Penta, A.; Carmona, U.; Yang, F.; Schöps, R.; Brandsch, M.; Zugaza, J.L.; Knez, M. H‐chain ferritin: A natural nuclei targeting and bioactive delivery nanovector. Adv. Healthc. Mater., 2015, 4(9), 1305-1310.

[http://dx.doi.org/10.1002/adhm.201500226] [PMID: 25973730] [70] Kim, S.; Kim, G.S.; Seo, J.; Gowri Rangaswamy, G.; So, I.S.; Park,

R.W.; Lee, B.H.; Kim, I.S. Double-chambered ferritin platform: Dual-function payloads of cytotoxic peptides and fluorescent pro-tein. Biomacromolecules, 2016, 17(1), 12-19.

[http://dx.doi.org/10.1021/acs.biomac.5b01134] [PMID: 26646195] [71] Jing, Y.; Xiong, X.; Ming, Y.; Zhao, J.; Guo, X.; Yang, G.; Zhou,

S. A multifunctional micellar nanoplatform with pH‐triggered cell penetration and nuclear targeting for effective cancer therapy and inhibition to lung metastasis. Adv. Healthc. Mater., 2018, 7(7)e1700974

[http://dx.doi.org/10.1002/adhm.201700974] [PMID: 29334189]

Page 13: A Concise Review of Nanomaterials for Drug Delivery and ...unizar.es/gfgoya/index_archivos/PDF/CN_DrugRelease2019.pdf · 2 Current Nanoscience, 2020, Vol. 16, No. 00 Toro-Córdova

Nanomaterials for Drug Release Current Nanoscience, 2020, Vol. 16, No. 00 13

[72] Deshpande, P.; Jhaveri, A.; Pattni, B.; Biswas, S.; Torchilin, V. Transferrin and octaarginine modified dual-functional liposomes with improved cancer cell targeting and enhanced intracellular de-livery for the treatment of ovarian cancer. Drug Deliv., 2018, 25(1), 517-532.

[http://dx.doi.org/10.1080/10717544.2018.1435747] [PMID: 29433357]

[73] Patil, Y.; Shmeeda, H.; Amitay, Y.; Ohana, P.; Kumar, S.; Ga-bizon, A. Targeting of folate-conjugated liposomes with co-entrapped drugs to prostate cancer cells via prostate-specific mem-brane antigen (PSMA). Nanomedicine., 14(4), 1407-1416.

[74] Nguyen, V.D.; Zheng, S.; Han, J.; Le, V.H.; Park, J.O.; Park, S. Nanohybrid magnetic liposome functionalized with hyaluronic acid for enhanced cellular uptake and near-infrared-triggered drug re-lease. Colloids Surf. B Biointerfaces, 2017, 154, 104-114.

[http://dx.doi.org/10.1016/j.colsurfb.2017.03.008] [PMID: 28329728]

[75] Hardiansyah, A.; Yang, M-C.; Liu, T-Y.; Kuo, C-Y.; Huang, L-Y.; Chan, T-Y. Hydrophobic drug-loaded PEGylated magnetic liposomes for drug-controlled release. Nanoscale Res. Lett., 2017, 12(1), 355.

[http://dx.doi.org/10.1186/s11671-017-2119-4] [PMID: 28525950] [76] Hernández Montoto, A.; Montes, R.; Samadi, A.; Gorbe, M.; Ter-

rés, J.M.; Cao-Milán, R.; Aznar, E.; Ibañez, J.; Masot, R.; Marcos, M.D.; Orzáez, M.; Sancenón, F.; Oddershede, L.B.; Martínez-Máñez, R. Gold nanostars coated with mesoporous silica are effec-tive and nontoxic photothermal agents capable of gate keeping and laser-induced drug release. ACS Appl. Mater. Interfaces, 2018, 10(33), 27644-27656.

[http://dx.doi.org/10.1021/acsami.8b08395] [PMID: 30040374] [77] Khutale, G.V.; Casey, A. Synthesis and characterization of a multi-

functional gold-doxorubicin nanoparticle system for pH triggered intracellular anticancer drug release. Eur. J. Pharm. Biopharm., 2017, 119, 372-380.

[http://dx.doi.org/10.1016/j.ejpb.2017.07.009] [PMID: 28736333] [78] Rodrigues, R.O.; Baldi, G.; Doumett, S.; Garcia-Hevia, L.; Gallo,

J.; Bañobre-López, M.; Dražić, G.; Calhelha, R.C.; Ferreira, I.C.F.R.; Lima, R.; Gomes, H.T.; Silva, A.M.T. Multifunctional graphene-based magnetic nanocarriers for combined hyperthermia and dual stimuli-responsive drug delivery. Mater. Sci. Eng. C, 2018, 93, 206-217.

[http://dx.doi.org/10.1016/j.msec.2018.07.060] [PMID: 30274052] [79] Nosrati, H.; Sefidi, N.; Sharafi, A.; Danafar, H.; Kheiri Manjili, H.

Bovine Serum Albumin (BSA) coated iron oxide magnetic nanoparticles as biocompatible carriers for curcumin-anticancer drug. Bioorg. Chem., 2018, 76, 501-509.

[http://dx.doi.org/10.1016/j.bioorg.2017.12.033] [PMID: 29310081]

[80] Maeda, H.; Nakamura, H.; Fang, J. The EPR effect for macromo-lecular drug delivery to solid tumors: Improvement of tumor up-take, lowering of systemic toxicity, and distinct tumor imaging in vivo. Adv. Drug Deliv. Rev., 2013, 65(1), 71-79.

[http://dx.doi.org/10.1016/j.addr.2012.10.002] [PMID: 23088862] [81] Kumari, P.; Ghosh, B.; Biswas, S. Nanocarriers for cancer-targeted

drug delivery. J. Drug Target., 2016, 24(3), 179-191. [http://dx.doi.org/10.3109/1061186X.2015.1051049] [PMID:

26061298] [82] Karthivashan, G.; Ganesan, P.; Park, S.Y.; Kim, J.S.; Choi, D.K.

Therapeutic strategies and nano-drug delivery applications in man-agement of ageing Alzheimer’s disease. Drug Deliv., 2018, 25(1), 307-320.

[http://dx.doi.org/10.1080/10717544.2018.1428243] [PMID: 29350055]

[83] Saraiva, C.; Praça, C.; Ferreira, R.; Santos, T.; Ferreira, L.; Bernardino, L. Nanoparticle-mediated brain drug delivery: Over-coming blood-brain barrier to treat neurodegenerative diseases. J. Control. Release, 2016, 235, 34-47.

[http://dx.doi.org/10.1016/j.jconrel.2016.05.044] [PMID: 27208862]

[84] Dos Santos Rodrigues, B.; Oue, H.; Banerjee, A.; Kanekiyo, T.; Singh, J. Dual functionalized liposome-mediated gene delivery across triple co-culture blood brain barrier model and specific in vivo neuronal transfection. J. Control. Release, 2018, 286, 264-278.

[http://dx.doi.org/10.1016/j.jconrel.2018.07.043] [PMID: 30071253]

[85] Shahin, S.A.; Wang, R.; Simargi, S.I.; Contreras, A.; Parra Echa-varria, L.; Qu, L.; Wen, W.; Dellinger, T.; Unternaehrer, J.; Tamanoi, F.; Zink, J.I.; Glackin, C.A. Hyaluronic acid conjugated nanoparticle delivery of siRNA against TWIST reduces tumor bur-den and enhances sensitivity to cisplatin in ovarian cancer. Nanomedicine (Lond.), 2018, 14(4), 1381-1394.

[http://dx.doi.org/10.1016/j.nano.2018.04.008] [PMID: 29665439] [86] Song, S.; Mao, G.; Du, J.; Zhu, X. Novel RGD containing, temo-

zolomide-loading nanostructured lipid carriers for glioblastoma multiforme chemotherapy. Drug Deliv., 2016, 23(4), 1404-1408.

[http://dx.doi.org/10.3109/10717544.2015.1064186] [PMID: 26203687]

[87] Kalaydina, R.V.; Bajwa, K.; Qorri, B.; Decarlo, A.; Szewczuk, M.R. Recent advances in “smart” delivery systems for extended drug release in cancer therapy. Int. J. Nanomedicine, 2018, 13, 4727-4745.

[http://dx.doi.org/10.2147/IJN.S168053] [PMID: 30154657] [88] Lim, E.K.; Chung, B.H.; Chung, S.J. Recent advances in pH-

sensitive polymeric nanoparticles for smart drug delivery in cancer therapy. Curr. Drug Targets, 2018, 19(4), 300-317.

[http://dx.doi.org/10.2174/1389450117666160602202339] [PMID: 27262486]

[89] Mai, B.T.; Fernandes, S.; Balakrishnan, P.B.; Pellegrino, T. Nano-systems based on magnetic nanoparticles and thermo- or pH-responsive polymers: An update and future perspectives. Acc. Chem. Res., 2018, 51(5), 999-1013.

[http://dx.doi.org/10.1021/acs.accounts.7b00549] [PMID: 29733199]

[90] Allen, T.M.; Cullis, P.R. Liposomal drug delivery systems: from concept to clinical applications. Adv. Drug Deliv. Rev., 2013, 65(1), 36-48.

[http://dx.doi.org/10.1016/j.addr.2012.09.037] [PMID: 23036225] [91] Toro-Cordova, A.; Flores-Cruz, M.; Santoyo-Salazar, J.; Carrillo-

Nava, E.; Jurado, R.; Figueroa-Rodriguez, P.A.; Lopez-Sanchez, P.; Medina, L.A.; Garcia-Lopez, P. Liposomes loaded with cis-platin and magnetic nanoparticles: Physicochemical characteriza-tion, pharmacokinetics, and in-vitro efficacy. Molecules, 2018, 23(9)E2272

[http://dx.doi.org/10.3390/molecules23092272] [PMID: 30200551] [92] Di Corato, R.; Béalle, G.; Kolosnjaj-Tabi, J.; Espinosa, A.;

Clément, O.; Silva, A.K.; Ménager, C.; Wilhelm, C. Combining magnetic hyperthermia and photodynamic therapy for tumor abla-tion with photoresponsive magnetic liposomes. ACS Nano, 2015, 9(3), 2904-2916.

[http://dx.doi.org/10.1021/nn506949t] [PMID: 25695371] [93] Yang, R.; Tang, Q.; Miao, F.; An, Y.; Li, M.; Han, Y.; Wang, X.;

Wang, J.; Liu, P.; Chen, R. Inhibition of heat-shock protein 90 sen-sitizes liver cancer stem-like cells to magnetic hyperthermia and enhances anti-tumor effect on hepatocellular carcinoma-burdened nude mice. Int. J. Nanomedicine, 2015, 10, 7345-7358.

[http://dx.doi.org/10.2147/IJN.S93758] [PMID: 26677324] [94] Gao, M.; Meng, X.; Guo, X.; Zhu, J.; Fan, A.; Wang, Z.; Zhao, Y.

All-active antitumor micelles via triggered lipid peroxidation. J. Control. Release, 2018, 286, 381-393.

[http://dx.doi.org/10.1016/j.jconrel.2018.08.003] [PMID: 30098375]

[95] Shi, J.; Liu, W.; Fu, Y.; Yin, N.; Zhang, H.; Chang, J.; Zhang, Z. “US-detonated nano bombs” facilitate targeting treatment of resistant breast cancer. J. Control. Release, 2018, 274, 9-23.

[http://dx.doi.org/10.1016/j.jconrel.2018.01.030] [PMID: 29408184] [96] Liu, M.; Du, H.; Khan, A.R.; Ji, J.; Yu, A.; Zhai, G. Redox/enzyme

sensitive chondroitin sulfate-based self-assembled nanoparticles load-ing docetaxel for the inhibition of metastasis and growth of mela-noma. Carbohydr. Polym., 2018, 184, 82-93.

[http://dx.doi.org/10.1016/j.carbpol.2017.12.047] [PMID: 29352946] [97] Marchal, S.; El Hor, A.; Millard, M.; Gillon, V.; Bezdetnaya, L.

Anticancer drug delivery: An update on clinically applied nanothera-peutics. Drugs, 2015, 75(14), 1601-1611.

[http://dx.doi.org/10.1007/s40265-015-0453-3] [PMID: 26323338] [98] Park, K. Controlled drug delivery systems: past forward and future

back. J. Control. Release, 2014, 190, 3-8. [http://dx.doi.org/10.1016/j.jconrel.2014.03.054] [PMID: 24794901] [99] Ulbrich, K.; Holá, K.; Šubr, V.; Bakandritsos, A.; Tuček, J.; Zbořil,

R. Targeted drug delivery with polymers and magnetic nanoparticles: Covalent and noncovalent approaches, release control, and clinical studies. Chem. Rev., 2016, 116(9), 5338-5431.

Page 14: A Concise Review of Nanomaterials for Drug Delivery and ...unizar.es/gfgoya/index_archivos/PDF/CN_DrugRelease2019.pdf · 2 Current Nanoscience, 2020, Vol. 16, No. 00 Toro-Córdova

14 Current Nanoscience, 2020, Vol. 16, No. 00 Toro-Córdova et al.

[http://dx.doi.org/10.1021/acs.chemrev.5b00589] [PMID: 27109701] [100] Kralj, S.; Potrc, T.; Kocbek, P.; Marchesan, S.; Makovec, D. De-

sign and fabrication of magnetically responsive nanocarriers for drug delivery. Curr. Med. Chem., 2017, 24(5), 454-469.

[http://dx.doi.org/10.2174/0929867323666160813211736] [PMID: 27528059].