Skip to main content

REVIEW article

Front. Chem., 25 June 2018
Sec. Analytical Chemistry
Volume 6 - 2018 | https://doi.org/10.3389/fchem.2018.00237

Application of Light Scattering Techniques to Nanoparticle Characterization and Development

  • Instituto de Medicina Molecular, Faculdade de Medicina, Universidade de Lisboa, Lisbon, Portugal

Over the years, the scientific importance of nanoparticles for biomedical applications has increased. The high stability and biocompatibility, together with the low toxicity of the nanoparticles developed lead to their use as targeted drug delivery systems, bioimaging systems, and biosensors. The wide range of nanoparticles size, from 10 nm to 1 μm, as well as their optical properties, allow them to be studied using microscopy and spectroscopy techniques. In order to be effectively used, the physicochemical properties of nanoparticle formulations need to be taken into account, namely, particle size, surface charge distribution, surface derivatization and/or loading capacity, and related interactions. These properties need to be optimized considering the final nanoparticle intended biodistribution and target. In this review, we cover light scattering based techniques, namely dynamic light scattering and zeta-potential, used for the physicochemical characterization of nanoparticles. Dynamic light scattering is used to measure nanoparticles size, but also to evaluate their stability over time in suspension, at different pH and temperature conditions. Zeta-potential is used to characterize nanoparticles surface charge, obtaining information about their stability and surface interaction with other molecules. In this review, we focus on nanoparticle characterization and application in infection, cancer and cardiovascular diseases.

Introduction

Nanotechnology research and development have increased over the last three decades. The concern about the bioavailability and efficacy of conventional therapeutics by their suboptimal results on targeted cells and high toxicity in normal cells have lead the scientific community to reshape the vision of drug development (Geszke-Moritz and Moritz, 2016). Nanoparticles (NPs) have been developed to overcome the problems of targeting and efficiency, with reduced toxicity. In the last decade, their applicability has been focused on the biomedical and pharmaceutical fields, used as drug delivery systems, diagnostic tools, and implants (Zhang, 2015; Geszke-Moritz and Moritz, 2016; Alegret et al., 2017; Jurj et al., 2017; Ramos et al., 2017; Wong et al., 2017). Nanoparticles can be made of different materials, organic or inorganic, such as metal, polymers, carbon nanotubes, and liposomes (Liu et al., 2016). The use of nanoparticle-based drug delivery systems has increased due to their controlled release of reservoir content, leading to a decrease in undesirable side effects (Cosco et al., 2011; Mahmoodi et al., 2016; Jurj et al., 2017; Panahi et al., 2017; Singh et al., 2017). At the same time, the use of nanoparticles in drug development reduces the usage of additional components on the formulation to protect therapeutics from degradation and increase circulation time.

Nanoparticle formulation requires full characterization of its size, surface charge, shape, and distribution (Oberdörster, 2010). It is often technically challenging to obtain reproducible suspensions of nanoparticles with low polydispersion and desired shape and size. The tight control of mixing and separation of particles is crucial to obtain a homogeneous nanoparticle suspension (Cosco et al., 2015b). Usually, only a small fraction of the nanoparticles injection dose (<0.7%) reaches the target (Schmidt and Storsberg, 2015). This shows that NPs have some organism barriers to overcome, such as unspecific distribution, interstitial fluid pressure, cellular internalization, and drug efflux pumps, before achieving therapeutic effect (Park and Na, 2015).

Nanoparticles have size-related properties influencing their mode of action and in vivo lifetime. The optimal size for drug delivery systems is considered to be broadly between 10 and 1000 nm (Ramos et al., 2017). Low sizes allow NPs to cross cell membranes and avoid detection by the reticuloendothelial system (RES), increasing the drug circulation lifetime (Schmidt and Storsberg, 2015; Hare et al., 2017; Jahan et al., 2017). However, they must not be too small, in order to avoid rapid distribution into lymph nodes, being eliminated by fast renal clearance. On the other hand, nanoparticles larger than 100 nm are more prone to accumulate at the site of injection or trapped by the spleen, lung, and liver macrophages (Jurj et al., 2017). In conclusion, size must be optimized taking into account the amount of cargo to be delivered and the desirable biodistribution (Figure 1).

FIGURE 1
www.frontiersin.org

Figure 1. Comparison of capillaries and different nanoparticles size described in literature for different therapeutic applications. Nanoparticles were designed in terms of size and material considering the therapeutic target desired, with the size being determined by dynamic light scattering.

In terms of surface charge, neutrality may lead to nanoparticle instability, with aggregation and precipitation after long-term storage. The surface charge characterization is an important parameter to measure in a NPs suspension, because the first interaction is with the body fluids before reaching a target. In physiological media, the nanoparticle is covered by plasma proteins leading to surface charge alterations and, concomitantly, changing its biological activity and affinities (Ramos et al., 2017). Positive surface charges may facilitate the binding of nanoparticles to cell membranes and might promote unspecific binding to normal tissues, promoting platelet accumulation and hemolytic events (Licciardi et al., 2016; Jahan et al., 2017; Jurj et al., 2017; Jiang et al., 2018; Peretz et al., 2018).

The unique physicochemical properties and nanoscale effects have drawn interest on nanoparticle as drug delivery systems for the treatment of diseases such as cancer, cardiovascular diseases, pathogenic infections, and diabetes. Despite the raised interest in nanoparticle development, not so many have been approved for therapeutic use (Wang et al., 2017). Here, we will focus on the light scattering approaches to characterize nanoparticle suspensions and their applicability on nanoparticle development against infectious and cardiovascular diseases.

Light Scattering Techniques

Dynamic Light Scattering

The detection of the light scattered from the interaction of light with matter gives information related to the physical characteristics of the sample. Typically, in light scattering experiments, a monochromatic beam is directed to the sample and then a detector records the scattered light at a certain angle. Early light scattering experiments started in late nineteenth century, with John Tyndall's research in colloidal suspensions (Tyndall, 1868). Lord Rayleigh (John William Strutt) reported another important effect of the light scattering by particles smaller than its wavelength, by explaining the blue color of the sky and the effect of the atmospheric particles (Strutt, 1871). For larger particles relative to the wavelength of light, Gustav Mie developed a theory to study the light scattering from absorbing and non-absorbing particles, considering particle shape and the difference in refractive index between particles and the medium where they are dispersed (Mie, 1908). Taking into account the differences of the light scattering at different angles of detection from large particles (Mie theory) with the more homogeneous light scattering at each angle for small particles (Rayleigh theory), we hereby use the Rayleigh particle for theoretical purposes.

In static light scattering, the intensity of the light detected is averaged over time, and from this we can obtain information about the molecular weight of the particle and its radius of gyration (Rg). On the other hand, dynamic light scattering (DLS), by measuring over time the fluctuations of the light intensity, due to particle Brownian motion, allows to determine the diffusion coefficient (D), which relates to the hydrodynamic radius (Rh) of the particle through the Stokes-Einstein equation (Pusey, 1974),

D= kbT6πηRh    (1)

where κb is the Boltzmann constant (1.380 × 10−23 kg.m2.s−2.K−1), T is the absolute temperature, and η is the viscosity of the medium.

As it shows up in Equation (1), the particle diffusion depends on the temperature, viscosity of the media and size of the particle. DLS measures the intensity of the light scattered over time. When the intensity is correlated at several time points, in the beginning the scattered intensities are similar, losing this similarity over time due to particle's movement. Then, for small particles, the diffusion is much faster, photon correlation is lost faster and the correlation decays at early time points of the measurement (Figures 2A,B). However, as large particles diffuse more slowly, the similarity of the intensities over time persists for longer periods, leading to a longer time for the photon correlation to decay (Figures 2C,D). A digital correlation measures the intensity fluctuation and their correlation in respect to time frames (on the ns and μs timescale). The measured parameter is a normalized integration of the intensities at the beginning and a delayed time τ (Chu, 1974),

g2(τ)= I(t) . I(t+τ)I(t)2    (2)
FIGURE 2
www.frontiersin.org

Figure 2. Dynamic light scattering intensity signal and correlation function for small (A,B) and large particles (C,D). The scattering intensity signal over time is obtained directly from the particle's Brownian motion. The correlation function is obtained from the intensity fluctuation in the respective time frames. Small particles (A,B) diffuse faster, with the correlation decaying at early time points. Large particles (C,D) diffuse more slowly, which implies a longer time for the photon correlation to decay.

However, the measurement of each particle position in the scattered volume is not possible under the experimental apparatus. For this reason, there is a measurement of the normalized electrical field generated by the volume of the particles under an incident beam (Berne and Pecora, 1976),

g1(τ)= E(t) . E(t+τ)E(t)2    (3)

The normalized intensity integration is correlated with the normalized electrical field measured by the Siegert relation (Siegert, 1943),

g2(τ)= B+ β|g1(τ)|2    (4)

where, B is the baseline (~1) and β is the coherence factor, which depends on detector area, optical alignment and scattering properties of macromolecules or supramolecular aggregates. Considering a monodisperse sample, the normalized intensity integration decays exponentially and is dependent on a decay constant, Γ, for macromolecules undergoing a Brownian motion (Einstein, 1905, 1906),

g2(τ)= 1+ βe-2Γτ    (5)

where Γ is related to diffusion coefficient of the sample particles, D, by (Berne and Pecora, 1976),

Γ= Dq2    (6)

where q is the scattering vector, directly proportional to the refractive index, n0, and inversely proportional to the wavelength, λ (Harding, 1997),

q=4πn0λsin(θ/2)    (7)

where θ is the angle of the detector's position. However, when considering a polydisperse sample, the normalized intensity integration cannot be described by a single exponential decay (Briggs and Nicoli, 1980). Instead, there is a sum of exponential decays rates G(Γ) corresponding to each particle in the sample (Berne and Pecora, 1976),

g2(τ)= 1+ β(0G(Γ)e-ΓτdΓ)2    (8)

Data can be analyzed from the fitting of the correlation function. However, it is possible to distinguish two types of methods of fitting: assuming a monomodal distribution or a non-monomodal distribution. The common monomodal approach is the cumulants fitting, where a Taylor expansion with a mean decay rate is fitted to the correlation function, obtaining a mean diffusion coefficient (Koppel, 1972). From the relation of the second cumulant to the mean decay rate, it is possible to obtain the polydispersity index (PDI), informing about the monodispersity tendency of the sample. Regarding non-monomodal distribution methods, the fitting of the correlation function is based on multiple decay rates, which is more suitable for polydisperse samples. The common methodologies are non-negative least squares (NNLS), where the decay rates are constants from the list of G(Γ) in a determined range, but spaced linearly or logarithmically (Morrison et al., 1985). The exponential sampling uses the decay rates in a determined range but spaced exponentially. The most common methodology applied to non-monomodal distribution is the constrained regularization method for inverting data (CONTIN) (Provencher, 1982a,b). The CONTIN method is similar to NNLS, but instead of the minimization of residuals in the NNLS methodology, it works by the minimization of regularized residuals and an appropriate weighing function. For more details on the mathematical approach used in the methods, please refer to Fischer and Schmidt (2016) and Stetefeld et al. (2016).

Zeta-Potential

The zeta-potential is the potential measured at the slipping plane of a particle under an electrical field. It reflects the potential difference between the electric double layer (EDL) of electrophoretic mobile particles and the layer of dispersant around them (aqueous or organic environment) at the slipping plane (Figure 3) (Montes Ruiz-Cabello et al., 2014). The EDL surface of a particle in solution develops instantaneously and is formed of two layers. The inner layer, the so-called Stern layer, is composed of opposite charged particles tightly coupled to the core of the central particle. The second and outermost layer is a diffusive layer consisting of both opposite and same charged ions/molecules. When an electrical field is applied to the sample, the particles move to the opposite electrode. Within the diffuse layer there is a hypothetical plane that acts as the interface between the moving particles and the layer of the surrounding dispersant while in the electrical field. This plane is the characteristic slipping/shear plane and zeta-potential is the potential at this particle-fluid interface (Kaszuba et al., 2010; Bhattacharjee, 2016). The zeta-potential is measured by the electrophoretic mobility of charged particles under an applied electric field. The electrophoretic mobility (μe) of the particles is calculated by Henry's equation (Kaszuba et al., 2010),

μe=2εrε0ζf(Ka)3η    (9)

where εr is the relative permittivity/dielectric constant, ε0 is the permittivity of vacuum, ζ is the zeta-potential value, f (Ka) is the Henry's or Helmholtz-Smoluchowski function, and η is the viscosity at the experimental temperature. Depending on the solvent where the particles are dispersed, the value of f (Ka) is assumed to be 1 or 1.5, for organic medium or aqueous medium, respectively (Domingues et al., 2008).

FIGURE 3
www.frontiersin.org

Figure 3. Schematic representation of the double layer that surrounds the nanoparticle in aqueous medium, considering that it has negative charge. The nanoparticle represented as example is composed by negatively charged phospholipids, implying a first layer (Stern-potential) mainly composed by positively charged counterions after application of an electric field. The second layer (zeta-potential) is a diffusive layer that consists of both counterions and ions of the same charge as the nanoparticle, which contact the organic or aqueous environment.

Nanoparticles in Therapeutics

Nanoparticles are widely used in biomedical sciences for different therapies, due to their high biocompatibility and chemical stability, either by direct activity or by encapsulating poorly soluble drugs/surface incorporation (Arakha et al., 2015; Elzoghby et al., 2015). Among the most notorious examples are magnetic nanoparticles, with a metal core of Zn, Ni, Cu, Ag, or Au, synthetically obtained or naturally isolated (Bilal et al., 2017; El-Batal et al., 2018). Some of these were shown to have antimicrobial activity, and were considered perfect candidates for magnetic resonance imaging techniques, presenting a dual activity: therapeutic and diagnostic (Niemirowicz et al., 2015; Dinali et al., 2017). Their use in bandages, implants or prostheses is already becoming common, but overproduction of reactive oxygen species (ROS) in long-term usage has raised concerns regarding the toxicity of magnetic NPs (Bilberg et al., 2012; Casciaro et al., 2017). Different authors have explored this issue, even in polymeric-coated magnetic NPs, which were consider less toxic than the uncoated, but high dosages during a larger period of time increase cytotoxic and genotoxic effects on macrophages (Jena et al., 2012; Mohanty et al., 2012). With nanoparticles activity being dependent of their physicochemical properties, namely size, shape, and surface, their toxicity toward cells is also dependent of these properties (Bera et al., 2014; Sun et al., 2014; Rajchakit and Sarojini, 2017). A strategy followed to deal with these problems has been the development of different types of nanoparticles, including polymeric nanoparticles, micelles, or liposomes, with the advantage of being possible to shape their properties to increase the efficacy in targeting or drug delivering (Xie et al., 2014; Bilal et al., 2017; Solairaj et al., 2017). With the objective of reducing toxicity without reducing NP activity, another adopted strategy was the incorporation or surface derivatization with different ligands, such as antibodies, small organic molecules, or proteins/peptides (Chen et al., 2012; Gao et al., 2017). This last hypothesis was shown to reduce toxicity, improving peptide properties/activity, and enhancing solubility, leading to a general improvement of the pharmacokinetic profile and therapeutic index (Molinaro et al., 2013; Gao et al., 2014, 2017; Cosco et al., 2015a; Libralato et al., 2017). As a matter of fact, different proteins have already been tested for different activities, including albumin, casein or elastin-like polypeptides, exploring either an active targeting (direct activity on target cells) or a passive targeting (prolonged blood circulation and activity) (Sneharani et al., 2010; Zhao et al., 2010; Bachar et al., 2012; Kratz, 2014; MacEwan and Chilkoti, 2014). With different possibilities of surface modification, these protein nanoparticles rapidly evolved to peptide nanoparticles, due to an easier manufacture process and reduced production costs (Elzoghby et al., 2015).

Peptide therapeutics is a field that is fast growing since the beginning of the century, with a large number of scientific papers exploring their potential use in healthcare (Albericio and Kruger, 2012). Moreover, with resistance increase in different pathologies, including infectious diseases and cancer, the urgency for new alternatives has promoted a significant number of studies aiming at improving the efficacy of peptides as drugs or applied in diagnostic techniques (Hamilton et al., 2015; Gomes et al., 2018). Several in vitro and in vivo studies have been published focusing on the efficiency of particular peptide classes, namely antimicrobial and anticancer peptides (AMPs and ACPs, respectively), due to their promising applications as drugs in the market (Hancock et al., 2016; Felício et al., 2017). Even so, downsides of their use have been pointed out, including low enzymatic stability, low permeability across biological barriers, low solubility, rapid metabolic excretion, and high toxicity (Tam et al., 2002; Rajchakit and Sarojini, 2017; Serna et al., 2017). Strategies to overcome these peptide therapeutic applicability problems include in silico structure design (considering their sequence, using natural and non-natural amino acids), peptidomimetics, lipidation and, naturally, nanoparticle conjugation, which will be further explored (Rajchakit and Sarojini, 2017; Primavera et al., 2018).

Nanoparticles With Antimicrobial Activity

As mentioned above, an increase in multiresistant pathogens (bacteria, fungi, and viruses) has been reported on the last decades, with several reasons already explored being held responsible for this (Dickey et al., 2017; Llewelyn et al., 2017). The World Health Organization has inclusively pointed out different bacteria strains where researchers should focus on, due to the high incidence of resistance in patients (World Health Organization, 2015). AMPs are considered one of the major promises to overcome this growing public healthcare problem. Due to this, studies on their isolation, purification, design, and applicability, both in vitro and in vivo, have increased in recent years (Dias et al., 2017; Unubol et al., 2017). These peptides are usually characterized by a short amino acid sequence (less than 50 amino acid residues), high amphipathic and hydrophobic content, and a positive net charge (de la Fuente-Núñez et al., 2017; Haney et al., 2017). Their mechanisms of action, frequently at the membrane level, are not well-defined, but it is clear that their physicochemical properties are essential for the peptide-membrane interaction (Neelay et al., 2017). Initially, they were thought to target specifically different pathogens but, nowadays, it is clear that their action is more complex than that, participating in the recruitment of immune cells to the site of infection, or modulating the immune response by promoting pathogen cell death (Hancock et al., 2016). Also, they have a broad-spectrum activity (Vigant et al., 2015), being active toward bacteria (including biofilms), fungi or viruses, with properties of the target membrane driving the interactions (Ribeiro et al., 2016). Even so, their limitations became also notorious in a large number of studies, limiting their potential as therapeutic molecules (Gomes et al., 2018).

At the same time, nanotechnology (particularly using nanoparticles) has also focused its research in these applications, having reached a higher success in clinical applications. With the advantage of the possible use of different metals, magnetic nanoparticles with intrinsic antimicrobial activity were developed and medically applied (Bilal et al., 2017; Dinali et al., 2017; Pham et al., 2018). The fact that these NPs have in their core a metal predisposing to electrostatic interactions, promotes their attachment to bacterial membranes, leading to the loss of integrity and bacteria cell death (Fang et al., 2015; Bilal et al., 2017). A high number of systems have been tested with reported activity toward pathogens, using different antibiotic molecules conjugated either on the surface or by encapsulation (Park et al., 2011; Gaspar et al., 2016; Morales et al., 2017). An important example is silver nanoparticles (AgNPs) conjugated with polymixin B or gold nanoparticles (AuNPs) conjugated with vancomycin, both showing a synergistic effect, with improved activity (Fayaz et al., 2011; Park et al., 2011). Metal nanoparticles chosen for antibiotic conjugation include titanium, zinc or cooper, and as for antibiotic molecules, gentamicin, streptomycin, cecropin-melittin, among others, have shown improved activity (Gu et al., 2003; Birla et al., 2009; Allahverdiyev et al., 2011; Lai et al., 2015). However, as already stated, the use of metals for nanoparticle development raised some doubts due to inherent toxicity toward healthy cells, forcing researchers to find alternatives. An example was testing NPs for local/topic applications, lowering the dosage amount and toxicity effects (Gao et al., 2014; Arakha et al., 2015; de Oliveira et al., 2017). In a pH-sensitive system, Pichavant et al. developed antibiotic (gentamicin sulfate and/or vancomycin) functionalized nanoparticles that were covalently grafted into titanium surfaces (Pichavant et al., 2016). The nanoparticle characterization was achieved using nuclear magnetic resonance (NMR) and dynamic light scattering measurements, confirming their size and stability in different media (Pichavant et al., 2011, 2012). Besides the enhanced antimicrobial activity, these NPs also presented other advantages, such as the possibility to be used in other surfaces and promoting an increase in the target tissue/cell drug density (Pichavant et al., 2011, 2012). Another study, by Di Francesco et al. showed the advantages of using pH-sensitive nanoparticles as a fusogenic drug delivery system (Di Francesco et al., 2017). The nanoparticles were formulated according to their target cells, and physicochemical properties were measured by DLS and fluorescence spectroscopy.

The application of nanolipid systems (like liposomes or micelles) or polymeric NPs (chitosan based or conjugated with polyethylene glycol, PEG) had special success (Hann and Prentice, 2001; Allen and Cullis, 2013; Cosco et al., 2014; Paolino et al., 2017). These NPs have the advantage of being more biocompatible, with the effects toward healthy cells being reduced and having an improved targeted-oriented activity (Solairaj et al., 2017). Water et al. developed poly(lactic-co-glycolic acid) nanoparticles (PLGaNPs) that were used as drug delivery system for plectasin, an antibiotic specific for airway Staphylococcus aureus infection (Water et al., 2015). They used DLS and zeta-potential measurements to assure that plectasin was efficiently loaded on the NP. Other authors used chitosan-sodium phytate NPs and tested them against Gram-negative and Gram-positive bacteria, showing a high antimicrobial activity, with the advantage of these NPs could also be used for drug delivery, combining their efficacy with an antibiotic (Yang et al., 2017). In order to identify the optimal chitosan/sodium phytate ratio for their activity, DLS and zeta-potential measurements were performed to determine the NPs size, surface charge, and stability at different pH values. As for liposomes, other authors have developed lipid NPs composed of phosphatidylcholine (zwitterionic phospholipid) and phosphatidylserine (negatively charged phospholipid), intercalated with Pluronic-P85 (HLB 16), a polymer that favors the uptake of the NP (Fidler, 1988; Zhang et al., 1998). By incorporating gentamycin in their core, this system was tested for drug delivery, with a high efficiency rate (Xie et al., 2014).

Despite all the strategies studied along the years, one has gained special attention nowadays, when conventional therapeutic molecules are facing a new resistance paradigm. This strategy consists in the combination of nanoparticles (liposomes, polymeric, or metallic) with antimicrobial peptides, either for peptide delivery or for a direct action toward the target cells (Niemirowicz et al., 2015; Water et al., 2015). The objective was to overcome the limitations on AMPs application, but, later on, it was stated that the nature of nanoparticle-AMP interaction is essential for the system activity (Pal et al., 2016). Actually, weaker interactions between the AMP and the nanoparticle promote a decrease in NP toxicity and, at the same time, increase AMP activity, because it allows the peptide to adopt favorable structure and/or charge properties essential for the interaction with biomembranes (Liu et al., 2013; Rajchakit and Sarojini, 2017). These AMP-NP complexes also allow a higher concentration of the drug in the site of action, with a selective activity, including a differential interaction between the complex and the outer and inner-membranes of the target bacterial pathogens, implying a drug-delivery and direct activity system (Park et al., 2017; Rajchakit and Sarojini, 2017).

Different AMP-NP complexes have been tested throughout the years, trying to establish one with high activity toward the target pathogens, without having a significant toxicity for the other cells, a common flaw for AMPs and metal NPs (Galdiero and Gomes, 2017). Different metals were tested, as already described, such as iron oxide, coupled with LL-37, a natural host-defense peptide with antimicrobial activity (Niemirowicz et al., 2015). Other examples include silver nanoparticles surrounded by AMPs, or gold NPs with bactenecin molecules on their surface (Allahverdiyev et al., 2011; Golubeva et al., 2011). All these systems were shown to have less toxicity and higher efficiency, including against clinical isolated multiresistant pathogens, but their pharmacokinetic and pharmacodynamic profiles still need to be improved (Ruden et al., 2009). Considering this scenario, methods to improve these properties were designed, including the use of natural isolated nanoparticles from biomass (Mohanty et al., 2013). Their biogenic AgNPs combined with two different AMPs (NK-2 and LLKK-18), characterized by DLS and zeta-potential, were shown to have synergistic effect and improved applicability in clinical scenarios (Mohanty et al., 2013). In another study, polymeric nanoparticles (chitosan-alginate polyelectrolyte complex NPs) combined with pexiganan (a synthetic AMP) had an improved profile for therapeutic application (Zhang et al., 2015). Another strategy tested was the use of PEG: nanoparticle surface was covered with PEG and AMPs, increasing biocompatibility and antimicrobial properties (Pal et al., 2016; Casciaro et al., 2017). Zeta-potential was used to confirm that the peptide was attached to the NP surface after coupling synthesis, with an overall charge increase after interaction with positive peptides such as AMPs.

Besides antimicrobial peptides, nanoparticles can also be combined with cell-penetrating peptides (CPPs) (Guidotti et al., 2017). There is not a rigid boundary between these two classes of peptides, with reported AMPs having a CPP function, as well as CPPs with described antimicrobial activity, besides the capacity to deliver cargo into different cells (Bahnsen et al., 2015; Kristensen et al., 2016). One example is the combination of micelles with TAT, a HIV-derived CPP with antimicrobial activity, conjugated with cholesterol, a spacer and six arginine residues (Liu et al., 2009). These self-assembly micelles, besides enhanced activity and low toxicity, were able to cross the blood brain barrier, which introduced a great advantage for brain infection diseases (Liu et al., 2009). On another study from the same authors, they used the same CPP, with a spacer of three glycine and six arginine residues, but conjugated to colloid AgNPs surface (Liu et al., 2013). Improved antimicrobial activity and reduced hemolysis were observed. In both cases, DLS and zeta-potential were essential to characterize the NPs, regarding size and surface charge, but also to assess colloidal stability (Liu et al., 2009, 2013).

It is important to refer that these complexes of AMPs/CPPs-NPs have high potential for the treatment of bacterial infection, including those leading to biofilm formation (Ribeiro et al., 2016). Biofilms are complex pathogen aggregates, encased in a matrix composed of extracellular polymeric substances (EPS), that normally tend to form when bacteria faces stress adaptation (Flemming et al., 2016). Due to this matrix, AMPs efficient against planktonic (free) bacteria can be ineffective against biofilms (Batoni et al., 2016). Nanoparticles by themselves have small size, with an enormous surface area and easy penetrability properties, including on biofilms. These properties and association with AMP introduce advantages to tackle biofilm infection, and should be considered in future works (Qayyum and Khan, 2016).

The use of peptides in nanotechnology has been largely increasing, as described above. At this level, other structures with promising results are self-assembling peptide NPs, which are formed by small peptides that self-aggregate, forming clusters, or oligomers (Serna et al., 2017). The idea came from dendrimeric peptides, small nanosystems with a size range from 2 to 50 nm, with great advantages in terms of biocompatibility, structural/functional versatility and drug delivery efficiency (Tam, 1988; Tam et al., 2002; Serna et al., 2017). These systems are characterized by a hyper-branched and almost perfect geometrical 3D architecture, that grow from the core into a globular shape, with reported activity against infectious pathogens and cancer cells (Tam et al., 2002; Ionov et al., 2013; García-Gallego et al., 2017). Examples of systems already studied are diverse, with each author exploring different mechanisms to promote the assembling or activity toward the target cells. They include AMPs conjugated to the N-terminal of histidine-tagged proteins, forming oligomers with antimicrobial activity (Serna et al., 2017). As the synthesis of self-assembly NPs starts with small aggregates, DLS was used to determine the evolution of the size distribution, confirming the oligomers formation (Serna et al., 2017). Lipidation of AMPs, besides the increased activity already explored, can also promote the formation of dendrimeric peptide NPs (Siriwardena et al., 2018). Using parental systems, these authors developed a new one, with higher antimicrobial activity and pro-angiogenic properties in biological burn-wound bandages, named TNS18 (Siriwardena et al., 2018). Finally, other authors recently focused in self-assembling peptide nanoparticles that only act on the target cell after activation, using for that specific characteristics of the target tissue, such as overexpressed membrane proteins or enriched proteases concentration (Yu et al., 2018; Zhang et al., 2018). This field is now expanding and, therefore, more research is needed to understand how this strategy can benefit current therapies relative to other systems that are easier to manipulate.

Nanoparticles With Anticancer Activity

Therapies to deal with cancer have evolved in response to the human need, but resistance to therapy is a public health concern (Arnold et al., 2015). Nanotechnology has for long tried to fight this burden, by improving the pharmacokinetic and pharmacodynamics of the chemotherapeutic agents that target solid tumors. For that, drug encapsulation was studied and tested in vivo, with the first molecules being FDA approved in the middle of the 1990s, namely Doxil and DaunoXome (Eertwegh et al., 2006). Both therapeutics consist of liposomes with encapsulated drugs, doxorubicin (DOX) and a mixture of anthracycline and daunorubicin, respectively (Eertwegh et al., 2006; Allen and Cullis, 2013). Cancer drugs face diverse challenges, creating the need of developing new drugs according to the type of target: solid tumors or circulating cancer cells (Pearce et al., 2012). For solid tumors, evading the mononuclear phagocyte system (MPS) and remaining in the tumor tissue is essential for drug efficacy, while for circulating cancer cells there is the need for the drug to be internalized to ensure its action at the target site (Stylianopoulos and Jain, 2015).

Considering the current scenario, different strategies have been followed trying to overcome these limitations. Metal nanoparticles with gold or silver core have been tested and showed to have natural anticancer activity, either in vitro or in vivo against tumors and cancer cells (Shanmugasundaram et al., 2017; Shmarakov et al., 2017). Following the improvements on the development of nanoparticles, combinations of copper and chitosan were also tested, with observable anticancer activity and less toxic effects (Solairaj et al., 2017). DLS was used here not only as a mere characterization technique, but as a tool to identify metal structures with higher colloidal stability and better size distribution (Shmarakov et al., 2017). Di Francesco et al., using non-ionic surfactant vesicles (NSVs) loaded with DOX, developed nanosystems with different ratios of Tween21/Tween80, promoting a pH-responsive approach with anticancer properties (Di Francesco et al., 2017). These NSVs showed a fusogenic behavior and an increased targeting efficiency, which translated in higher anticancer activity. Nanoparticles with direct activity can also be used as drug carriers, as mentioned before. Zakerzadeh et al. designed silica NPs with encapsulated tetrazole, a cyclic/aromatic molecule with antimicrobial, antifungal and anticancer activity (Zakerzadeh et al., 2017).

Despite previous advances, improvements in the targeting were still necessary. As in infection therapies, also here the use of peptides was considered, either to increase activity or to promote specific targeting to cancer cells and solid tumors (Pearce et al., 2012). As an example, Chang et al. designed NPs that were able to bind to the tumor mass (oral, breast, lung, colon, or pancreatic tumors) by coating them with the small antimicrobial peptides PIVO-8 and PIVO-24 (also acting at the vascularization process), which are significantly increased around tumors (Chang et al., 2009). To confirm NP coating with both peptides, authors used DLS, evaluating afterwards the differences in activity (Lee et al., 2004). Also targeting tumors, iron oxide NPs coated with an heptapeptide that recognize fibrin-fibronectin complexes or chitosan NPs with antiangiogenic peptide endostatin (ES) improved anticancer activity by targeting the vascularization of the tumor (Agemy et al., 2010; Xie et al., 2017). Coating nanoparticle surfaces with two or more different peptides was also reported (Colombo et al., 2002; Marchiò et al., 2004). Even so, ideally, anticancer therapies would be able to eliminate tumors and malignant cancer cells, including those that are no longer associated with the main tumor, without toxicity toward healthy cells. NPs that act as drug carriers (for drugs like doxorubicin, 5-fluororacil or cisplatin), with good pharmacokinetic and pharmacodynamic profiles (using PEG on their surface or polymer NPs), specific (by using small peptides) and with enhanced cellular uptake would be the desired candidates (Safra, 2003; Paolino et al., 2013; Ribeiro et al., 2016; Gomes et al., 2018). For this system, the missing point is the enhanced uptake, which was possible with the attachment of CPPs to the nanoparticle surface, besides the AMPs necessary for their activity. Authors tested the use of TAT, the HIV-1 derived CPP, by coupling it to NPs with PEG on their surface, and demonstrated the improved cellular uptake (Kuai et al., 2010, 2011). Besides the CPP, these authors also tested the possible applicability of different PEG molecules, due to their concern for increasing NPs distribution near the tumor, but loss of internalization ability (Kuai et al., 2010). Using DLS and zeta-potential measurements, Kuai et al. studied the optimal proportion of cleavable PEG to maintain their accessibility and activity. Takara et al. also tested the incorporation of a CPP (STR-R8) on the nanoparticle surface, that was already coated with NGR motif peptides (recognizes CD13 presence in endothelial tumor cells) and PEG, showing that a synergic effect between all the molecules incorporated occurred (Takara et al., 2010). For that, DLS and zeta-potential were used to evaluate the best CPP amino acid residue to use for the anchoring, considering that size should be stable, and that surface charge is essential for NP targeting membrane interaction. Recently, Xia et al. further increased the complexity with a high efficiency construct: using selenium NPs, which have advantages in terms of dosage, biocompatibility, toxicity, and drug delivery, they coated them with an anticancer peptide (RGDFC heptapeptide) and incorporated DOX and siRNA (anti-Nanog, a human homeobox protein that is essential for cancer cell proliferation) (Xia et al., 2018). This SeNPs@DOX/siRNA system showed to be very effective on the targeting and treatment of cancer, presenting a new hypothesis as synergistic system. Nevertheless, regarding cancer therapies, there is a lot to improve in terms of targeting and efficiency of cancer eradication in vivo, because most of the systems tested in vitro have been failing on clinical trials (Pearce et al., 2012).

Nanoparticles in Cardiovascular Diseases

Although areas like cancer and antimicrobial resistance draw most of the attention from the public and scientific community, cardiovascular diseases (CVD) are the major epidemic of the modern era, claiming a higher number of deaths than cancer, malaria, AIDS, or tuberculosis. Indeed, CVD remains the most common cause of death worldwide (Park et al., 2008). Just in Europe, CVD are responsible for 45% of all deaths, reaching 4 million deaths per year (Townsend et al., 2015). Coronary heart disease is the most common single cause of death, resulting in 19% of deaths in men and 20% of deaths in women, much higher than breast cancer in women (2%) and lung cancer in men (6%) (Townsend et al., 2015). Most conventional therapeutics and clinical approaches are outdated, and researchers are putting their efforts into fast employing all the potential of “nano” in the CVD management, approaching strategies for both imaging and treatment of these conditions.

Developing New Agents for CVD Imaging

Conventional medical tools still fail on the detection of atherosclerotic lesions and plaque rupture, while interventions with a catheter ultrasound or magnetic resonance imaging (MRI) give purely morphological information, without stating the progression of inflammation and the occurring of functional changes (Park et al., 2008). New imaging techniques and agents are in high demand. Contrast agents incorporating nanoparticles and peptides have significantly evolved and are now capable of detecting and quantifying microthrombus. Nonetheless, they mainly consist of hard particles, which present excretion difficulties and slow or inexistent metabolization. The tendency is to look for more compliant particles, like self-assembling and small molecules, capable of flowing through the microvasculature of clearance organs (Pan et al., 2009), with low toxicity, good biodegradability, and biocompatibility (Park et al., 2008). There have been advances in the development of fibrin-specific manganese nanocolloids, that successfully reach the low nanomolar range of detection and present a high relaxitivity (Pan et al., 2009). These results are directly compared to the micromolar range only of the mostly used gadolinium-based agents.

In recent years, some approaches previously used mainly for oncology imaging have been adapted to cardiovascular imaging, as it is the case of iron oxide nanoparticles, especially in the form of ultra-small supermagnetic iron oxide (USPIO) nanoparticles (<50 nm) (Ploussi et al., 2015). Early use of these NPs for medical imaging was described as a solution for the limiting factor in MRI, the background signal produced by the host tissue, but they can also be used for magnetic particle imaging (MPI), being capable of providing a higher sensitivity and a better spatial resolution (Gleich and Weizenecker, 2005). Due to the high interest in these particles, several variations of superparamagnetic iron oxide nanoparticles (SPIONs) can be found, as well as the characterization of their behavior in different situations. Park et al. have subjected three formulations of SPIONs to pH variations (5, 7, 9, and 11) and time progression (30 days) (Park et al., 2012). By light scattering analysis at pH 11, a significant increase in hydrodynamic diameter was observed, leading to the conclusion that nanoparticle aggregation is occurring, especially when PEG was one of the components (Figure 4A). Under further analysis, authors concluded that the PEG coating was desorbed from the surface, leading to an unstable NP suspension and triggering aggregation. At pH 7, there were no alterations in measured sizes for the particle.

FIGURE 4
www.frontiersin.org

Figure 4. Example of dynamic light scattering applications to study nanoparticle stability at different pH values (A,B) and temperatures (C). (A) Iron nanoparticles (FeNPs) aggregation stability studied over time at four different pH values. At the top, FeNPs without surface polymer; in the middle FeNPs, coupled with PEG2000 and at the bottom FeNPs couple with PEG5000. Adapted with permission from Park et al. (2012). Copyright 2018 American Chemical Society. (B) SPIONs aggregation stability studied over time at two pH values. Adapted from Oberle and Lüdtke-Buzug (2013). (C) Perfluoropentane (PFP) micelles size stability studied at different temperatures. Micelles were prepared with different percentages of PFP (Rapoport et al., 2007).

SPIONs can have multiple coating options. Thus, authors can play with either PEG or other biocompatible molecules, like chitosan. Szpak et al. studied the stability of iron oxide nanoparticles coated with a thin layer of charged chitosan derivatives (Szpak et al., 2013). Performing DLS measurements, they concluded that the diameter for the negatively charged NPs was slightly smaller, indicating an effect of the charge in the behavior of its milieu. Further characterization of the coating charge was performed by zeta-potential measurements. Authors emphasized that, for biological applications, SPIONs must be resistant to adsorption of biomacromolecules and that chitosan might be the ideal candidate for facilitating several degrees of physical properties manipulation, such as the tailoring of surface charge. Another study has analyzed the stability of SPIONs at 37°C, but this time with a dextran coating (Oberle and Lüdtke-Buzug, 2013). The nanoparticles were stable at pH 7.2 for as long as 6 weeks, while at pH 6.2 the hydrodynamic diameter strongly increased, denoting particle aggregation, also visible by precipitation (Figure 4B). This strongly suggests that dextran is not only a biocompatible polymer, but also an excellent solution to keep a SPION formulation stable at physiological conditions.

Iron oxide agents have been tested for the detection of abdominal aortic aneurism (Richards et al., 2011; Sadat et al., 2011), atherosclerotic plaques (Schmitz et al., 2001; Kooi et al., 2003), and acute myocardial infarction (Alam et al., 2012; Yilmaz et al., 2012, 2013a,b). Tang et al. also extensively studied the use of ferumoxtran-10 for imaging carotid plaques (Tang et al., 2009a,b), carotid stenosis (Tang et al., 2006) and carotid atheromas (Tang et al., 2007, 2008). Moreover, several of these tests are basically giving a new use for ferucarbotran (Resovist), an agent firstly used to detect either benign or malign hepatic lesions (Namkung et al., 2007). The particles usually have a core of magnetite (Fe3O4)/maghemite (γFe2O3) coated with carboxydextran, and an overall hydrodynamic diameter of 62 nm (Reimer et al., 1995). By 2015, only Resovist was available in very limited countries, with other agents being stopped for further development. This is the case of ferumoxtran-10 (Sinerem), also widely tested in the cited studies for cardiovascular conditions, although initially developed for lymph node imaging. Here, the core is a crystalline inverse spinel structure of magnetite, coated with dextran, with 20 nm diameter (Shen et al., 1993).

At the same time, the field is also actively looking for new disease biomarkers and intensively exploring agents involved in the inflammatory response in CVD. In an effort to provide better quantitative macrophage imaging in vascular tissue, Keliher et al. developed a class of modified polyglucose nanoparticles, with a size below the limit for renal excretion (Keliher et al., 2017). When macrophages fail in removing cholesterol deposits from the arterial wall, an inflammatory response is triggered, with recruitment of more cells, which enhances inflammation and compromises blood flow and tissue integrity. Animal studies succeeded in detecting atherosclerotic regions with practically no interaction with other lymphocytes. The same happened when using mice with permanent coronary ligation (Keliher et al., 2017).

With CVD prevention being the focus of significant attention from the scientific community, and studies pointing to new disease biomarkers steadily reaching publications (Hijazi et al., 2016; Walters et al., 2016), some authors explored CVD relationship with other medical conditions (Gerdes et al., 2014; Pavo et al., 2015). Following the work on screening for a peptide to bind to atherosclerotic plaques (Hong et al., 2008), other authors have carried out its incorporation as a targeting moiety in chitosan nanoparticles (Park et al., 2008). After working on hydrophobic modified glycol chitosan (HGC) nanoparticles as cancer imaging probes (Park et al., 2007) and for other therapeutic purposes (Kwon et al., 2003; Park et al., 2004; Kim et al., 2006), the team was able to conjugate the peptide on the NP surface and detect the selective binding to atherosclerotic plaques in vivo, by adhering to the IL-4 receptor on endothelial cells, macrophages and smooth muscle cells (Park et al., 2008). The authors highlighted that these 270 nm self-assembled nanoparticles have a long residence time even in flow conditions. In fact, the fluorescence in the aortic arch of the Ldlr−/− mice exhibited a more prominent fluorescence signal than the aortic arch of healthy mice, even after 6 h from intravenous administration.

Drug Delivering Nanoparticles for CVD Treatment

The delivery of a therapeutic drug through a nanoparticle vehicle allows high drug concentrations in the intended local environments, while the total drug concentration and side effects are significantly reduced (Chen et al., 2015). In CVD, the introduction of these therapeutic agents can be done either with surgical intervention or through systemic administration. In cases of coronary artery disease, a common approach is a percutaneous coronary intervention. This procedure is performed under local anesthesia and involves the insertion of a guidewire into the aorta, to then pass other therapeutic tools, such as inflatable balloons, stents, and catheters (Chen et al., 2015). An usual side effect is restenosis, which is a narrowing of the artery, either by remodeling and recoiling of the vessel lumen, or by proliferation of smooth muscle cells in response to the injury caused by the inserted devices (Cyrus et al., 2008). The insertion of a stent is indicated to prevent the situation, but it may itself be a trigger to a proliferative response, and also a vehicle for cell migration, decreasing the internal diameter of the treated vessel. For this reason, it is important to develop modified coatings. The real advantages of using nanoparticle infused polymers are still under evaluation, with some studies concluding that the use of drug-eluting stents has a risk for thrombosis at least as great as with bare-metal stents, showing no significant effect in long-term survival and a reduction in the need for re-intervention (Kastrati et al., 2007). Other authors have shown that although successfully preventing restenosis, drug-eluting stents have a major impact delaying endothelial healing (Cyrus et al., 2008). That occurs because the coatings are made using cytostatic agents, like sirolimus and paclitaxel (Brito and Amiji, 2007). Nonetheless, despite growing expertise in manipulating surfaces, structures and materials containing nanoparticles, it is also important to keep in mind that the nanotopography also plays a role in promoting cell mobility, adhesion, and differentiation (González-Béjar et al., 2016).

The increased variety of innovative materials for stent manufacture as also raised some awareness to determine, not only which have a better drug releasing performance, but also which are the safest to use in in vivo magnetic particle imaging. Wegner et al. explored the heating patterns of stents made from stainless steel, nitinol, platinum-chromium, and cobalt-chromium, from several diameters and lengths (Wegner et al., 2018). The study concluded that temperature increase is a real concern in larger stents, with diameter playing a leading role. In addition, the authors suggest that a combination of geometries, and conductive and non-conductive materials would most probably prove to be the best approach for stent design.

On another approach, we can also find the direct targeting of blood clots by vehiculation of thrombolytic agents. Recombinant tissue plasminogen activator (tPA) or streptokinase are administered in case of ischemic events, despite frequent complications, such as severe hemorrhage. Both agents act by activating plasmin, which will then lyse the fibrin from the clot structure. By directing the action to the specific clot site, it is possible to diminish the administered doses, maintaining or even improving the outcome, and avoiding a systemic effect that leads to the unwanted symptoms (Elbayoumi and Torchilin, 2008; Kim et al., 2009; Koudelka et al., 2016).

Perfluorocarbons have long been a target of scientific interest for their properties of transport and delivery, being firstly studied for their ability to dissolve oxygen, and later modified with several less toxic variations to produce emulsions, as a substitute of blood components (Clark and Gollan, 1966; Mitsuno et al., 1984; Riess, 1984). Nowadays, perfluorocarbons are still being explored for their drug carrying properties, with applications in both prevention and treatment of medical conditions (Chang et al., 1988; Schad and Hynynen, 2010; Song et al., 2016; Vemuri et al., 2016). Rapoport et al., for example, have already studied the influence of temperature on the stability of PEGylated perfluorochemical formulations (Rapoport et al., 2007). Following size evolution up to 42°C, by DLS, maintaining temperature for 5 min and cooling the sample before size measurements, the authors were able to observe the transformation of the nanodroplets in microbubbles within the prepared formulations (Figure 4C). Perfluorocarbon nanoparticles were also derivatized for fibrin targeting in blood clot (Lanza et al., 1996). Authors used a biotinylated form of the emulsion to target the NPs to thrombin. As a way of ensuring the presence of functional biotin at the surface, they performed an avidin titration while measuring particle size by DLS. The method revealed a steady increase in size with increasing concentrations of avidin, which demonstrates the successful functionalization of the emulsion. In another application, in vitro studies have demonstrated the lytic activity of perfluorocarbon NP formulations directed to clot dissolution. Marsh et al. successfully conjugated streptokinase on the surface of NPs made mainly of fluorooctylbromide, egg yolk lecithin, cholesterol and MPB-PE (1,2-dioleoyl-sn-glycero-3-phosphoethanolamine-N-[4-(p-maleimidophenyl)butyramide]) for streptokinase conjugation (Marsh et al., 2007). The in vitro assays showed an almost complete lysis of the human plasma clots in less than 60 min. In fact, Banai et al. demonstrated that the administration of a nanoencapsulated drug, in this case tyrphostin AGL-2043, can even be more clinically interesting than its surface adsorbed or even free form, in reducing in-stent neointimal formation (Banai et al., 2005). The in vitro characterization of the particles used by Banai et al. on their studies was previously studied by other authors (Chorny et al., 2002). These authors focused on describing a modified nanoprecipitation method for optimization of the particles' size, drug recovery yield, and release kinetics. NPs intended for intravascular delivery must be developed under optimal conditions. The authors used DLS measurements to evaluate the influence of the polymer poly(D,L-lactide) (PLA) concentration and ethanol presence in production varying sizes of particles. The size of the NPs increased with the increase of the PLA concentration used and decreased by increasing the concentration of ethanol. Ethanol impairs the solubility of PLA, decreasing the time for precipitation when in contact with an aqueous phase, producing smaller droplets. Authors also stress the importance of size in the release of the therapeutic agent, with their data showing that smaller particles had higher release rates, as a result of a greater surface area exposed to the medium (Chorny et al., 2002).

Combining Imaging and First-Line Treatment—Theranostics

Theranostics is a field of individualized medicine that arises from combining diagnostics and therapy, being possible due to the capacity of nanoplatforms to carry cargo and target a specific agent. Being a major aim on the area of cancer research, on the specific context of cardiovascular diseases the ultimate goal is to non-invasively define atherosclerotic burden, to deliver effective targeted drug at a fraction of previous levels, and to quantify local response to treatment (Winter et al., 2006). Although still far from meeting clinical standards, this is fast progressing, with in vivo studies showing high success. A clot-binding peptide was already used in the surface of micelles to target blood clots, both concentrating an imaging dye and specifically delivering a thrombin inhibitor (Peters et al., 2009). Other authors have used a formulation with a perfluorocarbon core surrounded by a lipid coat, which was derivatized with PPACK (phenylalanine-proline-arginine-chloromethylketone), delivering that thrombin inhibitor to the kidney (Chen et al., 2015). Due to the chemical properties of the core of the NP, it was possible to run quantitative molecular imaging in vivo with fluorine MRI, confirming the concentration of particles in the kidney, thrombin binding, and perfusion recovery. The same combined approach was used twice, in the first demonstrating that paramagnetic perfluorocarbon nanoparticles could be used for the non-invasive detection and delineation of a marker of aortic plaque angiogenesis, as well as the local delivery of an effective single treatment of fumagillin, inhibiting plaque angiogenesis at a dose several orders of magnitude lower than previously reported (Winter et al., 2006, 2008). This type of approach becomes especially relevant when the disease severity is rapidly advancing, as the targeted local administration of antiangiogenic agents delays plaque progression and enlarges the window of opportunity for clinical intervention through other conventional methods. The same authors demonstrated that αVβ3-targeted fumagillin NPs could also work synergistically with other therapeutic agents, greatly increasing a continuous clinically relevant antiangiogenic effect (Winter et al., 2008).

This class of combined-effect NPs are now tailoring the future of new therapeutics, with most significance in the administration of therapeutic agents as the disease is being diagnosed, providing a first line of care to the patient. In fact, both the NPs mentioned in the imaging and in the treatment sections may be further manipulated to also acquire the other applicability.

Conclusion

Nanomedicine is considered as, at least, one of the most relevant paths for the future of therapeutics. This perception has dramatically increased with the new paradigm of personalized medicine. Inserted in this category, NPs with activity toward the diseases responsible for the major death tolls worldwide have deserved special attention. A myriad of systems has been proposed in recent years, some of them described above. However, just a small number has reached clinical trials. More studies are necessary to assess the real potential of these nanosystems, and even different formulations need to be considered if we want to tackle cardiovascular diseases, cancer, or multi-resistant infections.

Common to all systems described are the methods necessary to characterize the proposed nanoparticles. In this field, light scattering spectroscopy techniques have a considerable number of roles to play, for different purposes. Regarding DLS, it is mostly used to determine the size distribution of the NPs, but some authors use this technique in different ways (Water et al., 2015; Xie et al., 2017; Xia et al., 2018). Confirm surface functionalization, characterize long term stability in different media or pH values, and identification of the aggregation profile are just examples of other possible applications of this technique (Figure 5) (Mohanty et al., 2013; Casciaro et al., 2017; Shmarakov et al., 2017; Zhang et al., 2018). As for zeta-potential, optimization of peptide anchoring profile to the nanoparticle, confirmation of surface charge modification, and validation of electrostatic interaction between the NP and the target cells are some of the processes where it could be essential (Kuai et al., 2010; Takara et al., 2010; Pal et al., 2016). In future studies, light scattering should be essential for the characterization and development of nanoparticles applied to therapeutics, which do not invalidate the fact that other techniques should be also used to further confirm the conclusions obtained.

FIGURE 5
www.frontiersin.org

Figure 5. Schematic representation of nanoparticle application in cancer therapy, considering the different size distribution profiles obtained after surface derivatization. NPs may have their surface derivatized with different materials,considering their intended purpose. If the coupled NPs do not aggregate, the size distribution will only reflect an increase corresponding to the coupling material, passing through the blood circulation and acting at the desired targets. Surface derivatization that promotes NP aggregation will be identified in the size distribution. On the bloodstream, NP aggregates will be recognized by macrophages, which will be responsible for their elimination.

Author Contributions

All authors listed have made a substantial, direct and intellectual contribution to the work, and approved it for publication.

Conflict of Interest Statement

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Acknowledgments

This work was supported by Fundação para a Ciência e a Tecnologia–Ministério da Ciência, Tecnologia e Ensino Superior (FCT-MCTES, Portugal) projects PTDC/BBB-BQB/3494/2014 and PTDC/BBB-BMD/6307/2014. This work was also supported by LISBOA-01-0145-FEDER-007391 project, cofunded by FEDER, through POR Lisboa 2020—Programa Operacional Regional de Lisboa, PORTUGAL 2020, and Fundação para a Ciência e a Tecnologia. PMC, MRF and MMD acknowledge FCT-MCTES fellowships SFRH/BD/108077/2015, SFRH/BD/100517/2014, and SFRH/BPD/122779/2016, respectively.

References

Agemy, L., Sugahara, K. N., Kotamraju, V. R., Gujraty, K., Girard, O. M., Kono, Y., et al. (2010). Nanoparticle-induced vascular blockade in human prostate cancer. Blood 116, 2847–2856. doi: 10.1182/blood-2010-03-274258

PubMed Abstract | CrossRef Full Text | Google Scholar

Alam, S. R., Shah, A. S., Richards, J., Lang, N. N., Barnes, G., Joshi, N., et al. (2012). Ultrasmall superparamagnetic particles of iron oxide in patients with acute myocardial infarction: early clinical experience. Circ. Cardiovasc. Imaging 5, 559–565. doi: 10.1161/CIRCIMAGING.112.974907

PubMed Abstract | CrossRef Full Text | Google Scholar

Albericio, F., and Kruger, H. G. (2012). Therapeutic peptides. Future Med. Chem. 4, 1527–1531. doi: 10.4155/fmc.12.94

PubMed Abstract | CrossRef Full Text | Google Scholar

Alegret, N., Criado, A., and Prato, M. (2017). Recent advances of graphene-based hybrids with magnetic nanoparticles for biomedical applications. Curr. Med. Chem. 24, 529–536. doi: 10.2174/0929867323666161216144218

PubMed Abstract | CrossRef Full Text | Google Scholar

Allahverdiyev, A. M., Kon, K. V., Abamor, E. S., Bagirova, M., and Rafailovich, M. (2011). Coping with antibiotic resistance: combining nanoparticles with antibiotics and other antimicrobial agents. Expert Rev. Anti Infect. Ther. 9, 1035–1052. doi: 10.1586/eri.11.121

PubMed Abstract | CrossRef Full Text | Google Scholar

Allen, T. M., and Cullis, P. R. (2013). Liposomal drug delivery systems: from concept to clinical applications. Adv. Drug Deliv. Rev. 65, 36–48. doi: 10.1016/j.addr.2012.09.037

PubMed Abstract | CrossRef Full Text | Google Scholar

Arakha, M., Pal, S., Samantarrai, D., Panigrahi, T. K., Mallick, B. C., Pramanik, K., et al. (2015). Antimicrobial activity of iron oxide nanoparticle upon modulation of nanoparticle-bacteria interface. Sci. Rep. 5:14813. doi: 10.1038/srep14813

PubMed Abstract | CrossRef Full Text | Google Scholar

Arnold, M., Karim-Kos, H. E., Coebergh, J. W., Byrnes, G., Antilla, A., Ferlay, J., et al. (2015). Recent trends in incidence of five common cancers in 26 European countries since 1988: analysis of the European Cancer Observatory. Eur. J. Cancer 51, 1164–1187. doi: 10.1016/j.ejca.2013.09.002

PubMed Abstract | CrossRef Full Text

Bachar, M., Mandelbaum, A., Portnaya, I., Perlstein, H., Even-Chen, S., Barenholz, Y., et al. (2012). Development and characterization of a novel drug nanocarrier for oral delivery, based on self-assembled β-casein micelles. J. Control Release 160, 164–171. doi: 10.1016/j.jconrel.2012.01.004

PubMed Abstract | CrossRef Full Text | Google Scholar

Bahnsen, J. S., Franzyk, H., Sayers, E. J., Jones, A. T., and Nielsen, H. M. (2015). Cell-penetrating antimicrobial peptides – prospectives for targeting intracellular infections. Pharm. Res. 32, 1546–1556. doi: 10.1007/s11095-014-1550-9

PubMed Abstract | CrossRef Full Text | Google Scholar

Banai, S., Chorny, M., Gertz, S. D., Fishbein, I., Gao, J., Perez, L., et al. (2005). Locally delivered nanoencapsulated tyrphostin (AGL-2043) reduces neointima formation in balloon-injured rat carotid and stented porcine coronary arteries. Biomaterials 26, 451–461. doi: 10.1016/j.biomaterials.2004.02.040

PubMed Abstract | CrossRef Full Text | Google Scholar

Batoni, G., Maisetta, G., and Esin, S. (2016). Antimicrobial peptides and their interaction with biofilms of medically relevant bacteria. Biochim. Biophys. Acta 1858, 1044–1060. doi: 10.1016/j.bbamem.2015.10.013

PubMed Abstract | CrossRef Full Text | Google Scholar

Bera, R. K., Mandal, S. M., and Raj, C. R. (2014). Antimicrobial activity of fluorescent Ag nanoparticles. Lett. Appl. Microbiol. 58, 520–526. doi: 10.1111/lam.12222

PubMed Abstract | CrossRef Full Text | Google Scholar

Berne, B. J., and Pecora, R. (eds.) (1976). Dynamic Light Scattering: With Applications to Chemistry, Biology, and Physics. 1st Edn. New York, NY: John Wiley & Sons, Inc Available online at: https://books.google.pt/books?id=vBB54ABhmuEC

Google Scholar

Bhattacharjee, S. (2016). DLS and zeta potential - What they are and what they are not? J. Control Release 235, 337–351. doi: 10.1016/j.jconrel.2016.06.017

PubMed Abstract | CrossRef Full Text | Google Scholar

Bilal, M., Rasheed, T., Iqbal, H. M. N., Hu, H., Wang, W., and Zhang, X. (2017). Macromolecular agents with antimicrobial potentialities: a drive to combat antimicrobial resistance. Int. J. Biol. Macromol. 103, 554–574. doi: 10.1016/j.ijbiomac.2017.05.071

PubMed Abstract | CrossRef Full Text | Google Scholar

Bilberg, K., Hovgaard, M. B., Besenbacher, F., and Baatrup, E. (2012). In vivo toxicity of silver nanoparticles and silver ions in zebrafish (Danio rerio). J. Toxicol. 2012:293784. doi: 10.1155/2012/293784

PubMed Abstract | CrossRef Full Text

Birla, S. S., Tiwari, V. V., Gade, A. K., Ingle, A. P., Yadav, A. P., and Rai, M. K. (2009). Fabrication of silver nanoparticles by Phoma glomerata and its combined effect against Escherichia coli, Pseudomonas aeruginosa and Staphylococcus aureus. Lett. Appl. Microbiol. 48, 173–179. doi: 10.1111/j.1472-765X.2008.02510.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Briggs, J., and Nicoli, D. F. (1980). Photon correlation spectroscopy of polydisperse systems. J. Chem. Phys. 72, 6024–6030. doi: 10.1063/1.439057

CrossRef Full Text | Google Scholar

Brito, L., and Amiji, M. (2007). Nanoparticulate carriers for the treatment of coronary restenosis. Int. J. Nanomedicine 2, 143–61.

PubMed Abstract | Google Scholar

Casciaro, B., Moros, M., Rivera-Fernández, S., Bellelli, A., de la Fuente, J. M., and Mangoni, M. L. (2017). Gold-nanoparticles coated with the antimicrobial peptide esculentin-1a(1-21)NH2 as a reliable strategy for antipseudomonal drugs. Acta Biomater. 47, 170–181. doi: 10.1016/j.actbio.2016.09.041

PubMed Abstract | CrossRef Full Text | Google Scholar

Chang, D.-K., Chiu, C.-Y., Kuo, S.-Y., Lin, W.-C., Lo, A., Wang, Y.-P., et al. (2009). Antiangiogenic targeting liposomes increase therapeutic efficacy for solid tumors. J. Biol. Chem. 284, 12905–12916. doi: 10.1074/jbc.M900280200

PubMed Abstract | CrossRef Full Text | Google Scholar

Chang, S., Ozmert, E., and Zimmerman, N. J. (1988). Intraoperative perfluorocarbon liquids in the management of proliferative vitreoretinopathy. Am. J. Ophthalmol. 106, 668–674. doi: 10.1016/0002-9394(88)90698-8

PubMed Abstract | CrossRef Full Text | Google Scholar

Chen, C., Gunawan, P., Lou, X. W. D., and Xu, R. (2012). Silver nanoparticles deposited layered double hydroxide nanoporous coatings with excellent antimicrobial activities. Adv. Funct. Mater. 22, 780–787. doi: 10.1002/adfm.201102333

CrossRef Full Text | Google Scholar

Chen, J., Vemuri, C., Palekar, R. U., Gaut, J. P., Goette, M., Hu, L., et al. (2015). Antithrombin nanoparticles improve kidney reperfusion and protect kidney function after ischemia-reperfusion injury. Am. J. Physiol. Renal Physiol. 308, F765–F773. doi: 10.1152/ajprenal.00457.2014

PubMed Abstract | CrossRef Full Text | Google Scholar

Chorny, M., Fishbein, I., Danenberg, H. D., and Golomb, G. (2002). Lipophilic drug loaded nanospheres prepared by nanoprecipitation: effect of formulation variables on size, drug recovery and release kinetics. J. Control Release 83, 389–400. doi: 10.1016/S0168-3659(02)00211-0

PubMed Abstract | CrossRef Full Text | Google Scholar

Chu, B. (1974). Laser Light Scattering, 2nd Edn Elsevier Science. Available online at: https://books.google.pt/books?id=yKvGOyAHQlkC

Clark, L. C., and Gollan, F. (1966). Survival of mammals breathing organic liquids equilibrated with oxygen at atmospheric pressure. Science 152, 1755–1756. doi: 10.1126/SCIENCE.152.3730.1755

PubMed Abstract | CrossRef Full Text | Google Scholar

Colombo, G., Curnis, F., De Mori, G. M., Gasparri, A., Longoni, C., Sacchi, A., et al. (2002). Structure-activity relationships of linear and cyclic peptides containing the NGR tumor-homing motif. J. Biol. Chem. 277, 47891–47897. doi: 10.1074/jbc.M207500200

PubMed Abstract | CrossRef Full Text | Google Scholar

Cosco, D., Cilurzo, F., Maiuolo, J., Federico, C., Di Martino, M. T., Cristiano, M. C., et al. (2015a). Delivery of miR-34a by chitosan/PLGA nanoplexes for the anticancer treatment of multiple myeloma. Sci. Rep. 5:17579. doi: 10.1038/srep17579

PubMed Abstract | CrossRef Full Text | Google Scholar

Cosco, D., Federico, C., Maiuolo, J., Bulotta, S., Molinaro, R., Paolino, D., et al. (2014). Physicochemical features and transfection properties of chitosan/poloxamer 188/poly(D,L-lactide-co-glycolide) nanoplexes. Int. J. Nanomedicine 9, 2359–2372. doi: 10.2147/IJN.S58362

PubMed Abstract | CrossRef Full Text | Google Scholar

Cosco, D., Paolino, D., De Angelis, F., Cilurzo, F., Celia, C., Di Marzio, L., et al. (2015b). Aqueous-core PEG-coated PLA nanocapsules for an efficient entrapment of water soluble anticancer drugs and a smart therapeutic response. Eur. J. Pharm. Biopharm. 89, 30–39. doi: 10.1016/j.ejpb.2014.11.012

PubMed Abstract | CrossRef Full Text | Google Scholar

Cosco, D., Paolino, D., Maiuolo, J., Russo, D., and Fresta, M. (2011). Liposomes as multicompartmental carriers for multidrug delivery in anticancer chemotherapy. Drug Deliv. Transl. Res. 1, 66–75. doi: 10.1007/s13346-010-0007-x

PubMed Abstract | CrossRef Full Text | Google Scholar

Cyrus, T., Zhang, H., Allen, J. S., Williams, T. A., Hu, G., Caruthers, S. D., et al. (2008). Intramural delivery of rapamycin with v 3-targeted paramagnetic nanoparticles inhibits stenosis after balloon injury. Arterioscler. Thromb. Vasc. Biol. 28, 820–826. doi: 10.1161/ATVBAHA.107.156281

CrossRef Full Text | Google Scholar

de la Fuente-Núñez, C., Silva, O. N., Lu, T. K., and Franco, O. L. (2017). Antimicrobial peptides: role in human disease and potential as immunotherapies. Pharmacol. Ther. 178, 132–140. doi: 10.1016/j.pharmthera.2017.04.002

PubMed Abstract | CrossRef Full Text | Google Scholar

de Oliveira, J. F. A., Saito, Â., Bido, A. T., Kobarg, J., Stassen, H. K., and Cardoso, M. B. (2017). Defeating bacterial resistance and preventing mammalian cells toxicity through rational design of antibiotic-functionalized nanoparticles. Sci. Rep. 7:1326. doi: 10.1038/s41598-017-01209-1

PubMed Abstract | CrossRef Full Text | Google Scholar

Di Francesco, M., Celia, C., Primavera, R., D'Avanzo, N., Locatelli, M., Fresta, M., et al. (2017). Physicochemical characterization of pH-responsive and fusogenic self-assembled non-phospholipid vesicles for a potential multiple targeting therapy. Int. J. Pharm. 528, 18–32. doi: 10.1016/j.ijpharm.2017.05.055

PubMed Abstract | CrossRef Full Text | Google Scholar

Dias, S. A., Freire, J. M., Pérez-Peinado, C., Domingues, M. M., Gaspar, D., Vale, N., et al. (2017). New potent membrane-targeting antibacterial peptides from viral capsid proteins. Front. Microbiol. 8:775. doi: 10.3389/fmicb.2017.00775

PubMed Abstract | CrossRef Full Text | Google Scholar

Dickey, S. W., Cheung, G. Y. C., and Otto, M. (2017). Different drugs for bad bugs: antivirulence strategies in the age of antibiotic resistance. Nat. Rev. Drug Discov. 16, 457–471. doi: 10.1038/nrd.2017.23

PubMed Abstract | CrossRef Full Text | Google Scholar

Dinali, R., Ebrahiminezhad, A., Manley-Harris, M., Ghasemi, Y., and Berenjian, A. (2017). Iron oxide nanoparticles in modern microbiology and biotechnology. Crit. Rev. Microbiol. 43, 493–507. doi: 10.1080/1040841X.2016.1267708

PubMed Abstract | CrossRef Full Text | Google Scholar

Domingues, M. M., Santiago, P. S., Castanho, M. A., and Santos, N. C. (2008). What can light scattering spectroscopy do for membrane-active peptide studies? J. Pept. Sci. 14, 394–400. doi: 10.1002/psc.1007

PubMed Abstract | CrossRef Full Text | Google Scholar

Eertwegh, A. J. M., van den Pinedo, H. M., and Smorenburg, C. H. (2006). Drugs Affecting Growth of Tumours, 1st Edn. Basel: Birkhäuser.

Einstein, A. (1905). Über einen die erzeugung und verwandlung des lichtes betreffenden heuristischen gesichtspunkt. Ann. Phys. 322, 132–148. doi: 10.1002/andp.19053220607

CrossRef Full Text | Google Scholar

Einstein, A. (1906). Zur theorie der brownschen bewegung. Ann. Phys. 324, 371–381. doi: 10.1002/andp.19063240208

CrossRef Full Text | Google Scholar

El-Batal, A. I., Mosalam, F. M., Ghorab, M. M., Hanora, A., and Elbarbary, A. M. (2018). Antimicrobial, antioxidant and anticancer activities of zinc nanoparticles prepared by natural polysaccharides and gamma radiation. Int. J. Biol. Macromol. 107, 2298–2311. doi: 10.1016/j.ijbiomac.2017.10.121

PubMed Abstract | CrossRef Full Text | Google Scholar

Elbayoumi, T. A., and Torchilin, V. P. (2008). Liposomes for targeted delivery of antithrombotic drugs. Expert Opin. Drug Deliv. 5, 1185–1198. doi: 10.1517/17425240802497457

PubMed Abstract | CrossRef Full Text | Google Scholar

Elzoghby, A. O., Elgohary, M. M., and Kamel, N. M. (2015). Implications of protein- and Peptide-based nanoparticles as potential vehicles for anticancer drugs. Adv. Protein Chem. Struct. Biol. 98, 169–221. doi: 10.1016/bs.apcsb.2014.12.002

PubMed Abstract | CrossRef Full Text | Google Scholar

Fang, B., Jiang, Y., Nüsslein, K., Rotello, V. M., and Santore, M. M. (2015). Antimicrobial surfaces containing cationic nanoparticles: how immobilized, clustered, and protruding cationic charge presentation affects killing activity and kinetics. Colloids Surf. B. Biointerfaces 125, 255–263. doi: 10.1016/j.colsurfb.2014.10.043

PubMed Abstract | CrossRef Full Text | Google Scholar

Fayaz, A. M., Girilal, M., Mahdy, S. A., Somsundar, S. S., Venkatesan, R., and Kalaichelvan, P. T. (2011). Vancomycin bound biogenic gold nanoparticles: a different perspective for development of anti VRSA agents. Process Biochem. 46, 636–641. doi: 10.1016/j.procbio.2010.11.001

CrossRef Full Text | Google Scholar

Felício, M. R., Silva, O. N., Gonçalves, S., Santos, N. C., and Franco, O. L. (2017). Peptides with dual antimicrobial and anticancer activities. Front. Chem. 5:5. doi: 10.3389/fchem.2017.00005

PubMed Abstract | CrossRef Full Text | Google Scholar

Fidler, I. J. (1988). Targeting of immunomodulators to mononuclear phagocytes for therapy of cancer. Adv. Drug Deliv. Rev. 2, 69–106. doi: 10.1016/0169-409X(88)90006-3

CrossRef Full Text | Google Scholar

Fischer, K., and Schmidt, M. (2016). Pitfalls and novel applications of particle sizing by dynamic light scattering. Biomaterials 98, 79–91. doi: 10.1016/j.biomaterials.2016.05.003

PubMed Abstract | CrossRef Full Text | Google Scholar

Flemming, H. C., Wingender, J., Szewzyk, U., Steinberg, P., Rice, S. A., and Kjelleberg, S. (2016). Biofilms: an emergent form of bacterial life. Nat. Rev. Microbiol. 14, 563–575. doi: 10.1038/nrmicro.2016.94

PubMed Abstract | CrossRef Full Text | Google Scholar

Galdiero, S., and Gomes, P. (2017). Peptide-based drugs and drug delivery systems. Molecules 22:2185. doi: 10.3390/molecules22122185

PubMed Abstract | CrossRef Full Text | Google Scholar

Gao, W., Chen, Y., Zhang, Y., Zhang, Q., and Zhang, L. (2017). Nanoparticle-based local antimicrobial drug delivery. Adv. Drug Deliv. Rev. doi: 10.1016/j.addr.2017.09.015

PubMed Abstract | CrossRef Full Text | Google Scholar

Gao, W., Thamphiwatana, S., and Angsantikul, P. (2014). Nanoparticle approaches against bacterial infections. Wiley Interdiscip. Rev. Nanomed. Nanobiotechnol. 6, 532–547. doi: 10.1002/wnan.1282

PubMed Abstract | CrossRef Full Text | Google Scholar

García-Gallego, S., Franci, G., Falanga, A., Gómez, R., Folliero, V., Galdiero, S., et al. (2017). Function oriented molecular design: dendrimers as novel antimicrobials. Molecules 22, 1581. doi: 10.3390/molecules22101581

PubMed Abstract | CrossRef Full Text | Google Scholar

Gaspar, D. P., Faria, V., Gonçalves, L. M., Taboada, P., Remuñán-López, C., and Almeida, A. J. (2016). Rifabutin-loaded solid lipid nanoparticles for inhaled antitubercular therapy: physicochemical and in vitro studies. Int. J. Pharm. 497, 199–209. doi: 10.1016/j.ijpharm.2015.11.050

PubMed Abstract | CrossRef Full Text | Google Scholar

Gerdes, S., Osadtschy, S., Buhles, N., Baurecht, H., and Mrowietz, U. (2014). Cardiovascular biomarkers in patients with psoriasis. Exp. Dermatol. 23, 322–325. doi: 10.1111/exd.12381

PubMed Abstract | CrossRef Full Text | Google Scholar

Geszke-Moritz, M., and Moritz, M. (2016). Solid lipid nanoparticles as attractive drug vehicles: composition, properties and therapeutic strategies. Mater. Sci. Eng. C Mater. Biol. Appl. 68, 982–994. doi: 10.1016/j.msec.2016.05.119

PubMed Abstract | CrossRef Full Text | Google Scholar

Gleich, B., and Weizenecker, J. (2005). Tomographic imaging using the nonlinear response of magnetic particles. Nature 435, 1214–1217. doi: 10.1038/nature03808

PubMed Abstract | CrossRef Full Text | Google Scholar

Golubeva, O. Y., Shamova, O. V., Orlov, D. S., Pazina, T. Y., Boldina, A. S., Drozdova, I. A., et al. (2011). Synthesis and study of antimicrobial activity of bioconjugates of silver nanoparticles and endogenous antibiotics. Glas. Phys. Chem. 37, 78–84. doi: 10.1134/S1087659611010056

CrossRef Full Text | Google Scholar

Gomes, B., Augusto, M. T., Felício, M. R., Hollmann, A., Franco, O. L., Gonçalves, S., et al. (2018). Designing improved active peptides for therapeutic approaches against infectious diseases. Biotechnol. Adv. 36, 415–429. doi: 10.1016/j.biotechadv.2018.01.004

PubMed Abstract | CrossRef Full Text | Google Scholar

González-Béjar, M., Francés-Soriano, L., and Pérez-Prieto, J. (2016). Upconversion nanoparticles for bioimaging and regenerative medicine. Front. Bioeng. Biotechnol. 4:47. doi: 10.3389/fbioe.2016.00047

PubMed Abstract | CrossRef Full Text | Google Scholar

Gu, H., Ho, P. L., Tong, E., Wang, L., and Xu, B. (2003). Presenting vancomycin on nanoparticles to enhance antimicrobial activities. Nano Lett. 3, 1261–1263. doi: 10.1021/nl034396z

CrossRef Full Text | Google Scholar

Guidotti, G., Brambilla, L., and Rossi, D. (2017). Cell-penetrating peptides: from basic research to clinics. Trends Pharmacol. Sci. 38, 406–424. doi: 10.1016/j.tips.2017.01.003

PubMed Abstract | CrossRef Full Text | Google Scholar

Hamilton, A. M., Aidoudi-Ahmed, S., Sharma, S., Kotamraju, V. R., Foster, P. J., Sugahara, K. N., et al. (2015). Nanoparticles coated with the tumor-penetrating peptide iRGD reduce experimental breast cancer metastasis in the brain. J. Mol. Med. 93, 991–1001. doi: 10.1007/s00109-015-1279-x

PubMed Abstract | CrossRef Full Text | Google Scholar

Hancock, R. E., Haney, E. F., and Gill, E. E. (2016). The immunology of host defence peptides: beyond antimicrobial activity. Nat. Rev. Immunol. 16, 321–334. doi: 10.1038/nri.2016.29

PubMed Abstract | CrossRef Full Text | Google Scholar

Haney, E. F., Mansour, S. C., and Hancock, R. E. (2017). Antimicrobial peptides: an introduction. Methods Mol. Biol. 1548, 3–22. doi: 10.1007/978-1-4939-6737-7_1

PubMed Abstract | CrossRef Full Text | Google Scholar

Hann, I. M., and Prentice, H. G. (2001). Lipid-based amphotericin B: a review of the last 10 years of use. Int. J. Antimicrob. Agents 17, 161–169. doi: 10.1016/S0924-8579(00)00341-1

PubMed Abstract | CrossRef Full Text | Google Scholar

Harding, S. E. (1997). “Protein hydrodynamics,” in Protein: A Comprehensive Treatise, ed G. Allen (Greenwich: Jai Press, Incorporated), 271–305. Available online at: https://books.google.pt/books?id=HZaslAEACAAJ

Google Scholar

Hare, J. I., Lammers, T., Ashford, M. B., Puri, S., Storm, G., and Barry, S. T. (2017). Challenges and strategies in anti-cancer nanomedicine development: an industry perspective. Adv. Drug Deliv. Rev. 108, 25–38. doi: 10.1016/j.addr.2016.04.025

PubMed Abstract | CrossRef Full Text | Google Scholar

Hijazi, Z., Aulin, J., Andersson, U., Alexander, J. H., Gersh, B., Granger, C. B., et al. (2016). Biomarkers of inflammation and risk of cardiovascular events in anticoagulated patients with atrial fibrillation. Heart 102, 508–517. doi: 10.1136/heartjnl-2015-308887

PubMed Abstract | CrossRef Full Text | Google Scholar

Hong, H. Y., Lee, H. Y., Kwak, W., Yoo, J., Na, M.-H., So, I. S., et al. (2008). Phage display selection of peptides that home to atherosclerotic plaques: IL-4 receptor as a candidate target in atherosclerosis. J. Cell. Mol. Med. 12, 2003–2014. doi: 10.1111/j.1582-4934.2008.00189.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Ionov, M., Ciepluch, K., Klajnert, B., Glinska, S., Gomez-Ramirez, R., de la Mata, F. J., et al. (2013). Complexation of HIV derived peptides with carbosilane dendrimers. Colloids Surf. B. Biointerfaces 101, 236–242. doi: 10.1016/j.colsurfb.2012.07.011

PubMed Abstract | CrossRef Full Text | Google Scholar

Jahan, S. T., Sadat, S. M. A., Walliser, M., and Haddadi, A. (2017). Targeted therapeutic nanoparticles: an immense promise to fight against cancer. J. Drug Deliv. 2017, 1–24. doi: 10.1155/2017/9090325

PubMed Abstract | CrossRef Full Text | Google Scholar

Jena, P., Mohanty, S., Mallick, R., Jacob, B., and Sonawane, A. (2012). Toxicity and antibacterial assessment of chitosan-coated silver nanoparticles on human pathogens and macrophage cells. Int. J. Nanomedicine 7, 1805–1818. doi: 10.2147/IJN.S28077

PubMed Abstract | CrossRef Full Text | Google Scholar

Jiang, Z., Dong, X., and Sun, Y. (2018). Charge effects of self-assembled chitosan-hyaluronic acid nanoparticles on inhibiting amyloid β-protein aggregation. Carbohydr. Res. 461, 11–18. doi: 10.1016/j.carres.2018.03.001

PubMed Abstract | CrossRef Full Text | Google Scholar

Jurj, A., Braicu, C., Pop, L.-A., Tomuleasa, C., Gherman, C. D., and Berindan-Neagoe, I. (2017). The new era of nanotechnology, an alternative to change cancer treatment. Drug Des. Devel. Ther. 11, 2871–2890. doi: 10.2147/DDDT.S142337

PubMed Abstract | CrossRef Full Text | Google Scholar

Kastrati, A., Mehilli, J., Pache, J., Kaiser, C., Valgimigli, M., Kelbaek, H., et al. (2007). Analysis of 14 trials comparing sirolimus-eluting stents with bare-metal stents. N. Engl. J. Med. 356, 1030–1039. doi: 10.1056/NEJMoa067484

PubMed Abstract | CrossRef Full Text | Google Scholar

Kaszuba, M., Corbett, J., Watson, F. M., and Jones, A. (2010). High-concentration zeta potential measurements using light-scattering techniques. Philos. Trans. A Math. Phys. Eng. Sci. 368, 4439–4451. doi: 10.1098/rsta.2010.0175

PubMed Abstract | CrossRef Full Text | Google Scholar

Keliher, E. J., Ye, Y.-X., Wojtkiewicz, G. R., Aguirre, A. D., Tricot, B., Senders, M. L., et al. (2017). Polyglucose nanoparticles with renal elimination and macrophage avidity facilitate PET imaging in ischaemic heart disease. Nat. Commun. 8:14064. doi: 10.1038/ncomms14064

PubMed Abstract | CrossRef Full Text | Google Scholar

Kim, J.-H., Kim, Y.-S., Kim, S., Park, J. H., Kim, K., Choi, K., et al. (2006). Hydrophobically modified glycol chitosan nanoparticles as carriers for paclitaxel. J. Control Release 111, 228–234. doi: 10.1016/j.jconrel.2005.12.013

PubMed Abstract | CrossRef Full Text | Google Scholar

Kim, J. Y., Kim, J. K., Park, J. S., Byun, Y., and Kim, C. K. (2009). The use of PEGylated liposomes to prolong circulation lifetimes of tissue plasminogen activator. Biomaterials 30, 5751–5756. doi: 10.1016/j.biomaterials.2009.07.021

PubMed Abstract | CrossRef Full Text | Google Scholar

Kooi, M. E., Cappendijk, V. C., Cleutjens, K. B., Kessels, A. G., Kitslaar, P. J., Borgers, M., et al. (2003). Accumulation of ultrasmall superparamagnetic particles of iron oxide in human atherosclerotic plaques can be detected by in vivo magnetic resonance imaging. Circulation 107, 2453–2458. doi: 10.1161/01.CIR.0000068315.98705.CC

PubMed Abstract | CrossRef Full Text | Google Scholar

Koppel, D. E. (1972). Analysis of macromolecular polydispersity in intensity correlation spectroscopy: the method of cumulants. J. Chem. Phys. 57, 4814–4820. doi: 10.1063/1.1678153

CrossRef Full Text | Google Scholar

Koudelka, S., Mikulik, R., Mašek, J., Raška, M., Turánek Knotigová, P., Miller, A. D., et al. (2016). Liposomal nanocarriers for plasminogen activators. J. Control Release 227, 45–57. doi: 10.1016/j.jconrel.2016.02.019

PubMed Abstract | CrossRef Full Text | Google Scholar

Kratz, F. (2014). A clinical update of using albumin as a drug vehicle - a commentary. J. Control Release 190, 331–336. doi: 10.1016/j.jconrel.2014.03.013

PubMed Abstract | CrossRef Full Text | Google Scholar

Kristensen, M., Birch, D., and Mørck Nielsen, H. (2016). Applications and challenges for use of cell-penetrating peptides as delivery vectors for peptide and protein cargos. Int. J. Mol. Sci. 17:185. doi: 10.3390/ijms17020185

PubMed Abstract | CrossRef Full Text | Google Scholar

Kuai, R., Yuan, W., Li, W., Qin, Y., Tang, J., Yuan, M., et al. (2011). Targeted delivery of cargoes into a murine solid tumor by a cell-penetrating peptide and cleavable poly(ethylene glycol) comodified liposomal delivery system via systemic administration. Mol. Pharm. 8, 2151–2161. doi: 10.1021/mp200100f

PubMed Abstract | CrossRef Full Text | Google Scholar

Kuai, R., Yuan, W., Qin, Y., Chen, H., Tang, J., Yuan, M., et al. (2010). Efficient delivery of payload into tumor cells in a controlled manner by TAT and thiolytic cleavable PEG co-modified liposomes. Mol. Pharm. 7, 1816–1826. doi: 10.1021/mp100171c

PubMed Abstract | CrossRef Full Text | Google Scholar

Kwon, S., Park, J. H., Chung, H., Kwon, I. C., Jeong, S. Y., and Kim, I.-S. (2003). Physicochemical characteristics of self-assembled nanoparticles based on glycol chitosan bearing 5β-cholanic acid. Langmuir 19, 10188–10193. doi: 10.1021/la0350608

CrossRef Full Text | Google Scholar

Lai, H. Z., Chen, W. Y., Wu, C. Y., and Chen, Y. C. (2015). Potent antibacterial nanoparticles for pathogenic bacteria. ACS Appl. Mater. Interfaces 7, 2046–2054. doi: 10.1021/am507919m

PubMed Abstract | CrossRef Full Text | Google Scholar

Lanza, G. M., Wallace, K. D., Scott, M. J., Cacheris, W. P., Abendschein, D. R., Christy, D. H., et al. (1996). A novel site-targeted ultrasonic contrast agent with broad biomedical application. Circulation 94, 3334–3340.

PubMed Abstract | Google Scholar

Lee, T. Y., Wu, H. C., Tseng, Y. L., and Lin, C. T. (2004). A novel peptide specifically binding to nasopharyngeal carcinoma for targeted drug delivery. Cancer Res. 64, 8002–8008. doi: 10.1158/0008-5472.CAN-04-1948

PubMed Abstract | CrossRef Full Text | Google Scholar

Libralato, G., Galdiero, E., Falanga, A., Carotenuto, R., de Alteriis, E., and Guida, M. (2017). Toxicity effects of functionalized quantum dots, gold and polystyrene nanoparticles on target aquatic biological models: a review. Molecules 22:E1439. doi: 10.3390/molecules22091439

PubMed Abstract | CrossRef Full Text | Google Scholar

Licciardi, M., Paolino, D., Mauro, N., Cosco, D., Giammona, G., Fresta, M., et al. (2016). Cationic supramolecular vesicular aggregates for pulmonary tissue selective delivery in anticancer therapy. ChemMedChem 11, 1734–1744. doi: 10.1002/cmdc.201600070

PubMed Abstract | CrossRef Full Text | Google Scholar

Liu, D., Yang, F., Xiong, F., and Gu, N. (2016). The smart drug delivery system and its clinical potential. Theranostics 6, 1306–1323. doi: 10.7150/thno.14858

PubMed Abstract | CrossRef Full Text | Google Scholar

Liu, L., Xu, K., Wang, H., Tan, P. K., Fan, W., Venkatraman, S. S., et al. (2009). Self-assembled cationic peptide nanoparticles as an efficient antimicrobial agent. Nat. Nanotechnol. 4, 457–463. doi: 10.1038/nnano.2009.153

PubMed Abstract | CrossRef Full Text | Google Scholar

Liu, L., Yang, J., Xie, J., Luo, Z., Jiang, J., Yang, Y. Y., et al. (2013). The potent antimicrobial properties of cell penetrating peptide-conjugated silver nanoparticles with excellent selectivity for gram-positive bacteria over erythrocytes. Nanoscale 5, 3834–3840. doi: 10.1039/c3nr34254a

PubMed Abstract | CrossRef Full Text | Google Scholar

Llewelyn, M. J., Fitzpatrick, J. M., Darwin, E., SarahTonkin-Crine, Gorton, C., Paul, J., et al. (2017). The antibiotic course has had its day. BMJ 358:j3418. doi: 10.1136/bmj.j3418

PubMed Abstract | CrossRef Full Text | Google Scholar

MacEwan, S. R., and Chilkoti, A. (2014). Applications of elastin-like polypeptides in drug delivery. J. Control Release 190, 314–330. doi: 10.1016/j.jconrel.2014.06.028

PubMed Abstract | CrossRef Full Text | Google Scholar

Mahmoodi, N. O., Ghavidast, A., and Amirmahani, N. (2016). A comparative study on the nanoparticles for improved drug delivery systems. J. Photochem. Photobiol. B 162, 681–693. doi: 10.1016/j.jphotobiol.2016.07.037

PubMed Abstract | CrossRef Full Text | Google Scholar

Marchiò, S., Lahdenranta, J., Schlingemann, R. O., Valdembri, D., Wesseling, P., Arap, M. A., et al. (2004). Aminopeptidase A is a functional target in angiogenic blood vessels. Cancer Cell 5, 151–162. doi: 10.1016/S1535-6108(04)00025-X

PubMed Abstract | CrossRef Full Text | Google Scholar

Marsh, J. N., Senpan, A., Hu, G., Scott, M. J., Gaffney, P. J., Wickline, S. A., et al. (2007). Fibrin-targeted perfluorocarbon nanoparticles for targeted thrombolysis. Nanomedicine 2, 533–543. doi: 10.2217/17435889.2.4.533

PubMed Abstract | CrossRef Full Text | Google Scholar

Mie, G. (1908). Beiträge zur Optik trüber Medien, speziell kolloidaler Metallösungen. Ann. Phys. 330, 377–445. doi: 10.1002/andp.19083300302

CrossRef Full Text | Google Scholar

Mitsuno, T., Ohyanagi, H., and Yokoyama, K. (1984). Development of a perfluorochemical emulsion as a blood gas carrier. Artif. Organs 8, 25–33. doi: 10.1111/j.1525-1594.1984.tb04240.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Mohanty, S., Jena, P., Mehta, R., Pati, R., Banerjee, B., Patil, S., et al. (2013). Cationic antimicrobial peptides and biogenic silver nanoparticles kill mycobacteria without eliciting DNA damage and cytotoxicity in mouse macrophages. Antimicrob. Agents Chemother. 57, 3688–3698. doi: 10.1128/AAC.02475-12

PubMed Abstract | CrossRef Full Text | Google Scholar

Mohanty, S., Mishra, S., Jena, P., Jacob, B., Sarkar, B., and Sonawane, A. (2012). An investigation on the antibacterial, cytotoxic, and antibiofilm efficacy of starch-stabilized silver nanoparticles. Nanomedicine 8, 916–924. doi: 10.1016/j.nano.2011.11.007

PubMed Abstract | CrossRef Full Text

Molinaro, R., Wolfram, J., Federico, C., Cilurzo, F., Di Marzio, L., Ventura, C. A., et al. (2013). Polyethylenimine and chitosan carriers for the delivery of RNA interference effectors. Expert Opin. Drug Deliv. 10, 1653–1668. doi: 10.1517/17425247.2013.840286

PubMed Abstract | CrossRef Full Text | Google Scholar

Montes Ruiz-Cabello, F. J., Trefalt, G., Maroni, P., and Borkovec, M. (2014). Electric double-layer potentials and surface regulation properties measured by colloidal-probe atomic force microscopy. Phys. Rev. E 90:012301. doi: 10.1103/PhysRevE.90.012301

CrossRef Full Text | Google Scholar

Morales, D. P., Wonderly, W. R., Huang, X., McAdams, M., Chron, A. B., and Reich, N. O. (2017). Affinity-based assembly of peptides on plasmonic nanoparticles delivered intracellularly with light activated control. Bioconjug. Chem. 28, 1816–1820. doi: 10.1021/acs.bioconjchem.7b00276

PubMed Abstract | CrossRef Full Text | Google Scholar

Morrison, I. D., Grabowski, E. F., and Herb, C. A. (1985). Improved techniques for particle size determination by quasi-elastic light scattering. Langmuir 1, 496–501. doi: 10.1021/la00064a016

CrossRef Full Text | Google Scholar

Namkung, S., Zech, C. J., Helmberger, T., Reiser, M. F., and Schoenberg, S. O. (2007). Superparamagnetic iron oxide (SPIO)-enhanced liver MRI with ferucarbotran: efficacy for characterization of focal liver lesions. J. Magn. Reson. Imaging 25, 755–765. doi: 10.1002/jmri.20873

PubMed Abstract | CrossRef Full Text | Google Scholar

Neelay, O. P., Peterson, C. A., Snavely, M. E., Brown, T. C., TecleMariam, A. F., Campbell, J. A., et al. (2017). Antimicrobial peptides interact with peptidoglycan. J. Mol. Struct. 1146, 329–336. doi: 10.1016/j.molstruc.2017.06.018

CrossRef Full Text | Google Scholar

Niemirowicz, K., Prokop, I., Wilczewska, A. Z., Wnorowska, U., Piktel, E., Watek, M., et al. (2015). Magnetic nanoparticles enhance the anticancer activity of cathelicidin LL-37 peptide against colon cancer cells. Int. J. Nanomedicine 10, 3843–3853. doi: 10.2147/IJN.S76104

PubMed Abstract | CrossRef Full Text | Google Scholar

Oberdörster, G. (2010). Safety assessment for nanotechnology and nanomedicine: concepts of nanotoxicology. J. Intern. Med. 267, 89–105. doi: 10.1111/j.1365-2796.2009.02187.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Oberle, A., and Lüdtke-Buzug, K. (2013). Stability analysis of superparamagnetic iron oxide nanoparticles (spions) at 37°C. Biomed. Tech. doi: 10.1515/bmt-2013-4099. [Epub ahead of print].

PubMed Abstract | CrossRef Full Text | Google Scholar

Pal, I., Brahmkhatri, V. P., Bera, S., Bhattacharyya, D., Quirishi, Y., Bhunia, A., et al. (2016). Enhanced stability and activity of an antimicrobial peptide in conjugation with silver nanoparticle. J. Colloid Interface Sci. 483, 385–393. doi: 10.1016/j.jcis.2016.08.043

PubMed Abstract | CrossRef Full Text | Google Scholar

Pan, D., Senpan, A., Caruthers, S. D., Williams, T. A., Scott, M. J., Gaffney, P. J., et al. (2009). Sensitive and efficient detection of thrombus with fibrin-specific manganese nanocolloids. Chem. Commun. 14:3234. doi: 10.1039/b902875g

CrossRef Full Text | Google Scholar

Panahi, Y., Farshbaf, M., Mohammadhosseini, M., Mirahadi, M., Khalilov, R., Saghfi, S., et al. (2017). Recent advances on liposomal nanoparticles: synthesis, characterization and biomedical applications. Artif. Cells Nanomedicine Biotechnol. 45, 788–799. doi: 10.1080/21691401.2017.1282496

PubMed Abstract | CrossRef Full Text | Google Scholar

Paolino, D., Accolla, M. L., Cilurzo, F., Cristiano, M. C., Cosco, D., Castelli, F., et al. (2017). Interaction between PEG lipid and DSPE/DSPC phospholipids: an insight of PEGylation degree and kinetics of de-PEGylation. Colloids Surf. B. Biointerfaces 155, 266–275. doi: 10.1016/j.colsurfb.2017.04.018

PubMed Abstract | CrossRef Full Text | Google Scholar

Paolino, D., Cosco, D., Celano, M., Moretti, S., Puxeddu, E., Russo, D., et al. (2013). Gemcitabine-loaded biocompatible nanocapsules for the effective treatment of human cancer. Nanomedicine 8, 193–201. doi: 10.2217/nnm.12.101

PubMed Abstract | CrossRef Full Text | Google Scholar

Park, J. H., Kwon, S., Nam, J.-O., Park, R.-W., Chung, H., Seo, S. B., et al. (2004). Self-assembled nanoparticles based on glycol chitosan bearing 5beta-cholanic acid for RGD peptide delivery. J. Control Release 95, 579–588. doi: 10.1016/j.jconrel.2003.12.020

PubMed Abstract | CrossRef Full Text | Google Scholar

Park, K., Hong, H. Y., Moon, H. J., Lee, B. H., Kim, I. S., Kwon, I. C., et al. (2008). A new atherosclerotic lesion probe based on hydrophobically modified chitosan nanoparticles functionalized by the atherosclerotic plaque targeted peptides. J. Control Release 128, 217–223. doi: 10.1016/j.jconrel.2008.03.019

PubMed Abstract | CrossRef Full Text | Google Scholar

Park, K., Kim, J.-H., Nam, Y. S., Lee, S., Nam, H. Y., Kim, K., et al. (2007). Effect of polymer molecular weight on the tumor targeting characteristics of self-assembled glycol chitosan nanoparticles. J. Control Release 122, 305–314. doi: 10.1016/j.jconrel.2007.04.009

PubMed Abstract | CrossRef Full Text | Google Scholar

Park, S., Chibli, H., Wong, J., and Nadeau, J. L. (2011). Antimicrobial activity and cellular toxicity of nanoparticle-polymyxin B conjugates. Nanotechnology 22:185101. doi: 10.1088/0957-4484/22/18/185101

PubMed Abstract | CrossRef Full Text | Google Scholar

Park, S.-C., Kim, Y.-M., Lee, J.-K., Kim, N.-H., Kim, E.-J., Heo, H., et al. (2017). Targeting and synergistic action of an antifungal peptide in an antibiotic drug-delivery system. J. Control Release 256, 46–55. doi: 10.1016/j.jconrel.2017.04.023

PubMed Abstract | CrossRef Full Text | Google Scholar

Park, W., and Na, K. (2015). Advances in the synthesis and application of nanoparticles for drug delivery. Wiley Interdiscip. Rev. Nanomed. Nanobiotechnol. 7, 494–508. doi: 10.1002/wnan.1325

PubMed Abstract | CrossRef Full Text | Google Scholar

Park, Y., Whitaker, R. D., Nap, R. J., Paulsen, J. L., Mathiyazhagan, V., Doerrer, L. H., et al. (2012). Stability of superparamagnetic iron oxide nanoparticles at different pH values: experimental and theoretical analysis. Langmuir 28, 6246–6255. doi: 10.1021/la204628c

PubMed Abstract | CrossRef Full Text | Google Scholar

Pavo, N., Raderer, M., Hülsmann, M., Neuhold, S., Adlbrecht, C., Strunk, G., et al. (2015). Cardiovascular biomarkers in patients with cancer and their association with all-cause mortality. Heart 101, 1874–1880. doi: 10.1136/heartjnl-2015-307848

PubMed Abstract | CrossRef Full Text | Google Scholar

Pearce, T. R., Shroff, K., and Kokkoli, E. (2012). Peptide targeted lipid nanoparticles for anticancer drug delivery. Adv. Mater. 24, 3803–3822, 3710. doi: 10.1002/adma.201200832

PubMed Abstract | CrossRef Full Text | Google Scholar

Peretz, Y., Malishev, R., Kolusheva, S., and Jelinek, R. (2018). Nanoparticles modulate membrane interactions of human Islet amyloid polypeptide (hIAPP). Biochim. Biophys. Acta. doi: 10.1016/j.bbamem.2018.03.029. [Epub ahead of print].

PubMed Abstract | CrossRef Full Text | Google Scholar

Peters, D., Kastantin, M., Kotamraju, V. R., Karmali, P. P., Gujraty, K., Tirrell, M., et al. (2009). Targeting atherosclerosis by using modular, multifunctional micelles. Proc. Natl. Acad. Sci. U.S.A. 106, 9815–9819. doi: 10.1073/pnas.0903369106

PubMed Abstract | CrossRef Full Text | Google Scholar

Pham, D. C., Nguyen, T. H., Ngoc, U. T. P., Le, N. T. T., Tran, T. V., and Nguyen, D. H. (2018). Preparation, characterization and antifungal properties of chitosan-silver nanoparticles synergize fungicide against Pyricularia oryzae. J. Nanosci. Nanotechnol. 18, 5299–5305. doi: 10.1166/jnn.2018.15400

PubMed Abstract | CrossRef Full Text | Google Scholar

Pichavant, L., Amador, G., Jacqueline, C., Brouillaud, B., Héroguez, V., and Durrieu, M.-C. (2012). pH-controlled delivery of gentamicin sulfate from orthopedic devices preventing nosocomial infections. J. Control. Release 162, 373–381. doi: 10.1016/j.jconrel.2012.06.033

PubMed Abstract | CrossRef Full Text | Google Scholar

Pichavant, L., Bourget, C., Durrieu, M.-C., and Héroguez, V. (2011). Synthesis of pH-Sensitive particles for local delivery of an antibiotic via dispersion ROMP. Macromolecules 44, 7879–7887. doi: 10.1021/ma2015479

CrossRef Full Text | Google Scholar

Pichavant, L., Carrié, H., Nguyen, M. N., Plawinski, L., Durrieu, M.-C., and Héroguez, V. (2016). Vancomycin functionalized nanoparticles for bactericidal biomaterial surfaces. Biomacromolecules 17, 1339–1346. doi: 10.1021/acs.biomac.5b01727

PubMed Abstract | CrossRef Full Text | Google Scholar

Ploussi, A. G., Gazouli, M., Stathis, G., Kelekis, N. L., and Efstathopoulos, E. P. (2015). Iron oxide nanoparticles as contrast agents in molecular magnetic resonance imaging: do they open new perspectives in cardiovascular imaging? Cardiol. Rev. 23, 229–235. doi: 10.1097/CRD.0000000000000055

PubMed Abstract | CrossRef Full Text | Google Scholar

Primavera, R., Palumbo, P., Celia, C., Cinque, B., Carata, E., Carafa, M., et al. (2018). An insight of in vitro transport of PEGylated non-ionic surfactant vesicles (NSVs) across the intestinal polarized enterocyte monolayers. Eur. J. Pharm. Biopharm. 127, 432–442. doi: 10.1016/j.ejpb.2018.03.013

PubMed Abstract | CrossRef Full Text | Google Scholar

Provencher, S. W. (1982a). A constrained regularization method for inverting data represented by linear algebraic or integral equations. Comput. Phys. Commun. 27, 213–227. doi: 10.1016/0010-4655(82)90173-4

CrossRef Full Text | Google Scholar

Provencher, S. W. (1982b). CONTIN: a general purpose constrained regularization program for inverting noisy linear algebraic and integral equations. Comput. Phys. Commun. 27, 229–242. doi: 10.1016/0010-4655(82)90174-6

CrossRef Full Text | Google Scholar

Pusey, P. (1974). “Correlation and light beating spectroscopy,” in Photon Correlation and Light Beating Spectroscopy Nato Science Series B, eds H. Z. Cummins and E. R. Pike (Boston, MA: Springer), 387–428.

Qayyum, S., and Khan, A. U. (2016). Nanoparticles vs. biofilms: a battle against another paradigm of antibiotic resistance. Medchemcomm 7, 1479–1498. doi: 10.1039/C6MD00124F

CrossRef Full Text | Google Scholar

Rajchakit, U., and Sarojini, V. (2017). Recent developments in antimicrobial-peptide-conjugated gold nanoparticles. Bioconjug. Chem. 28, 2673–2686. doi: 10.1021/acs.bioconjchem.7b00368

PubMed Abstract | CrossRef Full Text | Google Scholar

Ramos, A. P., Cruz, M. A. E., Tovani, C. B., and Ciancaglini, P. (2017). Biomedical applications of nanotechnology. Biophys. Rev. 9, 79–89. doi: 10.1007/s12551-016-0246-2

PubMed Abstract | CrossRef Full Text | Google Scholar

Rapoport, N., Gao, Z., and Kennedy, A. (2007). Multifunctional nanoparticles for combining ultrasonic tumor imaging and targeted chemotherapy. J. Natl. Cancer Inst. 99, 1095–1106. doi: 10.1093/jnci/djm043

PubMed Abstract | CrossRef Full Text | Google Scholar

Reimer, P., Rummeny, E. J., Daldrup, H. E., Balzer, T., Tombach, B., Berns, T., et al. (1995). Clinical results with Resovist: a phase 2 clinical trial. Radiology 195, 489–496. doi: 10.1148/radiology.195.2.7724772

PubMed Abstract | CrossRef Full Text | Google Scholar

Ribeiro, S. M., Felício, M. R., Boas, E. V., Gonçalves, S., Costa, F. F., Samy, R. P., et al. (2016). New frontiers for anti-biofilm drug development. Pharmacol. Ther. 160, 133–144. doi: 10.1016/j.pharmthera.2016.02.006

PubMed Abstract | CrossRef Full Text | Google Scholar

Richards, J. M., Semple, S. I., MacGillivray, T. J., Gray, C., Langrish, J. P., Williams, M., et al. (2011). Abdominal aortic aneurysm growth predicted by uptake of ultrasmall superparamagnetic particles of iron oxide: a pilot study. Circ. Cardiovasc. Imaging 4, 274–281. doi: 10.1161/CIRCIMAGING.110.959866

PubMed Abstract | CrossRef Full Text | Google Scholar

Riess, J. G. (1984). Reassessment of criteria for the selection of perfluorochemicals for second-generation blood substitutes: analysis of structure/property relationships. Artif. Organs 8, 44–56. doi: 10.1111/j.1525-1594.1984.tb04243.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Ruden, S., Hilpert, K., Berditsch, M., Wadhwani, P., and Ulrich, A. S. (2009). Synergistic interaction between silver nanoparticles and membrane-permeabilizing antimicrobial peptides. Antimicrob. Agents Chemother. 53, 3538–3540. doi: 10.1128/AAC.01106-08

PubMed Abstract | CrossRef Full Text | Google Scholar

Sadat, U., Taviani, V., Patterson, A. J., Young, V. E., Graves, M. J., Teng, Z., et al. (2011). Ultrasmall superparamagnetic iron oxide-enhanced magnetic resonance imaging of abdominal aortic aneurysms–a feasibility study. Eur. J. Vasc. Endovasc. Surg. 41, 167–174. doi: 10.1016/j.ejvs.2010.08.022

PubMed Abstract | CrossRef Full Text | Google Scholar

Safra, T. (2003). Cardiac safety of liposomal anthracyclines. Oncologist 8(Suppl. 2), 17–24. doi: 10.1007/s12012-007-0014-4

PubMed Abstract | CrossRef Full Text | Google Scholar

Schad, K. C., and Hynynen, K. (2010). In vitro characterization of perfluorocarbon droplets for focused ultrasound therapy. Phys. Med. Biol. 55, 4933–4947. doi: 10.1088/0031-9155/55/17/004

PubMed Abstract | CrossRef Full Text | Google Scholar

Schmidt, C., and Storsberg, J. (2015). Nanomaterials-tools, technology and methodology of nanotechnology based biomedical systems for diagnostics and therapy. Biomedicines 3, 203–223. doi: 10.3390/biomedicines3030203

PubMed Abstract | CrossRef Full Text | Google Scholar

Schmitz, S. A., Taupitz, M., Wagner, S., Wolf, K. J., Beyersdorff, D., and Hamm, B. (2001). Magnetic resonance imaging of atherosclerotic plaques using superparamagnetic iron oxide particles. J. Magn. Reson. Imaging 14, 355–361. doi: 10.1002/jmri.1194

PubMed Abstract | CrossRef Full Text | Google Scholar

Serna, N., Sánchez-García, L., Sánchez-Chardi, A., Unzueta, U., Roldán, M., Mangues, R., et al. (2017). Protein-only, antimicrobial peptide-containing recombinant nanoparticles with inherent built-in antibacterial activity. Acta Biomater. 60, 256–263. doi: 10.1016/j.actbio.2017.07.027

PubMed Abstract | CrossRef Full Text | Google Scholar

Shanmugasundaram, T., Radhakrishnan, M., Gopikrishnan, V., Kadirvelu, K., and Balagurunathan, R. (2017). Biocompatible silver, gold and silver/gold alloy nanoparticles for enhanced cancer therapy: in vitro and in vivo perspectives. Nanoscale 9, 16773–16790. doi: 10.1039/c7nr04979j

PubMed Abstract | CrossRef Full Text | Google Scholar

Shen, T., Weissleder, R., Papisov, M., Bogdanov, A., and Brady, T. J. (1993). Monocrystalline iron oxide nanocompounds (MION): physicochemical properties. Magn. Reson. Med. 29, 599–604. doi: 10.1002/mrm.1910290504

PubMed Abstract | CrossRef Full Text | Google Scholar

Shmarakov, I., Mukha, I., Vityuk, N., Borschovetska, V., Zhyshchynska, N., Grodzyuk, G., et al. (2017). Antitumor activity of alloy and core-shell-type bimetallic AgAu nanoparticles. Nanoscale Res. Lett. 12:333. doi: 10.1186/s11671-017-2112-y

PubMed Abstract | CrossRef Full Text | Google Scholar

Siegert, A. J. F. (1943). On the Fluctuations in Signals Returned by Many Independently Moving Scatterers. Cambridge, MA: Radiation Laboratory, Massachusetts Institute of Technology. Available online at: https://books.google.pt/books?id=5PzqGwAACAAJ

Singh, L., Kruger, H. G., Maguire, G. E. M., Govender, T., and Parboosing, R. (2017). The role of nanotechnology in the treatment of viral infections. Ther. Adv. Infect. Dis. 4, 105–131. doi: 10.1177/2049936117713593

PubMed Abstract | CrossRef Full Text | Google Scholar

Siriwardena, T. N., Stach, M., He, R., Gan, B.-H., Javor, S., Heitz, M., et al. (2018). Lipidated peptide dendrimers killing multidrug-resistant bacteria. J. Am. Chem. Soc. 140, 423–432. doi: 10.1021/jacs.7b11037

PubMed Abstract | CrossRef Full Text | Google Scholar

Sneharani, A. H., Karakkat, J. V., Singh, S. A., and Rao, A. G. A. (2010). Interaction of curcumin with β-lactoglobulin-stability, spectroscopic analysis, and molecular modeling of the complex. J. Agric. Food Chem. 58, 11130–11139. doi: 10.1021/jf102826q

PubMed Abstract | CrossRef Full Text | Google Scholar

Solairaj, D., Rameshthangam, P., and Arunachalam, G. (2017). Anticancer activity of silver and copper embedded chitin nanocomposites against human breast cancer (MCF-7) cells. Int. J. Biol. Macromol. 105, 608–619. doi: 10.1016/j.ijbiomac.2017.07.078

PubMed Abstract | CrossRef Full Text | Google Scholar

Song, G., Liang, C., Yi, X., Zhao, Q., Cheng, L., Yang, K., et al. (2016). Perfluorocarbon-loaded hollow Bi2Se3 nanoparticles for timely supply of oxygen under near-infrared light to enhance the radiotherapy of cancer. Adv. Mater. 28, 2716–2723. doi: 10.1002/adma.201504617

PubMed Abstract | CrossRef Full Text | Google Scholar

Stetefeld, J., McKenna, S. A., and Patel, T. R. (2016). Dynamic light scattering: a practical guide and applications in biomedical sciences. Biophys. Rev. 8, 409–427. doi: 10.1007/s12551-016-0218-6

PubMed Abstract | CrossRef Full Text | Google Scholar

Strutt, J. W. (1871). XXXVI. On the light from the sky, its polarization and colour. Lond. Edinburgh Dublin Philos. Mag. J. Sci. 41, 274–279. doi: 10.1080/14786447108640479

CrossRef Full Text | Google Scholar

Stylianopoulos, T., and Jain, R. K. (2015). Design considerations for nanotherapeutics in oncology. Nanomedicine 11, 1893–1907. doi: 10.1016/j.nano.2015.07.015

PubMed Abstract | CrossRef Full Text

Sun, T., Zhang, Y. S., Pang, B., Hyun, D. C., Yang, M., and Xia, Y. (2014). Engineered nanoparticles for drug delivery in cancer therapy. Angew. Chem. Int. Ed. 53, 12320–12364. doi: 10.1002/anie.201403036

PubMed Abstract | CrossRef Full Text | Google Scholar

Szpak, A., Kania, G., Skórka, T., Tokarz, W., Zapotoczny, S., and Nowakowska, M. (2013). Stable aqueous dispersion of superparamagnetic iron oxide nanoparticles protected by charged chitosan derivatives. J. Nanopart. Res. 15:1372. doi: 10.1007/s11051-012-1372-9

PubMed Abstract | CrossRef Full Text | Google Scholar

Takara, K., Hatakeyama, H., Ohga, N., Hida, K., and Harashima, H. (2010). Design of a dual-ligand system using a specific ligand and cell penetrating peptide, resulting in a synergistic effect on selectivity and cellular uptake. Int. J. Pharm. 396, 143–148. doi: 10.1016/j.ijpharm.2010.05.002

PubMed Abstract | CrossRef Full Text | Google Scholar

Tam, J. P. (1988). Synthetic peptide vaccine design: synthesis and properties of a high-density multiple antigenic peptide system. Proc. Natl. Acad. Sci. U.S.A. 85, 5409–5413. doi: 10.1073/pnas.85.15.5409

PubMed Abstract | CrossRef Full Text | Google Scholar

Tam, J. P., Lu, Y.-A., and Yang, J.-L. (2002). Antimicrobial dendrimeric peptides. Eur. J. Biochem. 269, 923–932. doi: 10.1046/j.0014-2956.2001.02728.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Tang, T. Y., Howarth, S. P. S., Miller, S. R., Graves, M. J., Patterson, A. J., U.-King-Im, J.-M., et al. (2009a). The ATHEROMA (Atorvastatin Therapy: Effects on Reduction of Macrophage Activity) Study. Evaluation using ultrasmall superparamagnetic iron oxide-enhanced magnetic resonance imaging in carotid disease. J. Am. Coll. Cardiol. 53, 2039–2050. doi: 10.1016/j.jacc.2009.03.018

PubMed Abstract | CrossRef Full Text | Google Scholar

Tang, T. Y., Howarth, S. P., Miller, S. R., Graves, M. J., U.-King-Im, J.-M., Trivedi, R. A., et al. (2007). Comparison of the inflammatory burden of truly asymptomatic carotid atheroma with atherosclerotic plaques contralateral to symptomatic carotid stenosis: an ultra small superparamagnetic iron oxide enhanced magnetic resonance study. J. Neurol. Neurosurg. Psychiatry 78, 1337–1343. doi: 10.1136/jnnp.2007.118901

PubMed Abstract | CrossRef Full Text | Google Scholar

Tang, T. Y., Howarth, S. P. S., Miller, S. R., Graves, M. J., U.-King-Im, J. M., Li, Z. Y., et al. (2008). Comparison of the inflammatory burden of truly asymptomatic carotid atheroma with atherosclerotic plaques in patients with asymptomatic carotid stenosis undergoing coronary artery bypass grafting: an ultrasmall superparamagnetic iron oxide enhanced magnet. Eur. J. Vasc. Endovasc. Surg. 35, 392–398. doi: 10.1016/j.ejvs.2007.10.019

CrossRef Full Text | Google Scholar

Tang, T. Y., Patterson, A. J., Miller, S. R., Graves, M. J., Howarth, S. P., U.-King-Im, J. M., et al. (2009b). Temporal dependence of in vivo USPIO-enhanced MRI signal changes in human carotid atheromatous plaques. Neuroradiology 51, 457–465. doi: 10.1007/s00234-009-0523-x

PubMed Abstract | CrossRef Full Text | Google Scholar

Tang, T., Howarth, S. P., Miller, S. R., Trivedi, R., Graves, M. J., King-Im, J. U., et al. (2006). Assessment of inflammatory burden contralateral to the symptomatic carotid stenosis using high-resolution ultrasmall, superparamagnetic iron oxide-enhanced MRI. Stroke 37, 2266–2270. doi: 10.1161/01.STR.0000236063.47539.99

PubMed Abstract | CrossRef Full Text | Google Scholar

Townsend, N., Nichols, M., Scarborough, P., and Rayner, M. (2015). Cardiovascular disease in Europe-epidemiological update 2015. Eur. Heart J. 36, 2696–2705. doi: 10.1093/eurheartj/ehv428

CrossRef Full Text | Google Scholar

Tyndall, J. (1868). On the blue colour of the sky, the polarization of skylight, and on the polarization of light by cloudy matter generally. Proc. R Soc. Lond.17, 223–233. doi: 10.1098/rspl.1868.0033

CrossRef Full Text | Google Scholar

Unubol, N., Selim Cinaroglu, S., Elmas, M. A., Akcelik, S., Ozal Ildeniz, A. T., Arbak, S., et al. (2017). Peptide antibiotics developed by mimicking natural antimicrobial peptides. Clin. Microbiol. 6:1000291. doi: 10.4172/2327-5073.1000291

CrossRef Full Text | Google Scholar

Vemuri, C., Upadhya, A., Arif, B., Jia, J., Gaut, J., Manning, P., et al. (2016). Preservation of transplant organ function and recipient survival with thrombin-targeted perfluorocarbon nanoparticles perfused ex vivo. Circulation 134:A20869.

Google Scholar

Vigant, F., Santos, N. C., and Lee, B. (2015). Broad-spectrum antivirals against viral fusion. Nat. Rev. Microbiol. 13, 426–437. doi: 10.1038/nrmicro3475

PubMed Abstract | CrossRef Full Text | Google Scholar

Walters, J. W., Amos, D., Ray, K., and Santanam, N. (2016). Mitochondrial redox status as a target for cardiovascular disease. Curr. Opin. Pharmacol. 27, 50–55. doi: 10.1016/j.coph.2016.01.006

PubMed Abstract | CrossRef Full Text | Google Scholar

Wang, Y.-F., Liu, L., Xue, X., and Liang, X.-J. (2017). Nanoparticle-based drug delivery systems: what can they really do in vivo? F1000Res. 6:681. doi: 10.12688/f1000research.9690.1

PubMed Abstract | CrossRef Full Text | Google Scholar

Water, J. J., Smart, S., Franzyk, H., Foged, C., and Nielsen, H. M. (2015). Nanoparticle-mediated delivery of the antimicrobial peptide plectasin against Staphylococcus aureus in infected epithelial cells. Eur. J. Pharm. Biopharm. 92, 65–73. doi: 10.1016/j.ejpb.2015.02.009

PubMed Abstract | CrossRef Full Text | Google Scholar

Wegner, F., Friedrich, T., Panagiotopoulos, N., Valmaa, S., Goltz, J. P., Vogt, F. M., et al. (2018). First heating measurements of endovascular stents in magnetic particle imaging. Phys. Med. Biol. 63:045005. doi: 10.1088/1361-6560/aaa79c

PubMed Abstract | CrossRef Full Text | Google Scholar

Winter, P. M., Caruthers, S. D., Zhang, H., Williams, T. A., Wickline, S. A., and Lanza, G. M. (2008). Antiangiogenic Synergism of Integrin-Targeted Fumagillin Nanoparticles and Atorvastatin in Atherosclerosis. JACC Cardiovasc. Imaging 1, 624–634. doi: 10.1016/j.jcmg.2008.06.003

PubMed Abstract | CrossRef Full Text | Google Scholar

Winter, P. M., Neubauer, A. M., Caruthers, S. D., Harris, T. D., Robertson, J. D., Williams, T. A., et al. (2006). Endothelial alpha(v)beta3 integrin-targeted fumagillin nanoparticles inhibit angiogenesis in atherosclerosis. Arterioscler. Thromb. Vasc. Biol. 26, 2103–2109. doi: 10.1161/01.ATV.0000235724.11299.76

PubMed Abstract | CrossRef Full Text

Wong, A., Liu, Q., Griffin, S., Nicholls, A., and Regalbuto, J. R. (2017). Synthesis of ultrasmall, homogeneously alloyed, bimetallic nanoparticles on silica supports. Science 358, 1427–1430. doi: 10.1126/science.aao6538

PubMed Abstract | CrossRef Full Text | Google Scholar

World Health Organization (2015). Global Action Plan on Antimicrobial Resistance. Geneva: WHO Document Production Services.

Xia, Y., Xu, T., Wang, C., Li, Y., Lin, Z., Zhao, M., et al. (2018). Novel functionalized nanoparticles for tumor-targeting co-delivery of doxorubicin and siRNA to enhance cancer therapy. Int. J. Nanomedicine 13, 143–159. doi: 10.2147/IJN.S148960

PubMed Abstract | CrossRef Full Text | Google Scholar

Xie, F., Ding, R.-L., He, W.-F., Liu, Z.-J., Fu, S.-Z., Wu, J.-B., et al. (2017). In vivo antitumor effect of endostatin-loaded chitosan nanoparticles combined with paclitaxel on Lewis lung carcinoma. Drug Deliv. 24, 1410–1418. doi: 10.1080/10717544.2017.1378938

PubMed Abstract | CrossRef Full Text | Google Scholar

Xie, S., Tao, Y., Pan, Y., Qu, W., Cheng, G., Huang, L., et al. (2014). Biodegradable nanoparticles for intracellular delivery of antimicrobial agents. J. Control. Release 187, 101–117. doi: 10.1016/j.jconrel.2014.05.034

PubMed Abstract | CrossRef Full Text | Google Scholar

Yang, J., Lu, H., Li, M., Liu, J., Zhang, S., Xiong, L., et al. (2017). Development of chitosan-sodium phytate nanoparticles as a potent antibacterial agent. Carbohydr. Polym. 178, 311–321. doi: 10.1016/j.carbpol.2017.09.053

PubMed Abstract | CrossRef Full Text | Google Scholar

Yilmaz, A., Dengler, M. A., van der Kuip, H., Yildiz, H., Rösch, S., Klumpp, S., et al. (2013a). Imaging of myocardial infarction using ultrasmall superparamagnetic iron oxide nanoparticles: a human study using a multi-parametric cardiovascular magnetic resonance imaging approach. Eur. Heart J. 34, 462–475. doi: 10.1093/eurheartj/ehs366

PubMed Abstract | CrossRef Full Text | Google Scholar

Yilmaz, A., Rösch, S., Klingel, K., Kandolf, R., Helluy, X., Hiller, K.-H., et al. (2013b). Magnetic resonance imaging (MRI) of inflamed myocardium using iron oxide nanoparticles in patients with acute myocardial infarction - preliminary results. Int. J. Cardiol. 163, 175–182. doi: 10.1016/j.ijcard.2011.06.004

PubMed Abstract | CrossRef Full Text | Google Scholar

Yilmaz, A., Rösch, S., Yildiz, H., Klumpp, S., and Sechtem, U. (2012). First multiparametric cardiovascular magnetic resonance study using ultrasmall superparamagnetic iron oxide nanoparticles in a patient with acute myocardial infarction: new vistas for the clinical application of ultrasmall superparamagnetic iron oxide. Circulation 126, 1932–1934. doi: 10.1161/CIRCULATIONAHA.112.108167

PubMed Abstract | CrossRef Full Text | Google Scholar

Yu, X., Gou, X., Wu, P., Han, L., Tian, D., Du, F., et al. (2018). Activatable protein nanoparticles for targeted delivery of therapeutic peptides. Adv. Mater. 30:1705383. doi: 10.1002/adma.201705383

PubMed Abstract | CrossRef Full Text | Google Scholar

Zakerzadeh, E., Salehi, R., and Mahkam, M. (2017). Smart tetrazole-based antibacterial nanoparticles as multifunctional drug carriers for cancer combination therapy. Drug Dev. Ind. Pharm. 43, 1963–1977. doi: 10.1080/03639045.2017.1357730

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhang, C., Liu, L.-H., Qiu, W.-X., Zhang, Y.-H., Song, W., Zhang, L., et al. (2018). A transformable chimeric peptide for cell encapsulation to overcome multidrug resistance. Small. 14:e1703321. doi: 10.1002/smll.201703321

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhang, Q., Liao, G., Wei, D., and Nagai, T. (1998). Increase in gentamicin uptake by cultured mouse peritoneal macrophages and rat hepatocytes by its binding to polybutylcyanoacrylate nanoparticles. Int. J. Pharm. 164, 21–27. doi: 10.1016/S0378-5173(97)00395-5

CrossRef Full Text | Google Scholar

Zhang, X. (2015). Gold nanoparticles: recent advances in the biomedical applications. Cell Biochem. Biophys. 72, 771–775. doi: 10.1007/s12013-015-0529-4

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhang, X.-L., Jiang, A.-M., Ma, Z.-Y., Li, X.-B., Xiong, Y.-Y., Dou, J.-F., et al. (2015). The synthetic antimicrobial peptide pexiganan and its nanoparticles (PNPs) exhibit the anti-helicobacter pylori activity in vitro and in vivo. Molecules 20, 3972–3985. doi: 10.3390/molecules20033972

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhao, D., Zhao, X., Zu, Y., Li, J., Zhang, Y., Jiang, R., et al. (2010). Preparation, characterization, and in vitro targeted delivery of folate-decorated paclitaxel-loaded bovine serum albumin nanoparticles. Int. J. Nanomedicine 5, 669–677. doi: 10.2147/IJN.S12918

PubMed Abstract | CrossRef Full Text | Google Scholar

Keywords: nanoparticles, dynamic light scattering, zeta-potential, antimicrobial peptides, anticancer peptides, cardiovascular diseases

Citation: Carvalho PM, Felício MR, Santos NC, Gonçalves S and Domingues MM (2018) Application of Light Scattering Techniques to Nanoparticle Characterization and Development. Front. Chem. 6:237. doi: 10.3389/fchem.2018.00237

Received: 17 April 2018; Accepted: 04 June 2018;
Published: 25 June 2018.

Edited by:

José Morais Catita, Fernando Pessoa University, Portugal

Reviewed by:

Felisa Cilurzo, Università degli Studi G. d'Annunzio Chieti e Pescara, Italy
Victor M. Bolanos-Garcia, Oxford Brookes University, United Kingdom

Copyright © 2018 Carvalho, Felício, Santos, Gonçalves and Domingues. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Sónia Gonçalves, sabreu@medicina.ulisboa.pt
Marco M. Domingues, madomingues@medicina.ulisboa.pt

These authors have contributed equally to this work.

Download