Skip to content
Publicly Available Published by De Gruyter June 6, 2015

The bio-corona and its impact on nanomaterial toxicity

  • Dana Westmeier

    Dana Westmeier received her Master’s degree of Biomedicine from Johannes-Gutenberg University Mainz in 2013. She is currently a PhD candidate at the Institute for Molecular and Cellular Oncology/Nanobiomedicine at the University Medical Center Mainz with Prof Dr. Roland H. Stauber. Her research is focused on the (patho) biological influence of nanoparticles on human cells and pathogens as well as the impact of the NP corona on the microbiome-host interaction.

    , Chunying Chen

    Chunying Chen is a principal investigator at CAS Key Laboratory for Biomedical Effects of Nanomaterials and Nanosafety, National Center for Nanoscience and Technology of China. She received her Bachelor’s degree in Chemistry (1991) and PhD degree in Biomedical Engineering (1996) from Huazhong University of Science and Technology of China. Her research interests include the interaction of nanoparticles with biological systems, therapies for malignant tumors using theranostic nanomedicine systems and vaccine nanoadjuvants using nanomaterials, which are supported by the China MOST 973 Programs, EU-FP6 and FP7 and IAEA. She was awarded China Outstanding Young Female Scientists and the National Science Fund for Distinguished Young Scholars.

    , Roland H. Stauber

    Roland H. Stauber is currently a Professor at the Molecular and Cellular Oncology/Nanobiomedicine department at the ENT Medical University Mainz. His research interests focus on the understanding of molecular mechanisms at the nano-bio interface, the development of nano- and microtechnology for cancer treatment and diagnosis, as well as drug development. He has received the Alexander Karl Award for Cancer Research and a President’s Fellowship of the Chinese Academy of Sciences. Dr. Stauber received his PhD degree from Würzburg University in 1994 and did his post-doctoral training at the National Cancer Institute (USA).

    and Dominic Docter

    Dominic Docter is currently a junior group leader in the Department of Nanobiomedicine at the ENT University Medical Center of Mainz and supported by the Peter and Traudl Engelhorn Foundation. He received his Diploma degree in Biology in 2008 and his PhD degree at the Johannes Gutenberg University Mainz in 2014 after participating in the “Schwerpunktprogramm: Biological response to Nanoscale Particles” of the German Research Foundation. He has received the NMFZ-research award from the University Medical Center of Mainz. His research is focused on understanding events at the host-nano-bio interface to optimize and explore possible nano-diagnostic and -therapeutic applications.

    EMAIL logo

Abstract

The rapidly growing application of nano-sized materials and nano-scaled processes will result in increased exposure of humans and the environment. The small size of nanomaterials (NM) comparable with molecular building blocks of cells raises concerns that their toxic potential cannot be extrapolated from studies of larger particles due to their unique physico-chemical properties. These properties are also responsible that NM rapidly adsorb various (bio)molecules when introduced into complex physiological or natural environments. As the thus formed protein/biomolecule ‘corona’ seems to affect the NM’ in situ identity, an understanding of its toxicological relevance and the biophysical forces regulating corona formation is needed but not yet achieved. This review introduces our current concept of corona formation and evolution and present analytical methods for corona profiling. We discuss toxicity mechanisms potentially affected by the biomolecule corona, including NM cellular uptake and impact on components of the blood system. Further, we comment on pending knowledge gaps and challenges, which need to be resolved by the field. We conclude by presenting a tiered systems biology-driven approach recommended to mechanistically understand the coronas’ nanotoxicological relevance and predictive potential.

Introduction

Throughout evolution, humans have been exposed to airborne nanosized particles, which are currently defined as materials with at least one external dimension in the size range of approximately 1–100 nm. However, such exposure has increased dramatically over the last century due to anthropogenic sources and is expected to further increase due to the rapidly developing field of nanotechnology. The applications of engineered NM are not only increasingly used in technical products but also more and more in biotechnology and biomedicine (1–3). Engineered nanoparticles (NPs) can be generated by a variety of physical, chemical, and/or biotechnological methods, and additionally tailored by specific functionalization procedures to optimally meet their intended technological or biomedical applications (3, 4). As such, NPs have a wide range of potential applications in biomedical sciences to include targeted drug delivery systems and contrast agents for imaging methods such as magnetic resonance imaging or fluorescence spectroscopy and even as theranostics (5–16).

Though, besides the current enthusiasm for nanotechnology, the wide use of nanomaterials may pose unknown risks to human health and the environment (7, 17–21). Particularly, experiences with ‘new’ materials in the past, such as asbestos, have sensitized the public also for the risk of ‘similar’ NM, such as single-walled carbon nanotubes (SWCNT). Issues relevant to pulmonary toxicity of SWCNT, their direct cytotoxic effects, their ability to cause inflammation, and induce oxidative stress upon inhalation are currently actively investigated (7, 17–21). As NPs are able to highjack the endocytosis machinery and may enter almost any cell type (22), their recognition and engulfment by macrophages, phagocytosis and bio-distribution in tissues and the circulation are still under intense discussion (7, 17–21). Once inside a cell, NPs may cause adverse effects (23) resulting even in permanent cell damage (24, 25). As potential mechanisms, oxidative stress, inflammation, genetic instability, and the inhibition of proper cell division have been suggested (26–29). Moreover, immunosuppressive effects of NM and their impact on increased sensitivity of exposed individuals to additional microbial infections as well as their effect on the microbiome are currently investigated.

The discussion about nanosafety upon intended as well as unintended exposure of humans and the environment is thus clearly important and still ongoing (7, 17–21). As society at large is now aware of the use of NM in ever growing quantities in consumer products and their presence in the environment, critical interest in the impact of this emerging technology has grown. The main concern is whether the unknown risks of engineered NP, in particular their impact on health and environment, outweighs their established benefits for society. Therefore, a key issue in this field is to evaluate their potential toxicity, identify responsible pathways of toxicity (PoTs), and nano-structure activity relationships (nanoSARs). As the biomolecule corona forms rapidly in exposure relevant complex physiological and natural environments, an understanding of its relevance for nanotoxicity is mandatory.

Protein corona takes shape

When NPs are designed for specific biomedical applications, such as drug delivery and imaging purposes, these often requires intravenous injections of the NPs (2, 30). It is not surprising that for this intended human exposure, the detailed knowledge about physical and chemical aspects associated with the behavior of NPs in physiological systems in general has been recognized as an important factor especially for understanding nano-toxicology (19, 31). In physiological environments NPs are exposed not only to relatively high ion concentrations or drastic pH changes (32), but also to a huge variety of complex biomolecules (33). Many body fluids (e.g. blood, lung lining fluid, saliva, intestinal juice, etc.) have one common characteristics, their composition is highly complex (34, 35). Thus, whereas the physico-chemical properties and behavior of NPs can be engineered and controlled in technically stable, protected environments, such as technical products, this is no longer the case in complex physiological or natural environments. Such ‘complex environments’ are not only represented by simple and higher organisms, including humans, but also by complex solid and liquid interfaces to which NPs are exposed during their technical application and/or their intended/unintended exposure of humans. Moreover, complex organisms are again composed of various additional ‘complex (micro)environments’, such as organs and cells, which also differ very dramatically in their physico-chemical composition.

In this complex (micro)environments, NPs adsorb various (bio)molecules due to their high surface energy (36–38). The physical and chemical interactions with proteins and/or other biomolecules (e.g. phospholipids, sugars, nucleic acids, etc.) will in most cases significantly affect the NPs’ behavior and fate. By examining databases, it is obvious that the ‘nanoparticle corona’ represents a still unresolved ‘hot topic’, with high scientific and economic relevance. Over the last 10 years the number of publications dealing with ‘nanoparticles’ and ‘corona’ increased almost 15-fold (Figure 1). The adsorbed proteins clearly defines NPs surface and mediates further interactions between the NPs and the biological environment (39–42). This coating marks the biological identity of NPs and indirectly causes its ‘transformation’ by drastically altering the NPs’ colloidal stability. Here, the protein corona can either have a stabilizing effect by inducing steric stabilization (43) or a de-stabilization impact, caused by protein mediated bridging, charge compensation and/or by introduction of charge inhomogeneity onto the NPs surface. Upon aggregation, multiple interactions may result in stronger affinities compared to proteins binding to single NPs, which is more likely to occur in a biological solution in which particles are highly diluted. Moreover, there could even be a trapping of (abundant) proteins in such aggregates with low or no affinity for single NPs. Depending on the NPs, such aggregation may also require a certain time and thus, additional may impact the kinetics of corona formation. Ex vivo, the sub-fractionation of this aggregates by centrifugation techniques are possible during NP synthesis, but these effects are almost impossible to control in vivo, not to mention to predict. In summary, the aggregation of NPs will add an additional level of complexity, which has to be considered in the description and application of NPs within physiological systems (36, 37, 44).

Figure 1: Timeline of PubMed entries matching the search criteria ‘nanoparticles’ AND ‘corona’ from 2005 to 2014.
Figure 1:

Timeline of PubMed entries matching the search criteria ‘nanoparticles’ AND ‘corona’ from 2005 to 2014.

The term ‘protein corona’ was introduced to the nanoparticle community by the study of Cedervall and colleagues (40), and stimulated a whole field of investigations. Until today, the ‘hard corona’ represents an analytically accessible protein/biomolecule signature of a NP in a certain environment, and this term is often used to describe the long-lived equilibrium state, representing a protein signature of a NP in a certain environment (36, 37, 45, 46). On top of this ‘hard corona’ some models suggest the existence of a ‘soft corona’, a putative more loosely associated and rapidly exchanging layer of biomolecules (37, 40, 46–48). However, since this ‘soft corona’ desorbs during current purification processes, its existence and (patho)biological relevance remains vage. Hence, as inspection of the current literature revealed that the term ‘soft’ versus ‘hard’ corona seems to mostly create more confusion than helping to describe and resolve scientific questions, we suggest to generally refer to analytically accessible NP-protein complexes simply as the ‘protein corona’.

Factors influencing NM-corona composition, -fate, and -toxicity

The biomolecular corona primarily interacts with biological systems and thereby constitutes a major element of the NP’ biological identity affecting multiple molecular-scale interactions (36, 37, 46, 49, 50). A significant difference in the bio-physical properties between such corona-covered NPs and those of the formulated pristine particle during manufacturing is observed (17, 36, 37, 46, 51, 52). Ccientist and eventually regulators therefore consider (bio)molecule-coated NMs as novel materials with different properties compared to the pristine nanomaterials during production (17, 36). The NP-protein complexes significantly influence the the particles biodistribution in patients as well as the subsequent biological responses of the body, potentially contributing not only to favorable nanomedical reactions but also to unwanted (patho)biological side-effects (17, 36, 53–55). Clearly, for the rational development of NM for any kind of biological or biomedical application but also for the unwanted uptake of NPs it is thus the key to understand the formation and kinetic evolution of the protein corona (37, 38, 56–59). Numerous studies have been conducted to generally dissect and mechanistically understand the biomolecule corona on nanoscale materials, its dependence on the nanoparticles’ physico-chemical properties and its biomedical and/or (patho)biological relevance (17, 36, 44, 53–55). Typically, corona profiles differ significantly from the protein composition of the (biological) fluid investigated (36, 43, 44, 46, 60, 61). Distinct proteins will be either enriched or display only weak affinity for the nanoparticle surface. The determination of the corona by the protein source is also an important factor for the so called ‘personalized protein corona’ (PPC) (62). Therefore, humans with a specific disease may have specific NP coronas, which could be a determinant factor in nano-biomedical science (62–64). As we are still at the beginning of understanding the role of the protein corona for a biomedical application, the issue ‘PPC’ adds an additional level of complexity to the field, which certainly has to be addressed and most importantly confirmed in future studies.

However, the relation between original surface functionality of the NPs and the nature of the corona is far from being trivial and currently still remains impossible to predict in complex physiological environments (36, 43, 44, 46, 55, 60, 61). The different physic-chemical properties, such as material size and surface properties, but also the relative ratio of the physiological fluid to the nanoparticle dispersion and the exposure time play an important role for the composition and evolution of the protein corona, although the underlying physical mechanism are not yet resolved in detail (Table 1, and references therein) (36, 37, 43, 44, 46, 55, 60, 61, 84). Moreover, when NPs move from one biological (micro)environment to another, e.g. from the blood system via different cellular uptake mechanisms into cells (e.g. monocytes or macrophages), a key issue is whether the original corona remains stable or is subjected to substantial changes (44), which adds an additional level of complexity. Untill today it is assumed that after passing through several ‘biological (micro)environments’, the final corona may still contains a fingerprint of its history and keeps a memory of its prior journey through the body (44, 70).

Table 1

Modulators of protein corona signatures.

FactorsReported by
NP size/surface curvature/topology(40, 42, 45, 48, 56, 59, 65–69)
Exposure time(36, 48, 70–75)
NPs’ hydrophobicity(33, 40, 68, 76–79)
NPs’ surface charge(36, 59, 69, 75, 76, 79, 80)
Exposure temperature(64, 81–83)
NPs’ surface functionalization(40, 45, 76, 80)
Relative ratio protein/NP concentration(33, 84)

Though, some studies still tend to suggest of having identified ‘the major NM factor’ controlling protein/biomolecule corona formation. However, as convincingly shown by recent comprehensive studies, none of the above-mentioned factors, such as the NPs’ physicochemical properties or exposure time, alone is able to determine formation and composition of the protein/biomolecule corona (36, 46). In the study by Tenzer and colleagues (36), bioinformatic unsupervised hierarchical cluster analysis revealed that protein corona profiles were correctly grouped according to exposure time as well as the NPs’ physicochemical properties. The relation between the pristine NPs characteristics and the nature of the corona in complex environments is thus far from being trivial (36, 43, 44, 46, 55, 60, 61). Despite the complexity and analytical challenges of the biomolecule corona during its ex situ characterization, researchers are facing additional challenges during its in situ analysis in both, physiological and environmental systems.

However, as recent data demonstrate that the plasma protein corona is surprisingly stable and matures only quantitatively rather than qualitatively (36), one might however hypothesize that the corona may not be subjected to significant changes, even when passing through several ‘biological (micro)environments’, unless processing is performed by enzymatic cellular machineries. Particularly for metal and metal oxide NPs, dissolution processes have been recognized of being essential for the NPs’ fate, biodistribution and also ROS-mediated toxicity (46, 85–89). Here, proteins play central roles not only by forming a protein corona, but also by binding released toxic metal ions, as has been shown for silver NPs in both, physiological and environmental systems (86, 90). Close inspection of the literature indeed reveals that the detailed fate of the original corona, as it travels through membranes and barriers, thereby interacting with the extracellular matrix and various cellular enzymatic machineries is still not resolved in detail, and may differ in various body organs, such as the liver or the brain. By developing nanomaterials for in vitro diagnostic sensors or drug delivery-/cell-targeting vehicles, one has to at least consider the formation and potential impact of the biomolecule corona on the biomedical performance of the product.

As described above, the protein corona is surprisingly stable, however, it is important to clarify the presentation of functional biomolecular motifs of the individual proteins in the corona, as the nanoparticle–biomolecule complex interacts with cells and biological barriers, and thus, engaging with different biological pathways. This question so far, has only been addressed in single-protein models, but needs to be clarified in the future. Kelly et al. have recently presented an approach to map protein binding sites on the biomolecular corona which will help to understand the spatial location of the proteins and their binding sites after binding to NPs (91).

Summarizing our current knowledge, a multi-parameter classifier will be required to generally model and predict nanoparticle-protein/biomolecule interaction profiles in complex physiological and environmental systems in the future.

The blood system – Model for corona formation and toxicological impact

Aiming at therapeutic drug delivery and/or imaging systems, NPs are expected to be most commonly intravenously injected into the blood stream, at which point they immediately begin to interact with a very complex biological milieu. Also, if NPs are capable of overcomming biological barriers, such as the lung, the skin or the gut, they will likewise face the bloodsystems prior to their potential transport to distict organs. Here, a variety of biomolecules such as lipids, sugars and especially proteins will adsorb onto the surface of NPs. Such coronas may change how the body interacts with a NP because the nanoparticle’s size and surface characteristics such as charge or targeting molecules can be altered (36, 44, 55, 60, 92). Indeed, particularly the plasma protein corona has been shown to be highly complex (36). Proteins involved in physiological as well as toxicological relevant processes in the blood system, such as complement activation and coagulation have been identified in the coronas of various NPs (Figure 2) (36, 44, 55, 60, 66, 92, 93). Identified proteins span about three to four orders of magnitude dynamic range, most likely covering most of biologically relevant corona proteins (36). Notably, the respective abundance of all of these proteins was affected by the NPs’ characteristics, such as size and surface functionalization, and plasma exposure time. As NPs’ uptake is also an important determinant for nanopathology and targeted delivery (25, 44, 94), several reports showed that the protein corona has a major impact on the NPs’ cellular uptake. Particularly, the recent study by Walkey and colleagues indicates that indeed distinct protein corona signatures are able to predict cellular uptake of gold and silver NPs (46). Hence, covering NPs with a ‘physiological coating’ can indeed promote or inhibit their interaction with the cellular uptake machinery, whereas the surface charge of the bare NPs appears to be less important (36, 94–96).

Figure 2: NM- and exposure time-dependent adsorption of pathophysilogically relevant human plasma proteins. Profiles of proteins involved in complement activation and coagulation are shown. Exposure time-dependent functional classification of corona proteins on negatively charged silica NPs (SNP). SNP of various sizes and various surface-functionalization were analyzed. Abbreviations: CO8B, Complement component C8 beta chain; CO6, Complement factor C6; CO7, Complement factor C7; CO9, Complement factor C9; CO8A, Complement component C8 alpha chain; C4BPA, C4b binding protein alpha chain; C1S, Complement C1s subcomponent; CFAB, Complement factor B; CO4B, Complement factor C4B; C1R, Complement C1r subcomponent; CO5, Complement C5; CO4A, Complement factor C4A; CFAH, Complement factor H; CO3, Complement C3; HABP2, Hyaluronan binding protein 2; ANT3, Antithrombin III; FIBA, Fibrinogen alpha chain; KLKB1, Kininogen 1; MMRN1, Multimerin 1; FA12, Coagulation factor XII; IC1, Plasma protease C1 inhibitor; HRG, Histidine rich glycoprotein; PLMN, Plasminogen; KNG1, Kininogen 1; THRB, Prothrombin; VWF, von Willebrand factor; FA5, Coagulation factor V; TSP1, Thrombospondin 1.
Figure 2:

NM- and exposure time-dependent adsorption of pathophysilogically relevant human plasma proteins. Profiles of proteins involved in complement activation and coagulation are shown. Exposure time-dependent functional classification of corona proteins on negatively charged silica NPs (SNP). SNP of various sizes and various surface-functionalization were analyzed. Abbreviations: CO8B, Complement component C8 beta chain; CO6, Complement factor C6; CO7, Complement factor C7; CO9, Complement factor C9; CO8A, Complement component C8 alpha chain; C4BPA, C4b binding protein alpha chain; C1S, Complement C1s subcomponent; CFAB, Complement factor B; CO4B, Complement factor C4B; C1R, Complement C1r subcomponent; CO5, Complement C5; CO4A, Complement factor C4A; CFAH, Complement factor H; CO3, Complement C3; HABP2, Hyaluronan binding protein 2; ANT3, Antithrombin III; FIBA, Fibrinogen alpha chain; KLKB1, Kininogen 1; MMRN1, Multimerin 1; FA12, Coagulation factor XII; IC1, Plasma protease C1 inhibitor; HRG, Histidine rich glycoprotein; PLMN, Plasminogen; KNG1, Kininogen 1; THRB, Prothrombin; VWF, von Willebrand factor; FA5, Coagulation factor V; TSP1, Thrombospondin 1.

Consequently, one may use this information to rationally engineer the uptake-properties of NPs by modulating corona fingerprints. Different apolipoproteins have been described to promote transport across the blood-brain barrier (97) and different immunoglobulins and complement factors, known as opsonins, enable uptake into monocytes (30) while dysopsonins like albumin and again apolipoproteins (40) inhibit uptake. However, these findings are mostly based on the prior knowledge of selected proteins based on their biological function in isolation. As also functionalization of NPs with proteins seems not to (completely) prevent corona formation, the complexity of the protein corona with more than hundred different proteins, makes it difficult to predict the impact of individual proteins in vivo (36, 44, 46, 98). Hence, the engineering of modified coronas by depletion or enrichment of protein groups is required as the next step to identify corona components causally involved in (cell type specific) NPs’ uptake. Cleary, obtaining comprehensive quantitative and qualitative protein corona signatures, and ideally the implementation of an international standardized corona profile database resource, will finally allow the bioinformatic analysis and exploitation of signatures to guide a subsequent rational in vitro/in vivo investigation of the potential (adverse) impact of corona proteins in physiological systems (36, 44, 46, 60).

Albeit several studies reported impacts of corona formation on NPs’ exposure of the blood system, most of these effects were described to occur at rather late exposure time points (66, 99, 100). Employing primary human cell models of the blood system, it was though shown that already the early corona formation affected toxicological processes at the nano-bio interface (36). The study showed that although the studied pristine NPs existed only for a short period in the blood system, these were still able to affect vitality of endothelial cells, trigger thrombocyte activation and induced hemolysis (36).

Hence, it was concluded that formation of the biomolecule corona rapidly modulated the NPs’ decoration with bioactive proteins, thereby protecting cellular components of the blood system against NP-induced (patho)biological processes, and in addition also influenced cellular uptake of NPs (36). However, whether these findings are valid for other NP formulations as well as for NM of other shapes and materials remains to be investigated.

Prevention strategies to inhibit corona formation

The protein and most likely also other biomolecular coronas are currently a still unpredictable complex factor, potentially triggering not only desired reactions but also undesired toxicological biological responses (36, 44, 46, 55). Hence, there are currently numerous attempts to chemically (completely) prevent protein adsorption, which also have been reviewed previously (16, 44, 61, 74, 101–103). In this models the NPs are functionalized with certain polymer chains, such as the addition of various polyethylene glycol-based chains (‘PEGylation’) onto the NPs’ surface, which are often referred to be highly ‘biocompatible’, as unspecific interactions with biological components are minimized (74, 101, 102). Under physiological conditions, NPs functionalized with PEG confers colloidal stability caused by interparticular repulsion (74, 101, 102). However, even complex ‘PEGylation’ is unable to completely prevent protein/biomolecule corona formation, albeit the extent of protein adsorption is clearly reduced (16, 44, 61, 74, 101–103). As protein adsorption is reduced, it is assumed that numerous cellular responses are affected, including opsonization by cells of the reticuloendothelial system (RES) (3, 44, 61). Thus, the circulation time in the blood system as well as the biodistribution of NPs may be modulated via PEGylation, although the detailed mechanisms are not yet resolved (66, 99, 100).

An alternative method to the PEGylation is the functionalization of the NP surface with zwitterions (104). As recently discussed, NPs with tunable hydrophobicity were designed which do neither adsorb proteins at moderate levels of serum proteins nor do they form hard coronas at physiological serum concentration. This innovative strategy may lead to new options for analyzing the interaction of NM with biosystems without any interference from protein binding. However, further development of the fabrication of ‘corona-free’ NPs is sufficient (104).

Again, standardized corona profiling combined with data mining and subsequent experimental verification is required to answer these questions. Also, extensive surface functionalization with biomolecules, such as antibodies tailored to achieve specific targeting of cell types and/or organs, or ‘cellular uptake proteins’ does not completely prevent corona formation, albeit such modifications clearly affect corona profiles. In some cases, protein corona formation was even considered a major factor significantly reduced cell-targeting efficacy in vitro as well as in vivo (97, 105–107).

Collectively, although sophisticated surface modifications reduce the adsorption of biomolecules to NPs, association with biomolecules does still occur. To our knowledge, there is no existing nanomaterial functionalization strategy, which will completely prevent the formation of a biomolecule corona in complex environments. The design and synthesis of such nanodevices represents certainly one of the key challenges required not only to finally understand the regulation and impact of the biomolecule corona but also to allow novel nanomaterial applications.

Corona complexity and evolution

Indeed, not only physico-chemical properties of NPs, but also the exposure time of NPs to biological environments as well as additional factors has been reported to affect the protein and most likely other biomolecule coronas. Such in situ transformation of NM may lead to altered biodistribution and directly or in directly impacts the efficacy of desired therapeutic and (patho)physiological responses (36, 38, 44, 105, 108) (Table 1). A current model suggests that particularly at short exposure times the ‘soft’ protein corona is initially formed around the NP, which is highly dynamic and subsequently rather slowly matures to the ‘hard’ corona by significantly changing its composition over time (36, 44, 72). Most studies however focused on corona formation upon prolonged nanomaterial exposure to complex biological environments (36, 44, 46, 72). These studies often failed to recognize that physiological systems are highly dynamic and might need to react instantly to external stimuli. In the blood system, flow velocities are heterogenous, such as high in the ascending aorta or slow within tumors. Also, processes controlling hemostasis and thrombosis need to be triggered within minutes or even seconds (66, 99, 109, 110).

The formation and evolution of protein layers on flat surfaces was first analyzed by Vroman (111), describing a time-dependent composition of the bio-coating, in which highly abundant proteins adsorbing only weakly dominate the early state. These adsorbed proteins are later replaced by less abundant proteins, which however bind with higher affinity, resulting in adsorption and displacement step (75, 111, 112). However, one has to keep in mind that the ‘Vroman-effect’ was demonstrated only for a mixture of a few proteins and is unable to predict protein binding kinetics to NPs in complex mixtures (111, 112). Nevertheless, several models used the ‘Vroman-effect’ to directly explain the evolution of the protein corona around NPs even in complex environments, resulting in the concept of a ‘dynamic protein corona evolution’ (44, 61, 70, 72). However, in complex physiological liquids, such as blood, containing more than two thousand different proteins, a high-resolution time-resolved knowledge of NP-specific protein/biomolecules adsorption is required, as various protein/biomolecules are expected to display increased or reduced binding over time. Indeed, snapshot kinetic proteomic profiling recently demonstrated the existence of complex protein adsorption kinetics (36). As predicted from the ‘Vroman-effect’, protein groups displaying increased or reduced binding over time were observed also in the complexity of human plasma (Figure 3) (36). Interestingly, novel protein binding kinetics, such as ‘cup-‘ or ‘peak’ shaped binding kinetics were observed (Figure 3), which cannot be solely explained by the ‘Vroman-effect’ (36).

Figure 3: Correlation analysis to demonstrate distinct kinetic protein-binding modalities during the temporal evolution of the plasma protein corona. Time-smoothed normalized protein abundance profiles of negatively charged polystyrene NPs. NPs’ coronae were classified into four groups by correlation analysis and relative values were normalized to the maximum amount (set to 1) across all time points for each protein. Protein groups “rise” and “fall” showed increasing or decreasing binding over time, respectively. “Peak2” proteins, display low abundance at the beginning of plasma exposure and at later time points, but higher (peak) abundance at intermediate time points. “Cup” proteins show the opposite behavior, with a high abundance at early and late time points, but low abundance at intermediate time points. A selection of representatives is displayed. t: Plasma exposure time.
Figure 3:

Correlation analysis to demonstrate distinct kinetic protein-binding modalities during the temporal evolution of the plasma protein corona. Time-smoothed normalized protein abundance profiles of negatively charged polystyrene NPs. NPs’ coronae were classified into four groups by correlation analysis and relative values were normalized to the maximum amount (set to 1) across all time points for each protein. Protein groups “rise” and “fall” showed increasing or decreasing binding over time, respectively. “Peak2” proteins, display low abundance at the beginning of plasma exposure and at later time points, but higher (peak) abundance at intermediate time points. “Cup” proteins show the opposite behavior, with a high abundance at early and late time points, but low abundance at intermediate time points. A selection of representatives is displayed. t: Plasma exposure time.

A potential reason why such complex binding kinetics have been unnoticed so far is the fact that most kinetic studies did not employ sensitive quantitative LC-MS-based proteomics. Thus, a protein, which was no ‘longer detectable’ at a certain time point, was classified as being ‘absent’ or ‘disappearing’ in previous studies, thereby contributing to the model of a highly dynamic protein corona (44, 61, 70, 72). Hence, we again want to emphasize that it is highly important to use the highest technological standards and standardized experimental procedures (SOPs) for the determination of protein corona binding kinetics (82), allowing inter-laboratory comparison and model building by further systematic studies. The same study also demonstrated that the plasma protein corona is highly complex, containing over 200 different proteins, and surprisingly established in less than one minute (36). In contrast, previous studies suggested that the protein corona has a rather low complexity, consisting of only a few tens of proteins, even when NPs are introduced into highly complex environments, such as the human blood, and evolved rather slowly (33, 44, 61, 72). Moreover, the study also showed that the corona composition changed almost exclusively quantitatively but not qualitatively over time (36). Previous models however proposed a highly dynamic protein corona, changing significantly in its composition over time due to continuous protein association and dissociation events (44, 61, 72).

Impact of NM-coronas on nano-toxicity

Nanotoxicology was proposed as a new branch of toxicology to address the gaps in knowledge and to specifically address potential adverse health effects caused by NM (19). Due to their small size, NM are characterized by a high surface area to volume ratio, rendering them highly reactive. The latter potentially results in yet unknown toxicity mechanims due to novel interactions of NM with biological systems, including the environment (19). In biological environments, NPs bound with proteins can result in physiological and pathological changes, but the mechanisms and potential NM-specific stress and toxicity pathways remain to be fully elucidated. Several studies in the past have addressed the possible impact of the corona on subsequent cellular responses. Liu et al. could show that in the absence of proteins the level of p-p38 was significantly elevated by the positively charged mesoporous silica NPs (MSNs), whereas negatively charged MSNs resulted in marked reactive oxygen species (ROS) production (113). Surprisingly their study showed that the presence of protein efficiently mitigated the potential nano-hazard (113). Shannahan et al. showed that AgNPs with or without a protein corona were able to induce a concentration-dependent cytoxicity and that all corona-coated AgNPs were found to activate cells by inducing IL-6 mRNA expression (114). Ge et al. demonstrated that the competitive binding of blood proteins on single-wall carbon nanotubes (SWCNT) influenced cellular pathways and resulted in reduced cytotoxicity that depended on the presence of protein adsorption and the highly competitive binding of blood proteins on the SWCNT surface (115). Their study suggested that the binding of various protein types onto the SWCNT surface can elicit different cytotoxic cellular response (115). Hu et al. reported a protein corona-mediated reduction of cytotoxicity of graphene oxide (GO). Here the interaction with different concentrations of serum proteins resulted in a decrease of cytotoxicity, whereas the direct interaction with bare GO nanosheets led to a physical damage of the cell membrane (116).

Although protein corona formation can mitigate nanotoxicity, the immune cell response, e.g. the complement systems, can be activated, depending on the type of corona formation, which may cause inflammation and damage to the host (117). For such a scenario, different groups demonstrated that by modifying the surface functionality and availability of reactive functional groups of NPs, the complement activation was attenuated (118, 119). Although NP-protein coronas in most cases appear to reduce cytotoxicity, immunotoxicity can be mitigated or activated depending on the type of NM and/or adsorbed biomolecules.

Systematic strategies to dissect the impact of NM-coronas on nano-toxicity

Determination of qualitative and quantitative (kinetic) corona profiles is a prime and critical step in order to understand the biological and toxicological impact of the nanoparticle corona on living systems. However, these data have to be complemented by biological assay systems of varying complexity. Albeit in traditional toxicology a ‘top down’ approach is often prefered, we feel that a rational hypothesis and knowledge-driven ‘bottom up’ strategy is needed to mechanistically understand (toxicologically) relevant processes at the nano-bio interface. Hence, we here introduce a suitable tiered systems biology-driven approach.

TIER 1: The routes by which NPs may enter the human body, and potentially elicit (adverse) effects, are understood to include inhalation, injection, ingestion and permeation through (diseased) skin. Conventional cell culture models representing these targets were developed almost a century ago and have demonstrated a significant value in biomedical research and safety testing of chemicals. Thus, simple or highly advanced in vitro models mimicking major exposure and application routes of NPs in mammals need to be used for identifying and dissecting basic structure activity relationships (SARs).

Currently, cell models are exposed to pristine or corona-covered NPs often in low throughput applications. Assays are performed either by pre-coating the NPs with biomolecules or by performing the experiments in biomolecule containing liquids, such as plasma- or FCS-containing cell culture medium. Albeit such studies provided basic insights into corona-mediated (toxic) effects, their low throughput and lack of standardization make inter-laboratory comparisons often difficult, sometimes even leading to contradictory results (3, 44, 46).

To overcome these problems, high-throughput screening (HTS)/high-content screening (HCS) experimental systems should be used. Cell based HCS has evolved dramatically, allowing HTS applications to measure the responses of cells to chemicals and, as recently described, also to NPs (36, 85, 120–122). The concept and technological execution of such assay systems have been reviewed (120, 121, 123), albeit the impact of the biomolecule corona was not specifically addressed so far.

The workflows used predominantly for such HCS/HTS screening approaches mostly focused on classical plate reader-based assays using biochemical readouts (120, 121, 123). However, the ability to automate the capture and analysis of fluorescent images of thousands of cells has made fluorescence microscopy an additional tool of systematic cell biology, applicable also for investigating nanobiology. Importantly, HTS assays can be automatically performed in microtiter plates, and thus, the analysis is highly economical regarding cells and NPs, and can be executed under stable SOPs.

As an example for such a HTS workflow to investigate the impact of the plasma protein corona on cell-death, the results are shown in Figure 4. By using a HTS-fluorescence microscopy imaging platform (108), we established a dual-color fluorescence cell vitality assay. By employing fluorescent probes that recognize cell viability by measuring intracellular esterase activity (calcein-AM; green) as well as plasma membrane integrity (ethidium homodimer-1/EthD-1; red), the assay allows for the simultaneous quantitation of live and dead cells.

Figure 4: HTS-quantification of the silica NP-corona’s on toxicity. Cell viability was determined by the ratio of the average calcein (living cells; green) to ethidium homodimer-1 (dead cells; red) signals. Only living cells are able to convert the virtually non-fluorescent cell-permeable calcein-AM to the intensely fluorescent calcein, resulting in an intense uniform green fluorescence of living cells. EthD-1 is however excluded by the intact plasma membrane of living cells, and only enters cells with damaged membranes. Here, it undergoes a 40 fold enhancement of fluorescence upon binding to nucleic acids, thereby producing a bright red fluorescence characteristic for dead cells. Colon epithelial CaCo-2 cells were exposed to AmorSil30 NPs suspended in DMEM (open rectangles) or DMEM containing 10% human plasma (filled rectangles) for 4 h and analyzed by automated microscopy using the Cellomics ArrayScan® VTI. A minimum of 500 cells were analyzed per well, and each treatment was done in triplicate. Whereas the ratio of living to dead cells remained almost unchanged for the untreated Ctrl. as well as after the treatment with 0.6 μg/mL or 6 μg/mL AmorSil30 in the presence and absence of proteins, incubation with 60 μg/mL or 600 μg/mL in absence of proteins led to a significant decrease of this ratio, indicative of cell death. Columns, mean; bars, ± SD from three independent experiments.
Figure 4:

HTS-quantification of the silica NP-corona’s on toxicity. Cell viability was determined by the ratio of the average calcein (living cells; green) to ethidium homodimer-1 (dead cells; red) signals. Only living cells are able to convert the virtually non-fluorescent cell-permeable calcein-AM to the intensely fluorescent calcein, resulting in an intense uniform green fluorescence of living cells. EthD-1 is however excluded by the intact plasma membrane of living cells, and only enters cells with damaged membranes. Here, it undergoes a 40 fold enhancement of fluorescence upon binding to nucleic acids, thereby producing a bright red fluorescence characteristic for dead cells. Colon epithelial CaCo-2 cells were exposed to AmorSil30 NPs suspended in DMEM (open rectangles) or DMEM containing 10% human plasma (filled rectangles) for 4 h and analyzed by automated microscopy using the Cellomics ArrayScan® VTI. A minimum of 500 cells were analyzed per well, and each treatment was done in triplicate. Whereas the ratio of living to dead cells remained almost unchanged for the untreated Ctrl. as well as after the treatment with 0.6 μg/mL or 6 μg/mL AmorSil30 in the presence and absence of proteins, incubation with 60 μg/mL or 600 μg/mL in absence of proteins led to a significant decrease of this ratio, indicative of cell death. Columns, mean; bars, ± SD from three independent experiments.

Combining systematic analytical and cell-based technologies with the advances in bioinformatics is accepted as a powerful approach to rationally dissect and understand cause-effect relationships of NPs in living systems. Thus, analytical and experimental data delivered by proteomic and in vitro HTS profiling experiments, can be collected in a data repository. This data repository can be used for the identification of corona-dependent structure activity relationships (CoroNanoSARs) linking protein coronas with NPs’ characteristics and biological and toxicological effects. Importantly, recent developments intend to combine and to mine data obtained by methodological divergent ‘omics-technologies’, such as proteomics, transcriptomics, metabolomics, cellomics and epigenomics, already allow theoretically to build a multi classifier score for each type of NP under investigation (124, 125). Moreover, biological pathway exploitation will provide rational information, which type of NP affects corona-dependent cellular responses for the subsequent identification of key corona proteins causally involved in the observed (adverse) biological effects.

Nevertheless, the design and execution of such HTS assays is demanding, concerning technology platforms, experimental workflow and experience in data acquisition and handling.

TIER 2: Conventional 2D-cell monocultures are ideal for HCS/HTS approaches and already provided first important information regarding NanoSARs and/or pathways of toxicity (PoTs) (36, 85, 108, Nel, 2012 #4332, 121, 122). Nevertheless, the relevance of such results needs to be subsequently confirmed by using cell models mimicking more closely the in vivo exposure situation. Limitations of cell monocultures lay in the differentiated functions of many cell types in communication with their natural neighboring (micro)environment and thus, in the accurate prediction of in vivo tissue function. Shortcomings of 2D setup are the lack of mimicking the complex three-dimensional (3D) (micro)environment, in which the cells and an extracellular matrix exist in an organized structure. In this regard, efforts were shifted toward developing multiple 3D culture systems that can better recapitulate in vivo tissue functions. Compared with 2D cell cultures, 3D models are expected to achieve better capturing of signaling pathways and response to NPs (126). Hence, in vitro co-culture systems of higher complexity are valuable tools to verify the relevance of mechanistic insights obtained from more simple experimental systems and to further evaluate the effects of NPs on organs and human health, allowing cell-to-cell communications, inter-/intra-signaling regulations, as well as inter-/intra-NPs trafficking absent in monocultures (127, 128). Albeit such systems have not been applied so far to analyze the impact of the biomolecule corona on PoTs, these systems are in principle suitable, albeit only with low-throughput. In a transwell insert, different cell types, like endothelial, epithelial and immune cells can be cultivated in the same well allowing cellular communication via soluble second messengers (127, 128). The communication between endothelial cells and cells that have the direct contact with pristine or corona-covered NPs may play an important role in the systemic effects of NPs and thus, may result in a more realistic judgement regarding the impact of the protein corona. In addition, recent technological developments allow to develop ‘organ-on-a-chip’ systems to move further towards mimicking complete tissues (129). Albeit being far from trivial, such models have a high potential to unravel involved toxicity mechanisms. Also, they will serve as a valuable tool for the rational planning of subsequent in vivo studies, and also allow to avoid unnecessary animal experiments as requested by the 3R rule.

TIER 3: Similar to assessing the toxicity of ‘simple’ chemicals, animal experiments will ultimately be needed to verify the in vivo relevance of corona-nano-structure-activity mechanisms and also to predicte their relevance for human exposure and/or biomedical applications (17, 19, 31, 130). To date, various animal models are in principle available to investigate and dissect the relevance of the NPs’ biomolecule corona for human health. These are ranging from rodents to fish and other organisms, and have been reviewed before (17, 19, 31, 131–135). Notably, recent developments also applied successfully HTS/HCS approaches for whole animals, such as zebrafish (121, 136). However, comprehensive reports on studies focusing particularly on the biomolecule corona are missing so far. Clearly, the researchers are facing various experimental and analytical challenges, when investigating not only ‘simple’ endpoints such as vitality, but when analyzing the corona-mediated fate and biotransformation of NPs in vivo. Here, the recovery of NPs from various organs under ‘corona-preserving experimental conditions’ will require the development of novel extraction protocols and imaging techniques.

Conclusion and outlook

The biomolecule corona is far from being an already resolved topic in basic as well as in applied nanoscience, including nanotoxicology. Even regulatory agencies begin to accept that corona signatures might be relevant for (improved) NM’ risk assessment and prediction. Although the relationship between nanomaterial design and physiological responses has been studied intensely for almost two decades, only some general principles have emerged. Fortunately, the biomolecule corona is now no longer ignored as an ‘unknown factor’, but acknowledged as a potential but yet fully unexploited opportunity to understand and predict the impact of NPs on physiological environments. Also, the corona complexity increases once we are considering additional ‘small inhabitants’ of the human body, i.d., the human microbiome. Such interactions between NM, bacteria, and cells as well as the role of biomolecule coronas have not been addressed so far.

Currently, we cannot yet predict how the synthetic identity of a nanomaterial influences the structure, composition and evolution of the protein corona. However, presenting a cell or an organ with a NP, we do understand that it does not see the bare NP, but the particle with an entire surrounding biomolecule corona profile. Still unresolved are the questions which corona proteins are involved in the cell recognition, uptake and toxicity, and which mechanism governing the interaction of the NM, pristine or corona-covered, with cells. Furthermore, it has to be resolved whether every corona protein or only a certain subset is accessible to and/or biologically active in the physiological and toxicologically relevant environment. Without such a detailed knowledge, a rational design of NMs interacting only with defined corona proteins and cells in a controlled and predictable way remains challenging.

It is now accepted that biomolecule coronas are established rapidly, and formation most likely occurs to varying degrees for all nano-sized materials in general. Hence, the (long term) existence of pristine NPs in complex physiological environments appears to be rather an exception to the rule. Corona profiles are highly complex, consisting of a variety of adsorbed proteins potentially capable of modulating biological responses, and are ex situ rather stable. Nevertheless, long-term effects and the relevance of coronas on the fate and transformation of NPs deposited in vivo, such as organs, require future attention. Quantitative, high-resolution LC-MS/MS is now capable to dramatically reduce biomolecule corona characterization time and can provide quantitative data from large libraries of NM. In order to comprehensively analyze corona profiles and to mechanistically understand the coronas’ biological/toxicological impact, a tiered multidisciplinary approach is clearly mandatory/requisite, including not only comprehensive analytical methods but also the involvement of high throughput/high content ‘omics’ technologies together with bioinformatic data mining to reach the next level. Despite the achievements of the past years, the establishment of nanostructure activity-relationships linking NP/corona properties to physiological or toxicological responses remains still a distant goal. However, such knowledge is ultimately needed not only to understand and minimize nanotoxicity but also to develop NM allowing an improved and safe application of nanotechnology in general.


Corresponding author: Dominic Docter, Department of Nanobiomedicine, ENT, University Medical Center of Mainz, Langenbeckstr. 1, 55101 Mainz, Germany, E-mail:

About the authors

Dana Westmeier

Dana Westmeier received her Master’s degree of Biomedicine from Johannes-Gutenberg University Mainz in 2013. She is currently a PhD candidate at the Institute for Molecular and Cellular Oncology/Nanobiomedicine at the University Medical Center Mainz with Prof Dr. Roland H. Stauber. Her research is focused on the (patho) biological influence of nanoparticles on human cells and pathogens as well as the impact of the NP corona on the microbiome-host interaction.

Chunying Chen

Chunying Chen is a principal investigator at CAS Key Laboratory for Biomedical Effects of Nanomaterials and Nanosafety, National Center for Nanoscience and Technology of China. She received her Bachelor’s degree in Chemistry (1991) and PhD degree in Biomedical Engineering (1996) from Huazhong University of Science and Technology of China. Her research interests include the interaction of nanoparticles with biological systems, therapies for malignant tumors using theranostic nanomedicine systems and vaccine nanoadjuvants using nanomaterials, which are supported by the China MOST 973 Programs, EU-FP6 and FP7 and IAEA. She was awarded China Outstanding Young Female Scientists and the National Science Fund for Distinguished Young Scholars.

Roland H. Stauber

Roland H. Stauber is currently a Professor at the Molecular and Cellular Oncology/Nanobiomedicine department at the ENT Medical University Mainz. His research interests focus on the understanding of molecular mechanisms at the nano-bio interface, the development of nano- and microtechnology for cancer treatment and diagnosis, as well as drug development. He has received the Alexander Karl Award for Cancer Research and a President’s Fellowship of the Chinese Academy of Sciences. Dr. Stauber received his PhD degree from Würzburg University in 1994 and did his post-doctoral training at the National Cancer Institute (USA).

Dominic Docter

Dominic Docter is currently a junior group leader in the Department of Nanobiomedicine at the ENT University Medical Center of Mainz and supported by the Peter and Traudl Engelhorn Foundation. He received his Diploma degree in Biology in 2008 and his PhD degree at the Johannes Gutenberg University Mainz in 2014 after participating in the “Schwerpunktprogramm: Biological response to Nanoscale Particles” of the German Research Foundation. He has received the NMFZ-research award from the University Medical Center of Mainz. His research is focused on understanding events at the host-nano-bio interface to optimize and explore possible nano-diagnostic and -therapeutic applications.

Acknowledgments

Grant support: BMBF-MRCyte/NanoBEL/DENANA, Zeiss-ChemBioMed, Stiftung Rheinland-Pfalz (NanoScreen), Peter und Traudl Engelhorn Foundation, Fonds der chemischen Industrie. We apologize to all colleagues whose work could not be cited due to space limitations.

References

1. Reese M. Nanotechnology: using co-regulation to bring regulation of modern technologies into the 21st century. Health Matrix 2013;23:537–72.Search in Google Scholar

2. Webster TJ. Interview: Nanomedicine: past, present and future. Nanomed: Nanotechnol, Biol Med 2013;8:525–9.10.2217/nnm.13.37Search in Google Scholar PubMed

3. Docter D, Strieth S, Westmeier D, Hayden O, Gao M, Knauer SK, et al. No king without a crown – impact of the nanomaterial-protein corona on nanobiomedicine. Nanomedicine (Lond) 2015;10:503–19.10.2217/nnm.14.184Search in Google Scholar PubMed

4. Rauscher H, Sokull-Kluttgen B, Stamm H. The European Commission’s recommendation on the definition of nanomaterial makes an impact. Nanotoxicology 2013;7:1195–7.10.3109/17435390.2012.724724Search in Google Scholar PubMed

5. Pautler M, Brenner S. Nanomedicine: promises and challenges for the future of public health. Int J Nanomed 2012;5:803–9.Search in Google Scholar

6. Ranganathan R, Madanmohan S, Kesavan A, Baskar G, Krishnamoorthy YR, Santosham R, et al. Nanomedicine: towards development of patient-friendly drug-delivery systems for oncological applications. Int J Nanomed 2012;7:1043–60.Search in Google Scholar

7. Ventola CL. The nanomedicine revolution: part 3: regulatory and safety challenges. P & T: A Peer-Rev J Formulary Manage 2012;37:631–9.Search in Google Scholar

8. Rizzo LY, Theek B, Storm G, Kiessling F, Lammers T. Recent progress in nanomedicine: therapeutic, diagnostic and theranostic applications. Curr Opin Biotech 2013;24:1159–66.10.1016/j.copbio.2013.02.020Search in Google Scholar PubMed PubMed Central

9. Seeney C, Ojwang JO, Weiss RD, Klostergaard J. Magnetically vectored platforms for the targeted delivery of therapeutics to tumors: history and current status. Nanomed: Nanotechnol, Biol Med 2012;7:289–99.10.2217/nnm.11.183Search in Google Scholar PubMed

10. Prabhu P, Patravale V. The upcoming field of theranostic nanomedicine: an overview. J Biomed Nanotechnol 2012;8: 859–82.10.1166/jbn.2012.1459Search in Google Scholar PubMed

11. Murday JS, Siegel RW, Stein J, Wright JF. Translational nanomedicine: status assessment and opportunities. Nanomed: Nanotechnol, Biol Med 2009;5:251–73.10.1016/j.nano.2009.06.001Search in Google Scholar PubMed

12. Ferrari M. Nanogeometry: beyond drug delivery. Nat Nanotechnol 2008;3:131–2.10.1038/nnano.2008.46Search in Google Scholar PubMed

13. Qian X, Peng XH, Ansari DO, Yin-Goen Q, Chen GZ, Shin DM, et al. In vivo tumor targeting and spectroscopic detection with surface-enhanced Raman nanoparticle tags. Nat Biotechnol 2008;26:83–90.10.1038/nbt1377Search in Google Scholar PubMed

14. Ventola CL. The nanomedicine revolution: part 2: current and future clinical applications. P & T: A Peer-Rev J Formulary Manage 2012;37:582–91.Search in Google Scholar

15. Hou L, Zhu D, Wang X, Wang L, Zhang C, Chen W. Adsorption of phenanthrene, 2-naphthol, and 1-naphthylamine to colloidal oxidized multiwalled carbon nanotubes: effects of humic acid and surfactant modification. Environ Toxicol Chem/SETAC 2013;32:493–500.10.1002/etc.2088Search in Google Scholar PubMed

16. Hu FQ, Wei L, Zhou Z, Ran YL, Li Z, Gao MY. Preparation of biocompatible magnetite nanocrystals for in-vivo magnetic resonance detection of cancer. Adv Mater 2006;18:2553–6.10.1002/adma.200600385Search in Google Scholar

17. Nystrom AM, Fadeel B. Safety assessment of nanomaterials: implications for nanomedicine. J Control Release: Off J Controll Release Soc 2012;161:403–8.10.1016/j.jconrel.2012.01.027Search in Google Scholar PubMed

18. Bawa R. Regulating nanomedicine-can the FDA handle it? Curr Drug Del 2011;8:227–34.10.2174/156720111795256156Search in Google Scholar PubMed

19. Oberdorster G. Safety assessment for nanotechnology and nanomedicine: concepts of nanotoxicology. J Intern Med 2010;267:89–105.10.1111/j.1365-2796.2009.02187.xSearch in Google Scholar PubMed

20. Baun A, Hansen SF. Environmental challenges for nanomedicine. Nanomed: Nanotechnol, Biol Med 2008;3:605–8.10.2217/17435889.3.5.605Search in Google Scholar PubMed

21. Rahman M, Ahmad MZ, Kazmi I, Akhter S, Afzal M, Gupta G, et al. Emergence of nanomedicine as cancer targeted magic bullets: recent development and need to address the toxicity apprehension. Curr Drug Discov Technol 2012;9:319–29.10.2174/157016312803305898Search in Google Scholar PubMed

22. Zhu M, Nie G, Meng H, Xia T, Nel A, Zhao Y. Physicochemical properties determine nanomaterial cellular uptake, transport, and fate. Accounts Chem Res 2012;46:622–31.10.1021/ar300031ySearch in Google Scholar PubMed PubMed Central

23. Manke A, Wang L, Rojanasakul Y. Mechanisms of nanoparticle-induced oxidative stress and toxicity. BioMed Res Int 2013;2013:942916.10.1155/2013/942916Search in Google Scholar PubMed PubMed Central

24. AshaRani PV, Low Kah Mun G, Hande MP, Valiyaveettil S. Cytotoxicity and genotoxicity of silver nanoparticles in human cells. ACS Nano 2009;3:279–90.10.1021/nn800596wSearch in Google Scholar PubMed

25. Lunov O, Syrovets T, Rocker C, Tron K, Nienhaus GU, Rasche V, et al. Lysosomal degradation of the carboxydextran shell of coated superparamagnetic iron oxide nanoparticles and the fate of professional phagocytes. Biomaterials 2010;31:9015–22.10.1016/j.biomaterials.2010.08.003Search in Google Scholar PubMed

26. Johnston HJ, Hutchison G, Christensen FM, Peters S, Hankin S, Stone V. A review of the in vivo and in vitro toxicity of silver and gold particulates: particle attributes and biological mechanisms responsible for the observed toxicity. Crit Rev Toxicol 2010;40:328–46.10.3109/10408440903453074Search in Google Scholar PubMed

27. Ju-Nam Y, Lead JR. Manufactured nanoparticles: an overview of their chemistry, interactions and potential environmental implications. Sci Tot Environm 2008;400:396–414.10.1016/j.scitotenv.2008.06.042Search in Google Scholar PubMed

28. Li N, Xia T, Nel AE. The role of oxidative stress in ambient particulate matter-induced lung diseases and its implications in the toxicity of engineered nanoparticles. Free Radic Biol Med 2008;44:1689–99.10.1016/j.freeradbiomed.2008.01.028Search in Google Scholar PubMed PubMed Central

29. Stone V, Johnston H, Clift MJ. Air pollution, ultrafine and nanoparticle toxicology: cellular and molecular interactions. IEEE Trans Nanobiosci 2007;6:331–40.10.1109/TNB.2007.909005Search in Google Scholar

30. Riehemann K, Schneider SW, Luger TA, Godin B, Ferrari M, Fuchs H. Nanomedicine–challenge and perspectives. Angew Chem Int Ed Engl 2009;48:872–97.10.1002/anie.200802585Search in Google Scholar PubMed PubMed Central

31. Oberdorster G. Nanotoxicology: in vitro-in vivo dosimetry. Environ Health Persp 2012;120:A13; author reply A.10.1289/ehp.1104320Search in Google Scholar PubMed PubMed Central

32. Li J, Chen YC, Tseng YC, Mozumdar S, Huang L. Biodegradable calcium phosphate nanoparticle with lipid coating for systemic siRNA delivery. J Controll Release: Official J Controll Release Soc 2010;142:416–21.10.1016/j.jconrel.2009.11.008Search in Google Scholar PubMed PubMed Central

33. Monopoli MP, Walczyk D, Campbell A, Elia G, Lynch I, Bombelli FB, et al. Physical-chemical aspects of protein corona: relevance to in vitro and in vivo biological impacts of nanoparticles. J Am Chem Soc 2011;133:2525–34.10.1021/ja107583hSearch in Google Scholar PubMed

34. Jeong SK, Kwon MS, Lee EY, Lee HJ, Cho SY, Kim H, et al. BiomarkerDigger: a versatile disease proteome database and analysis platform for the identification of plasma cancer biomarkers. Proteomics 2009;9:3729–40.10.1002/pmic.200800593Search in Google Scholar PubMed

35. Anderson NL. The clinical plasma proteome: a survey of clinical assays for proteins in plasma and serum. Clin Chem 2011;56:177–85.10.1373/clinchem.2009.126706Search in Google Scholar PubMed

36. Tenzer S, Docter D, Kuharev J, Musyanovych A, Fetz V, Hecht R, et al. Rapid formation of plasma protein corona critically affects nanoparticle pathophysiology. Nat Nanotechnol 2013;8:772–81.10.1038/nnano.2013.181Search in Google Scholar PubMed

37. Monopoli MP, Bombelli FB, Dawson KA. Nanobiotechnology: nanoparticle coronas take shape. Nat Nanotechnol 2011;6: 11–2.10.1038/nnano.2010.267Search in Google Scholar

38. Nel AE, Madler L, Velegol D, Xia T, Hoek EM, Somasundaran P, et al. Understanding biophysicochemical interactions at the nano-bio interface. Nat Mater 2009;8:543–57.10.1038/nmat2442Search in Google Scholar PubMed

39. Klein J. Probing the interactions of proteins and nanoparticles. Proc Natl Acad Sci USA 2007;104:2029–30.10.1073/pnas.0611610104Search in Google Scholar PubMed PubMed Central

40. Cedervall T, Lynch I, Lindman S, Berggård T, Thulin E, Nilsson H, et al. Understanding the nanoparticle-protein corona using methods to quantify exchange rates and affinities of proteins for nanoparticles. Proc Natl Acad Sci USA 2007;104:2050–5.10.1073/pnas.0608582104Search in Google Scholar PubMed PubMed Central

41. Treuel L, Malissek M. Interactions of nanoparticles with proteins: determination of equilibrium constants. Method Mole Biol 2013;991:225–35.10.1007/978-1-62703-336-7_21Search in Google Scholar PubMed

42. Lacerda SH, Park JJ, Meuse C, Pristinski D, Becker ML, Karim A, et al. Interaction of gold nanoparticles with common human blood proteins. ACS Nano 2010;4:365–79.10.1021/nn9011187Search in Google Scholar PubMed

43. Gebauer JS, Malissek M, Simon S, Knauer SK, Maskos M, Stauber RH, et al. Impact of the nanoparticle-protein corona on colloidal stability and protein structure. Langmuir: ACS J Surface Colloids 2012;28:9673–9.10.1021/la301104aSearch in Google Scholar PubMed

44. Monopoli MP, Aberg C, Salvati A, Dawson KA. Biomolecular coronas provide the biological identity of nanosized materials. Nat Nanotechnol 2012;7:779–86.10.1038/nnano.2012.207Search in Google Scholar PubMed

45. Lundqvist M, Stigler J, Elia G, Lynch I, Cedervall T, Dawson KA. Nanoparticle size and surface properties determine the protein corona with possible implications for biological impacts. Proc Natl Acad Sci USA 2008;105:14265–70.10.1073/pnas.0805135105Search in Google Scholar PubMed PubMed Central

46. Walkey CD, Olsen JB, Song F, Liu R, Guo H, Olsen DW, et al. Protein corona fingerprinting predicts the cellular interaction of gold and silver nanoparticles. ACS Nano 2014;8:2439–55.10.1021/nn406018qSearch in Google Scholar PubMed

47. Lynch I, Cedervall T, Lundqvist M, Cabaleiro-Lago C, Linse S, Dawson KA. The nanoparticle-protein complex as a biological entity; a complex fluids and surface science challenge for the 21st century. Adv Colloid Interface Sci 2007.10.1016/j.cis.2007.04.021Search in Google Scholar PubMed

48. Walczyk D, Bombelli FB, Monopoli MP, Lynch I, Dawson KA. What the cell “sees” in bionanoscience. J Am Chem Soc 2010;132:5761–8.10.1021/ja910675vSearch in Google Scholar PubMed

49. Treuel L, Brandholt S, Maffre P, Wiegele S, Shang L, Nienhaus GU. Impact of protein modification on the protein corona on nanoparticles and nanoparticle-cell interactions. ACS Nano 2014;8:503–13.10.1021/nn405019vSearch in Google Scholar PubMed

50. Treuel L, Eslahian KA, Docter D, Lang T, Zellner R, Nienhaus K, et al. Physicochemical characterization of nanoparticles and their behavior in the biological environment. Phys Chem Chem Phys 2014;16:15053–67.10.1039/C4CP00058GSearch in Google Scholar

51. Toke ER, Lorincz O, Somogyi E, Lisziewicz J. Rational development of a stable liquid formulation for nanomedicine products. Int J Pharmaceutics 2010;392:261–7.10.1016/j.ijpharm.2010.03.048Search in Google Scholar PubMed

52. Wolfram J, Yang Y, Shen J, Moten A, Chen C, Shen H, et al. The nano-plasma interface: Implications of the protein corona. Colloid surface 2014;124:17–24.10.1016/j.colsurfb.2014.02.035Search in Google Scholar PubMed PubMed Central

53. Leszczynski J. Bionanoscience: Nano meets bio at the interface. Nat Nanotechnol 2010;5:633–4.10.1038/nnano.2010.182Search in Google Scholar PubMed

54. Adiseshaiah PP, Hall JB, McNeil SE. Nanomaterial standards for efficacy and toxicity assessment. Wiley Interdiscip Rev Nanomed Nanobiotechnol 2010;2:99–112.10.1002/wnan.66Search in Google Scholar PubMed

55. Dobrovolskaia MA, Neun BW, Man S, Ye X, Hansen M, Patri AK, et al. Protein corona composition does not accurately predict hematocompatibility of colloidal gold nanoparticles. Nanomed: Nanotechnol, Biol Med, 2014;10:1453–63.10.1016/j.nano.2014.01.009Search in Google Scholar PubMed PubMed Central

56. Tenzer S, Docter D, Rosfa S, Wlodarski A, Kuharev J, Rekik A, et al. Nanoparticle size is a critical physicochemical determinant of the human blood plasma corona: a comprehensive quantitative proteomic analysis. ACS Nano 2011;5:7155–67.10.1021/nn201950eSearch in Google Scholar PubMed

57. Thomas CR, George S, Horst AM, Ji Z, Miller RJ, Peralta-Videa JR, et al. Nanomaterials in the environment: from materials to high-throughput screening to organisms. ACS Nano 2011;5:13–20.10.1021/nn1034857Search in Google Scholar PubMed

58. Xia XR, Monteiro-Riviere NA, Riviere JE. An index for characterization of nanomaterials in biological systems. Nat Nanotechnol 2010;5:671–5.10.1038/nnano.2010.164Search in Google Scholar PubMed

59. Zhang H, Burnum KE, Luna ML, Petritis BO, Kim JS, Qian WJ, et al. Quantitative proteomics analysis of adsorbed plasma proteins classifies nanoparticles with different surface properties and size. Proteomics 2011;11:4569–77.10.1002/pmic.201100037Search in Google Scholar PubMed PubMed Central

60. Walkey CD, Olsen JB, Guo H, Emili A, Chan WC. Nanoparticle size and surface chemistry determine serum protein adsorption and macrophage uptake. J Am Chem Soc 2012;134:2139–47.10.1021/ja2084338Search in Google Scholar PubMed

61. Walkey CD, Chan WC. Understanding and controlling the interaction of nanomaterials with proteins in a physiological environment. Chem Soc Rev 2012;41:2780–99.10.1039/C1CS15233ESearch in Google Scholar PubMed

62. Hajipour MJ, Laurent S, Aghaie A, Rezaee F, Mahmoudi M. Personalized protein coronas: a “key” factor at the nanobiointerface. Biomater Sci 2014;2:1210–21.10.1039/C4BM00131ASearch in Google Scholar

63. Caracciolo G, Pozzi D, Capriotti AL, Cavaliere C, Piovesana S, Amenitsch H, et al. Lipid composition: a “key factor” for the rational manipulation of the liposome-protein corona by liposome design. RSC Adv 2015;5:5967–75.10.1039/C4RA13335HSearch in Google Scholar

64. Caracciolo G. Liposome-protein corona in a physiological environment: Challenges and opportunities for targeted delivery of nanomedicines. Nanomed: Nanotechnol, Biol Med 2015;11:543–57.10.1016/j.nano.2014.11.003Search in Google Scholar PubMed

65. Chakraborty S, Joshi P, Shanker V, Ansari ZA, Singh SP, Chakrabarti P. Contrasting effect of gold nanoparticles and nanorods with different surface modifications on the structure and activity of bovine serum albumin. Langmuir: ACS J Surface Colloids 2011;27:7722–31.10.1021/la200787tSearch in Google Scholar PubMed

66. Dobrovolskaia MA, Patri AK, Zheng J, Clogston JD, Ayub N, Aggarwal P, et al. Interaction of colloidal gold nanoparticles with human blood: effects on particle size and analysis of plasma protein binding profiles. Nanomed: Nanotechnol, Biol Med 2009;5:106–17.10.1016/j.nano.2008.08.001Search in Google Scholar PubMed PubMed Central

67. Dutta D, Sundaram SK, Teeguarden JG, Riley BJ, Fifield LS, Jacobs JM, et al. Adsorbed proteins influence the biological activity and molecular targeting of nanomaterials. Toxicol Sci: An Off J Soc Toxicol 2007;100:303–15.10.1093/toxsci/kfm217Search in Google Scholar PubMed

68. Lindman S, Lynch I, Thulin E, Nilsson H, Dawson KA, Linse S. Systematic investigation of the thermodynamics of HSA adsorption to N-iso-propylacrylamide/N-tert-butylacrylamide copolymer nanoparticles. Effects of particle size and hydrophobicity. Nano Lett 2007;7:914–20.10.1021/nl062743+Search in Google Scholar PubMed

69. Mahmoudi M, Sant S, Wang B, Laurent S, Sen T. Superparamagnetic iron oxide nanoparticles (SPIONs): development, surface modification and applications in chemotherapy. Adv Drug Deliv Rev 2011;63:24–46.10.1016/j.addr.2010.05.006Search in Google Scholar PubMed

70. Lundqvist M, Stigler J, Cedervall T, Berggård T, Flanagan MB, Lynch I, et al. The evolution of the protein corona around nanoparticles: a test study. ACS Nano 2011;5:7503–9.10.1021/nn202458gSearch in Google Scholar PubMed

71. Barran-Berdon AL, Pozzi D, Caracciolo G, Capriotti AL, Caruso G, Cavaliere C, et al. Time evolution of nanoparticle-protein corona in human plasma: relevance for targeted drug delivery. Langmuir: The ACS J Surfaces and Colloids 2013;29:6485–94.10.1021/la401192xSearch in Google Scholar PubMed

72. Casals E, Pfaller T, Duschl A, Oostingh GJ, Puntes V. Time evolution of the nanoparticle protein corona. ACS Nano 2010;4:3623–32.10.1021/nn901372tSearch in Google Scholar PubMed

73. Dell’Orco D, Lundqvist M, Oslakovic C, Cedervall T, Linse S. Modeling the time evolution of the nanoparticle-protein corona in a body fluid. PLoS One 2010;5:e10949.10.1371/journal.pone.0010949Search in Google Scholar PubMed PubMed Central

74. Natte K, Friedrich JF, Wohlrab S, Lutzki J, von Klitzing R, Österle W, et al. Impact of polymer shell on the formation and time evolution of nanoparticle-protein corona. Colloids Surfaces B, Biointerfaces 2013;104:213–20.10.1016/j.colsurfb.2012.11.019Search in Google Scholar PubMed

75. Rocker C, Potzl M, Zhang F, Parak WJ, Nienhaus GU. A quantitative fluorescence study of protein monolayer formation on colloidal nanoparticles. Nat Nanotechnol 2009;4:577–80.10.1038/nnano.2009.195Search in Google Scholar PubMed

76. Ehrenberg MS, Friedman AE, Finkelstein JN, Oberdorster G, McGrath JL. The influence of protein adsorption on nanoparticle association with cultured endothelial cells. Biomaterials 2009;30:603–10.10.1016/j.biomaterials.2008.09.050Search in Google Scholar

77. Goppert TM, Muller RH. Protein adsorption patterns on poloxamer- and poloxamine-stabilized solid lipid nanoparticles (SLN). Eur J Pharm Biopharm 2005;60:361–72.10.1016/j.ejpb.2005.02.006Search in Google Scholar

78. Labarre D, Vauthier C, Chauvierre C, Petri B, Muller R, Chehimi MM. Interactions of blood proteins with poly(isobutylcyanoacrylate) nanoparticles decorated with a polysaccharidic brush. Biomaterials 2005;26:5075–84.10.1016/j.biomaterials.2005.01.019Search in Google Scholar

79. Owens DE, 3rd, Peppas NA. Opsonization, biodistribution, and pharmacokinetics of polymeric nanoparticles. Int J Pharm 2006;307:93–102.10.1016/j.ijpharm.2005.10.010Search in Google Scholar

80. Gessner A, Lieske A, Paulke B, Muller R. Influence of surface charge density on protein adsorption on polymeric nanoparticles: analysis by two-dimensional electrophoresis. Eur J Pharm Biopharm 2002;54:165–70.10.1016/S0939-6411(02)00081-4Search in Google Scholar

81. Mahmoudi M, Abdelmonem AM, Behzadi S, Clement JH, Dutz S, Ejtehadi MR, et al. Temperature: the “ignored” factor at the NanoBio interface. ACS Nano 2013;7:6555–62.10.1021/nn305337cSearch in Google Scholar PubMed

82. Docter D, Distler U, Storck W, Kuharev J, Wünsch D, Hahlbrock A, et al. Quantitative profiling of the protein coronas that form around nanoparticles. Nat Protocols 2014;9:2030–44.10.1038/nprot.2014.139Search in Google Scholar PubMed

83. Mahmoudi M, Lohse SE, Murphy CJ, Fathizadeh A, Montazeri A, Suslick KS. Variation of protein corona composition of gold nanoparticles following plasmonic heating. Nano Lett 2014;14:6–12.10.1021/nl403419eSearch in Google Scholar PubMed

84. Caracciolo G, Pozzi D, Capriotti AL, Cavaliere C, Foglia P, Amenitsch H, et al. Evolution of the protein corona of lipid gene vectors as a function of plasma concentration. Langmuir: ACS J Surfaces Colloids 2011;27:15048–53.10.1021/la202912fSearch in Google Scholar PubMed

85. Zhang H, Pokhrel S, Ji Z, Meng H, Wang X, Lin S, et al. PdO doping tunes band-gap energy levels as well as oxidative stress responses to a Co(3)O(4) p-type semiconductor in cells and the lung. J Am Chem Soc 2014;136:6406–20.10.1021/ja501699eSearch in Google Scholar PubMed PubMed Central

86. Wang X, Ji Z, Chang CH, Zhang H, Wang M, Liao Y-P, et al. Use of coated silver nanoparticles to understand the relationship of particle dissolution and bioavailability to cell and lung toxicological potential. Small 2014;10:385–98.10.1002/smll.201301597Search in Google Scholar PubMed PubMed Central

87. Gilbert B, Fakra SC, Xia T, Pokhrel S, Madler L, Nel AE. The fate of ZnO nanoparticles administered to human bronchial epithelial cells. ACS Nano 2012;6:4921–30.10.1021/nn300425aSearch in Google Scholar PubMed PubMed Central

88. Konduru N, Keller J, Ma-Hock L, Gröters S, Landsiedel R, Donaghey TC, et al. Biokinetics and effects of barium sulfate nanoparticles. Particle and Fibre Toxicology 2014;11:55.10.1186/s12989-014-0055-3Search in Google Scholar PubMed PubMed Central

89. Li R, Ji Z, Chang CH, Dunphy DR, Cai X, Meng H, Zhang H, et al. Surface interactions with compartmentalized cellular phosphates explain rare earth oxide nanoparticle hazard and provide opportunities for safer design. ACS Nano 2014;8:1771–83.10.1021/nn406166nSearch in Google Scholar PubMed PubMed Central

90. Anderson JW, Semprini L, Radniecki TS. Influence of water hardness on silver ion and silver nanoparticle fate and toxicity toward. Environ Eng Sci 2014;31:403–9.10.1089/ees.2013.0426Search in Google Scholar PubMed PubMed Central

91. Kelly PM, Aberg C, Polo E, O’Connell A, Cookman J, Fallon J, et al. Mapping protein binding sites on the biomolecular corona of nanoparticles. Nat Nanotechnol 2015;10:472–9.10.1038/nnano.2015.47Search in Google Scholar PubMed

92. Barkam S, Saraf S, Seal S. Fabricated micro-nano devices for in vivo and in vitro biomedical applications. Wiley Interdisciplin Rev Nanomed Nanobiotechnol 2013;5:544–68.10.1002/wnan.1236Search in Google Scholar PubMed

93. Moghimi SM, Farhangrazi ZS. Nanomedicine and the complement paradigm. Nanomed: Nanotechnol, Biol Med 2013;9:458–60.10.1016/j.nano.2013.02.011Search in Google Scholar PubMed

94. Lesniak A, Fenaroli F, Monopoli MP, Aberg C, Dawson KA, Salvati A. Effects of the presence or absence of a protein corona on silica nanoparticle uptake and impact on cells. ACS Nano 2012;6:5845–57.10.1021/nn300223wSearch in Google Scholar PubMed

95. Huhn D, Kantner K, Geidel C, Brandholt S, De Cock I, Soenen SJ, et al. Polymer-coated nanoparticles interacting with proteins and cells: focusing on the sign of the net charge. ACS Nano 2013;7:3253–63.10.1021/nn3059295Search in Google Scholar PubMed

96. Lesniak A, Salvati A, Santos-Martinez MJ, Radomski MW, Dawson KA, Aberg C. Nanoparticle adhesion to the cell membrane and its effect on nanoparticle uptake efficiency. J Am Chem Soc 2013;135:1438–44.10.1021/ja309812zSearch in Google Scholar PubMed

97. Meister S, Zlatev I, Stab J, Docter D, Baches S, Stauber RH, et al. Nanoparticulate flurbiprofen reduces amyloid-beta42 generation in an in vitro blood-brain barrier model. Alz Res Therapy 2013;5:51.10.1186/alzrt225Search in Google Scholar PubMed PubMed Central

98. Zhu M, Nie G, Meng H, Xia T, Nel A, Zhao Y. Physicochemical properties determine nanomaterial cellular uptake, transport, and fate. Accounts Chem Res 2013;46:622–31.10.1021/ar300031ySearch in Google Scholar PubMed PubMed Central

99. Aggarwal P, Hall JB, McLeland CB, Dobrovolskaia MA, McNeil SE. Nanoparticle interaction with plasma proteins as it relates to particle biodistribution, biocompatibility and therapeutic efficacy. Adv Drug Deliver Rev 2009;61:428–37.10.1016/j.addr.2009.03.009Search in Google Scholar PubMed PubMed Central

100. Dobrovolskaia MA, Germolec DR, Weaver JL. Evaluation of nanoparticle immunotoxicity. Nat Nanotechnol 2009;4:411–4.10.1038/nnano.2009.175Search in Google Scholar PubMed

101. Sacchetti C, Motamedchaboki K, Magrini A, Palmieri G, Mattei M, Bernardini S, et al. Surface polyethylene glycol conformation influences the protein corona of polyethylene glycol-modified single-walled carbon nanotubes: potential implications on biological performance. ACS Nano 2013;7:1974–89.10.1021/nn400409hSearch in Google Scholar PubMed

102. Murthy AK, Stover RJ, Borwankar AU, Nie GD, Gourisankar S, Truskett TM, et al. Equilibrium gold nanoclusters quenched with biodegradable polymers. ACS Nano 2013;7:239–51.10.1021/nn303937kSearch in Google Scholar PubMed PubMed Central

103. Pozzi D, Colapicchioni V, Caracciolo G, Piovesana S, Capriotti AL, Palchetti S, et al. Effect of polyethyleneglycol (PEG) chain length on the bio-nano-interactions between PEGylated lipid nanoparticles and biological fluids: from nanostructure to uptake in cancer cells. Nanoscale 2014;6:2782–92.10.1039/c3nr05559kSearch in Google Scholar PubMed

104. Moyano DF, Saha K, Prakash G, Yan B, Kong H, Yazdani M, et al. Fabrication of corona-free nanoparticles with tunable hydrophobicity. ACS Nano 2014;8:6748–55.10.1021/nn5006478Search in Google Scholar PubMed PubMed Central

105. Salvati A, Pitek AS, Monopoli MP, Prapainop K, Bombelli FB, Hristov DR, et al. Transferrin-functionalized nanoparticles lose their targeting capabilities when a biomolecule corona adsorbs on the surface. Nat Nanotechnol 2013;8:137–43.10.1038/nnano.2012.237Search in Google Scholar PubMed

106. von Maltzahn G, Park JH, Lin KY, Singh N, Schwöppe C, Mesters R, et al. Nanoparticles that communicate in vivo to amplify tumour targeting. Nat Mater 2012;10:545–52.10.1038/nmat3049Search in Google Scholar PubMed PubMed Central

107. Helou M, Reisbeck M, Tedde SF, Richter L, Bär L, Bosch JJ, et al. Time-of-flight magnetic flow cytometry in whole blood with integrated sample preparation. Lab on a chip 2013;13:1035–8.10.1039/c3lc41310aSearch in Google Scholar PubMed

108. Docter D, Bantz C, Westmeier D, Galla HJ, Wang Q, Kirkpatrick JC, et al. The protein-corona protects against size- and dose-dependent toxicity of amorphous silica nanoparticles. Beilstein J Nanotechnol 2014;5:1380–92.10.3762/bjnano.5.151Search in Google Scholar PubMed PubMed Central

109. Strieth S, Nussbaum CF, Eichhorn ME, Fuhrmann M, Teifel M, Michaelis U, et al. Tumor-selective vessel occlusions by platelets after vascular targeting chemotherapy using paclitaxel encapsulated in cationic liposomes. Int J Cancer 2008;122:452–60.10.1002/ijc.23088Search in Google Scholar PubMed

110. Dobrovolskaia MA, McNeil SE. Immunological properties of engineered nanomaterials. Nat Nanotechnol 2007;2:469–78.10.1038/nnano.2007.223Search in Google Scholar PubMed

111. Vroman L. Effect of absorbed proteins on the wettability of hydrophilic and hydrophobic solids. Nature 1962;196:476–7.10.1038/196476a0Search in Google Scholar PubMed

112. Vogler EA. Protein adsorption in three dimensions. Biomaterials 2012;33:1201–37.10.1016/j.biomaterials.2011.10.059Search in Google Scholar PubMed PubMed Central

113. Liu TP, Wu SH, Chen YP, Chou CM, Chen CT. Biosafety evaluations of well-dispersed mesoporous silica nanoparticles: towards in vivo-relevant conditions. Nanoscale 2015;7:6471–80.10.1039/C4NR07421ASearch in Google Scholar

114. Shannahan JH, Podila R, Aldossari AA, Emerson H, Powell BA, Ke PC, et al. Formation of a protein corona on silver nanoparticles mediates cellular toxicity via scavenger receptors. Toxicol Sci: An Off J Soc Toxicol 2015;143:136–46.10.1093/toxsci/kfu217Search in Google Scholar PubMed PubMed Central

115. Ge C, Du J, Zhao L, Wang L, Liu Y, Li D, et al. Binding of blood proteins to carbon nanotubes reduces cytotoxicity. Proc Natl Acad Sci USA 2011;108:16968–73.10.1073/pnas.1105270108Search in Google Scholar PubMed PubMed Central

116. Hu W, Peng C, Lv M, Li X, Zhang Y, Chen N, et al. Protein corona-mediated mitigation of cytotoxicity of graphene oxide. ACS Nano 2011;5:3693–700.10.1021/nn200021jSearch in Google Scholar PubMed

117. Sim RB, Wallis R. Surface properties: Immune attack on nanoparticles. Nat Nanotechnol 2011;6:80–1.10.1038/nnano.2011.4Search in Google Scholar PubMed

118. Hulander M, Lundgren A, Berglin M, Ohrlander M, Lausmaa J, Elwing H. Immune complement activation is attenuated by surface nanotopography. Int J Nanomed 2011;6:2653–66.10.2147/IJN.S24578Search in Google Scholar PubMed PubMed Central

119. Bertholon I, Vauthier C, Labarre D. Complement activation by core-shell poly(isobutylcyanoacrylate)-polysaccharide nanoparticles: influences of surface morphology, length, and type of polysaccharide. Pharmaceut Res 2006;23:1313–23.10.1007/s11095-006-0069-0Search in Google Scholar PubMed

120. Nel A, Xia T, Meng H, Wang X, Lin S, Ji Z, et al. Nanomaterial toxicity testing in the 21st century: use of a predictive toxicological approach and high-throughput screening. Accounts Chem Res 2012;46:607–21.10.1021/ar300022hSearch in Google Scholar PubMed PubMed Central

121. Damoiseaux R, George S, Li M, Pokhrel S, Ji Z, France B, et al. No time to lose – high throughput screening to assess nanomaterial safety. Nanoscale 2012;3:1345–60.10.1039/c0nr00618aSearch in Google Scholar PubMed PubMed Central

122. Roach P, Farrar D, Perry CC. Surface tailoring for controlled protein adsorption: effect of topography at the nanometer scale and chemistry. J Am Chem Soc 2006;128:3939–45.10.1021/ja056278eSearch in Google Scholar PubMed

123. Patel T, Telesca D, Rallo R, George S, Xia T, Nel AE. Hierarchical rank aggregation with applications to nanotoxicology. J Agric Biol Environ Stat 2013;18:159–77.10.1007/s13253-013-0129-ySearch in Google Scholar PubMed PubMed Central

124. Liu R, Zhang HY, Ji ZX, Rallo R, Xia T, Chang CH, et al. Development of structure-activity relationship for metal oxide nanoparticles. Nanoscale 2013;5:5644–53.10.1039/c3nr01533eSearch in Google Scholar PubMed

125. Liu R, Rallo R, Weissleder R, Tassa C, Shaw S, Cohen Y. Nano-SAR development for bioactivity of nanoparticles with considerations of decision boundaries. Small 2013;9:1842–52.10.1002/smll.201201903Search in Google Scholar PubMed

126. Lancaster MA, Renner M, Martin CA, Wenzel D, Bicknell LS, Hurles ME, et al. Cerebral organoids model human brain development and microcephaly. Nature 2013;501:373–9.10.1038/nature12517Search in Google Scholar PubMed PubMed Central

127. Kasper J, Hermanns MI, Bantz C, Maskos M, Stauber R, Pohl C, et al. Inflammatory and cytotoxic responses of an alveolar-capillary coculture model to silica nanoparticles: comparison with conventional monocultures. Particle Fibre Toxicol 2011;8:6.10.1186/1743-8977-8-6Search in Google Scholar PubMed PubMed Central

128. Lehmann AD, Daum N, Bur M, Lehr CM, Gehr P, Rothen-Rutishauser BM. An in vitro triple cell co-culture model with primary cells mimicking the human alveolar epithelial barrier. Eur J Pharm Biopharm 2011;77:398–406.10.1016/j.ejpb.2010.10.014Search in Google Scholar PubMed

129. Nguyen TA, Yin TI, Reyes D, Urban GA. Microfluidic chip with integrated electrical cell-impedance sensing for monitoring single cancer cell migration in three-dimensional matrixes. Anal Chem 2013;85:11068–76.10.1021/ac402761sSearch in Google Scholar PubMed

130. Nel A, Zhao Y, Madler L. Environmental health and safety considerations for nanotechnology. Accounts Chem Res 2013;46:605–6.10.1021/ar400005vSearch in Google Scholar PubMed

131. Lin S, Zhao Y, Nel AE. Zebrafish: an in vivo model for nano EHS studies. Small 2013;9:1608–18.10.1002/smll.201202115Search in Google Scholar PubMed PubMed Central

132. Gomes SI, Soares AM, Scott-Fordsmand JJ, Amorim MJ. Mechanisms of response to silver nanoparticles on Enchytraeus albidus (Oligochaeta): survival, reproduction and gene expression profile. J Hazard Mater 2013;254–255:336–44.10.1016/j.jhazmat.2013.04.005Search in Google Scholar PubMed

133. Zhang Y, Bai Y, Jia J, Gao N, Li Y, Zhang R, et al. Perturbation of physiological systems by nanoparticles. Chem Soc Rev 2014;43:3762–809.10.1039/C3CS60338ESearch in Google Scholar PubMed

134. Fernandez MD, Alonso-Blazquez MN, Garcia-Gomez C, Babin M. Evaluation of zinc oxide nanoparticle toxicity in sludge products applied to agricultural soil using multispecies soil systems. Sci Tot Environm 2014;497–498:688–96.10.1016/j.scitotenv.2014.07.085Search in Google Scholar PubMed

135. Sakulkhu U, Maurizi L, Mahmoudi M, Motazacker M, Vries M, Gramoun A, et al. Ex situ evaluation of the composition of protein corona of intravenously injected superparamagnetic nanoparticles in rats. Nanoscale 2014;6:11439–50.10.1039/C4NR02793KSearch in Google Scholar PubMed

136. Lin S, Zhao Y, Ji Z, Ear J, Chang CH, Zhang H, et al. Zebrafish high-throughput screening to study the impact of dissolvable metal oxide nanoparticles on the hatching enzyme, ZHE1. Small 2013;9:1776–85.10.1002/smll.201202128Search in Google Scholar PubMed PubMed Central

Received: 2015-4-1
Accepted: 2015-5-19
Published Online: 2015-6-6
Published in Print: 2015-6-1

©2015 by De Gruyter

Downloaded on 23.4.2024 from https://www.degruyter.com/document/doi/10.1515/ejnm-2015-0018/html
Scroll to top button