Paper 1

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 22

Biochimica et Biophysica Acta 1858 (2016) 1688–1709

Contents lists available at ScienceDirect

Biochimica et Biophysica Acta

journal homepage: www.elsevier.com/locate/bbamem

Computational and experimental approaches for investigating


nanoparticle-based drug delivery systems☆
M. Ramezanpour a, S.S.W. Leung b, K.H. Delgado-Magnero a, B.Y.M. Bashe c, J. Thewalt b,c, D.P. Tieleman a
a
Centre for Molecular Simulation, Department of Biological Sciences, University of Calgary, Calgary, AB T2N 1N4, Canada
b
Department of Physics, Simon Fraser University, Burnaby, BC V5A 1S6, Canada
c
Department of Molecular Biology and Biochemistry, Simon Fraser University, Burnaby, BC V5A 1S6, Canada

a r t i c l e i n f o a b s t r a c t

Article history: Most therapeutic agents suffer from poor solubility, rapid clearance from the blood stream, a lack of targeting,
Received 23 January 2016 and often poor translocation ability across cell membranes. Drug/gene delivery systems (DDSs) are capable of
Received in revised form 20 February 2016 overcoming some of these barriers to enhance delivery of drugs to their right place of action, e.g. inside cancer
Accepted 23 February 2016
cells. In this review, we focus on nanoparticles as DDSs. Complementary experimental and computational studies
Available online 27 February 2016
have enhanced our understanding of the mechanism of action of nanocarriers and their underlying interactions
Keywords:
with drugs, biomembranes and other biological molecules. We review key biophysical aspects of DDSs and dis-
Drug delivery system cuss how computer modeling can assist in rational design of DDSs with improved and optimized properties.
Gene nanocarrier We summarize commonly used experimental techniques for the study of DDSs. Then we review computational
Nanoparticle studies for several major categories of nanocarriers, including dendrimers and dendrons, polymer-, peptide-,
Biological membrane nucleic acid-, lipid-, and carbon-based DDSs, and gold nanoparticles. This article is part of a Special Issue entitled:
Computer modeling Membrane Proteins edited by J.C. Gumbart and Sergei Noskov.
Experimental techniques © 2016 Elsevier B.V. All rights reserved.

1. Introduction [1,2]. Practically, an ideal DDS is cheap and straightforward to make,


as well as stable prior to its administration. Intense research into DDS
In medicine, the delivery of a drug can be as important as the drug design is providing increasingly effective disease treatment.
itself. Physiology poses key challenges to effective drug delivery; an ad- A central aim for specialized DDS development is the optimization of
ministered drug must penetrate obstacles such as endo- or epithelial tiny drug encapsulation vehicles known as nanoparticles (NPs). NPs typ-
membranes and also survive the host's defenses in order to be effective. ically have diameters in the range of 10 to 100 nm [3]. These small DDS
Addressing such challenges requires some form of drug encapsulation, can circulate freely even in capillaries [4], and are intrinsically better at
and the entity forming the capsule, which has a defined molecular ar- traversing biological barriers than larger DDS. It is worth mentioning
chitecture, is known as a “drug delivery system (DDS).” From a func- that nanomedicines, however, have a more complicated and time con-
tional point of view, an ideal DDS is easy to administer, non-toxic, suming FDA approval process than parent unimolecular therapeutics
carries the drug to its desired destination, and then releases it (Fig. 1) [5]. In addition to efficacy and side effect evaluations common between
Abbreviations: 5-FU, 5-fluorouracil; AuNPs, gold nanoparticles; BSA, bovine serum
both small molecules and nanomedicines, there are other concerns
albumin; CD, cyclodextrin; CG, coarse-grained; CNP, carbon-based nanoparticle; CNT, about nanocarriers which need further evaluation. Nanocarrier aggrega-
carbon nanotube; CPe, inverse-phosphatidylcholine; CPNT, cyclic peptide based nano- tion in vivo and their drug release, as well as evaluation of each compo-
tube; CPP, cell penetrating peptide; DDS, drug delivery system; DLS, dynamic light scat- nent of the formulation are some of those concerns. FDA regulations for
tering; DOX, doxorubicin; DPD, Dissipative Particle Dynamics; DPPC,
nanomedicines, the complexity of nanocarriers, and the prospect of
dipalmitoylphosphatidylcholine; DSC, differential scanning calorimetry; ESR, electron
spin resonance spectroscopy; FT-IR, Fourier transform-infrared spectroscopy; only a small increase in performance upon reformulating small drugs in
gamma-PGA, poly(gamma-glutamic acid); HSA, human serum albumin; IDP, intrinsi- nanocarriers can discourage pharmaceutical companies from investing
cally disordered proteins; IFN, interferon alpha-1b; IR, infrared; ITC, isothermal titra- in these systems. In fact, it seems that investigations into DDS are much
tion calorimetry; LD, laser diffraction; LNP, lipid nanoparticle; MD, Molecular more successful in generating papers than in developing new treatment
Dynamics; NMR, nuclear magnetic resonance spectroscopy; NP, nanoparticle; p-
methods [5]. However, considering the number of approved drugs and
THPP, 5,10,15,20-tetrakis(4-hydroxyphenyl)porphyrin; PAMAM, poly(amido amine);
PEI, polyethylenimine; PPI, poly(propylene imine); QD, quantum dot; SAXS, small current ones in clinical trials [6,7], as well as clearer guidelines for
angle X-ray scattering; SCPs, star copolymers; SEM, scanning electron microscopy; nanomedicine approval by the FDA, there is growing momentum in
SSMs, sterically stabilized micelles; TAT, trans-activator of transcription; TEM, trans- transferring these systems from publication to clinical development [5].
mission electron microscopy; VIP, vasoactive intestinal peptide; WAXS, wide angle
Many different types of NPs exist, including liposomes [7,8],
X-ray scattering; XRD, X-ray diffraction.
☆ This article is part of a Special Issue entitled: Membrane Proteins edited by J.C. dendrimers [9–11], carbon nanotubes (CNTs) [12–14], inorganic
Gumbart and Sergei Noskov. [15–20], and polymer-based [21,22] NPs, with each having its own

http://dx.doi.org/10.1016/j.bbamem.2016.02.028
0005-2736/© 2016 Elsevier B.V. All rights reserved.
M. Ramezanpour et al. / Biochimica et Biophysica Acta 1858 (2016) 1688–1709 1689

Fig. 1. Drug delivery pathway for a drug-nanocarrier system. After the DDSs enter the bloodstream (1), drug (green circle) leakage might occur if the drug–DDS complex is unstable (2).
Depending on the surface chemistry, DDSs might interact with biomolecules (3) and ions (4) in circulation. Eventually, they will extravasate to the extracellular space (5). After reaching
the target cell, they interact with the cell membrane (6) and can be internalized either via direct penetration (7) or endocytosis (8). In the latter case, the nanocarrier will be entrapped in
endosomes and must be released (9) before endosomal maturation. Drug–DDS dissociation (10) is necessary for drugs to be effective in the cytoplasm or nucleus (11). This dissociation can
be mediated by interactions with endogeneous biomolecules (12). Not drawn to scale.

unique properties (Fig. 2). Physicochemical properties including size, common computational approaches. Section 4 highlights recent com-
shape, deformability, surface charge and chemical composition affect putational studies of what currently are the major classes of NP-based
how well they evade phagocytosis and how they interact with vascula- DDSs. We conclude with a brief outlook on the role of computer simula-
ture, traverse cell membranes and escape from endosomes prior to drug tions in drug delivery technology.
degradation [1]. Thus, physical characterization of NPs is an important
step in the DDS design process. 2. Commonly used experimental techniques for nanoparticle
Despite the increasing sophistication of experimental efforts to mea- characterization
sure, design and optimize the structure and dynamics of NPs, such re-
search inevitably faces intrinsic and practical limitations. Resolution of Here we provide a brief summary of techniques regularly used to
structural details can be difficult or impossible, and systematic variation characterize NP size, zeta potential, surface morphology, and the exis-
of properties such as composition, size and surface charge can be pro- tence of colloidal structures (Table 2). We first describe the importance
hibitively time consuming and expensive. As a result, general biophysi- of each of these properties and the experimental techniques used to
cal principles connecting the effectiveness of a DDS with its composition measure them. Next we describe some sophisticated physical chemistry
are difficult to elucidate. Predicting optimum DDS design solely on the tools that are used in DDS development to provide data that can param-
basis of experimental research is therefore unlikely. Another challenge eterize and validate computational simulations. Experiments used to vi-
in DDS development is that many DDSs show promise in vitro but fail sualize cellular uptake, NP–cell interactions, as well as those used to
in vivo [6]. This is mainly because of the lack of mechanistic insight determine cell viability will be omitted, even though they are vital to
obtainable by experiments which are based on trial and error. Theoret- pharmaceutical development and are also commonly performed, since
ical methods, both analytical and computational, can allow primary details on these can be found elsewhere (e.g. [25]).
screening of variables in order to predict suitable conditions for further Particle size is an important property because the size of the carrier
experiments [23]. Computer modeling techniques provide detailed in- can affect circulation time, encapsulation efficiency, and cellular uptake
formation about molecular interactions and other physicochemical [1,26,27]. For example, in the case of cancer treatment, nano-drug car-
properties of drugs and carriers (Table 1) and have found broad applica- riers of the appropriate sizes can extravasate from the bloodstream to
tions in biology, biochemistry, and biophysics [24]. Computer simula- tumor tissues (380–780 nm) [1,28], but not from the tighter capillaries
tion is capable of complementing experiments and assisting in the (50–200 nm) in healthy tissues [29]. This contributes to the well-known
rational design of new formulations with improved efficacies. In this re- enhanced permeability and retention effect [30–33]. Light scattering
view, we describe the main experimental approaches to characterize techniques such as laser diffraction (LD) and dynamic light scattering
NP-based DDSs and focus in more detail on current applications of com- (DLS) can be used to measure NP size [34]. In both of these experiments,
puter simulations aimed at understanding molecular aspects of DDSs. a laser beam is directed at a dilute sample (e.g. a suspension or solution
In the next sections, we first give a brief overview of commonly used of NPs). In LD, the diffraction angle is used to determine the particle size
experimental techniques, followed by a section that describes several [35]. In DLS, the time dependence of the scattered light intensity is
1690 M. Ramezanpour et al. / Biochimica et Biophysica Acta 1858 (2016) 1688–1709

Fig. 2. Examples of nanoparticulate DDS. From top-left to bottom-right, monolayer-protected gold nanoparticle, polymeric micelle, DNA cage, carbon nanotube, solid lipid nanoparticle,
polymer, dendrimer and dendron, liposome, and protein-based DDS. For dendrimer: orange, green, red, and blue represents G0, G1, G2, and G3, respectively. Dendron has also been
shown as black branch. Not drawn to scale.

detected at a known scattering angle. The Brownian motion of scattering techniques with a separation technique, such as field flow
suspended particles causes the scattered light intensity to fluctuate. fractionation [37,38].
These fluctuations are correlated with the particle's velocity, and can Microscopy techniques such as transmission electron microscopy
be used to determine particle size via an appropriate theoretical (TEM), scanning electron microscopy (SEM), and atomic force micros-
model (e.g. Stokes–Einstein equation). Most models assume the parti- copy can be used to measure NP size, as well as size distribution
cles are spherical and monodisperse; artifacts may arise if they are (dispersity), shape, and surface morphology. Dispersity – also known
not. DLS can be used to measure the size of particles in the 2 nm–3 μm as polydispersity in older literature – in a formulation can result in vary-
range, and LD can be used in the 20 nm–2 mm range [34,36]. Resolution ing body-residence time and immunogenicity [39,40]. Particle shape
of small differences in particle size can be improved by combining light can affect cellular uptake mechanisms and kinetics [41–45] as well as
margination dynamics — the lateral drift of NPs to endothelial walls
Table 1
[1]. The particle surface is the first point of contact between the NP
Questions relevant to drug delivery system design that can be answered by computational and its environment, and surface morphology is known to affect drug
simulations. release kinetics [46]. In electron microscopy, an electron beam is direct-
ed at the sample, and differences in electron density in the sample gives
Drug delivery step Properties of interest
contrast to structures. Since electron microscopy can give high-
Physicochemical characterization Self-assembly of DDS components
resolution images approaching molecular resolution, all motions on
Structure and dynamics of drug–DDS complex
Drug loading the supramolecular scale have to be halted. For SEM, the NPs are depos-
Drug distribution ited onto a carbon conductive tape, dried, and covered with a thin layer
Drug–DDS interactions of metal [47]. In cryo-TEM, samples are applied onto an electron micros-
Functionalization copy grid, plunge-frozen rapidly in ethane to prevent the formation of
Drug leakage
Aggregation
damaging cubic ice and then imaged under cryogenic conditions to pre-
Circulation Interaction with biomolecules vent radiation damage to the sample during imaging [48]. Magnifica-
Stability tions on the order of 50,000 ×, with a resolution of 0.3 nm, are
Interaction with ions possible [49]. Atomic force microscopy probes surface terrain using a
Cell-membrane Interactions with membrane
nanoscale mechanical probe with a vertical resolution as small as
Cellular internalization
Interaction with membrane proteins 0.01 nm [35].
Intracellular Drug–DDS dissociation A NP's shape can influence the physiological fate of its cargo [1].
Drug release from endosome Particle shape can be an indication of undesirable aggregate structures.
Interaction with endogenous molecules Aggregate structures can also cause inaccurate particle size determina-
Nucleus translocation
tion by LD or DLS [34]. Light microscopy can be used for determining
M. Ramezanpour et al. / Biochimica et Biophysica Acta 1858 (2016) 1688–1709 1691

if microparticles are aggregates of smaller particles [36,50], even though spectroscopy where all frequencies are sampled simultaneously. It has
it cannot be used for directly observing NP structure due to limitations been used to monitor cholesterol-induced changes in liposomal mem-
in optical resolution (~200 nm). Isothermal titration calorimetry (ITC), brane structure [67] and to verify conjugation of block copolymers
which will be described in greater detail below, has been used to onto iron oxide NPs [68]. UV–Vis spectroscopy has been used extensive-
study aggregate formation in polycationic nucleic acid carriers [51]. ly in NP characterization. For example it was used to confirm the pres-
Aggregation of NPs can be prevented via electrostatic repulsion [31, ence of cleavable disulfide bond in CNT-nucleic acid conjugates [69];
52]. Surface charge measurements can be a predictor of aggregation be- monitor the metal to ligand charge transfer band (5d(Pt11) to
havior, and consequently a predictor of colloidal dispersion stability π*(diimine)) in dendrimer synthesis [70]; measure dye and drug encap-
during storage. sulation in poly(lactic-co-glycolic acid) polymeric carriers [71,72] and
Surface charge can affect cellular uptake, cytotoxicity, and circula- study drug release kinetics of liposomal formulations [67].
tion times [1,31,53,54]. Zeta potential is used to characterize the surface NMR and ESR both rely on spin manipulations in the presence of an
charge of NPs. When a particle moves in a liquid, a thin layer of liquid applied external magnetic field. NMR monitors the nuclear spin of
moves with it. The zeta potential is defined as the electrical potential nuclei with non-zero nuclear spins, and ESR monitors the electron
at the boundary of this moving liquid, defined as the shear plane. It is spin of excited unpaired electrons [73]. NMR can provide a wealth of in-
a function of surface charge density, shear plane location, and surface formation on molecular arrangements. NMR is sensitive to small chang-
structure [55]. Zeta potential is determined by measuring light scatter- es in the local environment: it can be used to differentiate between
ing caused by particle motion in an applied electric field. Charged parti- various layers up to the fourth generation in dendrimers [74], and be-
cles with larger zeta potential (|zeta| N 25 mV) are less likely to form tween neighboring carbon atoms in lipids [75]. 2D NMR techniques
aggregates but neutral particles generally have longer circulation can be used to ascertain chemical connectivity (e.g. COSY, HMQC,
times [56]. Surface charge effects are very complex; positive and nega- HSQC) and distances (e.g. NOESY and ROESY) between specific atoms
tive zeta potentials correlate with higher cellular uptake in non- [74,76,77]. Dynamic information is also accessible to NMR. Pulsed
phagocytic and phagocytic cells, respectively [54]. Future DDS designs field-gradient spin echo 1H NMR and diffusion-ordered spectroscopy
may require NPs to switch zeta potential at the target site to maximize have been used to determine diffusion coefficients in aliphatic polyester
circulation time while targeting a particular type of cell [1]. It should and poly(propylene imine) (PPI) dendrimers [74,78]. 31P NMR has been
be noted that protein adsorption in cellular media can change the used to determine the mobility of siRNA encapsulated in LNPs [79].
surface charge [52]. Conclusions on surface-charge effects, therefore, NMR relaxation was used to monitor molecular tumbling activities
are only valid when comparing functionalized or non-functionalized [80] and this information, in turn, was used to determine the location
particles of similar sizes [54]. of a spin-labeled anticancer drug in micelles formed by PEG modified
Encapsulation efficiency – the ratio of encapsulated drug to the with long alkyl groups and of paclitaxel in cyclodextrin (CD) vesicles
amount of drug used during formulation preparation – is an important [77,80]. Most DDSs are invisible to ESR, requiring the addition of para-
parameter [46]. Low encapsulation efficiency implies that more carrier magnetic spin probes [81]. Lipophilic ESR probes have been used as
material is needed, heightening the risk of carrier material toxicity [2, model lipophilic drugs to help determine where these drugs localized
57]. Encapsulation efficiency can be determined in a number of ways, in solid LNPs [35]. In gadolinium poly(amido amine) (PAMAM) dendri-
depending on the particular type of NP. Drug loading can be determined mer containing spin probes, ESR was used to determine the location and
by weight [16]. Gel electrophoresis has been used to determine encap- concentration of the magnetic resonance imaging contrast agent gado-
sulation efficiency by analyzing protection from RNase degradation in linium [78]. ESR has also been used to prove that the terminal inter-
siRNA lipid nanoparticles (LNPs). Physical chemistry techniques can branch H-bonded groups preclude backfolding in PAMAM dendrimers
also be used to identify and quantify the degree to which components [78]. Since ESR can probe membrane fluidity, ESR has been used to
have been encapsulated. For example, fluorescence spectroscopy can study interactions between polymeric micelles and lipid membranes,
be used if the drug is fluorescent [16]. Ultraviolet–visible (UV–Vis) which are models for cell membranes [68]. ESR can also give
light spectroscopy can be used to determine the amount of cargo (e.g. microviscosity and micropolarity information [81]. Viscoelastic proper-
siRNA, anticancer drug doxorubicin (DOX)) present [58,59]. High per- ties of dispersions can affect DDS interactions with the body. For exam-
formance liquid chromatography can be used to separate, identify, and ple, capillary blockage can occur with solid LNPs, which are less
quantify the components of a mixture [46,60]. Novel single molecule deformable than nano-emulsions [64].
measurements have been applied to measure encapsulation efficiency Scattering techniques can be used to study the structure of colloidal
in individual lipid vesicles [61]. systems. In scattering studies, a monochromatic beam of light, neutrons,
An array of physical chemistry techniques has been instrumental in or X-ray is focused on the sample, and the intensity of the scattered
evaluating the chemical and physical properties of NP components. beam is measured. Light scattering was covered earlier in this section.
Dendrimer structures have been simulated based on experimental X-ray scattering results from variations in electron density and neutron
data from nuclear magnetic resonance spectroscopy (NMR) and X-ray scattering results from variations in the spatial distribution of atomic
[62]. LNP matrix state, polymorphism, and phase behavior have been nuclei [82]. The interference pattern formed by the scattered beam
characterized by differential scanning calorimetry (DSC), X-ray diffrac- can be used to determine characteristic lengths using Bragg's Law. Larg-
tion (XRD), and neutron scattering [63]. Lipid crystallinity of solid er angle measurements contain information for shorter length scales:
LNPs has been studied using X-ray scattering, DSC, NMR, and electron small and wide angle X-ray scattering (SAXS and WAXS) are used to
spin resonance spectroscopy (ESR) [64]. CNTs have been characterized look at liquids and solids with structures on the length scale of 1–
using XRD, UV–Vis spectroscopy, Fourier transform-infrared spectros- 200 nm and sub-nm, respectively. For example, SAXS is used to monitor
copy (FT-IR), ESR, SEM, and energy dispersive X-ray diffraction [65]. monolayer and bilayer repeat spacing, and hydrophobic thickness in
Iron oxide NPs, which are magnetic, can be studied using magnetome- lipid membranes, while WAXS is used to determine chain packing and
try, NMR relaxation dispersion profiles [55], and magnetic field flow extent of lipid motion [83,84]. SAXS has been used to study NPs with
fractionation [66]. Some more common physical chemistry tools will cubosome structures [85,86], to provide micelle shape, size and density
be briefly described here. of nonsteroidal anti-inflammatory drug celecoxib-loaded protein mi-
In optical spectroscopy, the absorption of electromagnetic radiation celles [50], and to observe in situ shell growth of polymeric
by NPs occurs at frequencies characteristic of the molecular structures nanocapsules [87]. WAXS has been used to study lipid polymorphic
present. UV–Vis light is absorbed if it supplies the right amount of ener- transformations in solid LNPs and crystallinity of encapsulated drugs
gy to excite valence electrons, while infrared (IR) light is absorbed if it [88], and crystallinity of polyelectrolyte complexes made with polysac-
excites bond vibrations. FT-IR is a more efficient variant of IR charides [89]. SANS, useful for studying structures of 1–100 nm [90], has
1692 M. Ramezanpour et al. / Biochimica et Biophysica Acta 1858 (2016) 1688–1709

been used to determine molecular weight and end group locations of changes, stoichiometry, and binding constants. ITC has been used to
dendrimers [74]. Neutron scattering can also be used to study drug dif- measure cyclodextran–guest molecule interactions, micelle–drug inter-
fusion and internal molecular motions in LNPs [63]. actions, and polyelectrolyte aggregation [98].
Fluorescence comprises a family of spectroscopy and microscopy NP DDSs are complex systems that can evolve over time. Dynamic
techniques relying on excitation of fluorescent species. An extrinsic processes cannot be captured in steady-state measurements, and can
fluorescent probe is added when no fluorescent moieties are present pose as a challenge to characterization [35]. For example, PEGylated
in the material of interest. Absorption of a photon excites the NPs are known to lose their PEG coating before reaching target cells
fluorophore and fluorescence occurs when a photon is later emitted. Be- [99–101]. Time-resolved DLS combined with TEM was used to inves-
tween excitation and emission, the fluorophore's motions and local en- tigate morphology transition kinetics of multiblock copolymer mi-
vironment (e.g. viscosity, polarity, temperature, pH, ion concentration) celles [102]. Time-resolved fluorescence lifetime measurements
influence the fluorophore's spectrum [91]. Polarity sensitive fluorescent have been used to follow the release of fluorescent compounds
probes are commonly used to determine the critical association concen- from polymeric carriers [71], and fluorescence lifetime imaging mi-
tration of polymeric micelles – analogous to the critical micellar concen- croscopy could extend drug release studies to cellular systems
tration of surfactant micelles [92]. Fluorescence quenching assays have [103,104].
been done to evaluate nucleic acid-polycation binding in polycationic
nucleic acid carriers [51] and dye leakage assays have been used to eval- 3. Computational approaches to study nanoparticles
uate drug release characteristics [93].
Thermal analysis is a versatile tool for DDS characterization [94,95]. Computer simulations in general describe the physics of materials at a
DSC measures enthalpy changes — the heat required for a sample to un- suitable level of detail for a particular application. A simulation is de-
dergo a physical phase transition (e.g. glass transition, melting, decom- scribed by the level of detail in the physics used to model the system of
position, isomerization or heats of solution, water sorption–desorption) interest, and the algorithms used to generate enough simulation data to
[95,96]. DSC was used to monitor the physical state of the antineoplastic draw statistically valid conclusions about the behavior of the system of in-
drug paclitaxel inside polymeric microspheres, for example [46]. More terest. NPs can be described at different levels of detail (Fig. 3), but for
specific to DDS design, DSC has been used to measure the gel-to-liquid most of the applications in this review the appropriate levels of detail
crystalline transition temperature of lipids in liposomal formulations, are limited to atomistic and semi-atomistic, slightly coarser, levels.
which can be used to predict release rates of encapsulated drugs [95, At the highest level of detail, a system is described by quantum me-
97]. In a closely related technique, ITC, the evolved heat is measured chanics in one of its approximations. Quantum mechanical calculations
as concentrated aliquots of one substance are added to a solution of a have been widely applied to CNTs and play a role in optimizing less de-
second substance. The released heat is directly proportional to the tailed simulations, but in practice they are only useful for small systems
amount of substance added, and can be used to measure enthalpy with limited degrees of freedom and generally not in solution. They are

Fig. 3. Categories of simulation methods and their respective spatio-temporal domains of applicability. The question of interest dictates the level of resolution required. Briefly, if electronic
motion play an important role in the property we are studying, quantum mechanics (QM) level is appropriate. However, if there is no bond formation and cleavage, all atom (AA) level
works well; we can safely ignore electronic motion by assigning point charges to each atom. If electrostatic and hydrophobic interactions are the dominant contributors, coarse grain
models (CG) can be used: individual atoms can be ignored and a group of atoms (e.g. 4 heavy atoms in MARTINI) can be treated as one interaction point/bead. CG models allow the
exploration of a larger area in phase space, at the expense of losing atomistic details. Coarser levels of simulations (e.g. DPD and continuum models) are appropriate for systems and
processes which require longer times and lengths.
M. Ramezanpour et al. / Biochimica et Biophysica Acta 1858 (2016) 1688–1709 1693

essential to correctly model electronic properties and can be used to cal- Table 2
culate electrical, optical, magnetic and mechanical properties of CNTs, Experimental methods used to study drug delivery systems.

quantum dots (QDs), magnetic NPs and similar systems. For soft- Experimental technique Abbreviation Properties
condensed matter systems such as liposomes or dendrimers in solution Dynamic light scattering DLS Particle size, aggregation
their usefulness is currently limited. Laser diffraction LD Particle size, aggregation
If we average over electronic properties by representing electrons as Transmission electron TEM Particle size, dispersity, shape, surface
partial charges on atoms rather than taking them into account explicitly, microscopy morphology
Scanning electron SEM Particle size, dispersity, shape, surface
we significantly simplify calculations. At this level of detail, atoms are
microscopy morphology
described as point charges that interact through simplified empirical Atomic force microscopy AFM Particle size, dispersity, shape, surface
potential terms including Coulomb interactions, Van der Waals interac- morphology
tions, and simplified terms describing bonds, angles, and dihedrals as Fluorescence FLIM Drug release
harmonic terms and cosine expansions [105]. The resulting technique Zeta potential – Surface charge
Gel electrophoresis – Encapsulation efficiency of nucleic acids
is called Molecular Dynamics (MD) and is widely used for biological Ultraviolet–visible light UV–vis Encapsulation efficiency
and soft-condensed matter systems. The vast majority of the papers spectroscopy
reviewed below use this approach, which accounts for a substantial Fourier FT-IR Bond vibrations
fraction of super computer time on the worlds' major academic comput- transform-infrared
light spectroscopy
er centers. MD is a mature technique that is implemented in a number
X-ray scattering SAXS and Repeat spacing, chemical connectivity
of widely-used software packages, including GROMACS [106], NAMD WAXS
[107], LAMPPS [108], CHARMM [109], and AMBER [110]. MD simula- Neutron scattering – Crystallinity
tions essentially generate a movie of a set of molecules over time, incor- High performance liquid HPLC Separate, identify and quantify
porating every degree or nearly every degree of freedom, from which chromatography components in mixture
Nuclear magnetic NMR Crystallinity, molecular mobility,
through statistical mechanics thermodynamic properties can be calcu-
resonance chemical connectivity
lated. Its main limitations include the computational cost and the corre- spectroscopy
sponding limited time scale (realistically, microseconds in most cases) Electron spin resonance ESR Carrier fluidity, gadolinium
and length scale (~10 nm per dimension) and the limitations implicit spectroscopy concentration, microviscosity,
micropolarity
in the physics that describes the interactions between atoms: point
Fluorescence – Viscosity, polarity, concentration, drug
charges mean electronic effects like polarization, pi-electron clouds spectroscopy release
possibly important for CNTs, and charge transfer are typically ignored. Differential scanning DSC Phase transitions
In addition, the simplified interaction function sometimes imposes calorimetry
other limitations. Isothermal calorimetry ITC Phase transitions

For many properties of interest, especially in mesoscopic-scale sys-


tems like NPs with a size of 20–200 nm, atomistic detail may not always
be necessary. Many papers below use coarse-grained (CG) simulations, characteristics of the resulting aggregates [117,118], drug loading ca-
often based on the MARTINI model [111]. Here CG means that several pacity, mechanism and rate [119–121], drug distribution/localization
atoms have been grouped into interaction sites, which are no longer rec- in DDS [121], complex stability [118,122], drug retention, release mech-
ognizable as atoms but instead are ‘beads’ that represent molecular anism and release rate [123–125], dominant drug–DDS interactions
fragments. Depending on the level of coarse-graining, these beads can [119,126,127], and to design or optimize DDSs targeting capabilities
still maintain a significant amount of chemical specificity, or they can [128,129]. Environmental conditions, e.g. pH, temperature, salt type
represent very generic properties such as those that represent a lipid and concentration, counterions [126,130], and external stimulus such
by one head group bead and two tail beads. The same physics applies as external magnetic fields [131–133], as well as interactions with
as in atomistic simulations, and sufficient conformations of a system of other biomolecules (e.g. serum proteins, miRNA, heparin), all might af-
interest have to be generated by a sampling algorithm to accurately fect the aforementioned aspects of DDSs [126,134–137] and can be
calculate thermodynamic and structural properties. This can be done studied computationally.
by MD, Monte Carlo, or various other algorithms including Langevin The way DDSs interact with cell membranes is one of the most com-
dynamics or Dissipative Particle Dynamics (DPD) [105]. monly studied steps in drug delivery by computational studies [138,
At the CG level, individual molecules and usually fragments of mol- 139]. Computer modeling can investigate the driving forces for NP-
ecules are still clearly recognizable. In material science, there exists a membrane interactions, as well as how factors such as design parameters
whole hierarchy of additional models. Beyond molecules, DPD solves and environmental conditions affect these. These parameters include size,
hydrodynamics equations for materials by representing materials as shape, surface chemistry, and concentration of NPs, as well as mechanical
discrete particles that can be much larger than individual molecules. and elastic properties of both NPs and membranes [138–143]. Surface
DPD blurs the line with CG MD a little by its choice of particle volume: chemistry refers to NPs hydrophobicity/hydrophilicity, charge density
if the volume is chosen to coincide with molecules or molecular frag- and distribution, as well as coating ligand's length, grafting density and
ments its resolution is similar to CG MD, although it typically treats distribution, rigidity, and their affinity to receptors [144–147]. Since, in
problems that are more generic. Beyond DPD there are many other pos- real systems the membrane often interacts with more than one NP simul-
sibilities, including field-based treatments where a system is described taneously, NP aggregation and interaction of multiple NPs with the mem-
by density fields. Simulations coarser than DPD are beyond the scope brane also is a relevant factor [138,139]. Membrane properties such as
of this review. lipid phase, membrane composition, surface tension, charge density, re-
ceptor types and density also play important roles [148–150]. In addition
4. Computational studies on drug and gene delivery systems to all of these, external macromolecules, e.g. proteins in the bloodstream,
and different environmental conditions in different cell types [134], as
Computer modeling, as complementary tools to experiments, is ex- well as differences between external and internal cellular environments,
cellent in shedding light into the structural and dynamical properties are also of great importance and worth further investigation. Of course,
of systems of interest at atomistic or molecular levels of detail [112, many of these factors are coupled and have to be taken into account to-
113]. Applied to DDSs, computer simulations have been used to address gether in the design of NPs [138,139].
a broad range of questions (Table 1). Computer simulations can be used NPs are usually functionalized by coating with polymers, lipids, or li-
to study self-assembly [114–116], the structural and dynamical gands, to increase their stability, targeting capability, cellular uptake
1694 M. Ramezanpour et al. / Biochimica et Biophysica Acta 1858 (2016) 1688–1709

efficiency, and also to reduce their toxicity [131]. Both the type of coating improve targeting and cellular uptake. Dendrimers have been studied
molecule and the coating pattern and density strongly affect the physico- in many computational papers to characterize their structure in solu-
chemical properties of these NPs, and as a result their interactions with tion, their interactions with drugs and biomolecules [62,175], and
their cargo and target lipid membranes [147,151–155]. Simulations can their response to external stimuli [178]. All parts of dendrimers can be
be used to study these coatings, how they affect NP properties including modified chemically; core, shell and surface. The surface can be opti-
their interactions with other molecules [152,156], and can be used to op- mized for different biological activities by functionalizing the terminal
timize these coatings towards DDS with improved delivery capabilities groups. Thus dendrimers are a flexible class of materials with the poten-
[151,154]. Functionalization with PEG polymers, a process called tial for almost infinite variation.
PEGylation, is a good example of functionalization [157,158]. Drugs/pro- Different dendrimers have been modeled including PPI, triazines
teins/nucleic acids and NPs' surfaces are usually PEGylated to improve and PAMAM. Simulations can be used to investigate their interactions
their water solubility, stability, and circulation lifetime, and to reduce with other molecules, the effect of solvents and counter-ions, pH, salt
their aggregation, toxicity, and immune response [157,159]. concentration, different generation number, the nature of terminal
Considering the importance of PEG and its wide usage in drug deliv- groups and chemical modifications. Computer simulations have also
ery applications, PEG is worth further discussion. PEG is the most com- been used to study deformation and conformational changes of
monly used non-ionic polymer with stealth behavior for drug delivery dendrimers upon interaction with nanopores [179], lipid membranes
purposes [160]. These biocompatible hydrophilic polymers can solubi- [180], and nucleic acids [181] as a function of generation number, pH,
lize hydrophobic molecules, and reduce renal clearance and toxicity of functionalization, and ionic strength. For example, Nandy et al. [181]
drugs and nanoparticles. They also inhibit the aggregation of PEGylated simulated a 38 base pair double stranded DNA and three different gen-
agents by steric stabilization, as well as mask therapeutic agents from erations of PAMAM dendrimer (G3, G4, G5). They found that DNA
the host immune system and consequently suppress immunogenicity caused deformations in lower-generation dendrimers (G3) while
and antigenicity. This protective polymer layer, which is known as higher generation PAMAM (G5) bent DNA (Fig. 4). Generation number,
stealth sheath, can reduce the fast recognition by the immune system pH, and architecture have been shown to affect dendron's flexibility and
and reduce the non-specific interactions with blood components, rigidity, and as a result influence their binding efficiency to nucleic acids
resulting in longer blood circulation times. These properties have [182,183]. Several studies have investigated the effect of chemical mod-
made it the gold standard polymer for drug delivery and other biomed- ification on self-assembly, conformation, and drug/gene binding
ical applications and have been the topic of several computational [184–186].
modeling studies reviewed in more detail below. Despite the wide- Self-assembly of Janus dendrimers, comprised of hydrophobic and
spread use of PEG, there are several drawback to PEG including immu- hydrophilic parts, has been studied both experimentally and computa-
nogenicity and antigenicity, e.g. hypersensitivity reactions and the tionally. They form onion-like dendrimers and other complex architec-
development of antibodies against PEGs [160], which adversely affects tures [187,188].
pharmacokinetics [161], and the search for alternatives, e.g. poly(amino Computer simulation has been used to study how chemical modifi-
acids), continues, in part aided by computer simulations. cation and surface functionalization (e.g. PEGylation, acetylation, folic
In addition to different aspects of DDS mechanism, different types of acid groups, peptides, and targeting ligands of the surface) influences
DDS have been simulated. There are many types of nanoparticulate DDS. the structure [189], and the interactions with cargos [190,191]. Comput-
DDSs based on polymers, peptides, nucleic acids, lipids, carbon, er modeling can also be used to determine the optimal grafting densities
dendrimers and dendrons, and gold have been investigated via comput- and patterns based on structural and drug-loading properties, and inter-
er modeling and are described in more detail below. Combined, these actions with target membranes [154,192–195]. Effect of PEGylation size
applications show that computer simulation has become a powerful and grafting density on the structure of dendrimers, and also the struc-
technique for rational design and optimization of DDS, and demon- ture of the PEG layer itself have been widely studied [154,196–198].
strates the growing importance of computational pharmaceutics. PEGylation expands the core of dendrimers, effectively reduces their
For the rest of this review, we will present examples of computation- surface charge density and increases the overall size of dendrimers
al studies on each aspect of interest for each nanoparticulate DDS type in [196]. The effect of PEGylation on structure and drug loading capacity
separate sections, and then comment on challenges in this field. We will of PAMAM-G4 dendrimers was investigated by an all-atom MD simula-
focus on atomistic, CG MD simulations, and DPD simulations of tion [154]. Different PEG densities (25%, 50%, 75% and 100%), were test-
nanoparticulate DDSs; these fall mainly in categories defined as ed for complexation and loading capacity for 5-fluorouracil (5-FU), as a
dendrimers/dendrons, polymer-, protein/peptide-, nucleic acid-, lipid-, model anticancer drug. The simulations suggested that 25% PEG is the
and carbon-based DDSs, as well as gold nanoparticles (AuNPs). We em- most suitable PEGylation density for drug delivery purposes. This
phasize studies published since 2010. There are computational studies grafting density can retain the internal drug complexation capability,
on other DDSs that are outside the scope of this review, e.g. graphene and also assist in retaining drugs. For higher PEGylation degrees than
oxides and reduced graphene oxides [19,162], silicon NPs [19,163], 25%, 5-FU molecules to a high extent interact with external PEG chains
nanodiamond [164,165], layered drug delivery carriers [166,167], zeo- than dendrimer branches. Pearson et al. [193] used MD simulation to
lites [168–170], metal–organic frameworks [171,172], and other porous understand why PEGylated dendron-based copolymers functionalized
materials. These are excluded because they are not NPs, there are limit- with different terminal groups, –NH2, –COOH, and –Ac, do not exhibit
ed computational studies available, or because their focus is different a charge-dependent cellular interaction. Simulations suggest that inter-
from the studies described below. actions between positively charged terminal groups with PEG mole-
cules sequester the charges, and cause this low level of cellular
4.1. Dendrimers and dendrons interaction for NH2 terminated micelles.
Interactions between dendrimers with lipid membranes, permeation
Dendrimers are a class of branched globular materials with broad of dendrimers as a function of generation number (size), protonation
application in nanotechnology and nanomedicine, including for drug/ state, functionalization, and dendrimer concentration, as well as the effect
gene delivery [9–11]. Their monodispersity, modifiable surface, shell of bilayer asymmetry, bilayer tension, lipid tail length, and lipid phase
and core characteristics, and their high loading capacity make them (temperature) all have been studied by simulations [124,180,195,
suitable carriers for therapeutic agents, with promising applications in 199–204]. Tian and Ma [124] studied interactions of G4-PAMAM
gene delivery through complex formation with nucleic acids, so-called dendrimers with both symmetric and asymmetric negatively charged bi-
dendriplexes [173–177]. Drugs can be covalently or noncovalently layers, with and without tension, at both neutral and low pH. It was
incorporated in dendrimers to overcome solubility problems and to shown that both membrane tension and electrostatic interactions
M. Ramezanpour et al. / Biochimica et Biophysica Acta 1858 (2016) 1688–1709 1695

Fig. 4. PAMAM Dendrimer-DNA interaction and deformation as a function of generation number. (a) For the G3 dendrimer, very little deformation in DNA was observed, while the
dendrimer was bent considerably by the DNA. (b) For G4, both DNA and dendrimer were deformed. (c) For G5, DNA was deformed, but the dendrimer behaved almost as a rigid body.
Adapted with permission from B. Nandy, P.K. Maiti, A. Bunker, Force Biased Molecular Dynamics Simulation Study of Effect of Dendrimer Generation on Interaction with DNA, (2013).
Copyright 2013 American Chemical Society.

between charged dendrimers and tensed membrane play important role dendrimer binding site. Higher generation cationic dendrimers are capa-
in dendrimer penetration through membrane. Based on their simulation ble of membrane disruption in both low and high temperatures [204].
results authors proposed a mechanism of endosomal escape for pH- Functionalization can influence these interactions since it affects the
responsive gene delivery vectors (Fig. 5). Xie et al. [204] found that low physicochemical properties of dendrimers [195].
generations leave gel-phase membranes intact while higher generations Dendrimers can condense DNA and siRNA, and stabilize small mole-
fluidize the membrane and cause a gel-fluid phase transition around the cules in their core, shell and interface. Several studies have considered

Fig. 5. Proposed mechanism of endosomal escape for pH-responsive dendrimers. In endosomes (pH ~ 5), dendrimers are protonated, leading to increasing osmotic pressure and dendrimer
swelling, putting the membrane under tension. Meanwhile, electrostatic interactions between charged dendrimers and asymmetric negatively charged endosomal membrane cause a
drop in the critical membrane tension required for membrane disruption, allowing NP escape from endosomes. Bottom row: snapshots of a MARTINI simulation of interactions
between a G4 dendrimer (orange beads) and a membrane (with lipid tails represented with cyan beads, and lipid headgroups represented with green and blue beads) at different pH
levels. Reproduced from [124] with permission from The Royal Society of Chemistry.
1696 M. Ramezanpour et al. / Biochimica et Biophysica Acta 1858 (2016) 1688–1709

the loading, release, interactions and distribution of small molecules in one (PCL). These SCPs were different in their PCL and PEG length and
dendrimers [174,205–213]. For example, Jain et al. [206] used all atom the resulting molecular weight. They form core–shell structures, in
MD simulations and molecular mechanics Poisson-Boltzmann surface which PCL blocks form a dense hydrophobic core while PEG blocks
area analysis to study the solubility and drug release profile of the hy- form a shell around the PCL core. The simulations suggested that partial
drophobic drugs famotidine and indomethacin in G5 ethylenediamine water exposure of the PCL core drives aggregation into multimolecular
cored PPI dendrimers under different pH conditions. They found fast micelles. Increasing PEG length protects the PCL core from exposure to
drug release in low pH, intermediate values in neutral pH and a slow re- the water but it also increases the size of micelles. Therefore, since
lease in high pH. These results and further simulations on the effect of smaller micelles are preferred in DDS applications, predicting the mini-
dendrimer chemistry and topology on drug loading and release suggest mum number of PEG units for a specific number of PCL blocks required
that both the pKa of the drug and pH-dependent changes in the dendri- to fully protect the PCL core is important. The authors found a quantita-
mer influence these dendrimer–drug interactions and complex stability. tive relation between hydration of PCL core and both number and mo-
Loading and release of the antibiotic rifampicin, from G4-PAMAM den- lecular weight of PEG and PCL blocks [122].
drimer has been studied by Bellini et al. [205] in a combined experimen- The internal structure of the thermoresponsive polymer blend NPs,
tal and simulation approach. Experimentally, approximately 20 drug and the phase transition between core–shell and Janus structures
molecules were found as the maximum potential loading capacity of were investigated by DPD simulations by Guo et al. [226]. Chen and
this dendrimer at neutral pH. In simulation, the model of the dendrimer Ruckenstein [230] looked at the structure and mechanism of formation
with 20 drug molecules in neutral pH was more stable than in low pH, and the degradation process of multicomponent multicore micelles via
where drug molecules got expelled rapidly and simultaneously to the DPD simulations. Taresco et al. [231] used a combined computational/
solvent. This suggests a pH dependent drug release desirable for drug experimental approach to study the self-assembly and structure of co-
delivery applications. Simulations can also be used to study the effect polymer micelles formed by two different types of monomers. Wang
of PEG lipids on loading capacity of dendrimers, and on surface et al. [233] studied the structural characterization of cholesterol-
ligand's targeting capability [154,211,214]. functionalized CD micelles by all-atom MD simulation. Tan et al. [102]
Based on a comparison of several generations of PAMAM by both experiment and DPD simulations showed that introducing a
dendrimers, G4 PAMAM might be optimal as vector for gene delivery, second hydrophilic phosphatidylcholine group into the polymer chains
based on PAMAM interactions with bilayers and siRNA complexation causes a phase transition from sphere to rod-like in multiblock polyure-
[203,215]. For gene delivery applications, generation number (size), thane micelles.
flexibility, hydrogen bonding capability, pH, ion concentration and salt Hybrid systems composed of polymers and other NPs, e.g. AuNPs
type have been considered in simulations [183,216–219]. Electrostatic and QDs, also show interesting structural and phase transition behavior
interactions and the entropy gain from counterion release upon binding [227,235]. Using both experiments and DPD simulations, Cai et al. [227]
have been proposed as important driving forces determining the inter- studied the effect of adding AuNPs on block copolymer self-assembly.
actions between dendrimers and nucleic acids [181,212,213]. Adding AuNPs caused a transformation from long cylindrical micelles
As one final example, we would like to design drug specific to short cylindrical micelles and finally to spherical micelles.
nanocarriers by including optimal drug-binding molecules in the struc- Interactions with and translocation through cell membranes for
ture. Shi et al. [220] used de novo design of dendrimers via optimizing polymer-based DDSs have been studied both by experiment and simu-
building blocks and including optimal drug-binding molecules with a lations [236–238]. Li et al. [236] studied the interactions between a
final goal the design of a drug-specific nanocarrier. They combined novel amphiphilic polymer, PMAL, and siRNAs, combining experiment
virtual screening with molecular docking and MD simulations on and CG MD simulations. Potential of mean force calculations showed
DOX-dendrimer complexes. Using Rhein, a molecule with strong DOX that siRNA by itself experiences a large free energy barrier for transloca-
binding, they achieved Rhein-containing nanocarriers with a suitable tion while the siRNA–PMAL complex entered the membrane spontane-
DOX profile. ously, consistent with experiment. The PMAL polymers induced pore
formation to facilitate siRNA translocation. Srinivas et al. [237] used
4.2. Polymer-based delivery systems CG MD simulation and showed that these patchy polymeric micelles
are capable of accommodating and transporting hydrophilic contents
Polymeric NPs are of interest in drug delivery applications and more across a lipid membrane into a lipid vesicle.
generally because of their stability and easily modifiable surface [21,31]. Ding and Ma [238], using DPD simulations, investigated receptor-
Block-copolymers [92] composed of hydrophobic and hydrophilic parts mediated endocytosis for a new type of pH-responsive DDS composed
can form a variety of self-assembled structures [221] of interest in drug of NP (radius of 4 nm) and some pH-sensitive polymers (of twelve
delivery applications. These structures, e.g. polymeric micelles [222], are beads length). Beads on the NP surface were treated as ligands and
capable of accumulating drug molecules in their core, interface and shell assigned +e charges. The polymer had N randomly distributed nega-
[121]. There are several reviews of computational studies on different tively charged (−e) beads, where N depends on the pKa of the polymer
aspects or specific type of polymers [23,221,223–225]. Here, we will dis- and the system's pH. Half of the model lipids in the membrane were
cuss some examples. treated as receptors by replacing charged beads in the lipid's head
Amphiphilic polymers form a variety of structures in solvents, e.g. group with neutral beads. They introduced a modified Lennard–Jones
core–shell, Janus particles, micelles, and rod-like micelles [102,226, potential to account for receptor–ligand interactions. Eighteen-
227]. Several factors affecting aggregate structures, stability and phase microsecond simulations for three different pH conditions combined
transitions between different structures have been studied: for exam- with potential of mean force calculations showed a triple-pH-response.
ple, solvent polarity [130], polymer architecture, concentration and The NP can only be engulfed by cell membrane when the pH is higher or
physicochemical properties [226,228,229], mixture components and lower than the polymer's pKa, whereas endocytosis is blocked when the
mixing ratios of comprising polymers [230,231], pH, temperature pH equals the polymer's pKa. It was also found that the properties of
[226,232], and the addition of components, e.g. lipids, cholesterol and li- NPs, pH-sensitive polymer and cell membranes, and the external envi-
gands [102,233,234]. Huynh et al. [122] performed all atom MD simula- ronment, all affect NP engulfment by the membrane, at least in this sim-
tions on star copolymers (SCPs) to understand why SCP micelles are plified DPD model (Fig. 6).
unstable and form multimolecular micelles in solution, and to assist in The effect of functionalization on interactions with the membrane,
the rational design of stable SCP unimolecular micelles in solution. uptake mechanism, as well as effect on the structure and conformation
Each SCP has a central connecting part and six attached arms, with of drugs/polymer complexes have been studied by simulation [118,
each arm comprised of a hydrophilic part (PEG) and a hydrophobic 239–243]. Liao et al. [240] used both experiment and all atom MD
M. Ramezanpour et al. / Biochimica et Biophysica Acta 1858 (2016) 1688–1709 1697

Fig. 6. Time evolution of nanoparticle-polymer complex endocytosis as a function of pH. The endocytosis process of a nanoparticle-polymer complex (nanoparticle with pH-sensitive
polymers absorbed on its surface) depends on the pH. For pH values lower (a) and higher (c) than the polymer's pKa, the nanoparticle can be fully engulfed by the membrane.
However, for pH = pKa (b), endocytosis is blocked. Membrane lipid head groups are shown in green, purple, and blue correspond to +e, −e, and neutral beads, respectively. Lipid
tails are in orange, receptor heads are in red, and nanoparticle beads are in yellow. Polymer beads are in cyan (−e) and pink. Reprinted by permission from Macmillan Publishers Ltd.:
Scientific reports H. Ding, Y. Ma, Controlling Cellular Uptake of Nanoparticles with pH-Sensitive Polymers., Sci. Rep. 3 (2013) 2804., Copyright 2013.

simulations to study the uptake of chitosan/DNA complexes coated with function of temperature [256], pH [228,232,257,258], concentration of
anionic poly(gamma-glutamic acid) (gamma-PGA). This coating en- drugs [232], physicochemical and structural properties of polymers
hanced the cellular uptake of chitosan/DNA complexes via a specific and cargos [259,260]. CD is a well-known family of cyclic oligosaccha-
protein-mediated endocytosis. Both the structure of this ternary com- rides with great promise as solubilizing agent for poorly soluble drugs
plex (CS/DNA/PGA), and the interaction between gamma-PGA and [246,261]. He et al. [246] performed MD simulations to investigate the
glutamyl transpeptidase proteins were studied by MD simulations. binding mode and affinities of the antifungal drug amphotericin B
Sun. et al. [239,244] studied the effect of polyethylenimine polymer drug with γ- and β-CD. Consistent with experiments, simulations
(PEI) modification by lipids on nucleic acid compaction and aggrega- showed a significantly higher binding affinity for γ-CD than β-CD. The
tion. They found that lipid association as an additional mechanism of ag- binding mechanism of the drugs bexarotene and human vasoactive
gregation resulted in more compact and stable structures. intestinal peptide (VIP) to highly PEGylated sterically stabilized micelles
Several studies have focused on interactions between polymeric (SSMs) has been investigated by Vukovic et al. [121] both experimental-
DDS and small molecule cargos, investigating driving forces for ly and computationally. Using free energy profiles, they predicted the
partitioning [60,245–249], complex stability [250], drug loading capac- drug distribution in SSMs. Single bexarotene, as a poorly water-
ity and rate, and drug distribution and release [121,247,251–255] as a soluble drug, resides in the ionic interface with its polar end exposed

Fig. 7. Solubilization of bexarotene in sterically stabilized micelles (SSM). (Left) Free energy profiles for systems composed of 1, 3, and 5 bexarotenes in SSMs. SSMs are either composed of
10 (small) or 90 (large) monomers, in water and 0.16 M NaCl, respectively. Single drugs prefer the ionic interface, represented with vertical arrows, in both SSMs. For multiple drugs
accumulated in the SSM core (shown as a gold surface), a deeper minimum develops in the core. This is because several bexarotene (e.g. 5) cluster together (Right) via a hydrogen
bond network between their –COOH groups. Reprinted (adapted) with permission from L. Vukovic, A. Madriaga, A. Kuzmis, A. Banerjee, A. Tang, K. Tao, et al., Solubilization of
Therapeutic Agents in Micellar Nanomedicines, (2013). Copyright 2013 American Chemical Society.
1698 M. Ramezanpour et al. / Biochimica et Biophysica Acta 1858 (2016) 1688–1709

to solvent. With multiple bexarotene molecules, the preferred distribu- There are also DDSs based on nucleic acids. RNA and DNA have been
tion is as clusters in the alkane core of SSM (Fig. 7). Dominated by elec- used as either building blocks or surface ligands to design a variety of
trostatic interactions, VIP molecules, with two regions with positively three-dimensional nanostructures, including rings, tubes, cubes, cages,
charged residues, interact with negatively charged − PO4 groups in and spheres [276–279]. As with other DDSs, the structural properties
the ionic interface. These simulations suggest the importance of the bal- of these systems, and their interactions with endogeneous biological
ance between hydrophobic and Coulomb interactions between drugs molecules, as a function of environmental conditions are of interest
and phospholipid polymers in stabilization of drugs in SSMs. [280–282]. Badu et al. [281] performed atomistic MD simulations on
The interaction and complexation of nucleic acids, including DNA and RNA nanotubes and investigated their structural properties. Juul et al.
siRNA with cationic polymers have been studied extensively [223]. PEI, as [282] studied DNA cages loaded with active enzymes both experimen-
an example of cationic polymers, is one of the most promising polymers tally and by all-atom MD simulation, and found a temperature-
which have been widely utilized in studies for gene delivery applications. dependent conformational change from an ‘open’ to a ‘closed’ state, in
Both experimental and computational studies have provided a better un- principle enabling controllable encapsulation and release (Fig. 8).
derstanding of PEIs' binding modes with DNA/siRNA [223]. Molecular Protein-based NPs are promising because of their unique character-
modeling has had a significant impact on our understanding of critical istics, including biocompatibility, biodegradability, low cytotoxicity, and
steps in gene delivery, including complexation of carriers with nucleic non-antigenicity [283,284]. Some proteins, e.g. albumin, can be used for
acids, as well as nucleic acid condensation and aggregation by gene car- targeted delivery to tumor cells and inflamed tissues since they are pref-
riers [177]. Simulations can be used to investigate the factors determining erentially taken up by these types of cells [284]. Luo et al. [272] using
interactions between cationic polymers and nucleic acids, including com- both theory and experimentations investigated the binding mode, affin-
plexation and decomplexation. Such factors, similar to those in other ity and interactions between bovine serum albumin (BSA) and interfer-
DDSs, include chemical surface composition as amine-to-phosphate on alpha-1b (IFN) as delivery system and protein drug, respectively.
ratio [51,262], charge density, protonation state, polymer molecular Simulations predicted domain III of BSA as the most probable binding
weight, polymer chemical structure and charge distribution [262–267], sites, as well as hydrogen bonds and salt bridges as the main contribu-
the effect of endogenous molecules [136], and multivalent ions, e.g. tors in binding between BSA and IFN. Cyclic peptide based nanotubes
Fe(III) [137,268]. miRNA as endogenous molecule might cause (CPNTs) can be used as a transporter for ions and small molecules
decomplexation. To test this hypothesis, Meneksedag-Erol et al. [136] [273,274]. Vijayaraj et al. [273] investigated the transport mechanism
used both experiment and all atom MD to study the interactions between of 5-FU through a variety of CPNTs and determined the driving forces
miRNA and pre-formed PEI-siRNA complexes. PEI was found to bind and free energy barriers affecting this transport. They found that trans-
more strongly to miRNA than siRNA. However, miRNA could not disrupt port was driven by direct or water-mediated hydrogen bonding and hy-
the integrity of PEI-siRNA but formed a layer on the complex, leading to drophobic interactions between CPNT and 5-FU.
a siRNA-PEI-miRNA complex. CPPs are of interest because of their intrinsic ability to enter cells,
Interactions between polymeric DDS and nucleic acids for gene de- their low cytotoxicity, and their versatility. They can facilitate the trans-
livery purposes can be studied indirectly using polymers. Polycations port of small molecules (e.g. drugs, imaging agents), macromolecules
and polyanions can be considered as models for DDS and nucleic (e.g. DNA, siRNA, proteins, peptides) and NPs/DDS. Therefore, CPPs
acids, respectively [269–271]. Zhao et al. [270] utilized polycations and can be used as drug/gene delivery systems, as well as carriers for imag-
polyanions and studied the effect of charge distribution along the ing agents, proteins and peptides [275]. Recently, it has been of great in-
chain on complexation behavior and the structure of the resulting com- terest to combine the benefits of nanomaterials and CPPs. CPPs can be
plexes. They found that charge distribution has a significant influence functionalized with other delivery systems, like QDs, liposomes, and
on aggregation and the complex structure. dendrimers, potentially improving CPPs intracellular drug release abili-
ty and decreasing their toxicity. DDSs such as polymeric micelles,
4.3. Peptide- and nucleic acid-based delivery systems dendrimers, liposomes, QDs, and inorganic nanocarriers (e.g. gold-,
silver-, and iron-based NPs) can be functionalized with CPPs to enable
There are several peptide-based DDSs, e.g. protein- [272] and cyclic their cellular uptake [275]. CPPs' unique physicochemical properties,
peptide-based [273,274], and cell penetrating peptides (CPPs) [275]. their uptake mechanisms, classifications, and applications in medicine

Fig. 8. DNA nanocage with temperature-controlled conformational transition capability. The cage is closed at 4 °C (a), and in a more open conformation at 37 °C (b). Heating the nanocage
to 37 °C in presence of horseradish peroxidase (HRP), followed by cooling to 4 °C, allows HRP enzymes to be encapsulated in nanocages. Reheating to 37 °C causes enzyme (orange) release
(not shown). Reprinted with permission from S. Juul, F. Iacovelli, M. Falconi, S.L. Kragh, B. Christensen, R. Frøhlich, et al., Encapsulation and Release of an Active Enzyme in the Cavity of a
Self-Assembled DNA Nanocage, (2013) 9724–9734. Copyright 2013 American Chemical Society.
M. Ramezanpour et al. / Biochimica et Biophysica Acta 1858 (2016) 1688–1709 1699

and biotechnology have recently been reviewed by Durzynska et al. study how factors such as solvent and temperature influence both
[275]. self-assembly and transitions between different morphologies
Key elements involved in CPPs' roles in drug delivery have been stud- [297–301].
ied computationally [285–288]. Atomistic MD simulations revealed the
importance of arginine, lysine and tryptophan in the interaction between 4.4. Carbon-based delivery systems
two CPPs, penetratin [285] and transportan [286], and
dipalmitoylphosphatidylcholine (DPPC) membranes, as well as their Several different types of carbon-based nanoparticles (CNPs) have
translocation mechanism. CPP conjugation can help translocate hydro- the potential to act as DDS [12,302]. Here we focus on DDS based on
phobic and hydrophilic drugs across lipid membranes [289,290]. Teixeira CNTs and fullerenes but DDS using graphene, graphene oxide, and
et al. [290], via both MD simulations and experiments, found that lysine- nanodiamond [19,162,165,168] are also of interest.
based surfactants enhanced the transdermal permeation of two topically Computational research has played a major role in exploring drug
administered hydrophilic drugs, tetracaine and ropivacaine hydrochlo- loading of [164,303] and release [132,133,304,305] from CNPs, as well
ride. As mentioned earlier, NPs can be functionalized with CPPs. The con- as interactions of cargos with carbon nanocarriers [306] as a function of
centration and distribution of CPPs on the NP's surface influence the internal and external stimulus, e.g. pH, temperature, torsion, and external
conformation of the peptide layer, which in turn affects the CPP- magnetic field. CNTs are capable of loading therapeutic agents on their
functionalized NPs' activity and the efficiency of cellular internalization. surface or inside their cylindrical hollow [19,307]. Saikia et al. [305] per-
So, designing a DDS of desired activity requires knowledge of the struc- formed MD simulation to study the C60 fullerene-mediated release of
ture and dynamics of CPPs on NPs in solution prior to interacting with multiple pyrazinamide molecules from the interior of a single walled
membranes. Todorova et al. [291] using both experimental and atomistic CNT as a function of temperature. Chaban et al. [304] investigated the ef-
MD simulations investigated the effect of the grafting density and surface fect of temperature and concentration on drug release from CNTs using
distribution of HIV-derived trans-activator of transcription (TAT) peptide MD simulations. The effect of external magnetic fields has been studied
on NPs cell internalization. They found a correlation between the surface on hybrid systems composed of CNTs with magnetic NP caps [132].
properties, e.g. positive charge distribution, of these TAT conjugated NPs Understanding the interactions of CNTs with biological macromole-
in solution, and their membrane permeation. cules is also of great interest [308–313]. Wu et al. [313] studied DNA
New DDS design strategies combine the benefits of several systems ejection from CNTs as a function of temperature, torsion loading and
such as block copolymers and CPPs [292], or mimic natural systems CNT size by MD simulation. Santosh et al. [309] studied the binding of
such as intrinsically disordered proteins (IDPs) [293]. Sanchez- siRNA and DNA to the surface of single walled CNTs via MD simulations.
Sanchez et al. [293], using both experiments and MD simulations, Chen et al. [311] studied the dynamic mechanism of encapsulation of
designed and characterized single chain NPs called artificial IDP HIV replication inhibitor peptide into CNTs using MD simulation. It is
mimetics. These IDP mimetics are capable of simultaneous release of also important to understand the interactions of carbon-based DDS
two dermal bioactive molecules, folic acid and hinokitiol, into aqueous with human serum albumin (HSA), the most abundant protein in
solution in a pH-dependent manner. blood. Interaction of DDS with HSA is crucial in DDS absorption, distri-
Computer simulations can provide insight into drug loading, distri- bution and metabolism [314].
bution/complex structure and release from peptide self-assemblies The membrane associations of carbon-based DDS have been studied
[294,295], as well as the influence of pH and temperature [282,294, by computer simulation [315–317]. Several factors affect these associa-
296] on these processes. Guo et al., using both experiments and DPD tions: DDS structural properties [318–321]; clustering [320–322];
simulations investigated the effect of pH on microstructure of peptidic functionalization [319,323,324]; concentration [319] and lipid mem-
micelles [296]. Micelles were self-assemblies of HR-20 (Histidine10-Ar- brane properties [321,325]. The effects of size and clustering of fullerene
ginine10) peptides conjugated with cholesterol, either loaded or not NPs, and fullerene concentration on a DPPC monolayer as a model for
with DOX. Micelles had more compact structure at pH N 6 (Fig. 9). For pulmonary surfactant was investigated by CG MD simulation by Chiu
pH b 6, channels in swollen micelles might cause drugs to diffuse out et al. [321]. Free energy calculations suggest that all fullerene systems
of micelles. This was verified experimentally as the DOX release rate in this study (C60, C180, C540, and cluster of five C60 fullerenes) spon-
was influenced by pH values, consistent with simulation results. taneously diffuse into the hydrophobic region of both monolayer and bi-
Finally, physicochemical properties of peptides and solvent condi- layer membranes. Large fullerene molecules were found to prefer
tions can govern self-assembly and structural properties of peptide- partitioning into bilayers rather than monolayers, however, which can
based carriers. Computer simulation can be used to design peptides influence the monolayer-to-bilayer transition in the respiratory cycle.
which can self-assemble, and to study the mechanism of self-assembly This may suggest a possible mechanism of internalization of CNPs
into nanostructures. Computational approaches can also be used to through lung inhalation.

Fig. 9. Effect of pH on microstructure of self-assembled micelles. Starting from a homogeneous state (a), cholesterol conjugated peptides form compact core/shell micelles at pH N 6.0 (b).
Decreasing the pH loosens these micelles (c) and the swelling facilitates DOX release. Arginine, histidine, and cholesterol, are shown in green, brawn, and black, respectively. Water has not
been shown for clarity. Reprinted with permission from X.D. Guo, L.J. Zhang, Z.M. Wu, Y. Qian, Dissipative Particle Dynamics Studies on Microstructure of pH-Sensitive Micelles for
Sustained Drug Delivery, Macromolecules. 43 (2010) 7839–7844. Copyright 2010 American Chemical Society.
1700 M. Ramezanpour et al. / Biochimica et Biophysica Acta 1858 (2016) 1688–1709

The effects of using polymers, peptides, and lipids as surface immune system [333]. However, the mechanism through which
functionalization agents in carbon-based DDS were studied by compu- PEGylated lipids interact with different components in the liposomes,
tational methods. Chehel Amirani and Tang [308] comprehensively its function in drug distribution, localization and retention as well as
reviewed studies, mainly at the quantum mechanical level, of graphene the influence of external environmental factors such as salt concentra-
and CNT functionalization, and their binding energies with nucleobases. tion are not completely understood.
Lai and Barnard [131] have reviewed recent studies on functionalized Since atomistic simulations of full liposomes are computationally ex-
nanodiamonds and their importance for biomedical applications. pensive, a common approach is to infer results on liposomes from lipid bi-
Functionalization can prevent NPs from aggregating [323,326], increase layer simulations. Dzieciuch et al., [334] for instance, performed atomistic
their stability [327], and influence their interactions with lipid mem- MD simulation to study the effect of liposome PEGylation on their drug-
branes [323,326,328]. Surface functionalization can also promote drug loading efficiency. They compared the effect of zwitterionic and
delivery through improving translocation across lipid bilayers [328]. PEGylated membranes on the location and orientation of a model hydro-
Sridhar et al. [328] used CG MD simulations to look at effects on fuller- phobic compound, 5,10,15,20-tetrakis(4-hydroxyphenyl)porphyrin (p-
ene (C60) translocation through the membrane due to polar and non- THPP). p-THPP enters both types of lipid bilayers, which agreed with ex-
polar functionalization, temperature, and fullerene concentration. perimental results, but in PEGylated liposomes p-THPP also localizes to
Although none of their models showed complete translocation, they the outer PEG corona where porphyrins are wrapped by PEG chains.
found that Janus particles having half the surface modified by polar Thus, in PEGylated liposomes p-THPP showed a greater exposure to the
groups were the most promising form of functionalized fullerenes in water than in zwitterionic liposomes. These results highlight that
terms of bilayer translocation. PEGylation enhances the drug-loading efficiency of membranes and sup-
A NP's behavior depends on the conformation of the functionalization port fluorescence experiments carried out in this study [334]. Simulations
layer on its surface. The coating mechanism used to functionalize the NP, also predicted strong interaction between PEG lipids and porphyrin [334].
e.g. covalent vs. non-covalent attachment, as well as the grafting density This interaction is particularly important in drug delivery studies as it may
are two factors which can affect this conformation [329,330]. For exam- be an extra barrier to drug release. This finding supports previous compu-
ple, Lee [330] found, using CG MD simulation, that the PEGylation method tational studies that showed that PEG interacts with hydrophobic mole-
can influence PEG distribution on CNTs. They also found that PEG size and cules [127,335,336]. Similarly, MD simulations combined with free
grafting density affected the PEG layer's conformation (Fig. 10). Designing energy calculations were used to elucidate binding mechanisms and di-
functionalized NPs with desired properties will require a detailed vulge the exact location and organization of two small drugs, bexarotene
understanding of these factors. and human VIP, within PEGylated micellar nanocarriers [121].
PEGs strongly interact with salts. This property has been utilized in
4.5. Lipid-based delivery system lithium ion batteries with PEGs as polymer electrolytes [337,338].
PEGs are soluble in a wide variety of both polar and nonpolar solvents
Recently, lipid-based DDS such as liposomes, micelles and LNPs have [339]. Although both of these general properties have been shown pre-
attracted much attention [7,8,331,332] due to their ability to encapsu- viously by computational studies [337,338,340] and experiments [341,
late and transport drugs as well as biomolecules, the versatility of 342], for DDS applications it is important to understand how the PEG
their structures and compositions, and their inherent selectivity to layer on DDS, e.g. PEGylated liposomes, interacts with salts in physiolog-
tumor cells and inflammation sites. There are currently more than ten ical conditions and how these interactions affect the structure of the
approved lipid-based drug formulations and many more in clinical trials PEG layer [118,343,344]. Stepniewski et al. [344] modeled the effect of
[7,8]. Specifically, sterically stabilized liposomes, i.e. PEGylated lipo- NaCl on the surface structure of PEGylated liposomes. They varied the
somes, are widely used as DDS. PEGylation achieves long circulation salt concentration in Langmuir monolayer film experiments and
times for liposomes in vivo due to the formation of a protective PEG added ions at the physiological level in atomistic MD simulation studies.
layer over the liposomal surface. This layer sterically prevents the coat- The results showed that the PEG surface layer should not be treated as
ing of liposomes by opsonins, thus reducing drug uptake by cells of the generic hydrophilic molecules completely outside the bilayer, nor

Fig. 10. Effect of PEGylation method, and PEG chain size and grafting density on the conformation of PEG layer on carbon nanotube. Non-covalently modified single walled CNT (SWNT)
(Left) is more exposed to water than covalently modified SWNT (Right). PEG (red) size and grafting density also influence the conformation of PEG layer on SWNT (gray cylinder). RF is the
mushroom radius and L is the thickness of brush state. SWNT is shown as gray cylinders. Reprinted with permission from H. Lee, Molecular Dynamics Studies of PEGylated Single-Walled
Carbon Nanotubes: The Effect of PEG Size and Grafting Density, (2013). Copyright 2013 American Chemical Society.
M. Ramezanpour et al. / Biochimica et Biophysica Acta 1858 (2016) 1688–1709 1701

should it be considered as totally neutral as was previously accepted. nanostructured lipid carriers [355]. A detailed model of the distribution
PEG molecules are able to penetrate into the liquid-crystalline lipid of drug molecules inside a liposome was obtained using the Martini
bilayer, which may affect the permeability and structure of the mem- force field [356]. Jämbeck et al. [356] performed large scale simulations
brane. It was also noted that Na+ ions bind to PEG chains, changing of hypericin, a photosensitizing drug, in a DPPC liposome (Fig. 11) and cal-
the surface charge of the liposome and enhance opsonization, whereas culated the distribution and orientation of the drugs in the lipid bilayer.
Ca2+ did not interact with PEG [343]. These results are not surprising, Hoof et al. [357] studied the encapsulation of proteins in spontaneously
however. In fact, when PEGs are attached on DDS surface, e.g. liposomes, forming vesicles using MD simulation with a CG force field. They showed
they seem to show similar properties as previously reported for other that the interactions of proteins with membranes govern the encapsula-
applications [337,338,340–342]. As another example, Pannuzo et al. tion efficiency, but the size of the encapsulated proteins and the speed
[345] performed CG simulations to study the effect of Ca2 + and PEG of the vesicle formation did not seem to have a significant effect.
molecules on membrane fusion. They showed that for a rapid fusion be- A promising gene therapy approach is based on siRNA. Sophisticated
tween negatively charged apposed membranes both Ca2 + and PEG delivery systems are required to protect and deliver siRNA effectively to
were required. the target tissues [358,359]. CG molecular simulations together with a va-
An important property governing drug release from liposomal DDS riety of experimental techniques were able to give a more detailed view of
is bilayer permeability [346,347]. Magarkar et al. [346] performed atom- the structure of a LNP containing siRNA (Fig. 12) [79,360]. Simulations of
istic simulations using the OPLS-AA force field to investigate the high small systems containing 8 duplex 12 base pair DNA complexes, DLinKC2-
permeability observed for “inverse-phosphatidylcholine” (CPe) lipo- DMA (an ionizable cationic lipid), DSPC, cholesterol, a PEG-lipid, water
somes in experimental studies [348]. CPe is a synthetic analog of phos- and ion molecules were performed. These molecules self-assembled and
phatidylcholine (PC) with a reversed zwitterionic headgroup and has organized in an ordered structure where, for instance, the PEG-lipids
been proposed for use in liposomal DDS formulations. The larger surface were distributed in the outer layer of the LNP. This was the first study
area observed for the CPe bilayer compared with the PC bilayer may to show that LNP siRNA systems likely have a nanostructured core, with
cause CPe's higher permeability. Adding cholesterol to liposomal formu- the encapsulated siRNA located in internalized inverted micelles com-
lations reduces leakage from liposomes due to its well-known role in plexed to cationic lipids. The in silico finding agreed with 31P NMR,
lipid packing and stability, but the mechanism for this is not understood. FRET, cryo-TEM, and RNase digestion data and it was possible to under-
Magarkar et al. [336], by performing MD simulation using OPLS force stand the mechanism whereby LNP siRNA systems are formed. This in-
field, proposed a model to explain the interaction between PEG mole- sight is being applied to the design and construction of new LNP
cules and cholesterol. PEG molecules tend to interact with the ß side systems. In addition, both CG and atomistic computer simulation have
of cholesterol, disrupting the membrane structure. Contrary to the ex- been used to study the DNA adsorption onto anionic lipid membranes in-
pectation that cholesterol should make the membranes more stable duced by multivalent cations [361].
and compact, adding cholesterol in the presence of PEG caused mem-
brane destabilization. This result may explain possible effects of choles- 4.6. Gold-nanoparticles
terol on the permeability and compressibility of PEGylated liposomes.
Self-assembly is another feature of special interest in nanotechnolo- AuNPs are of particular importance in drug delivery research due to
gy and nanomedicine as its understanding provides a framework for their unique properties including their stability, ease of synthesis, and
developing new nanoscale materials with desirable properties [116]. the ability to manufacture a range of sizes [15,362]. They can be loaded
CG and DPD simulations have been performed for the investigation of with various therapeutics including small molecules, peptides, proteins,
lipid-based aggregates relevant for drug delivery [115,349–352]. For ex- and nucleic acids [19], either conjugated to AuNPs covalently or
ample, Lee and Pastor [349] used the Martini CG force field to study mix- noncovalently. There are several computational studies on the properties
tures of lipids and PEGylated lipids in water at different sizes and of AuNPs and their interactions with other molecules [135,227,363–370].
concentrations of PEGylated lipids. They found that the mixtures self- The effect of AuNP conjugation on peptide conformational flexibility
assembled to liposomes, bicelles and micelles. The analyses and simula- and structure was studied by Lee and Ytreberg [366]. They examined
tions indicated that the average aggregate sizes decreased when the the structure and dynamics of six peptides that were either free or
concentration of PEGylated lipid increased, in agreement with experi- conjugated to AuNPs in water. Conjugation affected both structure and
mental results. Janke et al. [352] used CG MD simulations to study the dynamics in an amino acid sequence dependent way. Peptides with little
phase behavior of oleic acid aggregation at various concentration and or no secondary structure in solution were adsorbed on the AuNP surface.
protonation states. They observed a range of structures including This causes peptides to lose their specific interactions with cell compo-
micelles, vesicles and oil phases depending on the protonation state of nents. For drug delivery purposes, this suggests that peptides with signif-
the oleic acid head group. icant secondary structures in solution are suitable candidates for peptide-
Simulations can be used to calculate drug release rates from lipid- NP conjugation. Interactions between ubiquitin and AuNPs have also been
based DDS [353,354]. For instance, the release of encapsulated materials studied by Brancolini et al. [135] using computer simulation.
from systems including emulsions (100% liquid lipid phase) and nano- Van Lehn and co-workers performed a series of computational stud-
structured lipid carriers, which have a combination of solid and liquid ies on AuNPs coated by lipids. Among other properties, they investigat-
lipid domains, was investigated by Dan using Monte Carlo simulations ed the structure of the monolayer coating the AuNP under different
[353]. The results obtained suggest that the size of lipid domains does conditions [369], lipid composition and AuNP size [368], the interac-
not significantly affect the rate of release, but the location and distribution tions of a monolayer-coated AuNP with model membranes relevant
of the solid domains have a notable impact. When solid domains are con- for uptake [367], suggesting similarities in uptake process with
centrated near the solution interface, transport is inhibited due to a reduc- bilayer-bilayer fusion [371], and interactions between AuNPs [370].
tion in the accessible surface area. However, when the solid domain is
located in the center of the LNP the release rate is increased and may 5. Conclusions and future perspectives
even be higher than in the corresponding emulsion particle. In another
study by Dan [354], Monte Carlo simulations were performed to study We have reviewed common biophysical and computational ap-
drug release in response to electric field-induced liposome pores. proaches to studying DDSs, focusing in particular on computational
There are also other interesting studies, three of which are mentioned studies of several major classes of NPs used in drug delivery research.
briefly here. Computer simulations were combined with experiments to The strength of computer simulations in general is their ability to give
investigate the effect of penetration and membrane behavior of three ter- very detailed insight into the structure of NPs and their interactions
penes, which are effective chemical permeation enhancers in with drugs and model membranes under a range of conditions. Clearly,
1702 M. Ramezanpour et al. / Biochimica et Biophysica Acta 1858 (2016) 1688–1709

Fig. 11. Distribution and orientation of hypericin within liposome. a) Snapshots from MD simulations taken after 10 μs. Phosphorus groups (orange), choline groups (blue) and
hydrophobic tails (green) of the lipids are represented, with different numbers of drug molecules (yellow). b) Liposomes with some lipids removed to show the binding of hypericin
to DPPC; the hydrophilic parts of hypericin are shown in red and the hydrophobic parts in yellow. Reprinted with permission from J.P.M. Jämbeck, E.S.E. Eriksson, A. Laaksonen, A.P.
Lyubartsev, L.A. Eriksson, Molecular Dynamics Studies of Liposomes as Carriers for Photosensitizing Drugs: Development, Validation and Simulations with a Coarse-Grained Model,
(2013). Copyright 2013 American Chemical Society.

the efficiency of DDSs depends on a large number of other parameters, can characterize important aspects of DDSs, a major challenge for the
including stability in the bloodstream, targeting to the desired tissues, near future is to go beyond individual drug carriers and to investigate
uptake by endocytosis or other mechanisms, and final release of the at a mechanistic level the larger-scale processes that determine the suc-
drugs. Although experimental biophysical and computational studies cess of DDSs.

Fig. 12. Lipid nanoparticle. External and a cross-sectional view of the LNP (top). Components are represented by their molecular densities: protective PEG-lipid (magenta), therapeutic
siRNA (red), cholesterol (green), DSPC (yellow), ionizable cationic lipid (DLin-KC2-DMA) (blue) and water (cyan). Adapted from D. Rozmanov, S. Baoukina, D.P. Tieleman, Density
Based Visualization for Molecular Simulation., Faraday Discuss. 169 (2014) 225–43. doi:10.1039/c3fd00124e. Under a Creative Commons Attribution 3.0 Unported License.
M. Ramezanpour et al. / Biochimica et Biophysica Acta 1858 (2016) 1688–1709 1703

At the computational level, currently we are limited by some of the [14] A. Bianco, K. Kostarelos, C.D. Partidos, M. Prato, Biomedical applications of
functionalised carbon nanotubes, Chem. Commun. 571–577 (2005).
standard limitations of biomolecular simulation. In general, these in- [15] P. Ghosh, G. Han, M. De, C.K. Kim, V.M. Rotello, Gold nanoparticles in delivery ap-
volve limited sampling time, limited length scale, and expensive or dif- plications, Adv. Drug Deliv. Rev. 60 (2008) 1307–1315.
ficult access to important thermodynamic variables. Simulations can [16] T.K. Jain, M.A. Morales, S.K. Sahoo, D.L. Leslie-Pelecky, V. Labhasetwar, Iron oxide
nanoparticles for sustained delivery of anticancer agents, Mol. Pharm. 2 (2005)
access time scales of the order of microseconds or for coarser models 194–205.
tens or hundreds of microseconds, at length scales of ca. 10 × 10 nm [17] G. Bao, S. Mitragotri, S. Tong, Multifunctional nanoparticles for drug delivery and
for atomistic and up to 150 × 150 nm for CG simulations. Thus it is molecular imaging, Annu. Rev. Biomed. Eng. 15 (2013) 253–282.
[18] C. Sun, J.S.H. Lee, M. Zhang, Magnetic nanoparticles in MR imaging and drug deliv-
now possible to simulate entire nanocarriers, which will open up new ery, Adv. Drug Deliv. Rev. 60 (2008) 1252–1265.
areas of study for simulations. However, the process of NP formation [19] X. Sun, Z. Feng, T. Hou, Y. Li, Computational simulation of inorganic nanoparticle
by e.g. microfluidics remains outside reach of direct detailed simula- drug delivery systems at the molecular level, Computational Pharmaceutics: Appli-
cation of Molecular Modelling in Drug Delivery, John Wiley & Sons 2015,
tions. Questions regarding drug loading, the stability of complexes,
pp. 149–168.
and the interactions of NPs with membranes essentially involve free en- [20] C.S. Kim, G.Y. Tonga, D. Solfiell, V.M. Rotello, Inorganic nanosystems for therapeutic
ergies of binding, partitioning, and membrane pore formation. These delivery: status and prospects, Adv. Drug Deliv. Rev. 65 (2013) 93–99.
can all be studied by free energy methods, but the current state of the [21] R. Duncan, The dawning era of polymer therapeutics, Nat. Rev. Drug Discov. 2
(2003) 347–360.
literature remains somewhat qualitative. This is an obvious area for fu- [22] K.S. Soppimath, T.M. Aminabhavi, A.R. Kulkarni, W.E. Rudzinski, Biodegradable
ture work, which is already feasible with current methods and compu- polymeric nanoparticles as drug delivery devices, J. Control. Release 70 (2001)
tational resources. Some of the most interesting simulations currently 1–20.
[23] L. Huynh, C. Neale, R. Pomès, C. Allen, Computational approaches to the rational de-
available use CG MD or DPD. This is an exciting development, but also sign of nanoemulsions, polymeric micelles, and dendrimers for drug delivery,
comes with its own limitations; in particular, CG models need to be Nanomed. Nanotechnol. Biol. Med. 8 (2012) 20–36.
carefully tested for their ability to reproduce the effect of important fac- [24] W.F. van Gunsteren, D. Bakowies, R. Baron, I. Chandrasekhar, M. Christen, X. Daura,
et al., Biomolecular modeling: goals, problems, perspectives, Angew. Chem. Int. Ed.
tors in DDSs such as salt concentration, the distribution of chemical 45 (2006) 4064–4092.
functional groups over the surface of a DDS, and the effect of pH. [25] B. Kann, H.L. Offerhaus, M. Windbergs, C. Otto, Raman microscopy for cellular in-
DDSs are an interesting example of multi-scale computational prob- vestigations — from single cell imaging to drug carrier uptake visualization, Adv.
Drug Deliv. Rev. 89 (2015) 71–90.
lems, with a growing amount of data from biophysical characterization. [26] A. Sharma, U.S. Sharma, Liposomes in drug delivery: progress and limitations, Int. J.
They have an obvious biomedical and biotechnological importance, and Pharm. 154 (1997) 123–140.
bridge chemistry, nanoscience, materials, basic biology and medical ap- [27] R.W. Niven, M. Speer, H. Schreier, Nebulization of liposomes. II. The effects of
size and modeling of solute release profiles, Pharm. Res. 8 (1991) 217–221.
plications in a unique way. Although there already is a large body of lit-
[28] S.K. Hobbs, W.L. Monsky, F. Yuan, W.G. Roberts, L. Griffith, V.P. Torchilin, et al., Reg-
erature as reviewed in this paper, in our assessment the impact of ulation of transport pathways in tumor vessels: role of tumor type and microenvi-
computational work in this important area is likely to grow significantly ronment, Proc. Natl. Acad. Sci. U. S. A. 95 (1998) 4607–4612.
in the near future as model systems become more realistic and more [29] C.R. Dass, Drug delivery in cancer using liposomes, Drug Delivery Systems,
Humana Press 2008, pp. 177–182.
tightly coupled to experiment. [30] P.Y. Lu, M.C. Woodle, Delivering small interfering RNA for novel therapeutics, Drug
Delivery Systems, Humana Press 2008, pp. 93–107.
[31] R. Singh, J.W. Lillard, Nanoparticle-based targeted drug delivery, Exp. Mol. Pathol.
Conflict of interest
86 (2009) 215–223.
[32] S.H. Jang, M.G. Wientjes, D. Lu, J.L.S. Au, Drug delivery and transport to solid tu-
The authors declare no conflicts of interest. mors, Pharm. Res. 20 (2003) 1337–1350.
[33] A.R. Kirtane, R.A. Siegel, J. Panyam, A pharmacokinetic model for quantifying the ef-
fect of vascular permeability on the choice of drug carrier: a framework for person-
Acknowledgments alized nanomedicine, J. Pharm. Sci. 104 (2015) 1174–1186.
[34] S.K. Brar, M. Verma, Measurement of nanoparticles by light-scattering techniques,
Trends Anal. Chem. 30 (2011) 4–17.
This work was supported by a Natural Sciences and Engineering Coun-
[35] W. Mehnert, K. Mäder, Solid lipid nanoparticles: production, characterization and
cil (Canada) Strategic Project Grant (DPT, JT) (STPGP/463247-2014). DPT applications, Adv. Drug Deliv. Rev. 64 (2012) 83–101.
is an Alberta Innovates Health Solutions Scientist and Alberta Innovates [36] S.A. Kübart, C.M. Keck, Laser diffractometry of nanoparticles: frequent pitfalls &
Technology Futures Strategic Chair in (Bio)Molecular Simulation. overlooked opportunities, J. Pharmacol. Technol. Drug Res. 2 (2013) 17.
[37] A. Moquin, F.M. Winnik, The use of field-flow fractionation for the analysis of drug
and gene delivery systems, Field-Flow Fractionation in Biopolymer Analysis,
References Springer Vienna 2012, pp. 187–206.
[38] A. Zattoni, B. Roda, F. Borghi, V. Marassi, P. Reschiglian, Flow field-flow fraction-
[1] E. Blanco, H. Shen, M. Ferrari, Principles of nanoparticle design for overcoming bi- ation for the analysis of nanoparticles used in drug delivery, J. Pharm. Biomed.
ological barriers to drug delivery, Nat. Biotechnol. 33 (2015) 941–951. Anal. 87 (2014) 53–61.
[2] T.M. Allen, P.R. Cullis, Drug delivery systems: entering the mainstream, Science 303 [39] F.M. Veronese, G. Pasut, PEGylation, successful approach to drug delivery, Drug
(2004) 1818–1822. Discov. Today 10 (2005) 1451–1458.
[3] S. Parveen, R. Misra, S.K. Sahoo, Nanoparticles: a boon to drug delivery, therapeu- [40] N.S. Berchane, K.H. Carson, A.C. Rice-Ficht, M.J. Andrews, Effect of mean diameter
tics, diagnostics and imaging, Nanomed. Nanotechnol. Biol. Med. 8 (2012) and polydispersity of PLG microspheres on drug release: experiment and theory,
147–166. Int. J. Pharm. 337 (2007) 118–126.
[4] D.S. Kohane, Microparticles and nanoparticles for drug delivery, Biotechnol. [41] C.M. Beddoes, C.P. Case, W.H. Briscoe, Understanding nanoparticle cellular
Bioeng. 96 (2007) 203–209. entry: a physicochemical perspective, Adv. Colloid Interf. Sci. 218C (2015)
[5] V.J. Venditto, F.C. Szoka Jr., Cancer nanomedicines: so many papers and so few 48–68.
drugs! Adv. Drug Deliv. Rev. 65 (2013) 80–88. [42] M. Müllner, S.J. Dodds, T.-H. Nguyen, D. Senyschyn, C.J.H. Porter, B.J. Boyd, et al.,
[6] Y. Zhang, H.F. Chan, K.W. Leong, Advanced materials and processing for drug deliv- Size and rigidity of cylindrical polymer brushes dictate long circulating properties
ery: the past and the future, Adv. Drug Deliv. Rev. 65 (2013) 104–120. in vivo, ACS Nano 9 (2015) 1294–1304.
[7] B.S. Pattni, V.V. Chupin, V.P. Torchilin, New developments in liposomal drug deliv- [43] S.E.A. Gratton, P.A. Ropp, P.D. Pohlhaus, J.C. Luft, V.J. Madden, M.E. Napier, et al., The
ery, Chem. Rev. 115 (2015) 10938–10966. effect of particle design on cellular internalization pathways, Proc. Natl. Acad. Sci.
[8] T.M. Allen, P.R. Cullis, Liposomal drug delivery systems: from concept to clinical ap- U. S. A. 105 (2008) 11613–11618.
plications, Adv. Drug Deliv. Rev. 65 (2013) 36–48. [44] J.A. Champion, S. Mitragotri, Role of target geometry in phagocytosis, Proc. Natl.
[9] E.R. Gillies, J.M.J. Frechet, Dendrimers and dendritic polymers in drug delivery, Acad. Sci. U. S. A. 103 (2006) 4930–4934.
Drug Discov. Today 10 (2005) 35–43. [45] S. Fusco, H.W. Huang, K.E. Peyer, C. Peters, M. Häberli, A. Ulbers, et al., Shape-
[10] C.C. Lee, J.A. MacKay, J.M.J. Fréchet, F.C. Szoka, Designing dendrimers for biological switching microrobots for medical applications: the influence of shape in drug
applications, Nat. Biotechnol. 23 (2005) 1517–1526. delivery and locomotion, ACS Appl. Mater. Interfaces 7 (2015) 6803–6811.
[11] S. Svenson, D.A. Tomalia, Dendrimers in biomedical applications—reflections on [46] G. Ruan, S.S. Feng, Preparation and characterization of poly(lactic acid)–poly(ethyl-
the field, Adv. Drug Deliv. Rev. 57 (2005) 2106–2129. ene glycol)–poly(lactic acid) (PLA–PEG–PLA) microspheres for controlled release
[12] A. Bianco, K. Kostarelos, M. Prato, Applications of carbon nanotubes in drug deliv- of paclitaxel, Biomaterials 24 (2003) 5037–5044.
ery, Curr. Opin. Chem. Biol. 9 (2005) 674–679. [47] E. Fattal, H. Hillaireau, S. Mura, J. Nicolas, N. Tsapis, Targeted delivery using biode-
[13] L. Lacerda, A. Bianco, M. Prato, K. Kostarelos, Carbon nanotubes as nanomedicines: gradable polymeric nanoparticles, Fundamentals and Applications of Controlled
from toxicology to pharmacology, Adv. Drug Deliv. Rev. 58 (2006) 1460–1470. Release Drug Delivery, Springer US 2012, pp. 255–288.
1704 M. Ramezanpour et al. / Biochimica et Biophysica Acta 1858 (2016) 1688–1709

[48] J.L.S. Milne, M.J. Borgnia, A. Bartesaghi, E.E.H. Tran, L.A. Earl, D.M. Schauder, et al., catalysis, molecular electronics, photonics, and nanomedicine, Chem. Rev. 110
Cryo-electron microscopy —a primer for the non-microscopist, FEBS J. 280 (2010) 1857–1959.
(2013) 28–45. [79] A.K.K. Leung, I.M. Hafez, S. Baoukina, N.M. Belliveau, I.V. Zhigaltsev, E.
[49] A. Bartesaghi, A. Merk, S. Banerjee, D. Matthies, X. Wu, J.L.S. Milne, et al., 2.2 Å res- Afshinmanesh, et al., Lipid nanoparticles containing siRNA synthesized by
olution cryo-EM structure of β-galactosidase in complex with a cell-permeant in- microfluidic mixing exhibit an electron-dense nanostructured core, J. Phys.
hibitor, Science 348 (2015) 1147–1151. Chem. C 116 (2012) 18440–18450.
[50] T. Turovsky, R. Khalfin, S. Kababya, A. Schmidt, Y. Barenholz, D. Danino, Celecoxib [80] E.V. Mathias, X. Liu, O. Franco, I. Khan, Y. Ba, J.A. Kornfield, Model of drug-loaded
encapsulation in β-casein micelles: structure, interactions, and conformation, fluorocarbon-based micelles studied by electron-spin induced 19F relaxation
Langmuir 31 (2015) 7183–7192. NMR and molecular dynamics simulation, Langmuir 24 (2008) 692–700.
[51] M. Zheng, G.M. Pavan, M. Neeb, A.K. Schaper, A. Danani, G. Klebe, et al., Targeting [81] G. Martini, L. Ciani, Electron spin resonance spectroscopy in drug delivery, Phys.
the blind spot of polycationic nanocarrier-based siRNA delivery, ACS Nano 6 Chem. Chem. Phys. 11 (2009) 211–254.
(2012) 9447–9454. [82] H. Wennerström, D.F. Evans, The Colloidal Domain: Where Physics, Chemistry, Bi-
[52] A.M. Alkilany, C.J. Murphy, Toxicity and cellular uptake of gold nanoparticles: what ology, and Technology Meet, New York, Wiley-VCH, 1999.
we have learned so far? J. Nanoparticle Res. 12 (2010) 2313–2333. [83] C. Nicolini, J. Kraineva, M. Khurana, N. Periasamy, S.S. Funari, R. Winter, Tem-
[53] K. Murugan, Y.E. Choonara, P. Kumar, D. Bijukumar, L.C. du Toit, V. Pillay, Parame- perature and pressure effects on structural and conformational properties of
ters and characteristics governing cellular internalization and trans-barrier traf- POPC/SM/cholesterol model raft mixtures—an FT-IR, SAXS, DSC, PPC and
ficking of nanostructures, Int. J. Nanomedicine 10 (2015) 2191–2206. Laurdan fluorescence spectroscopy study, Biochim. Biophys. Acta 1758
[54] E. Fröhlich, The role of surface charge in cellular uptake and cytotoxicity of medical (2006) 248–258.
nanoparticles, Int. J. Nanomedicine 7 (2012) 5577–5591. [84] T.T. Mills, S. Tristram-Nagle, F.A. Heberle, N.F. Morales, J. Zhao, J. Wu, et al., Liquid–
[55] M. Di Marco, C. Sadun, M. Port, I. Guilbert, P. Couvreur, C. Dubernet, Physicochem- liquid domains in bilayers detected by wide angle X-ray scattering, Biophys. J. 95
ical characterization of ultrasmall superparamagnetic iron oxide particles (USPIO) (2008) 682–690.
for biomedical application as MRI contrast agents, Int. J. Nanomedicine 2 (2007) [85] V. Meli, C. Caltagirone, A.M. Falchi, S.T. Hyde, V. Lippolis, M. Monduzzi, et al.,
609–622. Docetaxel-loaded fluorescent liquid-crystalline nanoparticles for cancer
[56] K. Xiao, Y. Li, J. Luo, J.S. Lee, W. Xiao, A.M. Gonik, et al., The effect of surface charge theranostics, Langmuir 31 (2015) 9566–9575.
on in vivo biodistribution of PEG-oligocholic acid based micellar nanoparticles, Bio- [86] A. Angelova, B. Angelov, R. Mutafchieva, S. Lesieur, P. Couvreur, Self-assembled
materials 32 (2011) 3435–3446. multicompartment liquid crystalline lipid carriers for protein, peptide and nucleic
[57] L.P. Mendes, J.M.F. Delgado, A.D.A. Costa, M.S. Vieira, P.L. Benfica, E.M. Lima, et al., acid drug delivery, Acc. Chem. Res. 44 (2010) 147–156.
Biodegradable nanoparticles designed for drug delivery: the number of nanoparti- [87] R.H. Utama, M. Dulle, S. Förster, M.H. Stenzel, P.B. Zetterlund, SAXS analysis of shell
cles impacts on cytotoxicity, Toxicol. in Vitro 29 (2015) 1268–1274. formation during nanocapsule synthesis via inverse miniemulsion periphery RAFT
[58] M. Law, M. Jafari, P. Chen, Physicochemical characterization of SiRNA-peptide com- polymerization, Macromol. Rapid Commun. 1267-1271 (2015).
plexes, Biotechnol. Prog. 24 (2008) 957–963. [88] R.H. Müller, S.A. Runge, V. Ravelli, A.F. Thünemann, W. Mehnert, E.B. Souto,
[59] A.M. Chen, M. Zhang, D. Wei, D. Stueber, O. Taratula, T. Minko, et al., Co-delivery of Cyclosporine-loaded solid lipid nanoparticles (SLN®): drug–lipid physicochemical
doxorubicin and bcl-2 siRNA by mesoporous silica nanoparticles enhances the ef- interactions and characterization of drug incorporation, Eur. J. Pharm. Biopharm.
ficacy of chemotherapy in multidrug-resistant cancer cells, Small 5 (2009) 68 (2008) 535–544.
2673–2677. [89] A.F. Martins, A.G.B. Pereira, A.R. Fajardo, A.F. Rubira, E.C. Muniz, Characterization of
[60] X.Y. Wang, L. Zhang, X.H. Wei, Q. Wang, Molecular dynamics of paclitaxel encapsu- polyelectrolytes complexes based on N,N,N-trimethyl chitosan/heparin prepared
lated by salicylic acid-grafted chitosan oligosaccharide aggregates, Biomaterials 34 at different pH conditions, Carbohydr. Polym. 86 (2011) 1266–1272.
(2013) 1843–1851. [90] J. Pan, F.A. Heberle, R.S. Petruzielo, J. Katsaras, Using small-angle neutron
[61] B. Sun, D.T. Chiu, Determination of the encapsulation efficiency of individual vesi- scattering to detect nanoscopic lipid domains, Chem. Phys. Lipids 170-171
cles using single-vesicle photolysis and confocal single-molecule detection, Anal. (2013) 19–32.
Chem. 77 (2005) 2770–2776. [91] J.R. Lakowicz, Principles of Fluorescence SpectroscopySpringer Sceince & Business
[62] N. Martinho, H. Florindo, L. Silva, S. Brocchini, M. Zloh, T. Barata, Molecular model- Media 2006.
ing to study dendrimers for biomedical applications, Molecules 19 (2014) [92] K. Kataoka, A. Harada, Y. Nagasaki, Block copolymer micelles for drug delivery:
20424–20467. design, characterization and biological significance, Adv. Drug Deliv. Rev. 64
[63] H. Bunjes, T. Unruh, Characterization of lipid nanoparticles by differential scanning (2012) 37–48.
calorimetry, X-ray and neutron scattering, Adv. Drug Deliv. Rev. 59 (2007) [93] S.R. Pygall, J. Whetstone, P. Timmins, C.D. Melia, Pharmaceutical applications of
379–402. confocal laser scanning microscopy: the physical characterisation of pharmaceuti-
[64] W. Mehnert, Solid lipid nanoparticles: production, characterization and applica- cal systems, Adv. Drug Deliv. Rev. 59 (2007) 1434–1452.
tions, Adv. Drug Deliv. Rev. 47 (2001) 165–196. [94] C. Dubernet, Thermoanalysis of microspheres, Thermochim. Acta 248 (1995)
[65] M. Deborah, A. Jawahar, T. Mathavan, M.K. Dhas, A.M.F. Benial, Spectroscopic stud- 259–269.
ies on sidewall carboxylic acid functionalization of multi-walled carbon nanotubes [95] D. Giron, Applications of thermal analysis and coupled techniques in pharmaceuti-
with valine, Spectrochim. Acta Part A. 139 (2015) 138–144. cal industry, J. Therm. Anal. Calorim. 68 (2002) 335–357.
[66] P.S. Williams, F. Carpino, M. Zborowski, Magnetic nanoparticle drug carriers and [96] H. Reinl, T. Brumm, T.M. Bayerl, The physical properties of the liquid-ordered phase
their study by quadrupole magnetic field-flow fractionation, Molecules 6 (2009) with temperature in binary mixtures of DPPC with cholesterol: A 2H-NMR, FT-IR,
1290–1306. DSC, and neutron scattering study, Biophys. J. 61 (1992) 1025–1035.
[67] M.L. Briuglia, C. Rotella, A. McFarlane, D.A. Lamprou, Influence of cholesterol on [97] N.R. Labiris, M.B. Dolovich, Pulmonary drug delivery. Part I: Physiological factors
liposome stability and on in vitro drug release, Drug Deliv. Transl. Res. 5 (2015) affecting therapeutic effectiveness of aerosolized medications, Br. J. Clin.
231–242. Pharmacol. 56 (2003) 588–599.
[68] J. Wang, Y. Wang, W. Liang, Delivery of drugs to cell membranes by encapsulation [98] K. Bouchemal, New challenges for pharmaceutical formulations and drug delivery
in PEG–PE micelles, J. Control. Release 160 (2012) 637–651. systems characterization using isothermal titration calorimetry, Drug Discov.
[69] N.W.S. Kam, Z. Liu, H. Dai, Functionalization of carbon nanotubes via cleavable Today 13 (2008) 960–972.
disulfide bonds for efficient intracellular delivery of siRNA and potent gene silenc- [99] R. Gref, A. Domb, P. Quellec, T. Blunk, R.H. Müller, J.M. Verbavatz, et al., The con-
ing, J. Am. Chem. Soc. 127 (2005) 12492–12493. trolled intravenous delivery of drugs using PEG-coated sterically stabilized nano-
[70] S. Achar, R.J. Puddephatt, Organoplatinum dendrimers formed by oxidative addi- spheres, Adv. Drug Deliv. Rev. 64 (2012) 316–326.
tion, Angew. Chem. Int. Ed. 33 (1994) 847–849. [100] K. Avgoustakis, A. Beletsi, Z. Panagi, P. Klepetsanis, A.G. Karydas, D.S. Ithakissios,
[71] M.L. Viger, W. Sheng, C.L. McFearin, M.Y. Berezin, A. Almutairi, Application of time- PLGA-mPEG nanoparticles of cisplatin: in vitro nanoparticle degradation, in vitro
resolved fluorescence for direct and continuous probing of release from polymeric drug release and in vivo drug residence in blood properties, J. Control. Release 79
delivery vehicles, J. Control. Release 171 (2013) 308–314. (2002) 123–135.
[72] A.A. Metwally, R.M. Hathout, Computer-assisted drug formulation design: novel [101] B.L. Mui, Y.K. Tam, M. Jayaraman, S.M. Ansell, X. Du, Y.Y.C. Tam, et al., Influence of
approach in drug delivery, Mol. Pharm. 12 (2015) 2800–2810. polyethylene glycol lipid desorption rates on pharmacokinetics and pharmacody-
[73] M.B. Sankaram, T.E. Thompson, Interaction of cholesterol with various namics of siRNA lipid nanoparticles, Mol. Ther. Acids 2 (2013) e139.
glycerophospholipids and sphingomyelin, Biochemistry 29 (1990) [102] H. Tan, Z. Wang, J. Li, Z. Pan, M. Ding, Q. Fu, An approach for the sphere-to-rod tran-
10670–10675. sition of multiblock copolymer micelles, ACS Macro Lett. 2 (2013) 146–151.
[74] A. Caminade, R. Laurent, J. Majoral, Characterization of dendrimers, Adv. Drug [103] A. Almutairi, W.J. Akers, M.Y. Berezin, S. Achilefu, J.M.J. Frechet, Monitoring the bio-
Deliv. Rev. 57 (2005) 2130–2146. degradation of dendritic near-infrared nanoprobes by in vivo fluorescence imag-
[75] Y.W. Hsueh, M. Zuckermann, J. Thewalt, Phase diagram determination for phos- ing, Mol. Pharm. 5 (2008) 1103–1110.
pholipid/sterol membranes using deuterium NMR, Concepts Magn. Reson. Part A. [104] X. Dai, Z. Yue, M.E. Eccleston, J. Swartling, N.K.H. Slater, C.F. Kaminski, Fluorescence
26A (2005) 35–46. intensity and lifetime imaging of free and micellar-encapsulated doxorubicin in
[76] T. Io, T. Fukami, K. Yamamoto, T. Suzuki, J. Xu, K. Tomono, et al., Homogeneous living cells, Nanomed. Nanotechnol. Biol. Med. 4 (2008) 49–56.
nanoparticles to enhance the efficiency of a hydrophobic drug, antihyperlipidemic [105] D. Frenkel, B. Smit, Understanding Molecular Simulation: From Algorithms to Ap-
probucol, characterized by solid-state NMR, Mol. Pharm. 7 (2009) 299–305. plications, Academic Press, 2001.
[77] T. Sun, Q. Guo, C. Zhang, J. Hao, P. Xing, J. Su, et al., Self-assembled vesicles pre- [106] B. Hess, C. Kutzner, D. Van Der Spoel, E. Lindahl, GROMACS 4: algorithms for highly
pared from amphiphilic cyclodextrins as drug carriers, Langmuir 28 (2012) efficient, load-balanced and scalable molecular simulation, J. Chem. Theory
8625–8636. Comput. 4 (2008) 435–447.
[78] D. Astruc, E. Boisselier, C. Ornelas, Dendrimers designed for functions: from phys- [107] J.C. Phillips, R. Braun, W. Wang, J. Gumbart, E. Tajkhorshid, E. Villa, et al., Scalable
ical, photophysical, and supramolecular properties to applications in sensing, molecular dynamics with NAMD, J. Comput. Chem. 26 (2005) 1781–1802.
M. Ramezanpour et al. / Biochimica et Biophysica Acta 1858 (2016) 1688–1709 1705

[108] S. Plimpton, Fast parallel algorithms for short-range molecular dynamics, J. [142] R. Vácha, F.J. Martinez-Veracoechea, D. Frenkel, Receptor-mediated endocytosis of
Comput. Phys. 117 (1995) 1–19. nanoparticles of various shapes, Nano Lett. 11 (2011) 5391–5395.
[109] B.R. Brooks, C.L. Brooks, A.D. MacKerell, L. Nilsson, R.J. Petrella, B. Roux, et al., [143] X. Lin, Y. Li, N. Gu, Nanoparticle's size effect on its translocation across a lipid
CHARMM: the biomolecular simulation program, J. Comput. Chem. 30 (2009) bilayer: a molecular dynamics simulation, J. Comput. Theor. Nanosci. 7
1545–1614. (2010) 269–276.
[110] D.A. Case, T.E. Cheatham, T. Darden, H. Gohlke, R. Luo, K.M. Merz, et al., The Amber [144] E.L. da Rocha, G.F. Caramori, C.R. Rambo, Nanoparticle translocation through a lipid
biomolecular simulation programs, J. Comput. Chem. 26 (2005) 1668–1688. bilayer tuned by surface chemistry, Phys. Chem. Chem. Phys. 15 (2013)
[111] S.J. Marrink, D.P. Tieleman, Perspective on the Martini model, Chem. Soc. Rev. 42 2282–2290.
(2013) 6801–6822. [145] R.C. Van Lehn, A. Alexander-Katz, Penetration of lipid bilayers by nanoparticles
[112] M. Karplus, J.A. McCammon, Molecular dynamics simulations of biomolecules, Nat. with environmentally-responsive surfaces: simulations and theory, Soft Matter 7
Struct. Mol. Biol. 9 (2002) 646–652. (2011) 11392–11404.
[113] W.F. van Gunsteren, J. Dolenc, A.E. Mark, Molecular simulation as an aid to exper- [146] J. Lin, H. Zhang, Z. Chen, Y. Zheng, Penetration of lipid membranes by gold nano-
imentalists, Curr. Opin. Struct. Biol. 18 (2008) 149–153. particles: insights into cellular uptake, cytotoxicity, and their relationship, ACS
[114] M. Velinova, D. Sengupta, A.V. Tadjer, S.J. Marrink, Sphere-to-rod transitions of Nano 4 (2010) 5421–5429.
nonionic surfactant micelles in aqueous solution modeled by molecular dynamics [147] H.M. Ding, Y.Q. Ma, Role of physicochemical properties of coating ligands in
simulations, Langmuir 27 (2011) 14071–14077. receptor-mediated endocytosis of nanoparticles, Biomaterials 33 (2012)
[115] W. Shinoda, R. DeVane, M.L. Klein, Computer simulation studies of self-assembling 5798–5802.
macromolecules, Curr. Opin. Struct. Biol. 22 (2012) 175–186. [148] Y. Li, Computer investigations of influences of molar fraction and acyl chain length
[116] P.V. Messina, J. Miguel Besada-Porto, J.M. Ruso, Self-assembly drugs: from micelles of lipids on the nanoparticle–biomembrane interactions, RSC Adv. 5 (2015)
to nanomedicine, Curr. Top. Med. Chem 14 (2014) 555–571. 11049–11057.
[117] B. Aoun, V.K. Sharma, E. Pellegrini, S. Mitra, M. Johnson, R. Mukhopadhyay, Struc- [149] H.M. Ding, W. Tian, Y.Q. Ma, Designing nanoparticle translocation through mem-
ture and dynamics of ionic micelles: MD simulation and neutron scattering branes by computer simulations, ACS Nano 6 (2012) 1230–1238.
study, J. Phys. Chem. B. 119 (2015) 5079–5086. [150] S. Wang, E.E. Dormidontova, Selectivity of ligand–receptor interactions between
[118] L. Vukovic, F.A. Khatib, S.P. Drake, A. Madriaga, K.S. Brandenburg, P. Král, et al., nanoparticle and cell surfaces, Phys. Rev. Lett. 109 (2012) 238102.
Structure and dynamics of highly PEG-ylated sterically stabilized micelles in aque- [151] L. Zhang, M. Becton, X. Wang, Designing nanoparticle translocation through cell
ous media, J. Am. Chem. Soc. 133 (2011) 13481–13488. membranes by varying amphiphilic polymer coatings, J. Phys. Chem. B. 119
[119] J. Gupta, C. Nunes, S. Vyas, S. Jonnalagadda, Prediction of solubility parameters and (2015) 3786–3794.
miscibility of pharmaceutical compounds by molecular dynamics simulations, J. [152] Y. Li, X. Zhang, D. Cao, A spontaneous penetration mechanism of patterned nano-
Phys. Chem. B. 115 (2011) 2014–2023. particles across a biomembrane, Soft Matter 10 (2014) 6844–6856.
[120] S.M. Loverde, M.L. Klein, D.E. Discher, Nanoparticle shape improves delivery: ratio- [153] T. Yue, X. Zhang, Molecular understanding of receptor-mediated membrane re-
nal coarse grain molecular dynamics (rCG-MD) of taxol in worm-like PEG-PCL mi- sponses to ligand-coated nanoparticles, Soft Matter 7 (2011) 9104–9112.
celles, Adv. Mater. 24 (2012) 3823–3830. [154] L.F. Barraza, V.A. Jiménez, J.B. Alderete, Effect of PEGylation on the structure and
[121] L. Vuković, A. Madriaga, A. Kuzmis, A. Banerjee, A. Tang, K. Tao, et al., Solubilization drug loading capacity of PAMAM-G4 dendrimers: a molecular modeling approach
of therapeutic agents in micellar nanomedicines, Langmuir 29 (2013) on the complexation of 5-Fluorouracil with native and PEGylated PAMAM-G4,
15747–15754. Macromol. Chem. Phys. 216 (2015) 1689–1701.
[122] L. Huynh, C. Neale, R. Pomès, C. Allen, Systematic design of unimolecular star copol- [155] O.S. Lee, G.C. Schatz, Computational simulations of the interaction of lipid mem-
ymer micelles using molecular dynamics simulations, Soft Matter 6 (2010) branes with DNA-functionalized gold nanoparticles, Biomedical Nanotechnolgy,
5491–5501. Springer 2011, pp. 283–296.
[123] R. Vácha, F.J. Martinez-Veracoechea, D. Frenkel, Intracellular release of [156] Q. Mu, T. Hu, J. Yu, Molecular insight into the steric shielding effect of PEG on the
endocytosed nanoparticles upon a change of ligand–receptor interaction, ACS conjugated staphylokinase: biochemical characterization and molecular dynamics
Nano 6 (2012) 10598–10605. simulation, PLoS One 8 (2013) (e68559).
[124] W. Tian, Y.Q. Ma, Insights into the endosomal escape mechanism via investigation [157] H. Lee, Molecular modeling of PEGylated peptides, dendrimers, and single-walled
of dendrimer–membrane interactions, Soft Matter 8 (2012) 6378–6384. carbon nanotubes for biomedical applications, Polymers 6 (2014) 776–798.
[125] N. Arai, K. Yasuoka, X.C. Zeng, A vesicle cell under collision with a Janus or homo- [158] A. Bunker, Molecular modeling as a tool to understand the role of poly (ethylene)
geneous nanoparticle: translocation dynamics and late-stage morphology, Nano- glycol in drug delivery, Computational Pharmaceutics: Application of Molecular
scale 5 (2013) 9089–9100. Modelling in Drug Delivery, John Wiley & Sons 2015, pp. 217–233.
[126] A.J. Makarucha, N. Todorova, I. Yarovsky, Nanomaterials in biological environment: [159] Y. Li, M. Kröger, W.K. Liu, Endocytosis of PEGylated nanoparticles accompanied by
a review of computer modelling studies, Eur. Biophys. J. 40 (2011) 103–115. structural and free energy changes of the grafted polyethylene glycol, Biomaterials
[127] Y.C. Li, S. Rissanen, M. Stepniewski, O. Cramariuc, T. Róg, S. Mirza, et al., Study of 35 (2014) 8467–8478.
interaction between PEG carrier and three relevant drug molecules: piroxicam, [160] K. Knop, R. Hoogenboom, D. Fischer, U.S. Schubert, Poly(ethylene glycol) in drug
paclitaxel, and hematoporphyrin, J. Phys. Chem. B. 116 (2012) 7334–7341. delivery: pros and cons as well as potential alternatives, Angew. Chem. Int. Ed.
[128] F.J. Martinez-Veracoechea, D. Frenkel, Designing super selectivity in multivalent 49 (2010) 6288–6308.
nano-particle binding, Proc. Natl. Acad. Sci. U. S. A. 108 (2011) 10963–10968. [161] J.K. Armstrong, G. Hempel, S. Koling, L.S. Chan, T. Fisher, H.J. Meiselman, et al., An-
[129] S. Wang, E.E. Dormidontova, Nanoparticle design optimization for enhanced tibody against poly(ethylene glycol) adversely affects PEG-asparaginase therapy in
targeting: Monte Carlo simulations, Biomacromolecules 11 (2010) 1785–1795. acute lymphoblastic leukemia patients, Cancer 110 (2007) 103–111.
[130] L. Liu, S. Parameswaran, A. Sharma, S.M. Grayson, H.S. Ashbaugh, S.W. Rick, Molec- [162] L. Zhang, Z. Wang, Z. Lu, H. Shen, J. Huang, Q. Zhao, et al., PEGylated reduced
ular dynamics simulations of linear and cyclic amphiphilic polymers in aqueous graphene oxide as a superior ssRNA delivery system, J. Mater. Chem. B. 1 (2013)
and organic environments, J. Phys. Chem. B. 118 (2014) 6491–6497. 749–755.
[131] L. Lai, A.S. Barnard, Functionalized nanodiamonds for biological and medical appli- [163] X. Sun, Z. Feng, L. Zhang, T. Hou, Y. Li, The selective interaction between silica nano-
cations, J. Nanosci. Nanotechnol. 15 (2015) 989–999. particles and enzymes from molecular dynamics simulations, PLoS One 9 (2014)
[132] T. Panczyk, T.P. Warzocha, P.J. Camp, A magnetically controlled molecular (e107696).
nanocontainer as a drug delivery system: the effects of carbon nanotube and mag- [164] A. Adnan, R. Lam, H. Chen, J. Lee, D.J. Schaffer, A.S. Barnard, et al., Atomistic simula-
netic nanoparticle parameters from monte carlo simulations, J. Phys. Chem. C 114 tion and measurement of pH dependent cancer therapeutic interactions with
(2010) 21299–21308. nanodiamond carrier, Mol. Pharm. 8 (2011) 368–374.
[133] T. Panczyk, A. Jagusiak, G. Pastorin, W.H. Ang, J. Narkiewicz-Michalek, Molecular [165] L. Lai, A.S. Barnard, Molecular and analytical modeling of nanodiamond for drug
dynamics study of cisplatin release from carbon nanotubes capped by magnetic delivery applications, Computational Pharmaceutics: Application of Molecular
nanoparticles, J. Phys. Chem. C 117 (2013) 17327–17336. Modelling in Drug Delivery, John Wiley & Sons 2015, pp. 169–196.
[134] H.M. Ding, Y.Q. Ma, Computer simulation of the role of protein corona in cellular [166] M.L. Bello, A.M. Junior, B.A. Vieira, L.R.S. Dias, V.P. de Sousa, H.C. Castro, et al., Sodi-
delivery of nanoparticles, Biomaterials 35 (2014) 8703–8710. um montmorillonite/amine-containing drugs complexes: new insights on interca-
[135] G. Brancolini, D.B. Kokh, L. Calzolai, R.C. Wade, S. Corni, Docking of ubiquitin to gold lated drugs arrangement into layered carrier material, PLoS One 10 (2015)
nanoparticles, ACS Nano 6 (2012) 9863–9878. (e0121110).
[136] D. Meneksedag-Erol, T. Tang, H. Uludağ, Probing the effect of miRNA on siRNA–PEI [167] V. Murthy, Z.P. Xu, S.C. Smith, Molecular modeling of layered double hydroxide
polyplexes, J. Phys. Chem. B. 119 (2015) 5475–5486. nanoparticles for drug delivery, Computational Pharmaceutics: Application of Mo-
[137] A.F. Jorge, R.S. Dias, A.A. Pais, Enhanced condensation and facilitated release of DNA lecular Modelling in Drug Delivery, John Wiley & Sons 2015, pp. 197–216.
using mixed cationic agents: a combined experimental and Monte Carlo study, [168] P. Cox, Atomistic modelling of drug delivery systems, Drug Design Strategies: Com-
Biomacromolecules 13 (2012) 3151–3161. putational Techniques and Applications, Royal Society of Chemistry 2012,
[138] F. Tian, T. Yue, Y. Li, X. Zhang, Computer simulation studies on the interactions pp. 210–232.
between nanoparticles and cell membrane, Sci. China Chem. 57 (2014) [169] D.G. Fatouros, D. Douroumis, V. Nikolakis, S. Ntais, A.M. Moschovi, V. Trivedi, et al.,
1662–1671. In vitro and in silico investigations of drug delivery via zeolite BEA, J. Mater. Chem.
[139] H.M. Ding, Y.Q. Ma, Theoretical and computational investigations of nanoparticle- 21 (2011) 7789–7794.
biomembrane interactions in cellular delivery, Small 11 (2015) 1055–1071. [170] M. Spanakis, N. Bouropoulos, D. Theodoropoulos, L. Sygellou, S. Ewart, A.M.
[140] Y. Li, X. Zhang, D. Cao, Nanoparticle hardness controls the internalization pathway Moschovi, et al., Controlled release of 5-fluorouracil from microporous zeolites,
for drug delivery, Nanoscale 7 (2015) 2758–2769. Nanomed. Nanotechnol. Biol. Med. 10 (2014) 197–205.
[141] Y. Li, T. Yue, K. Yang, X. Zhang, Molecular modeling of the relationship between [171] M.C. Bernini, D. Fairen-Jimenez, M. Pasinetti, A.J. Ramirez-Pastor, R.Q. Snurr,
nanoparticle shape anisotropy and endocytosis kinetics, Biomaterials 33 (2012) Screening of bio-compatible metal–organic frameworks as potential drug carriers
4965–4973. using Monte Carlo simulations, J. Mater. Chem. B. 2 (2014) 766–774.
1706 M. Ramezanpour et al. / Biochimica et Biophysica Acta 1858 (2016) 1688–1709

[172] S.T. Meek, J.A. Greathouse, M.D. Allendorf, Metal-organic frameworks: a [200] R. Guo, J. Mao, L.T. Yan, Unique dynamical approach of fully wrapping dendrimer-
rapidly growing class of versatile nanoporous materials, Adv. Mater. 23 like soft nanoparticles by lipid bilayer membrane, ACS Nano 7 (2013)
(2011) 249–267. 10646–10653.
[173] G.M. Pavan, Modeling the interaction between dendrimers and nucleic acids: a [201] X. He, Z. Qu, F. Xu, M. Lin, J. Wang, X. Shi, et al., Molecular analysis of interactions
molecular perspective through hierarchical scales, ChemMedChem 9 (2014) between dendrimers and asymmetric membranes at different transport stages,
2623–2631. Soft Matter 10 (2014) 139–148.
[174] V. Jain, P.V. Bharatam, Pharmacoinformatic approaches to understand complexa- [202] C.K. Tu, K. Chen, W.D. Tian, Y.Q. Ma, Computational investigations of a peptide-
tion of dendrimeric nanoparticles with drugs, Nanoscale 6 (2014) 2476–2501. modified dendrimer interacting with lipid membranes, Macromol. Rapid
[175] D. Shcharbin, A. Shakhbazau, M. Bryszewska, Poly(amidoamine) dendrimer com- Commun. 34 (2013) 1237–1242.
plexes as a platform for gene delivery, Expert Opin. Drug Deliv. 10 (2013) [203] Y.Q. Ma, pH-responsive dendrimers interacting with lipid membranes, Soft Matter
1687–1698. 8 (2012) 2627–2632.
[176] P. Posocco, E. Laurini, V. Dal Col, D. Marson, K. Karatasos, M. Fermeglia, et al., Tell [204] L.Q. Xie, J.W. Feng, W.D. Tian, Y.Q. Ma, Generation-dependent gel-fluid phase tran-
me something I do not know. Multiscale molecular modeling of dendrimer/ sition of membrane caused by a PAMAM dendrimer, Macromol. Theory Simul. 23
dendron organization and self-assembly in gene therapy, Curr. Med. Chem. 19 (2014) 531–536.
(2012) 5062–5087. [205] R.G. Bellini, A.P. Guimarães, M.A. Pacheco, D.M. Dias, V.R. Furtado, R.B. De
[177] D. Meneksedag-Erol, T. Tang, H. Uludağ, Molecular modeling of polynucleotide Alencastro, et al., Association of the anti-tuberculosis drug rifampicin with a
complexes, Biomaterials 35 (2014) 7068–7076. PAMAM dendrimer, J. Mol. Graph. Model. 60 (2015) 34–42.
[178] M.R. Molla, P. Rangadurai, G.M. Pavan, S. Thayumanavan, Experimental and theo- [206] V. Jain, V. Maingi, P.K. Maiti, P.V. Bharatam, Molecular dynamics simulations of PPI
retical investigations in stimuli responsive dendrimer-based assemblies, Nanoscale dendrimer–drug complexes, Soft Matter 9 (2013) 6482–6496.
7 (2015) 3817–3837. [207] Y. Jiang, D. Zhang, Y. Zhang, Z. Deng, L. Zhang, The adsorption–desorption transi-
[179] E. Ficici, I. Andricioaei, S. Howorka, Dendrimers in nanoscale confinement: the in- tion of double-stranded DNA interacting with an oppositely charged dendrimer in-
terplay between conformational change and nanopore entrance, Nano Lett. 15 duced by multivalent anions, J. Chem. Phys. 140 (2014) 204912.
(2015) 4822–4828. [208] X.F. Wen, J.L. Lan, Z.Q. Cai, P.H. Pi, S.P. Xu, L.J. Zhang, et al., Dissipative particle dy-
[180] Y.L. Wang, Z.Y. Lu, A. Laaksonen, Specific binding structures of dendrimers on lipid namics simulation on drug loading/release in polyester-PEG dendrimer, J. Nano-
bilayer membranes, Phys. Chem. Chem. Phys. 14 (2012) 8348–8359. particle Res. 16 (2014) 2403.
[181] B. Nandy, P.K. Maiti, A. Bunker, Force biased molecular dynamics simulation study [209] V. Maingi, M.V.S. Kumar, P.K. Maiti, PAMAM dendrimer–drug interactions: effect of
of effect of dendrimer generation on interaction with DNA, J. Chem. Theory pH on the binding and release pattern, J. Phys. Chem. B. 116 (2012) 4370–4376.
Comput. 9 (2013) 722–729. [210] F. Avila-Salas, C. Sandoval, J. Caballero, S. Guiñez-Molinos, L.S. Santos, R.E. Cachau,
[182] G.M. Pavan, A. Danani, The influence of dendrons architecture on the “rigid” and et al., Study of interaction energies between the PAMAM dendrimer and nonsteroi-
“flexible” behaviour in binding DNA—a modelling study, Phys. Chem. Chem. dal anti-inflammatory drug using a distributed computational strategy and exper-
Phys. 12 (2010) 13914–13917. imental analysis by ESI-MS/MS, J. Phys. Chem. B. 116 (2012) 2031–2039.
[183] G.M. Pavan, L. Albertazzi, A. Danani, Ability to adapt: different generations of [211] J. Lim, G.M. Pavan, O. Annunziata, E.E. Simanek, Experimental and computational
PAMAM dendrimers show different behaviors in binding siRNA, J. Phys. Chem. B. evidence for an inversion in guest capacity in high-heneration triazine dendrimer
114 (2010) 2667–2675. hosts, J. Am. Chem. Soc. 134 (2012) 1942–1945.
[184] P. Posocco, S. Pricl, S. Jones, A. Barnard, D.K. Smith, Less is more — multiscale [212] D. Ouyang, H. Zhang, H.S. Parekh, S.C. Smith, The effect of pH on PAMAM
modelling of self-assembling multivalency and its impact on DNA binding and dendrimer–siRNA complexation—endosomal considerations as determined by
gene delivery, Chem. Sci. 1 (2010) 393–404. molecular dynamics simulation, Biophys. Chem. 158 (2011) 126–133.
[185] S.P. Jones, N.P. Gabrielson, C.H. Wong, H.F. Chow, D.W. Pack, P. Posocco, et al., [213] B. Nandy, P.K. Maiti, DNA compaction by a dendrimer, J. Phys. Chem. B. 115 (2010)
Hydrophobically modified dendrons: developing structure–activity relationships 217–230.
for DNA binding and gene transfection, Mol. Pharm. 8 (2011) 416–429. [214] D. Amado Torres, M. Garzoni, A.V. Subrahmanyam, G.M. Pavan, S. Thayumanavan,
[186] S.P. Jones, G.M. Pavan, A. Danani, S. Pricl, D.K. Smith, Quantifying the effect of sur- Protein-triggered supramolecular disassembly: insights based on variations in li-
face ligands on dendron–DNA interactions : insights into multivalency through a gand location in amphiphilic dendrons, J. Am. Chem. Soc. 136 (2014) 5385–5399.
combined experimental and theoretical approach, Chem. Eur. J. 16 (2010) [215] L.B. Jensen, G.M. Pavan, M.R. Kasimova, S. Rutherford, A. Danani, H.M. Nielsen, et al.,
4519–4532. Elucidating the molecular mechanism of PAMAM–siRNA dendriplex self-assembly:
[187] V. Percec, D.A. Wilson, P. Leowanawat, C.J. Wilson, A.D. Hughes, M.S. Kaucher, et al., effect of dendrimer charge density, Int. J. Pharm. 416 (2011) 410–418.
Self-assembly of Janus dendrimers into uniform dendrimersomes and other com- [216] K. Karatasos, P. Posocco, E. Laurini, S. Pricl, Poly (amidoamine)-based dendrimer/
plex architectures, Science 328 (2010) 1009–1014. siRNA complexation studied by computer simulations: effects of pH and genera-
[188] S. Zhang, H.J. Sun, A.D. Hughes, R.O. Moussodia, A. Bertin, Y. Chen, et al., Self-assembly tion on dendrimer structure and siRNA binding, Macromol. Biochem. 12 (2012)
of amphiphilic Janus dendrimers into uniform onion-like dendrimersomes with pre- 225–240.
dictable size and number of bilayers, Proc. Natl. Acad. Sci. U. S. A. 111 (2014) [217] G.M. Pavan, S. Monteagudo, J. Guerra, B. Carrion, V. Ocana, J. Rodriguez-Lopez, et al.,
9058–9063. Role of generation, architecture, pH and ionic strength on successful siRNA delivery
[189] S. Kavyani, S. Amjad-Iranagh, H. Modarress, Aqueous poly(amidoamine) dendri- and transfection by hybrid PPV-PAMAM dendrimers, Curr. Med. Chem. 19 (2012)
mer G3 and G4 generations with several interior cores at pHs 5 and 7: a molecular 4929–4941.
dynamics simulation study, J. Phys. Chem. B. 118 (2014) 3257–3266. [218] G.M. Pavan, M.A. Mintzer, E.E. Simanek, O.M. Merkel, T. Kissel, A. Danani, Compu-
[190] A. Barnard, P. Posocco, S. Pricl, M. Calderon, R. Haag, M.E. Hwang, et al., Degrad- tational insights into the interactions between DNA and siRNA with “rigid” and
able self-assembling dendrons for gene delivery: experimental and theoretical “flexible” triazine dendrimers, Biomacromolecules 11 (2010) 721–730.
insights into the barriers to cellular uptake, J. Am. Chem. Soc. 133 (2011) [219] G.M. Pavan, P. Posocco, A. Tagliabue, M. Maly, A. Malek, A. Danani, et al., PAMAM
20288–20300. dendrimers for siRNA delivery: computational and experimental insights, Chem.
[191] V. Carrasco-Sánchez, A. Vergara-Jaque, M. Zuñiga, J. Comer, A. John, F.M. Nachtigall, Eur. J. 16 (2010) 7781–7795.
et al., In situ and in silico evaluation of amine- and folate-terminated dendrimers as [220] C. Shi, D. Guo, K. Xiao, X. Wang, L. Wang, J. Luo, A drug-specific nanocarrier design
nanocarriers of anesthetics, Eur. J. Med. Chem. 73 (2014) 250–257. for efficient anticancer therapy, Nat. Commun. 6 (2015) 7449.
[192] C. Shi, D. Yuan, S. Nangia, G. Xu, K.S. Lam, J. Luo, A structure–property relationship [221] M. Kang, D. Lam, D.E. Discher, S.M. Loverde, Molecular modeling of block copoly-
study of the well-defined telodendrimers to improve hemocompatibility of mer self-assembly and micellar drug delivery, Computational Pharmaceutics: Ap-
nanocarriers for anticancer drug delivery, Langmuir 30 (2014) 6878–6888. plication of Molecular Modelling in Drug Delivery, John Wiley & Sons 2015,
[193] R.M. Pearson, N. Patra, H.J. Hsu, S. Uddin, P. Král, S. Hong, Positively charged pp. 53–80.
dendron micelles display negligible cellular interactions, ACS Macro Lett. 2 [222] A. Rösler, G.W.M. Vandermeulen, H.A. Klok, Advanced drug delivery devices via
(2012) 77–81. self-assembly of amphiphilic block copolymers, Adv. Drug Deliv. Rev. 64 (2012)
[194] H. Lee, J.S. Choi, R.G. Larson, Molecular dynamics studies of the size and internal 270–279.
structure of the PAMAM dendrimer grafted with arginine and histidine, Macro- [223] C. Sun, T. Tang, Study on the role of polyethylenimine as gene delivery carrier using
molecules 44 (2011) 8681–8686. molecular dynamics simulations, J. Adhes. Sci. Technol. 28 (2014) 399–416.
[195] H. Lee, R.G. Larson, Membrane pore formation induced by acetylated and polyeth- [224] S.M. Loverde, Computer simulation of polymer and biopolymer self-assembly for
ylene glycol-conjugated polyamidoamine dendrimers, J. Phys. Chem. C 115 (2011) drug delivery, Mol. Simul. 40 (2014) 794–801.
5316–5322. [225] D. Meneksedag-Erol, C. Sun, T. Tang, H. Uludag, Molecular dynamics simulations of
[196] L. Yang, S.R.P. Rocha, PEGylated, NH2-terminated PAMAM dendrimers: A micro- polyplexes and lipoplexes employed in gene delivery, Intracellular Delivery II,
scopic view from atomistic computer simulations, Mol. Pharm. 11 (2014) Springer 2014, pp. 277–311.
1459–1470. [226] H. Guo, X. Qiu, J. Zhou, Self-assembled core–shell and Janus microphase sepa-
[197] L. Albertazzi, F.M. Mickler, G.M. Pavan, F. Salomone, G. Bardi, M. Panniello, et al., rated structures of polymer blends in aqueous solution, J. Chem. Phys. 139
Enhanced bioactivity of internally functionalized cationic dendrimers with PEG (2013) 84907.
cores, Biomacromolecules 13 (2012) 4089–4097. [227] C. Cai, L. Wang, J. Lin, X. Zhang, Morphology transformation of hybrid micelles self-
[198] H. Lee, R.G. Larson, Effects of PEGylation on the size and internal structure of assembled from rod–coil block copolymer and nanoparticles, Langmuir 28 (2012)
dendrimers: self-penetration of long PEG chains into the dendrimer core, Macro- 4515–4524.
molecules 44 (2011) 2291–2298. [228] H.Y. Chang, Y.L. Lin, Y.J. Sheng, H.K. Tsao, Structural characteristics and fusion path-
[199] R. Bhattacharya, S. Kanchi, C. Roobala, A. Lakshminarayanan, O.H. Seeck, P.K. Maiti, ways of onion-like multilayered polymersome formed by amphiphilic comb-like
K.G. Ayappa, N. Jayaraman, J.K. Basu, A new microscopic insight into membrane graft copolymers, Macromolecules 46 (2013) 5644–5656.
penetration and reorganization by PETIM dendrimers, Soft Matter 10 (2014) [229] Y. Sheng, J. An, Y. Zhu, Self-assembly of ABA triblock copolymers under soft con-
7577–7587. finement, Chem. Phys. 452 (2015) 46–52.
M. Ramezanpour et al. / Biochimica et Biophysica Acta 1858 (2016) 1688–1709 1707

[230] H. Chen, E. Ruckenstein, Formation and degradation of multicomponent multicore [260] S.K. Patel, A. Lavasanifar, P. Choi, Molecular dynamics study of the encapsulation
micelles: insights from dissipative particle dynamics simulations, Langmuir 29 capability of a PCL-PEO based block copolymer for hydrophobic drugs with differ-
(2013) 5428–5434. ent spatial distributions of hydrogen bond donors and acceptors, Biomaterials 31
[231] V. Taresco, L. Gontrani, F. Crisante, I. Francolini, A. Martinelli, L. D'Ilario, et al., Self- (2010) 1780–1786.
assembly of catecholic moiety-containing cationic random acrylic copolymers, J. [261] S.S. Thakur, H.S. Parekh, C.H. Schwable, Y. Gan, D. Ouyang, Solubilization of poorly
Phys. Chem. B. 119 (2015) 8369–8379. soluble drugs, Computational Pharmaceutics: Application of Molecular Modelling
[232] S.Y. Nie, Y. Sun, W.J. Lin, W.S. Wu, X.D. Guo, Y. Qian, et al., Dissipative particle in Drug Delivery, John Wiley & Sons 2015, pp. 31–51.
dynamics studies of doxorubicin-loaded micelles assembled from four-arm star [262] S. Bagai, C. Sun, T. Tang, Potential of mean force of polyethylenimine-mediated
triblock polymers 4AS-PCL-b-PDEAEMA-b-PPEGMA and their pH-release mecha- DNA attraction, J. Phys. Chem. B. 117 (2012) 49–56.
nism, J. Phys. Chem. B. 117 (2013) 13688–13697. [263] H.S. Antila, M. Härkönen, M. Sammalkorpi, Chemistry specificity of DNA–
[233] T. Wang, C. Chipot, X. Shao, W. Cai, Structural characterization of micelles formed polycation complex salt response: a simulation study of DNA, polylysine and
of cholesteryl-functionalized cyclodextrins, Langmuir 27 (2010) 91–97. polyethyleneimine, Phys. Chem. Chem. Phys. 17 (2015) 5279–5289.
[234] L.X. Peng, A. Ivetac, A.S. Chaudhari, S. Van, G. Zhao, L. Yu, et al., Characterization of a [264] P. Pereira, A.F. Jorge, R. Martins, A.A. Pais, F. Sousa, A. Figueiras, Characterization of
clinical polymer-drug conjugate using multiscale modeling, Biopolymers 93 polyplexes involving small RNA, J. Colloid Interface Sci. 387 (2012) 84–94.
(2010) 936–951. [265] C. Sun, T. Tang, H. Uludag, Molecular dynamics simulations for complexation of
[235] P. Pi, D. Qin, J. Lan, Z. Cai, X. Yuan, S. Xu, et al., Dissipative particle dynamics simu- DNA with 2 kDa PEI reveal profound effect of PEI architecture on complexation,
lation on the nanocomposites delivery system of quantum dots and poly (styrene- J. Phys. Chem. B. 116 (2012) 2405–2413.
b-ethylene oxide) copolymer, Ind. Eng. Chem. Res. 54 (2015) 6123–6134. [266] D. Ouyang, H. Zhang, D. Herten, H.S. Parekh, S.C. Smith, Structure, dynamics and
[236] J. Li, Y. Ouyang, X. Kong, J. Zhu, D. Lu, Z. Liu, A multi-scale molecular dynamics energetics of siRNA-cationic vector complexation: a molecular dynamics study, J.
simulation of PMAL facilitated delivery of siRNA, RSC Adv. 5 (2015) Phys. Chem. B. 114 (2010) 9220–9230.
68227–68233. [267] D. Ouyang, H. Zhang, H.S. Parekh, S.C. Smith, Structure and dynamics of multiple
[237] G. Srinivas, R.V. Mohan, A.D. Kelkar, Polymer micelle assisted transport and deliv- cationic vectors — siRNA complexation by all-atomic molecular dynamics simula-
ery of model hydrophilic components inside a biological lipid vesicle: a coarse- tions, J. Phys. Chem. B. 114 (2010) 9231–9237.
grain simulation study, J. Phys. Chem. B. 117 (2013) 12095–12104. [268] A.F. Jorge, R.F.P. Pereira, S.C.C. Nunes, A.J.M. Valente, R.S. Dias, A.A.C.C. Pais,
[238] H.M. Ding, Y.Q. Ma, Controlling cellular uptake of nanoparticles with pH-sensitive Interpreting the rich behavior of ternary DNA-PEI-Fe(III) complexes,
polymers, Sci. Rep. 3 (2013) 2804. Biomacromolecules 15 (2014) 478–491.
[239] C. Sun, T. Tang, H. Uludag, A molecular dynamics simulation study on the effect of [269] B. Zhan, K. Shi, Z. Dong, W. Lv, S. Zhao, X. Han, et al., Coarse-grained simulation of
lipid substitution on polyethylenimine mediated siRNA complexation, Biomate- polycation/DNA-like complexes: role of neutral block, Mol. Pharm. 12 (2015)
rials 34 (2013) 2822–2833. 2834–2844.
[240] Z.X. Liao, S.F. Peng, Y.C. Ho, F.L. Mi, B. Maiti, H.W. Sung, Mechanistic study of [270] M. Zhao, J. Zhou, C. Su, L. Niu, D. Liang, B. Li, Complexation behavior of oppositely
transfection of chitosan/DNA complexes coated by anionic poly(γ-glutamic charged polyelectrolytes: effect of charge distribution, J. Chem. Phys. 142 (2015)
acid), Biomaterials 33 (2012) 3306–3315. 204902.
[241] S.F. Peng, M.T. Tseng, Y.C. Ho, M.C. Wei, Z.X. Liao, H.W. Sung, Mechanisms of cellu- [271] J. Ziebarth, Y. Wang, Coarse-grained molecular dynamics simulations of DNA con-
lar uptake and intracellular trafficking with chitosan/DNA/poly(γ-glutamic acid) densation by block copolymer and formation of core-corona structures, J. Phys.
complexes as a gene delivery vector, Biomaterials 32 (2011) 239–248. Chem. B. 114 (2010) 6225–6232.
[242] M. Kepczynski, D. Jamróz, M. Wytrwal, J. Bednar, E. Rzad, M. Nowakowska, Interac- [272] Q. Luo, Y. Wang, H. Yang, C. Liu, Y. Ding, H. Xu, et al., Modeling the interaction of
tions of a hydrophobically modified polycation with zwitterionic lipid membranes, interferon α-1b to bovine serum albumin as a drug delivery system, J. Phys.
Langmuir 28 (2011) 676–688. Chem. B. 118 (2014) 8566–8574.
[243] Y.L. Chiu, Y.C. Ho, Y.M. Chen, S.F. Peng, C.J. Ke, K.J. Chen, et al., The characteristics, [273] R. Vijayaraj, S. Van Damme, P. Bultinck, V. Subramanian, Theoretical studies on the
cellular uptake and intracellular trafficking of nanoparticles made of transport mechanism of 5-fluorouracil through cyclic peptide based nanotubes,
hydrophobically-modified chitosan, J. Control. Release 146 (2010) 152–159. Phys. Chem. Chem. Phys. 15 (2013) 1260–1270.
[244] C. Sun, T. Tang, H. Uludağ, Probing the effects of lipid substitution on polycation [274] H. Liu, J. Chen, Q. Shen, W. Fu, W. Wu, Molecular insights on the cyclic peptide
mediated DNA aggregation: a molecular dynamics simulations study, nanotube-mediated transportation of antitumor drug 5-fluorouracil, Mol. Pharm.
Biomacromolecules 13 (2012) 2982–2988. 7 (2010) 1985–1994.
[245] S. Samanta, D. Roccatano, Interaction of curcumin with PEO–PPO–PEO block copol- [275] J.Ł. Durzyńska, Ł. Przysiecka, R. Nawrot, J. Barylski, G. Nowicki, A. Warowicka, et al.,
ymers: a molecular dynamics study, J. Phys. Chem. B. 117 (2013) 3250–3257. Viral and other cell-penetrating peptides as vectors of therapeutic agents in med-
[246] J. He, C. Chipot, X. Shao, W. Cai, Cyclodextrin-mediated recruitment and delivery of icine, J. Pharmacol. Exp. Ther. 354 (2015) 32–42.
amphotericin B, J. Phys. Chem. C 117 (2013) 11750–11756. [276] Y. Krishnan, F.C. Simmel, Nucleic acid based molecular devices, Angew. Chem. Int.
[247] P. Shan, J.W. Shen, D.H. Xu, L.Y. Shi, J. Gao, Y.W. Lan, et al., Molecular dynamics Ed. 50 (2011) 3124–3156.
study on the interaction between doxorubicin and hydrophobically modified chi- [277] C.K. McLaughlin, G.D. Hamblin, H.F. Sleiman, Supramolecular DNA assembly, Chem.
tosan oligosaccharide, RSC Adv. 4 (2014) 23730–23739. Soc. Rev. 40 (2011) 5647–5656.
[248] M. Subashini, P.V. Devarajan, G.S. Sonavane, M. Doble, Molecular dynamics simula- [278] J.I. Cutler, E. Auyeung, C.A. Mirkin, Spherical nucleic acids, J. Am. Chem. Soc. 134
tion of drug uptake by polymer, J. Mol. Model. 17 (2011) 1141–1147. (2012) 1376–1391.
[249] R. Anand, S. Ottani, F. Manoli, I. Manet, S. Monti, A close-up on doxorubicin binding [279] S.N. Barnaby, T.L. Sita, S.H. Petrosko, A.H. Stegh, C.A. Mirkin, Therapeutic applica-
to γ-cyclodextrin: an elucidating spectroscopic, photophysical and conformational tions of spherical nucleic acids, Nanotechnology-Based Precision Tools for the De-
study, RSC Adv. 2 (2012) 2346–2357. tection and Treatment of Cancer, Springer 2015, pp. 23–50.
[250] R. Namgung, Y.M. Lee, J. Kim, Y. Jang, B.H. Lee, I.S. Kim, et al., Poly-cyclodextrin and [280] K.A. Afonin, W.K. Kasprzak, E. Bindewald, M. Kireeva, M. Viard, M. Kashlev, B.A.
poly-paclitaxel nano-assembly for anticancer therapy, Nat. Commun. 5 (2014) Shapiro, In silico design and enzymatic synthesis of functional RNA nanoparticles,
3702. Acc. Chem. Res. 47 (2014) 1731–1741.
[251] Z. Wang, J. Jiang, Dissipative particle dynamics simulation on paclitaxel loaded [281] S.R. Badu, R. Melnik, M. Paliy, S. Prabhakar, A. Sebetci, B.A. Shapiro, Modeling of
PEO–PPO–PEO block copolymer micelles, J. Nanosci. Nanotechnol. 14 (2014) RNA nanotubes using molecular dynamics simulation, Eur. Biophys. J. 43 (2014)
2644–2647. 555–564.
[252] Z. Wang, F. Wang, C. Su, Y. Zhang, Computer simulation of polymer delivery [282] S. Juul, F. Iacovelli, M. Falconi, S.L. Kragh, B. Christensen, R. Frøhlich, et al.,
system by dissipative particle dynamics, J. Comput. Theor. Nanosci. 10 (2013) Temperature-controlled encapsulation and release of an active enzyme in the cav-
2323–2327. ity of a self-assembled DNA nanocage, ACS Nano 7 (2013) 9724–9734.
[253] F. Razmimanesh, S. Amjad-Iranagh, H. Modarress, Molecular dynamics simulation [283] A.O. Elzoghby, W.M. Samy, N.A. Elgindy, Protein-based nanocarriers as promising
study of chitosan and gemcitabine as a drug delivery system, J. Mol. Model. 21 drug and gene delivery systems, J. Control. Release 161 (2012) 38–49.
(2015) 165. [284] A.O. Elzoghby, W.M. Samy, N.A. Elgindy, Albumin-based nanoparticles as potential
[254] S.Y. Nie, W.J. Lin, N. Yao, X.D. Guo, L.J. Zhang, Drug release from pH-sensitive poly- controlled release drug delivery systems, J. Control. Release 157 (2012) 168–182.
meric micelles with different drug distributions: insight from coarse-grained sim- [285] M. Pourmousa, M. Karttunen, Early stages of interactions of cell-penetrating pep-
ulations, ACS Appl. Mater. Interfaces 6 (2014) 17668–17678. tide penetratin with a DPPC bilayer, Chem. Phys. Lipids 169 (2013) 85–94.
[255] Z. Luo, J. Jiang, pH-sensitive drug loading/releasing in amphiphilic copolymer PAE- [286] M. Pourmousa, J. Wong-Ekkabut, M. Patra, M. Karttunen, Molecular dynamic stud-
PEG: integrating molecular dynamics and dissipative particle dynamics simula- ies of transportan interacting with a DPPC lipid bilayer, J. Phys. Chem. B. 117
tions, J. Control. Release 162 (2012) 185–193. (2012) 230–241.
[256] J. Hao, Y. Cheng, R.J.K.U. Ranatunga, S. Senevirathne, M.C. Biewer, S.O. Nielsen, et al., [287] X. He, M. Lin, B. Sha, S. Feng, X. Shi, Z. Qu, et al., Coarse-grained molecular dynamics
A combined experimental and computational study of the substituent effect on mi- studies of the translocation mechanism of polyarginines across asymmetric mem-
cellar behavior of γ-substituted thermoresponsive amphiphilic poly (ε- brane under tension, Sci. Rep. 5 (2015) 12808.
caprolactone), Macromolecules 46 (2013) 4829–4838. [288] M. Islami, F. Mehrnejad, F. Doustdar, M. Alimohammadi, M. Khadem-Maaref, M.
[257] L.S. Zheng, Y.Q. Yang, X.D. Guo, Y. Sun, Y. Qian, L.J. Zhang, Mesoscopic simulations Mir-Derikvand, et al., Study of orientation and penetration of LAH4 into lipid bilay-
on the aggregation behavior of pH-responsive polymeric micelles for drug deliv- er membranes: pH and composition dependence, Chem. Biol. Drug Des. 84 (2014)
ery, J. Colloid Interface Sci. 363 (2011) 114–121. 242–252.
[258] M.R. Rodríguez-Hidalgo, C. Soto-Figueroa, L. Vicente, Mesoscopic simulation of the [289] Z.L. Li, H.M. Ding, Y.Q. Ma, Translocation of polyarginines and conjugated nanopar-
drug release mechanism on the polymeric vehicle P(ST-DVB) in an acid environ- ticles across asymmetric membranes, Soft Matter 9 (2012) 1281–1286.
ment, Soft Matter 7 (2011) 8224–8230. [290] R.S. Teixeira, T.F. Cova, S.M. Silva, R. Oliveira, M.J. Araújo, E.F. Marques, et al., Lysine-
[259] R. Malik, J. Genzer, C.K. Hall, Proteinlike copolymers as encapsulating agents for based surfactants as chemical permeation enhancers for dermal delivery of local
small-molecule solutes, Langmuir 31 (2015) 3518–3526. anesthetics, Int. J. Pharm. 474 (2014) 212–222.
1708 M. Ramezanpour et al. / Biochimica et Biophysica Acta 1858 (2016) 1688–1709

[291] N. Todorova, C. Chiappini, M. Mager, B. Simona, I.I. Patel, M.M. Stevens, I. [322] A. Jusufi, R.H. DeVane, W. Shinoda, M.L. Klein, Nanoscale carbon particles and the
Yarovsky, Surface presentation of functional peptides in solution determines stability of lipid bilayers, Soft Matter 7 (2011) 1139–1146.
cell internalization efficiency of TAT conjugated nanoparticles, Nano Lett. 14 [323] H. Lee, Interparticle dispersion, membrane curvature, and penetration induced by
(2014) 5229–5237. single-walled carbon nanotubes wrapped with lipids and PEGylated lipids, J. Phys.
[292] Y. Li, D. Feng, X. Zhang, D. Cao, Design strategy of cell-penetrating copolymers for Chem. B. 117 (2013) 1337–1344.
high efficient drug delivery, Biomaterials 52 (2015) 171–179. [324] S. Kraszewski, A. Bianco, M. Tarek, C. Ramseyer, Insertion of short amino-
[293] A. Sanchez-Sanchez, S. Akbari, A.J. Moreno, F.L. Verso, A. Arbe, J. Colmenero, J.A. functionalized single-walled carbon nanotubes into phospholipid bilayer occurs
Pomposo, Design and preparation of single-chain nanocarriers mimicking disor- by passive diffusion, PLoS One 7 (2012) (e40703).
dered proteins for combined delivery of dermal bioactive cargos, Macromol. [325] S. Kraszewski, F. Picaud, I. Elhechmi, T. Gharbi, C. Ramseyer, How long a function-
Rapid Commun. 34 (2013) 1681–1686. alized carbon nanotube can passively penetrate a lipid membrane, Carbon 50
[294] L. Chen, T. Jiang, C. Cai, L. Wang, J. Lin, X. Cao, Polypeptide-based “smart” micelles (2012) 5301–5308.
for dual-drug delivery: a combination study of experiments and simulations, [326] H. Lee, Dispersion and bilayer interaction of single-walled carbon nanotubes mod-
Adv. Healthcare Mater. 3 (2014) 1508–1517. ulated by covalent and noncovalent PEGylation, Mol. Simul. 41 (2014) 1254–1263.
[295] E. Jabbari, X. Yang, S. Moeinzadeh, X. He, Drug release kinetics, cell uptake, and [327] A. Di Crescenzo, M. Aschi, A. Fontana, Toward a better understanding of steric sta-
tumor toxicity of hybrid VVVVVVKK peptide-assembled polylactide nanoparticles, bilization when using block copolymers as stabilizers of single-walled carbon
Eur. J. Pharm. Biopharm. 84 (2013) 49–62. nanotubes (SWCNTs) aqueous dispersions, Macromolecules 45 (2012)
[296] X.D. Guo, L.J. Zhang, Z.M. Wu, Y. Qian, Dissipative particle dynamics studies on mi- 8043–8050.
crostructure of pH-sensitive micelles for sustained drug delivery, Macromolecules [328] A. Sridhar, B. Srikanth, A. Kumar, A.K. Dasmahapatra, Coarse-grain molecular
43 (2010) 7839–7844. dynamics study of fullerene transport across a cell membrane, J. Chem. Phys. 143
[297] X. Tian, F. Sun, X.R. Zhou, S.Z. Luo, L. Chen, Role of peptide self-assembly in antimi- (2015) 24907.
crobial peptides, J. Pept. Sci. 21 (2015) 530–539. [329] E. Sarukhanyan, G. Milano, D. Roccatano, Coating mechanisms of single-walled car-
[298] I.W. Fu, C.B. Markegard, H.D. Nguyen, Solvent effects on kinetic mechanisms of self- bon nanotube by linear polyether surfactants: insights from computer simulations,
assembly by peptide amphiphiles via molecular dynamics simulations, Langmuir J. Phys. Chem. C 118 (2014) 18069–18078.
31 (2014) 315–324. [330] H. Lee, Molecular dynamics studies of PEGylated single-walled carbon nanotubes:
[299] I.W. Fu, C.B. Markegard, B.K. Chu, H.D. Nguyen, The role of electrostatics and tem- the effect of PEG size and grafting density, J. Phys. Chem. C 117 (2013)
perature on morphological transitions of hydrogel nanostructures self-assembled 26334–26341.
by peptide amphiphiles via molecular dynamics simulations, Adv. Healthcare [331] H. Shrestha, R. Bala, S. Arora, Lipid-based drug delivery systems, J. Pharm. 2014
Mater. 2 (2013) 1388–1400. (2014) 801820.
[300] N. Thota, Z. Luo, Z. Hu, J. Jiang, Self-assembly of amphiphilic peptide (AF) 6H5K15: [332] S. Mashaghi, T. Jadidi, G. Koenderink, A. Mashaghi, Lipid nanotechnology, Int. J. Mol.
coarse-grained molecular dynamics simulation, J. Phys. Chem. B. 117 (2013) Sci. 14 (2013) 4242–4282.
9690–9698. [333] A.L. Klibanov, K. Maruyama, V.P. Torchilin, L. Huang, Amphipathic
[301] R. García-Fandiño, J.R. Granja, Effect of organochloride guest molecules on the sta- polyethyleneglycols effectively prolong the circulation time of liposomes, FEBS
bility of homo/hetero self-assembled α, γ-cyclic peptide structures: a computa- Lett. 268 (1990) 235–237.
tional study toward the control of nanotube length, J. Phys. Chem. C 117 (2013) [334] M. Dzieciuch, S. Rissanen, N. Szydłowska, A. Bunker, M. Kumorek, D. Jamróz, et al.,
10143–10162. PEGylated liposomes as carriers of hydrophobic porphyrins, J. Phys. Chem. B. 119
[302] R. Bakry, R.M. Vallant, M. Najam-ul-Haq, M. Rainer, Z. Szabo, C.W. Huck, G.K. Bonn, (2015) 6646–6657.
Medicinal applications of fullerenes, Int. J. Nanomedicine 2 (2007) 639–649. [335] J. Lehtinen, A. Magarkar, M. Stepniewski, S. Hakola, M. Bergman, T. Róg, et al., Anal-
[303] M. Martincic, G. Tobias, Filled carbon nanotubes in biomedical imaging and drug ysis of cause of failure of new targeting peptide in PEGylated liposome: molecular
delivery, Expert Opin. Drug Deliv. 12 (2015) 563–581. modeling as rational design tool for nanomedicine, Eur. J. Pharm. Sci. 46 (2012)
[304] V.V. Chaban, T.I. Savchenko, S.M. Kovalenko, O.V. Prezhdo, Heat-driven release of a 121–130.
drug molecule from carbon nanotubes: a molecular dynamics study, J. Phys. Chem. [336] A. Magarkar, T. Róg, A. Bunker, Molecular dynamics simulation of PEGylated mem-
B. 114 (2010) 13481–13486. branes with cholesterol: building toward the DOXIL formulation, J. Phys. Chem. C
[305] N. Saikia, A.N. Jha, R.C. Deka, Dynamics of fullerene-mediated heat-driven release 118 (2014) 15541–15549.
of drug molecules from carbon nanotubes, J. Phys. Chem. Lett. 4 (2013) [337] K. Laasonen, M.L. Klein, Molecular dynamics simulations of the structure and ion
4126–4132. diffusion in poly(ethylene oxide), J. Chem. Soc. Faraday Trans. 91 (1995)
[306] U. Arsawang, O. Saengsawang, T. Rungrotmongkol, P. Sornmee, K. Wittayanarakul, 2633–2638.
T. Remsungnen, et al., How do carbon nanotubes serve as carriers for gemcitabine [338] F. Müller-Plathe, W.F. van Gunsteren, Computer simulation of a polymer electro-
transport in a drug delivery system? J. Mol. Graph. & Model. 29 (2011) 591–596. lyte: lithium iodide in amorphous poly(ethylene oxide), J. Chem. Phys. 103
[307] Y. Li, T. Hou, Computational simulation of drug delivery at molecular level, Curr. (1995) 4745–4756.
Med. Chem. 17 (2010) 4482–4491. [339] C. Özdemir Dinç, G. Kibarer, A. Güner, Solubility profiles of poly(ethylene glycol)/
[308] M. Chehel Amirani, T. Tang, Binding of nucleobases with graphene and carbon solvent systems. II. Comparison of thermodynamic parameters from viscosity mea-
nanotube: a review of computational studies, J. Biomol. Struct. Dyn. 33 (2015) surements, J. Appl. Polym. Sci. 117 (2010) 1100–1119.
1567–1597. [340] O. Borodin, G.D. Smith, R.L. Jaffe, Ab initio quantum chemistry and molecular dy-
[309] M. Santosh, S. Panigrahi, D. Bhattacharyya, A.K. Sood, P.K. Maiti, Unzipping and namics simulations studies of LiPF6/poly(ethylene oxide) interactions, J. Comput.
binding of small interfering RNA with single walled carbon nanotube: a platform Chem. 22 (2001) 641–654.
for small interfering RNA delivery, J. Chem. Phys. 136 (2012) 065106. [341] M. Nowakowska, K. Nawalany, M. Kepczynski, Z. Krawczyk, Novel nanostructural
[310] B. Nandy, M. Santosh, P.K. Maiti, Interaction of nucleic acids with carbon nanotubes hybride materials for photodynamic theraphy, Macromol. Symp. 279 (2009)
and dendrimers, J. Biosci. 37 (2012) 457–474. 132–137.
[311] B.D. Chen, C.L. Yang, J.S. Yang, M.S. Wang, X.G. Ma, Dynamic mechanism of HIV rep- [342] I.F. Hakem, J. Lal, M.R. Bockstaller, Binding of monovalent ions to PEO in solution:
lication inhibitor peptide encapsulated into carbon nanotubes, Curr. Appl. Phys. 13 relevant parameters and structural transitions, Macromolecules 37 (2004)
(2013) 1001–1007. 8431–8440.
[312] J.W. Shen, T. Tang, X.H. Wei, W. Zheng, T.Y. Sun, Z. Zhang, L. Liang, Q. Wang, On the [343] A. Magarkar, E. Karakas, M. Stepniewski, T. Róg, A. Bunker, Molecular dynamics
loading mechanism of ssDNA into carbon nanotubes, RSC Adv. 5 (2015) simulation of PEGylated bilayer interacting with salt ions: a model of the liposome
56896–56903. surface in the bloodstream, J. Phys. Chem. B. 116 (2012) 4212–4219.
[313] N. Wu, Q. Wang, B. Arash, Ejection of DNA molecules from carbon nanotubes, Car- [344] M. Stepniewski, M. Pasenkiewicz-Gierula, T. Róg, R. Danne, A. Orlowski, M.
bon 50 (2012) 4945–4952. Karttunen, et al., Study of PEGylated lipid layers as a model for PEGylated liposome
[314] J. Li, L. Jiang, X. Zhu, Computational studies of the binding mechanisms of fullerenes surfaces: molecular dynamics simulation and Langmuir monolayer studies, Lang-
to human serum albumin, J. Mol. Model. 21 (2015) 177. muir 27 (2011) 7788–7798.
[315] J. Wong-Ekkabut, S. Baoukina, W. Triampo, I.M. Tang, D.P. Tieleman, L. Monticelli, [345] M. Pannuzzo, D.H. De Jong, A. Raudino, S.J. Marrink, Simulation of polyethylene
Computer simulation study of fullerene translocation through lipid membranes, glycol and calcium-mediated membrane fusion, J. Chem. Phys. 140 (2014)
Nat. Nanotechnol. 3 (2008) 363–368. 124905.
[316] J. Barnoud, L. Urbini, L. Monticelli, C60 fullerene promotes lung monolayer collapse, [346] A. Magarkar, T. Róg, A. Bunker, Molecular dynamics simulation of inverse-
J. R. Soc. Interface 12 (2015) 20140931. phosphocholine lipids, J. Phys. Chem. C 118 (2014) 19444–19449.
[317] N. Nisoh, M. Karttunen, L. Monticelli, J. Wong-Ekkabut, Lipid monolayer disruption [347] N.D. Winter, R.K.J. Murphy, T.V. O′ Halloran, G.C. Schatz, Development and model-
caused by aggregated carbon nanoparticles, RSC Adv. 5 (2015) 11676–11685. ing of arsenic-trioxide-loaded thermosensitive liposomes for anticancer drug
[318] A.A. Skandani, R. Zeineldin, M. Al-Haik, Effect of chirality and length on the pene- delivery, J. Liposome Res. 21 (2011) 106–115.
trability of single-walled carbon nanotubes into lipid bilayer cell membranes, [348] E.K. Perttu, A.G. Kohli, F.C. Szoka, Inverse-phosphocholine lipids: a remix of a com-
Langmuir 28 (2012) 7872–7879. mon phospholipid, J. Am. Chem. Soc. 134 (2012) 4485–4488.
[319] H. Lee, Membrane penetration and curvature induced by single-walled carbon [349] H. Lee, R.W. Pastor, Coarse-grained model for PEGylated lipids: effect of PEGylation
nanotubes: the effect of diameter, length, and concentration, Phys. Chem. Chem. on the size and shape of self-assembled structures, J. Phys. Chem. B. 115 (2011)
Phys. 15 (2013) 16334–16340. 7830–7837.
[320] S. Baoukina, L. Monticelli, D.P. Tieleman, Interaction of pristine and functionalized car- [350] J.C. Shillcock, Spontaneous vesicle self-assembly: A mesoscopic view of membrane
bon nanotubes with lipid membranes, J. Phys. Chem. B. 117 (2013) 12113–12123. dynamics, Langmuir 28 (2011) 541–547.
[321] C.C. Chiu, W. Shinoda, R.H. DeVane, S.O. Nielsen, Effects of spherical fullerene nano- [351] W. Shinoda, D.E. Discher, M.L. Klein, S.M. Loverde, Probing the structure of
particles on a dipalmitoyl phosphatidylcholine lipid monolayer: a coarse grain PEGylated-lipid assemblies by coarse-grained molecular dynamics, Soft Matter 9
molecular dynamics approach, Soft Matter 8 (2012) 9610–9616. (2013) 11549–11556.
M. Ramezanpour et al. / Biochimica et Biophysica Acta 1858 (2016) 1688–1709 1709

[352] J.J. Janke, W.F.D. Bennett, D.P. Tieleman, Oleic acid phase behavior from molecular [363] A.C. Yang, C.I. Weng, Structural and dynamic properties of water near monolayer-
dynamics simulations, Langmuir 30 (2014) 10661–10667. protected gold clusters with various alkanethiol tail groups, J. Phys. Chem. C 114
[353] N. Dan, Nanostructured lipid carriers: effect of solid phase fraction and distribution (2010) 8697–8709.
on the release of encapsulated materials, Langmuir 30 (2014) 13809–13814. [364] G. Milano, G. Santangelo, F. Ragone, L. Cavallo, A. Di Matteo, Gold nanoparticle/
[354] N. Dan, Drug release through liposome pores, Colloids Surf. B 126 (2015) 80–86. polymer interfaces: all atom structures from molecular dynamics simulations, J.
[355] C. Vitorino, J. Almeida, L.M. Gonçalves, A.J. Almeida, J.J. Sousa, A.A.C.C. Pais, Co- Phys. Chem. C 115 (2011) 15154–15163.
encapsulating nanostructured lipid carriers for transdermal application: from [365] J. Yu, M.L. Becker, G.A. Carri, The influence of amino acid sequence and functional-
experimental design to the molecular detail, J. Control. Release 167 (2013) ity on the binding process of peptides onto gold surfaces, Langmuir 28 (2011)
301–314. 1408–1417.
[356] J.P.M. Jämbeck, E.S.E. Eriksson, A. Laaksonen, A.P. Lyubartsev, L.A. Eriksson, Molec- [366] K.H. Lee, F.M. Ytreberg, Effect of gold nanoparticle conjugation on peptide dynam-
ular dynamics studies of liposomes as carriers for photosensitizing drugs: develop- ics and structure, Entropy 14 (2012) 630–641.
ment, validation, and simulations with a coarse-grained model, J. Chem. Theory [367] R.C. Van Lehn, M. Ricci, P.H.J. Silva, P. Andreozzi, J. Reguera, K. Voïtchovsky, et al.,
Comput. 10 (2014) 5–13. Lipid tail protrusions mediate the insertion of nanoparticles into model cell mem-
[357] B. van Hoof, A.J. Markvoort, R.A. van Santen, P.A.J. Hilbers, Molecular simulation of branes, Nat. Commun. 5 (2014) 4482.
protein encapsulation in vesicle formation, J. Phys. Chem. B. 118 (2014) [368] R.C. Van Lehn, A. Alexander-Katz, Free energy change for insertion of charged,
3346–3354. monolayer-protected nanoparticles into lipid bilayers, Soft Matter 10 (2014)
[358] C.D. Novina, P.A. Sharp, The RNAi revolution, Nature 430 (2004) 161–164. 648–658.
[359] S.L. Wang, Optimizing drug delivery : Characterization of DLin-KC2-DMA/ [369] R.C. Van Lehn, A. Alexander-Katz, Structure of mixed-monolayer-protected nano-
distearoylphosphatidylserine by 31P and 2H NMR spectroscopy, Diss. Sci. Dep. particles in aqueous salt solution from atomistic molecular dynamics simulations,
Mol. Biol. Biochem. (2013). J. Phys. Chem. C 117 (2013) 20104–20115.
[360] D. Rozmanov, S. Baoukina, D.P. Tieleman, Density based visualization for molecular [370] R.C. Van Lehn, A. Alexander-Katz, Ligand-mediated short-range attraction drives
simulation, Faraday Discuss. 169 (2014) 225–243. aggregation of charged monolayer-protected gold nanoparticles, Langmuir 29
[361] A. Martín-Molina, G. Luque-Caballero, J. Faraudo, M. Quesada-Pérez, J. Maldonado- (2013) 8788–8798.
Valderrama, Adsorption of DNA onto anionic lipid surfaces, Adv. Colloid Interf. Sci. [371] S.J. Marrink, A.H. de Vries, D.P. Tieleman, Lipids on the move: simulations of mem-
206 (2014) 172–185. brane pores, domains, stalks and curves, Biochim. Biophys. Acta 1788 (2009)
[362] G. Ajnai, A. Chiu, T. Kan, C.C. Cheng, T.H. Tsai, J. Chang, Trends of gold nanoparticle- 149–168.
based drug delivery system in cancer therapy, J. Exp. Clin. Med. 6 (2014) 172–178.

You might also like