SlideShare una empresa de Scribd logo
1 de 50
Descargar para leer sin conexión
Department of Chemical and Bimolecular Engineering
CHEN90023 Chemical Engineering Research Project
Simulation of the Transmembrane
Transport of Poly(amidoamine)
Lachlan Russell 389374
Supervisors: Prof. Scales, Assoc. Prof. Li-Tang Yan
ii
Summary
The effect of modifications made to Yan and Yu’s existing dissipative particle dynamics (DPD) model
of an interacting generation five (G5) Poly(amidoamine) (PAMAM) dendrimer and a lipid bilayer were
examined.1, 2
The primitive bilayer was modified to include another double-chain lipid to simulate the
effect of ligand receptors on the transmembrane transport of dendrimer structures.
The structures formed by the interacting lipid bilayer and G5 dendrimer were analysed to see if they
resembled morphologies likely to be involved in clathrin-mediated endocytosis (CME). The effect of
changes to DPD soft-core repulsion forces acting between the dendrimer surface beads and the
receptor beads, as well as the internal dendrimer structure and the hydrophobic lipid tails, were
analysed.
It was found that self-assembly of the bilayer structure was not compromised when converting a
quarter of the simulated lipid molecules into receptor-simulating lipids. Binding and penetration were
found to be largely affected by the inclusion of an attractive interaction between receptors and ligands
coupled with changes to the magnitude of the repulsion between the bilayer lipid tails and the internal
structure of the dendrimer.
Encapsulation of the dendrimer by the bilayer was observed for a particular set of soft-core repulsion
parameters. The morphology closely matched that of a phagocytic mechanism. On the simulated
length scale such a result was unexpected. Further refinement of the model is required if CME, as it is
currently understood, is to be simulated and the unobservable dynamics conjectured. Due to
computational constraints, many of the results presented are preliminary in nature.
iii
Contents
Summary....................................................................................................................................................................................ii
1.0 Introduction.........................................................................................................................................................................4
2.0 Relevant Theory...................................................................................................................................................................5
2.1 Dissipative Particle Dynamics (DPD) ...............................................................................................................................5
2.2 Dendrimers .....................................................................................................................................................................9
2.3 Lipid Bilayers ...................................................................................................................................................................9
2.4 Transmembrane Transport: ..........................................................................................................................................10
3.0 Literature Review...............................................................................................................................................................12
3.1 A Case for Dissipative Particle Dynamics (DPD) ............................................................................................................12
3.2 Using Dissipative Particle Dynamics to Model Transmembrane Transport ..................................................................13
3.3 Ligand-Receptor Modelling...........................................................................................................................................14
3.4 Aim of Research ............................................................................................................................................................14
4.0 Materials and General Methodology.................................................................................................................................15
5.0 Results and Discussion.......................................................................................................................................................17
5.1 Self-Assembly of the Modified Lipid Bilayer..................................................................................................................17
5.2 Comparison with Yan and Yu’s Results .........................................................................................................................19
5.3 Simulation of the System with Dendrimer Heads modelled as Ligands........................................................................21
5.3.1 Initial Attempts to Generate Clathrin-mediated Endocytic Morphology ..............................................................21
5.3.2 Further Attempts to Generate Clathrin-mediated Endocytic Morphology ...........................................................24
6.0 Further Discussion .............................................................................................................................................................28
6.1 Known Shortcomings of the Model...............................................................................................................................28
6.2 Relationship of CME to Simulations..............................................................................................................................30
6.3 The Notable Shortage of Empirical Data.......................................................................................................................30
7.0 Conclusions........................................................................................................................................................................32
8.0 Suggestions for Further Work............................................................................................................................................33
8.1 Simple Immediate Revisions to the Current DPD Model ..............................................................................................33
8.2 Extension of Results......................................................................................................................................................33
9.0 References .........................................................................................................................................................................35
10.0 Appendices ......................................................................................................................................................................38
Appendix 1: Simulation Soft-core Repulsion Parameters ...................................................................................................38
Appendix 2: MATLAB Code for Bilayer Bead Density Profile...............................................................................................41
Appendix 3: Asphericity ......................................................................................................................................................43
Appendix 4: Radius of Gyration ..........................................................................................................................................45
Appendix 5: Penetrability....................................................................................................................................................47
Appendix 6: Confidence Bands ...........................................................................................................................................49
Appendix 7: Intermediate Data...........................................................................................................................................50
4
1.0 Introduction
Due to their complex nature, much of the science surrounding forces acting between nanoparticles is
not well understood.3
Dissipative particle dynamics (DPD) can be used to effectively model the
mesoscopic (10 nm - 100 mm) dynamics of interacting particles (refer to Figure 1).4, 5
DPD attempts
to incorporate the essence of many forces believed to be significant on this scale. The approximations
utilised by DPD methods allow for mesoscale simulation of many physical interactions that otherwise
could not be run at acceptable speeds.
Interactions between synthetic dendrimers and cellular membranes are significant due to the growing
use of dendrimers in medicine.6
These mesoscale interactions need to be studied in order to
synthesise drugs which allow for specific transmembrane transport whilst moderating cytotoxicity.7
DPD has been used to simulate dendritic particles interacting with lipid bilayers.1, 2, 8
Such simulations
can model the behaviour of different types of dendrimers reducing the need for costly drug testing by
synthesis and administration.
Figure 1: Schematic showing coarse-graining used to model the mesoscale. 9
Interactions between lipid bilayers and dendrimers have been studied extensively. However, modern
experimental techniques are currently unable to answer many fast-acting molecular-level mechanistic
questions.10
It is hoped, given DPD’s success with modelling other mesoscopic phenomena, that
simulations can provide additional insight until such a time when observational techniques are
sufficiently advanced to provide a more detailed description of interaction mechanisms and
morphology.
This project aims to extend previous DPD models of interacting lipid bilayers and dendrimers
developed by Yan and Yu.1, 2
Whilst amphiphilic lipid molecules comprise the bulk of cell membranes,
transport mechanisms often involve receptor proteins which are also present in the structure of the
membrane. The simulations performed incorporate primitive receptor-type molecules into Yan and
Yu’s simple bilayer. Such modifications have been utilised in other simulations however the effect of
such changes on the transmembrane transport of dendrimers such as poly(amidoamine) (PAMAM)
have yet to be rigorously studied.11-13
The effects of this change, along with the variation of DPD soft-
core repulsion parameters, was compared with Yan and Yu’s experiments, other simulation studies
and mechanisms proposed by biologists. Specifically, it was assessed whether clathrin-mediated
5
endocytic morphology could be replicated by these modifications as it has been shown that PAMAM
dendrimers and other particles of similar length scales are internalised via this mechanism.14, 15
The theory behind dissipative particle dynamics will first be introduced followed by a brief description
of dendrimers, cell membranes and transmembrane transport. Subsequently, a review of literature
concerning the use of computer simulation to model these transport processes will be presented. The
details of the modified DPD model will be covered and the results of simulations examined. Lastly,
conclusions and suggestions for further research will be presented.
2.0 Relevant Theory
2.1 Dissipative Particle Dynamics (DPD)
Hoogerbrugge and Koelman developed a coarse-grained/beaded simulation method combining
features of molecular dynamics (MD) and lattice-gas automata. Dissipative particle dynamics
simulates more quickly than MD and is more flexible than lattice-gas automata schemes.16
A reformulation of DPD was developed and applied to the modelling of bilayers by Groot and his
colleagues.17, 18, 19
The simulation equations adopted by Yan and Yu are largely consistent with the aforementioned
theory developed by Groot; these equations were utilised for the sake of consistent simulation.1
Newton’s equations of motion are solved for interacting particles:
(1)
For simplicity, the mass of particles are set to equal 1 and hence force equals acceleration. Forces
act within a dimensionless cut-off radius (rc) of 1.
The bilayer-dendrimer DPD system is governed by five forces: Conservative, Dissipative, Random,
Bond and Electrostatic. The forces acting on particle i are summed at each timestep:
(2)
Conservative force:
(3)
(4), (5), (6)
6
aij denotes the soft-core repulsion parameter between particles i and j and is related to the
compressibility of interacting particles.9
The repulsion parameter can often be related to Flory-Huggins χ parameters in the instance where
water is used as the solvent:
𝝌𝒊𝒋 ≈
𝜶 𝒊𝒋−𝜶 𝒊𝒊
𝟑.𝟐𝟕
(7)
• αii =25 is used in general DPD systems for all particle types.
• αij will be larger than 25 for strong bead-bead repulsion interactions; it will be smaller than
25 where there exists reasons for attraction between two types of beads (e.g. due to
electrostatic or entropic effects). For many of the cross terms, estimations are tested and
refined based on qualitative knowledge and simulation results. A complete set of accurate
Flory-Huggins χ-parameters often cannot be generated for complex systems with many
types of beads present.
Dissipative and Random forces:
The dissipative force and the random force act as the heat sink and source, respectively. They are
balanced to maintain a Boltzmann distribution and hence create a thermostat.
(8)
(9)
• wD
and wR
are r-dependent weight functions vanishing for r>rc. For simplicity they are chosen
to be:
(10)
• vij=vi-vj and vi denotes the velocity of bead i.
• ξij is a random number with zero mean and unit variance.
• The noise amplitude, σ, is fixed at σ =3.
• The forces are connected by Eq. (11).
(11)
• This thermostat conserves linear and angular momentum, resulting in a correct description of
hydrodynamics.
7
Bond forces:
Dendrimers and lipids are constructed by tying beads together using Hookean springs with the
potential:
(12)
For the purposes of simulation:
• The spring constant is set to ks=64.
• The unstretched length is set to l0=0.5rc.
• Hence, the average bond length is fixed.
A three-body potential acting between adjacent bead triples models hydrocarbon chain stiffness in
the lipids and dendrimer:
(13)
• The angle φ is defined by Eq. (14) where A, B denote the two bonds connecting beads i-1, i,
and i+1:
𝐜𝐨𝐬−𝟏 𝑨∙𝑩
‖𝑨‖‖𝑩‖
= 𝝓 (14)
• The bending constant ka=20 is used for both the lipid and dendrimer.
• The preferred angle, φ0, is specified as 00
and 1200
for lipids and dendrimer respectively.
From these potentials, corresponding forces can be calculated.
Electrostatic force:18
A lattice grid is constructed, over which the electrostatic field is spread. There is a trade-off made
between correct representation and fast implementation of the field. Each charged bead has charge
proportional to:
(15)
This charge is calculated for every grid node within a radius Re =1.6rc of the bead and normalised so
that the sum of all proximal charged nodes is equal to the charge of the bead.
8
The electric field is solved according to Eq. 16:
(16)
• ε, εr and εo, denote the dielectric permittivity of the medium, the vacuum and the water
respectively.
• 𝜌̅ 𝑒,𝑛 is the averaged charge density.
• ∇𝜑 𝑛 is the electric gradient at node n.
• Γ =13.87 is the coupling constant corresponding to the Bjerrum length of water at 300K.
At each timestep the electrostatic force on an ion can finally be determined by Eq. (17):
(17)
• Where qi is the charge of the ion.
Algorithm:
The velocity-Verlet algorithm is iterated for each bead at each timestep:
𝜆 = 0.65 has been found to reduce computational error accumulating whilst allowing for a relatively
large timestep to be chosen.
In the simulations, the radius of interaction, the bead mass, and the temperature are set such that
rc=m=kBT=1. A characteristic time scale can thus be defined:
(18)
9
2.2 Dendrimers
Dendrimers are snowflake-like macromolecules (refer to Figure 2). They have defined controllable
structures of known physical, chemical and biological properties. Their associated monodispersity and
reproducibility make them ideal for biological applications as structure-activity relationships can be
more accurately studied.20
They are defined by their core structure and, as additional shells
(generations) are added, approximately double in size and number of surface functional groups. The
chemical backbone and surface terminal groups have a large effect on the biological properties of
dendrimers.
Figure 2: Poly(amidoamine) (PAMAM): Structure comprises of repetitively branched amide and amine functional
groups.20
Dendrimers are increasingly being used for drug delivery as, when paired with pharmaceutically active
compounds, their defined properties can be used to improve drug specificity.21
Dendrimers are
capable of delivering drugs by many mechanisms. They can function as micelles which encapsulate
hydrophobic drugs; alternatively, the terminal functional groups of the dendrimer can covalently
attach to pharmaceutical agents.20
Dendrimers used for biomedical purposes usually have specific
ligands attached to their often cationic surface which interact with receptors on the surface of cellular
membranes.
2.3 Lipid Bilayers
Phospholipids are surfactant molecules with a hydrophilic head and two hydrophobic tails which can
self-assemble to form bilayers (refer to Figure 3). Complex lipid bilayers encapsulate cells and many
important organelles.2
Figure 3: Bilayer sheet.22
10
2.4 Transmembrane Transport:
The plasma membranes of cells serve to separate the intracellular cytoplasm from the extracellular
environment by controlling the transport of molecules across the membrane.23
Endocytosis is one of
many types of processes by which cells internalise molecules (refer to Figure 4). It involves the
deformation of the cell membrane and, subsequently, smaller membrane-bound carriers are
generated.24
Figure 4: Various methods of mass transport across cell membranes.25
Clathrin-mediated endocytosis (CME) is the chief receptor-mediated endocytic pathway.25
Hence, it is
the most widely studied endocytic mechanism. First observed by using electron microscopy, CME is
vital for such functions as nutrient uptake and receptor signalling.26
Mechanistically it involves
proteins with receptor sites binding to specific ligands, triggering the formation of ‘coated pits’ which
subsequently become fully enveloped by the cell (refer to Figure 5).24
Figure 5: Clathrin-mediated endocytosis requires specific ligand-receptor binding interactions.27
11
Different parameters affect the ability of nanoparticles to pass through the cellular membranes
including, size, shape and surface chemistry.6
Surface chemistry strongly affects the interactions
between nanoparticles and cells.24
Nonspecific binding forces caused by nanoparticle characteristics
such as cationic charge and roughness can also promote cellular uptake in conjuction with ligand-
receptor interactions.27
Consequently, it is believed that multiple transmembrane transport
mechanisms can work simultaneously (refer to Figure 6).6
Figure 6: Receptor-mediated uptake with additional nonspecific binding forces.27
12
3.0 Literature Review
In recent years, multiple studies have modelled nanoparticle lipid bilayer interactions using computer
simulation.7
This is because, despite extensive research in this area, experimentalists have largely
failed to elucidate the kinetic evolutions of these interactions. Dissipative particle dynamics (DPD) and
other coarse-graining techniques have been used widely to attempt to model these processes. Many
of these studies have concerned themselves with the effects of nanoparticle shape, size and surface
chemistry on the internalisation process.7
Few have concerned themselves with seeking to replicate
the true complexity of cell membranes instead modelling them simply as lipid bilayer sheets. Lately,
attempts have been made to model these membranes in increasingly representative ways. The effect
of the inclusion of ligand-receptor type molecules on a simulated bilayer-dendrimer system requires
further study in the wake of recent developments.12, 13, 28, 29
3.1 A Case for Dissipative Particle Dynamics (DPD)
At the nanoscale, interfacial forces dominate interactions. These include van der Waals, electrostatic,
depletion, hydration and hydrophobic forces.24
Compared with atomistic models which seek to
incorporate much of the complexity of modern physics, DPD utilises a different force potential
between interacting beads. In atomistic models, hard-core potentials such as Lennard–Jones (L–J)
potential, which incorporates van der Waals attraction and Pauli exclusion forces are used. In DPD
models, a soft-core conservative force is used.9
This soft-core repulsion is important as hard-core
modelling causes a ‘caging’ effect, where atoms undergo multiple collisions before any transport
occurs.18
Hence, models using soft repulsive spheres allow for the use of larger timesteps and length
scales.
Figure 7: Schematic illustrating the hard-core nature of Lennard–Jones potential and the soft-core nature of DPD inter-
particle potential.9
The DPD model attempts to compensate for missing physics by incorporating a dissipative force to
account for viscous effects; the molecular level randomness, which usually manifests as Brownian
motion, is modelled by the random force.9
Both DPD and MD methods can produce correct equilibrium distributions of polymer chains. However,
MD simulations do not include hydrodynamic interactions. Conversely, DPD is able to correctly
simulate compressible Navier-Stokes behaviour.30, 31
This result is important at the mesoscale. It is
currently possible to simulate novel scenarios of small dendrimers interacting with simple lipid bilayers
using atomistic MD models; however, in order to generate meaningful results at greater length and
time scales, coarse-graining is required.7
Coarse-graining enables significant speeding up of
simulations of interacting membranes and particles orders of magnitude larger.22, 32
13
The advantages and disadvantages of other coarse-graining techniques have been discussed at
length.33
Groot and Madden argue that what is chiefly required of a mesoscale physics model is that
it can reproduce thermodynamic properties on the relevant length scale. Namely, if the solubilities,
liquid compressibilities and the shape of interfaces are correct, then the model correctly represents
the physical system for the given length scale.4
Well-conditioned DPD simulations have been shown
to pass such tests.4
3.2 Using Dissipative Particle Dynamics to Model Transmembrane Transport
The fundamentals and technical subtleties of DPD are well established.4, 18, 19, 34
Traditionally DPD has
been used to model polymer-colloid systems where robust experimental results abound.4, 30, 34
Shilcock and Lipowsky were able to formulate self-assembling planar bilayers with experimentally
consistent density profiles and lateral stress distributions using the DPD model.22
These models have
been shown to describe the phase behaviour of phospholipids with accuracy.33, 35
Following the successful development of thermodynamically robust self-assembling bilayers many
have attempted to study their interactions with nanoscale particles. The modelling of the
nanoparticle-bilayer interactions have typically involved the use of the polyamidoamine (PAMAM),
which is the most commonly used dendrimer in biomedical applications, and a simple bilayer
comprised entirely of dipalmitoylphosphatidylcholine (DPPC).21
In addition to DPD, a variety of coarse-
grained and atomistic simulation models have been used for this purpose.10, 13, 36, 37
Yan and Yu used
the bilayer developed by Shilcock and Lipowsky to model interactions with cationic G5 PAMAM
dendrimers. The effect on morphology caused by changes to the soft-core repulsion parameters,
especially those concerning the outer-dendrimer hydrophilic components and the inner-dendrimer
hydrophobic components, was analysed extensively.2
Cationic dendrimers are known to be cytotoxic.37, 38
By introducing controlled surface tension into the
bilayer model, Yan and Yu were able to observe holes forming on the bilayer surface on interaction
with PAMAM dendrimers consistent with experimentally observed phenomena.1
They were
subsequently able to control the soft-core interaction parameters of the lipids used in the bilayer to
modify their shape factor (
𝜈
𝛼𝑙
< 1) to instead form vesicles of various sizes and surface tensions and
again interact these with PAMAM dendrimers.8
An attempt to better model the endocytic process was
incorporated by Guo, Mao and Yan.39
Canham-Helfrich theory was applied to incorporate contact and
membrane bending energies.40
Figure 8: DPD simulated endocytosis with incorporated membrane elasticity forces.39
14
3.3 Ligand-Receptor Modelling
The use of DPD to model endocytosis is still in its infancy. Until recently, the simplicity of the simulated
bilayers was such that meaningful comparisons to cell membranes were tenuous. Most transport
across cell membranes occurs with the involvement of receptor proteins. Very little work has been
attempted to model clathrin-mediated endocytosis using simulations as it can be difficult to
incorporate changes to the bilayer that do not compromise its self-assembling properties. Attempts
to circumvent this problem have been made by modifying the surface of interacting dendrimers to
have a proportion of surface beads exhibit a greater affinity for the bilayer.13
More direct attempts
modelled ligand-receptor interactions by embedding symmetrical plug-like receptors into a bilayer
and interacting them with spherical molecules with uniformly anchored ligand sites. CME mechanisms
resembling those proposed by biologists were generated.28
Another strategy involved the
incorporation of another type of lipid into the bilayer structure with a ligand-receptor headgroup. This
method preserved self-assembly.11-13
The addition of ligands to the model has been approached in a variety of ways. Yang and Ma model
their interacting particle as a single large bead which interacts with ligand receptors according to a
modified Lennard–Jones potential in order to incorporate the attractive force expected between
ligands and receptors.29
Other studies have all used different models to incorporate ligands into
simulations.11, 12, 29, 41
While many of these studies produce interesting findings they all fail to model
the interacting particle as anything more than solid polyhedrons or spheres.
3.4 Aim of Research
The aim of this study is to introduce a receptor-simulating double-chain lipid into the bilayer and
interact this modified bilayer with more complex particles such as PAMAM dendrimers. It has been
shown that PAMAM dendrimers are internalised primarily through clathrin-mediated endocytosis.14,
15
It is hypothesised that these structural changes to the bilayer combined with the balancing of soft-
core repulsion parameters can induce clathrin-mediated endocytic (CME) morphology. In order to test
this hypothesis a self-assembling bilayer must first be able to be generated. The stability of the DPD
model will be evaluated on comparing with Yan and Yu’s previous simulations.1, 2
Only then can it be
determined if structures resembling CME can be generated with or without the additional refinement
of some soft-core repulsion parameters. Any such refinements should be justifiable either with
empirical data or, less ideally, using qualitative arguments. Depending on the strength of these
justifications the hypothesis may be able to be supported. Consequently, it could be argued that the
simulation results might describe currently unobservable dynamics. If so, the model or its derivatives
may be able to assist in the future development of new dendrimers for drug delivery.
15
4.0 Materials and General Methodology
The DPD simulation was constructed using interacting coarse-grained molecules largely consistent
with previous work by Yan and Yu.1
The lipid bilayer was constructed by creating a coarse-grained representation of
dipalmitoylphosphatidylcholine, the phospholipid commonly used to study lipid bilayers. DPPC
consists of a phosphate group, a simple organic molecule choline and two palmitic acid chains.
Figure 9: Dipalmitoylphosphatidylcholine.17
The simulation model of DPPC is consists of a hydrophilic head with charge +1 (green), a hydrophilic
head with charge -1 (purple) a head without charge (blue) and a hydrophobic tail group (cyan). The
separation between the beads was set to 0.5rc.
Figure 10: Bead model of DPPC.1
Another type of lipid was introduced into the bilayer structure.11, 12
Primitive receptor-simulating
double-chain lipids were created using a similar structure consisting of two uncharged hydrophilic
heads (orange), another uncharged head (blue) and a hydrophobic tail group (cyan). The separation
between the beads was set to 0.5rc.
Figure 11: Bead model of primitive receptor-simulating double-chain lipid.
For each of the double-chained lipids introduced into the simulation box (approximately 2360 in total)
a random number generator gave the lipid a 25% chance of being a receptor-simulating double-chain
lipid (refer to Further Discussion). The DPD simulation was allowed to run in an aqueous environment
until a steady bilayer structure self-assembled.
16
A single G5 PAMAM dendrimer was subsequently introduced into the simulation box above the
bilayer. For illustrative purposes, the coarse-grained representation of a G1 PAMAM dendrimer is
depicted in Figure 12.
Figure 12: Bead model of G1 PAMAM dendrimer.1
The yellow beads represent the hydrophobic inner amine and amide functional groups and are
uncharged. The red beads represent the hydrophilic surface amine functional groups and carry a
charge of +1. The separation between the beads was set to 0.5rc.
Consistent with Yan and Yu’s modelling, counterions were also included in the simulation.2
Each simulation was run for more than 105
timesteps of length Δt=0.02τ in order to ensure accurate
temperature control. The approximate area per lipid in a tensionless DPPC membrane is 0.64 nm2
.
From this, the cut-off radius (rc) can be estimated to be approximately 0.7 nm. The time unit (τ) can
be related to physical time by known inplane lipid diffusion coefficients. Experiments have shown this
value to be approximately 5 μm2
/s. Relating this to Yan and Yu’s previous work τ≈7.7ns and, hence,
105
time steps equates to approximately 15 μs.2
Similarly, most of the soft-core repulsion parameters
used, (refer to Appendix 1: Table 4), were compiled by Yan and Yu from the literature and, in the
absence of sources, refined through qualitative arguments and iterative simulation.1
The introduction
of the new receptor-simulating lipids required additional interaction parameters to be generated. The
effect of varying some of these parameters, both new and old, was studied in order to attempt to
build a model that successfully simulated a bilayer-dendrimer system interacting in a way that
mimicked proposed clathrin-mediated endocytic mechanisms (refer to Results and Discussion).
The Fortran code used to perform these DPD simulations has been made available (refer also to
Theory: Dissipative Particle Dynamics (DPD)). Similarly, the MATLAB programs written to perform
subsequent analysis of binary output from these programs can be found in the appendices.
17
5.0 Results and Discussion
5.1 Self-Assembly of the Modified Lipid Bilayer
In an attempt to ensure self-assembly and an even distribution of receptors on the surface of the
bilayer; parameters for the new receptor beads were set such that their affinity for the other
headgroup beads was equivalent to their affinity for one another (α=25). As per other headgroup
beads, the receptor beads were designed to be similarly repelled by the hydrophobic lipid chains
(α=40).1
So as to achieve the slight protrusion of these receptors above the bilayer (refer to Figures 5,
6) receptor beads were given smaller repulsion parameters than other headgroup beads to govern
their interactions with water (α=15) (refer to Appendix 1: Table 5).
Consistent with expectations, for these carefully selected soft-core repulsion parameters, the
introduction of the new lipid molecule had no effect on the self-assembling property of the bilayer. It
was observed that an even surface distribution of this new coarse-grained molecule was achieved
(refer to Figure 13). The total lipid density was determined to be 1.47𝑟𝑐
2
=0.72 nm2
approximately 13%
greater than Yan and Yu’s literature value of 0.64 nm2
.1
This minor difference was due to a change to
the ‘target lipid density’ input variable in the code used to simulate the self-assembly of the bilayer.
The simulation box is continuous in the sense that beads cannot leave the box and are simply
transported to the other side if they move across a boundary. The targeted density is achieved by the
coded insertion or removal of more lipids once the bilayer structure is relatively stable. Simulation of
the steady-state self-assembly of a bilayer required significant computer processing. This minor
change to the ‘target lipid density’ was performed at the request of a colleague due to limited
computing resources in order to test the effect of such a change on the stability of the bilayer
structure. This increase in density lead to slightly less exposure of the lipid tails when compared to the
surfaces originally constructed by Yan and Yu.1, 2
This deviation was partially offset by the slight
protrusion of receptor beads into the water phase.
Figure 13: Self-assembling bilayer produced with 1:3 ratio of receptor-simulating double-chain lipids and DPPC lipids.
18
Figure 14: Bilayer bead density by displacement from bilayer beads’ average z-coordinate (refer to Appendix 2).
As illustrated by Figure 14, consistent with the DPD model, total bead density (solvent included) was
centered on 3 beads/nm3
.17
The density distribution for the bilayer beads was largely consistent with
previous studies with expected deviation arising from the 1:3 ratio of receptor-simulating double-
chain lipids and DPPC lipids.1, 22
The paired receptor beads protruded approximately 0.4 nm further
into the water than the two topmost combined headbeads on the DPPC lipids (refer to Figures 10,11).
Due to computational constraints it was not possible to revise the receptor-water interactions, lipid-
receptor split and lipid density to be consistent with literature (refer to Suggestions for Further Work).
19
5.2 Comparison with Yan and Yu’s Results
To further test the stability of the modified DPD model, attempts were made to replicate results from
Yan and Yu’s previous studies.1, 2
The affinity of all lipid headgroup beads for the external dendrimer
beads and the internal dendrimer beads were set to αRL=15 and αTU=28 respectively, consistent with
Yan and Yu’s soft-core interaction parameters (refer to Appendix 1: Table 6).
After 105
timesteps (approximately 15 μs), a contrasting structure was generated:
Figure 15: Comparison of dendrimer penetration after 15 μs. Note that the counterions are not shown in both figures;
however, their presence is incorporated into both models consistently.
Due to the presence of the receptor-simulating lipid in the bilayer structure, this simulation is subtly
different to Yan and Yu’s previous work. Figure 15 depicts a bilayer structure which is less planar, with
lipids capable of migrating onto the surface of the dendrimer; this was unexpected. The reasons for
these discrepancies require further investigation. However, it is worth noting that the bending of the
bilayer reduced as the simulation progressed and the migration of lipids onto the dendrimer was
observed by Yan and Yu in another simulation which used the same soft-repulsion parameters.2
This
finding potentially highlights the effect of DPD’s random force on simulations. The simulated
dendrimer was also slightly more spherical, with the asphericity calculated to be As=0.033 compared
with Yan and Yu’s figure centred on As=0.055 (refer to Appendix 3); the magnitude of this variation
was deemed not to be significant given the dynamic nature of the system. Importantly, the structure
of the simulated dendrimer was considerably more open in the modified system with the radius of
gyration equal to Rg=3.8 nm compared with Rg=2.7 nm (refer to Appendix 4). The magnitude of this
deviance was unexpected. The most likely causes of this deviation are unintentional changes from
historical soft-core repulsion parameters that govern the simulation environment of the dendrimer
equilibrating in water prior to being introduced into the simulation box with the lipid bilayer.
Table 1: Soft-core repulsion parameters governing isolated dendrimer-water system.
Solvent Bead (0) Surface
Dendrimer Bead
(+1)
Internal
Dendrimer Bead
(0)
Solvent Bead (0) 25 25 60
Surface Dendrimer Bead (+1) 25 25 30
Internal Dendrimer Bead (0) 60 30 25
(Yan & Yu 2009)
20
Table 2: Historic soft-core repulsion parameters used in the combined simulation.1, 2
Past simulation of the G5 PAMAM dendrimers in water using the parameters defined in Table 2 yielded
compact spherical structures with diameters consistent with empirical measurements.2, 42
Figure 16: Equilibrium configuration of a G5 PAMAM dendrimer in water 2.
The effect of the soft-core repulsion parameters changes, especially the reduction of the
hydrophobicity of internal dendrimer beads (from 80 to 60), caused the dendrimer to equilibrate to
a size approximately 40% larger than desired.
Figure 17: Dendrimer introduced into the combined simulations
The simulation codes which modelled the dendrimer isolated from the bilayer were not made
available and the artifacts were only detected after returning to Australia and hence could not be
rectified. In the combined simulation environment the parameters were restored to their original
values; however, the dendrimer did not have sufficient time to structurally equilibrate before
interacting with the bilayer in its more open form (refer to Suggestions for Further Work).
Minor structural variation was likely further compounded by other sources. As previously mentioned,
due to the inclusion of random forces in the DPD model an exact morphological match was not
expected. Furthermore, the incorporated hydrophilicity of the receptor beads caused the bilayer
surface to be more open. The internal dendrimer beads (yellow) were modelled as hydrophobic. In
addition the parameters defining the interaction between the hydrophobic lipid tails (cyan) and both
types of dendrimer beads (red, yellow) were less repulsive than the parameters defining the
interactions between the hydrophobic lipid tails (cyan) and the lipid headgroup beads (green, blue,
orange, purple). As a result penetrability of the dendrimer increased from fp=0.02 to fp=0.1109 when
interacting with the new bilayer with the more open surface (refer to Appendix 5).
Solvent Bead (0) Surface
Dendrimer Bead
(+1)
Internal
Dendrimer Bead
(0)
Solvent Bead (0) 25 20 80
Surface Dendrimer Bead (+1) 20 25 28
Internal Dendrimer Bead (0) 80 28 25
21
5.3 Simulation of the System with Dendrimer Heads modelled as Ligands
Despite the non-steady-state morphological discrepancies and discussed deficiencies of the model, it
was tested whether the generation of structures resembling those expected if clathrin-mediated
endocytosis were responsible for transmembrane transport was possible. Further refinement of soft-
core repulsion parameters was required. It was assumed that the surface amine functional groups of
the dendrimer could substitute for ligand complexes (refer to Further Discussion). The effect of this
assumption was tested by holding constant all soft-core repulsion parameters for the surface
dendrimer beads excluding the parameter governing the surface dendrimer bead-receptor bead
interaction (αRL). This parameter was, in many instances, modelled to be attractive in order to simulate
the strong binding relationship between ligands and receptors.11
5.3.1 Initial Attempts to Generate Clathrin-mediated Endocytic Morphology
The following structures were achieved after 105
timesteps (approximately 15 μs). The surface
dendrimer bead soft-core interaction parameter was varied from a repulsive αRL=15 through to an
attractive αRL=-30 (refer to Appendix 1: Table 7):
Figure 18: Penetration profiles of simulations. For the purposes of clear illustration of the counterions and part of the
bilayer structures have been omitted.
In these simulations, encapsulation of the dendrimer was not observed. The asphericity and radii of
gyration were larger than would be expected if encapsulation were responsible for penetration.
Figure 19: Radii of gyration of simulated dendrimer-
bilayer systems after approximately 105 timesteps.
Figure 20: Asphericity of simulated dendrimer-bilayer
systems after approximately 105 timesteps.
αRL=15 αRL=0 αRL=-15
3.6
3.7
3.8
3.9
4
4.1
-20 -10 0 10 20
Radiusof
GyrationRg(nm)
Soft-core Repulsion (αRL)
0
0.01
0.02
0.03
0.04
0.05
0.06
-20 -10 0 10 20
Asphericity(As)
Soft-core Repulsion (αRL)
22
Further data is required if the observed trends of increasing radii of gyration and asphericity with the
decrease in the soft-core interaction parameter αRL are to be substantiated. Due to computing
constraints, an inadequate number of different values of αRL were tested. Furthermore, sampling at
t=100,000 was not performed by calculating statistics from multiple simulation runs with the same αRL
value. Only one simulation per tested value of αRL was able to be run and statistics were calculated
from neighboring timesteps (98,000; 99,000; 100,000; 101,000; 102,000). As such, error bars of one
standard deviation from the mean are depicted by these graphs. Repeated simulations would allow
for the calculation of standard errors and greater confidence when interpreting results (refer to
Suggestions for Further Work). Such trends would be expected in these non-steady-state snapshots
as a reduction in αRL should allow the dendrimer structure to migrate through the bilayer more quickly.
This is because the dendrimer heads are more strongly attracted to receptor beads on both sides of
the membrane, stretching the dendrimer across the internal bilayer structure as encapsulation was
not observed. Furthermore, the increasing attraction between the receptor heads and outer
dendrimer beads led to a wider spread of dendrimer on the bilayer surfaces consistent with similar
simulations without the receptor-simulating lipid inclusion.2
These expectations were further
supported by penetration data. Penetration similarly appeared to increase with with a decrease in the
soft-core interaction parameter αRL:
Figure 21: Penetration of the dendrimer into the bilayer after approximately 105 timesteps.
Again, due to variability and incompleteness of data, further simulation is required in order to test
whether this observed and expected trend is statistically significant. This statistical insignificance of
the data is further demonstrated by Figure 22 using confidence bands calculated to bound 95% of
least-squares linear regressions that could be expected to be generated given the sampling
methodology (refer to Appendix 6). Note that a linearly-increasing relationship could be plotted
between αRL=-15 and αRL=15 within these confidence bands and that linearity is not expected
necessarily.
0
0.05
0.1
0.15
0.2
0.25
-20 -10 0 10 20
Penetration(fp)
Soft-core Repulsion (αRL)
23
Figure 22: Penetration data bounded by linear regression confidence bands.
In all simulations the dendrimers were able to penetrate the bilayer to varying degrees. In these
simulations, the penetration was possible due to the hydrophobicity of the internal dendrimer beads
(yellow) combined with the observation that the parameters defining the interaction between the
hydrophobic lipid tails (cyan) and both types of dendrimer beads (red, yellow) (α=28) were less
repulsive than the parameters defining the interactions between the hydrophobic lipid tails (cyan) and
the lipid headgroup beads (green, blue, orange, purple) (α=40).
As a result, the dendrimer did not simply adhere to the top surface of the bilayer. Instead the internal
dendrimer beads bridged the structure as they were only weakly repelled by the internal hydrophobic
section of the bilayer. The dendrimer head beads aggregated near the receptor beads on both sides
of the bilayer due to the introduced receptor-ligand interaction. The structures generated were
consistent with Yan and Yu’s findings that attraction between lipid tails and inner-dendrimer beads
has a large effect on penetrability.2
When the magnitude of the attractive force was too great (αRL=-30), the self-assembling bilayer
structure deteriorated (refer to Further Discussion).
Figure 23: Deterioration of bilayer structure with αRL=-30
With αRL<-25, the attractive force between receptors and dendrimer heads was greater than the
repulsive forces keeping like-beads apart from one another. This had the effect of creating physically
unrealistic and unstable regions of high bead density. The chaos of the system was further
24
compounded by these unstable regions crossing the boundaries of the simulation environment and
transporting to the other side of the box. Subsequent simulations used values of αRL equal to and
greater than -25.
5.3.2 Further Attempts to Generate Clathrin-mediated Endocytic Morphology
In accordance with Yan and Yu’s previous findings, to inhibit the observed bridging the repulsive
parameter governing the interaction between the internal dendrimer beads and the lipid tails was
increased from αTU=28 to αTU=40 (refer to Appendix 1: Table 8).2
Various values of αRL were trialled
from a repulsive αRL=15 through to an attractive αRL=-25. After longer simulations, 245
timesteps
(approximately 36 μs), the following morphologies were observed:
Figure 24: Penetration profiles of simulations. For the purposes of clear illustration of the counterions and part of the
bilayer structures have been omitted.
Very different structures were observed after this increase in αTU. For αRL=15 the repulsion was so
great that the dendrimer did not even adhere to the surface of the bilayer. Hence, its profile is not
shown. However, in the absence of direct interaction with the bilayer, the radius of gyration of the
dendrimer was reduced significantly to 2.85 nm. Some axial stretching (As≈ 0.06) was caused by the
migration of counterions which accumulated in the volume separating the dendrimer and bilayer.
Nevertheless, this observation of significant shrinkage was welcome due to better agreement with
empirical measurements and previously discussed concerns (refer to Figure 17). For αRL=0, the
dendrimer adhered to the surface of the bilayer. Although not made clear by the cross-section
exhibited in Figure 24, random forces permitted the jumping of some surface dendrimer beads, with
internal dendrimer beads attached, across the bilayer as was even more strongly observed for the
case where the receptor-ligand interaction was modelled as attractive with αRL=-15. Note for αRL=-15
this attractive force was large enough to cause lipids to migrate onto the dendrimer. For αRL=-25 this
migration was pronounced. Additionally, the strength of the ligand-receptor attraction was much
greater relative to the random forces and dendrimer beads could no longer jump across the
membrane. Consequently, a morphology resembling endocytosis evolved. It was also observed that
the ligand-simulating dendrimer beads became more locally concentrated as αRL decreased (refer to
Further Discussion).
αRL=0 αRL=-15 αRL=-25
25
Figure 25: Radii of gyration of simulated dendrimer-bilayer systems after approximately 245 timesteps.
As previously discussed, more simulations are required to make better supported assertions. The
gyration radius was smallest for the dendrimer that did not bind to the bilayer (αRL=15). It increased
with decreasing αRL as per the previous results generated (refer to Section 5.3.1) until it decreased
with the onset of encapsulation for αRL=-25 (refer to Figure 25).
Figure 26: Penetration of the dendrimer into the bilayer after approximately 245 timesteps.
Penetrability statistics were consistent with the profiles generated as well as the negative correlation
described in the analysis of earlier simulations (refer to Figure 26). Penetration was achieved by the
lipids moving to encapsulate the dendrimer for αRL=-25 by direct penetration for αRL=0 and likely by a
combination of both mechanisms for αRL=-15.
Figure 27: Asphericity of simulated dendrimer-bilayer systems after approximately 245 timesteps.
2
2.5
3
3.5
4
-30 -20 -10 0 10 20RadiusofGyration
Rg(nm) Soft-core Repulsion (αRL)
0
0.05
0.1
0.15
0.2
0.25
0.3
-30 -20 -10 0 10 20
Penetration(fp)
Soft-core Repulsion (αRL)
0
0.02
0.04
0.06
0.08
0.1
0.12
0.14
-30 -20 -10 0 10 20
Asphericity(As)
Soft-core Repulsion (αRL)
26
As discussed, the dendrimer that did not bind to the bilayer was axially stretched by the counterions.
For αRL=0 most of the dendrimer was bound to the surface of the bilayer in a comparatively spherical
shape. The asphericity increased for αRL=-15 as the dendrimer structure opened. Surprisingly, the
structure most closely resembling encapsulation developed the most aspherical form. The large
asphericity of the structure generated for αRL=-25 in Figure 24 can be explained by its morphological
development depicted by Figure 27:
Figure 28: Morphological progression for αRL=-25.
Due to the strong attraction between the receptors and the ligand-simulating dendrimer beads the
dendrimer initially flattened out over the surface of the bilayer. Subsequently, receptors migrated on
top of the dendrimer. The flat shape of the dendrimer was maintained and after approximately 27 µs
the dendrimer broke through to the other side of the bilayer due to a combination of random forces,
receptor-ligand attraction and increased pressure exerted by the growing number of repulsive lipid
tails on top of the dendrimer. Subsequently, more receptor beads migrated to cover this breach and
structures resembling encapsulation were observed. It is believed that if the simulation were able to
continue a steady-state would be reached with a more spherical dendrimer structure evolving and an
approximately equal number of dendrimer beads concentrated on either side of the midline of the
bilayer (fp≈0.5) (refer to Suggestions for Further Work). However, continued simulation was not
possible due to computing constraints. The relationship between these observed morphologies to
CME and other simulations will be discussed in Section 6.
0µs 9µs 18µs
27µs 36µs 45µs
27
A phase table describing the morphologies for different soft-core repulsion parameters was
generated:
Table 3: Phase table describing different morphologies as a result of changing soft-core repulsion parameters.
αRL
αTU
-25 -15 0 15
28 - Strong Direct
Penetration
Direct Penetration Direct Penetration
40 Endocytosis Direct Penetration
With Pronounced
Receptor
Migration
Adhesion with
Direct Penetration
No Adhesion
Scope exists to further develop this table in conjunction with improving an understanding for how
physical properties relate to the soft-repulsion parameters (refer to Suggestions Further Work).
28
6.0 Further Discussion
6.1 Known Shortcomings of the Model
Given the simplicity of the simulated membrane a true representation of the clathrin-mediated
endocytosis was unlikely to have been simulated.
Figure 29: Schematic illustrating more representative membrane complexity.43
This experiment used a tensionless membrane. In reality it has been shown that surface tension varies
and that locally, negative surface tension can be generated by the cytoskeleton, dynamin or the actin
filaments.39
Consequently, internalisation via invagination and cytotoxicity associated with hole
generation (refer to Figure 30) were unlikely to be well-modelled by these simulations.
Figure 30: Empirically observed hole inducing effects of cationic PAMAM dendrimers on tense lipid bilayer using atomic
force microscopy; 44 see also 1, 36-38, 45, 46.
Other simulations have regulated the tension of the bilayer. In order to maintain a constant surface
tension, frequent transformations (refer to Equation 19) to the dimensions of the simulation box can
be performed whilst keeping the simulation volume constant.2
(19)
29
Alternatively, the N-varied DPD method maintains a constant membrane tension by adding and
removing water and lipids from the edges of the simulation environment as required.39
Additionally, the receptor beads were too simple to form complex clathrin cages to encapsulate the
dendrimer:
Figure 31: A clathrin cage with a single triskelion (composed of three interlocking clathrin proteins) highlighted.47
In biological fluids, nanoparticles may interact with other circulated proteins and not with cellular
membranes. Hence, simulating in a pure solvent environment may not be representative of reality
even on such small length scales.7
The 1:3 ratio of lipid types was chosen hastily due to time pressures.
Other studies have incorporated 1:1 ratio in accordance with current knowledge of mammalian
membranes (refer to Suggestions for Further Work).13, 48
Also, the uniform distribution of receptors
and ligand-simulating dendrimer beads on the surfaces of the membrane and dendrimer was not
representative of a real system. It was assumed that the surface amine functional groups could
substitute for ligands yielding this uniform distribution. Chemical synthesis of dendrimers is not able
to achieve such uniformity. Synthesis of ligand coated PAMAM results in a distribution of the number
of attached ligands. The number of attached ligands is much fewer in number than the number of
surface amine groups.49
Simulating the effects of ligand distributions on cellular uptake has been
studied but only using polygons as interacting particles (refer to Suggestions for Further Work).12
Consistent with Yan and Yu’s modelling, a number of counterions were included in the simulation and
behaved according to the electrostatics detailed in Section 2.1
However, this implementation lacks
refinement, further compounding potential sources of inaccuracy. Similar concerns stemming from
the reuse of other parameters such as spring constants governing restorative forces exist. Finally, the
energies associated with receptor-ligand adhesion are largely unknown.50
αRL is related to these
unspecified energies. In order to observe endocytosis, αRL had to be made attractive enough to induce
locally dense regions populated by receptor and ligand head. This modification did not disturb the
thermostat; however, this clumping of beads may signify further deviation of the model from reality.
Yang and Ma were incorporated a classical Lennard–Jones potential to govern the receptor-ligand
interaction (refer to Figure 7) in order to prevent this issue.29
30
6.2 Relationship of CME to Simulations
Using this receptor model, encapsulation of the dendrimer was able to be simulated following changes
to the soft-core repulsion parameters. This encapsulation occurred via morphologies varying from
descriptions of CME and the hypothesis was consequently rejected. In the absence of invaginations
and actin-induced regions of negative surface tension, internalisation of the vesicle was not observed
via the ‘coated pit’ morphology depicted in Figure 5.51
Encapsulation was primarily observed to occur
by the migration of receptors to the adhesive front.50
Consequently, encapsulation appeared to occur
by a mechanism more closely resembling phagocytosis (refer to Figure 4). This mechanism is
associated with larger length scales than that which was simulated.25
Further investigation is required
to determine whether this model can be used for the purpose of simulating phagocytosis or revised
to produce simulations that mimic CME morphological progression.
Other simulations have been able to achieve similar structures resembling endocytosis. Macroscale
physics have been incorporated into these mesoscale models in the form of Canham-Helfrich theory.
Canham-Helfrich theory introduces membrane bending energies so as to limit direct penetration.39
It
was possible to achieve this outcome without the need for this added complexity by balancing soft
repulsion parameters.
6.3 The Notable Shortage of Empirical Data
Experimentalists have not been able to accurately observe many proposed mechanisms for
endocytosis due to technological limitations at these lengthscales and timescales (refer to Figure
32).52, 53
There is currently a pressing need to be able to effectively differentiate between pathways
and perform quantitative analysis of transport mechanisms.54
Figure 32:Currently there is little overlap between what can experimentally be observed and what can be simulated
when studying endocytosis.25
As mentioned previously, in order for the new model to be of value, it must reproduce solubilities,
liquid compressibilities and the shape of interfaces on the length scale of the course-grained
molecules.4
Such comparisons are currently limited by a deficiency in empirical data surrounding
endocytic processes. As a consequence, simulations do little more than attempt to generate
hypothesised morphological progression. Many attempts to incorporate the modest available data
31
into the DPD model have been made or discussed (e.g. isolated dendrimer size, lipid diffusion, hole
generation, hole size and lipid density). Due to the complexity of the system, many of the soft-core
repulsion parameters use in this simulation were estimated by qualitative arguments coupled with
iterative refinement or quoted from other studies which use similar methods.1, 17, 34
Some interaction
parameters are able to be generated from Flory-Huggins theory and empirical testing.19
Calculating a
complete set of Flory-Huggins 𝜒 parameters for complex systems, including those with multiple
nanoscale species with Janus properties would be extremely difficult given the number of interaction
pairs. Furthermore, energies associated with receptor-ligand binding, encapsulated gyration radii,
internalisation velocities are all currently unknown. If the key issues concerning this model are
addressed by reformulation, its future acceptance or rejection should be determined based on
agreement with available endocytic data and practical value however blind (refer to Suggestions for
Further Work).
32
7.0 Conclusions
Incorporation of a receptor-simulating lipid into the bilayer and subsequent tuning of soft-core
repulsion parameters generated endocytic morphology. It was found that self-assembly of the bilayer
structure was not compromised when converting a quarter of the simulated lipid molecules into a
receptor-simulating lipid. These receptor-simulating lipids extended on average 0.4 nm further into
the solvent than the DPPC simulating lipids. The lipid density of the bilayer increased by approximately
13% when compared with Yan and Yu’s previous simulations.2
The stability of the DPD model, on
comparing with Yan and Yu’s previous simulations, was found to be compromised by changes made
to the interaction parameters governing the generation of the G5 PAMAM dendrimer in solvent,
causing it to be 40% more open than empirically observed (Rg=3.4 nm > Rg=2.4 nm). These changes
were partially rectified for the simulations by reverting the interaction parameters to their historic
values; however, these changes could not impose the required structural changes before interaction
with the bilayer, increasing penetration.
Despite these issues, (resulting in enhanced direct penetration,) further refinement of the soft-core
repulsion parameters revealed that endocytosis could be modelled. Statistical analysis of changes to
the soft-core interaction parameter governing the behaviour of proximal dendrimer surface beads and
receptor beads suggested that penetrability could be increased if this repulsive force was instead
modelled as attractive. The magnitude of this attraction was limited such that for αRL>-25 the bilayer
structure became unstable due to the resultant creation of hyper-dense regions within the simulation
box. The observed structures of dendrimers generated by these changes were open and distributed
throughout the structure of the bilayer, again suggesting entry by direct penetration. Encapsulation
of the dendrimer was promoted after subsequently increasing the repulsive force between the
internal dendrimer beads and the lipid tails of the bilayer. From these results it was concluded that
the binding energy between receptors and ligands as well as the hydrophobic interaction between the
bilayer lipid tails and internal structure of the interacting particle has a large effect on binding and the
type of penetration.
Morphological changes in the simulation in which a form of endocytosis was observed may provide
some insight as to how such encapsulation progresses. However, due to the discussed limitations of
the model the morphology more closely matched a phagocytic mechanism. On the simulated length
scale such a result was unexpected indicating that further refinement of the model may be needed
(refer to Suggestions for Further Work) if CME, as it is currently understood, is to be simulated and
the presently unobservable dynamics conjectured. As discussed, due to computational constraints,
many of the results of the simulations run in this study were largely preliminary.
Continued effort in this area of research will have future implications for drug development. When
models can be used with confidence, molecular architectures will be able to be tested without costly
synthesis. In order to generate robust models and build this confidence, continued interest in
formulation and modification is required to better describe the intricacies of the very complex yet
fundamental process of endocytosis.
33
8.0 Suggestions for Further Work
Future experiments require access to appropriate computing resources similar to the cluster systems
housed at VLSCI.
8.1 Simple Immediate Revisions to the Current DPD Model
As uncovered by the results generated, revisions to enhance the representativeness of the DPD model
need to be made:
 The assembly of the bilayer structure needs to resimulated with the ‘target lipid density’ set
to 0.64 nm2
.
 The prevalence of receptor-simulating lipids needs to be increased to approximately 50% in
accordance with current knowledge of mammalian membranes.13, 48
 The protrusion of these receptor lipids into the solvent may also need to be revised in light of
information from empirical studies of bilayer structure.
 Reversion to the original parameters governing the dendrimer-water system need to be
readopted and simulated so that on introducing the dendrimer to the bilayer simulation box
the structure of the dendrimer agrees with empirical measurements.
 The introduction of an additional type of ligand bead distributed on the surface of the
dendrimer in accordance with Mullen and Banaszak Holl’s findings needs to be incorporated
to remove the erroneous assumption that surface amine functional groups can substitute for
ligand complexes.49
 To prevent the dense clumping of ligand and receptor beads, in accordance with Yang and
Ma, this attractive pairwise interaction should be governed by a classical Lennard–Jones
potential.29
 The diffusion constant for membrane receptors in a bilayer is estimated to be approximately
104
nm2
/s.50
If accurate tracking of these receptors can be performed, αRL can be tuned to
achieve expected diffusion.
 The concentration of counterions and spring coefficients may require revision following
examination of relevant literature.
8.2 Extension of Results
Subsequent to the introduction of these changes, extended simulation is required to see whether CME
can be observed. Multiple simulations running with the same set of soft-core interaction parameters
are required so that sampling techniques and data interpretation can be strengthened. Simulations
should ideally be run until steady-states have been reached so as to gain more a complete
understanding of morphological progression. From these results comprehensive phase diagrams could
be generated.
It may be that CME still cannot be simulated following these simple revisions. A new model
incorporating a more complex bilayer structure may be required. The distribution, size and structure
of the receptors may need to be revised so that they are able to form clathrin cage-like configurations
(refer to Figure 31). Similarly, strategies to control the surface tension of the bilayer on macro and
local scales may need to be incorporated to simulate the roles of actin and dynamin in promoting
invagination and vesicle release (refer to Figure 4).
34
If CME is not observed following the simple revisions this model could still be extended to simulate a
larger interacting particle instead of G5 PAMAM. It is known that phagocytosis occurs on larger length
scales. Subsequent to further evaluation as to whether the endocytic morphology generated by this
study resembles phagocytosis, it would be interesting to observe whether similar encapsulation could
be induced without the need to incorporate energies associated with membrane deformation into the
model, as is often done when simulating on larger length scales.
35
9.0 References
1. Yan, L.-T. and X. Yu, Enhanced Permeability of Charged Dendrimers across Tense Lipid Bilayer
Membranes. ACS Nano, 2009. 3(8): p. 2171-2176.
2. Yan, L.-T. and X. Yu, Charged Dendrimers on Lipid Bilayer Membranes: Insight through
Dissipative Particle Dynamics Simulations. Macromolecules, 2009. 42(16): p. 6277-6283.
3. Min, Y., et al., The role of interparticle and external forces in nanoparticle assembly. Nat
Mater, 2008. 7(7): p. 527-538.
4. Groot, R.D. and T.J. Madden, Dynamic Simulation of Diblock Copolymer Microphase
Separation. The Journal of Chemical Physics, 1998. 108(20): p. 8713-8724.
5. Qian, H.-J., et al., Computer Simulation of Cyclic Block Copolymer Microphase Separation.
Macromolecules, 2005. 38(4): p. 1395-1401.
6. Beddoes, C.M., C.P. Case, and W.H. Briscoe, Understanding Nanoparticle Cellular Entry: A
Physicochemical Perspective. Advances in Colloid and Interface Science, 2015. 218(0): p. 48-
68.
7. Tian, F., Computer Simulation Studies on the Interactions Between Nanoparticles and Cell
Membrane. Science China Chemistry, 2014. 57(12): p. 1662.
8. Yan, L.T. and X. Yu, Complexes Comprised of a Dendrimer and a Vesicle: Role of Vesicle Size
and the Surface Tension of the Vesicle Membrane. Nanoscale, 2011. 3(9): p. 3812-8.
9. Abraham, J., Dissipative Particle Dynamics, in Encyclopedia of Microfluidics and Nanofluidics,
D. Li, Editor. 2008, Springer US. p. 385-391.
10. Lee, H. and R. Larson, Multiscale Modeling of Dendrimers and Their Interactions with Bilayers
and Polyelectrolytes. Molecules, 2009. 14(1): p. 423.
11. Li, Y., et al., Molecular Modeling of the Relationship between Nanoparticle Shape Anisotropy
and Endocytosis Kinetics. Biomaterials, 2012. 33: p. 4965-4973.
12. Schubertova, V., F.J. Martinez-Veracoechea, and R. Vacha, Influence of Ligand Distribution on
Uptake Efficiency. Soft Matter, 2015. 11(14): p. 2726-2730.
13. Vácha, R., F.J. Martinez-Veracoechea, and D. Frenkel, Intracellular Release of Endocytosed
Nanoparticles Upon a Change of Ligand–Receptor Interaction. ACS Nano, 2012. 6(12): p.
10598-10605.
14. Kitchens, K.M., et al., Endocytosis and Interaction of Poly(amidoamine) Dendrimers with
Caco-2 Cells. Pharmaceutical Research, 2007. 24(11): p. 2138-2145.
15. Seib, F.P., A.T. Jones, and R. Duncan, Comparison of the Endocytic Properties of Linear and
Branched PEIs, and Cationic PAMAM Dendrimers in B16f10 Melanoma Cells. Journal of
Controlled Release, 2007. 117(3): p. 291-300.
16. Hoogerbrugge, P.J. and J.M.V.A. Koelman, Simulating Microscopic Hydrodynamic
Phenomena with Dissipative Particle Dynamics. EPL (Europhysics Letters), 1992. 19(3): p.
155.
17. Groot, R.D. and K.L. Rabone, Mesoscopic Simulation of Cell Membrane Damage, Morphology
Change and Rupture by Nonionic Surfactants. Biophysical Journal, 2001. 81(2): p. 725-736.
18. Groot, R.D., Electrostatic Interactions in Dissipative Particle Dynamics—Simulation of
Polyelectrolytes and Anionic Surfactants. The Journal of Chemical Physics, 2003. 118(24): p.
11265-11277.
19. Groot, R.D. and P.B. Warren, Dissipative Particle Dynamics: Bridging the Gap between
Atomistic and Mesoscopic Simulation. The Journal of Chemical Physics, 1997. 107(11): p.
4423-4435.
20. Mintzer, M.A. and M.W. Grinstaff, Biomedical Applications of Dendrimers: A Tutorial.
Chemical Society Reviews, 2011. 40(1): p. 173-190.
36
21. McNerny, D.Q., P.R. Leroueil, and J.R. Baker, Understanding Specific and Nonspecific
Toxicities: A Requirement for the Development of Dendrimer-Based Pharmaceuticals. Wiley
Interdisciplinary Reviews: Nanomedicine and Nanobiotechnology, 2010. 2(3): p. 249-259.
22. Shillcock, J.C. and R. Lipowsky, Equilibrium Structure and Lateral Stress Distribution of
Amphiphilic Bilayers from Dissipative Particle Dynamics Simulations. Journal of Chemical
Physics, 2002. 117(10): p. 5048.
23. Conner, S.D. and S.L. Schmid, Regulated Portals of Entry into the Cell. Nature, 2003.
422(6927): p. 37.
24. Canton, I. and G. Battaglia, Endocytosis at the Nanoscale. Chemical Society Reviews, 2012.
41(7): p. 2718-2739.
25. Qu, Z.G., et al., Advances in the Understanding of Nanomaterial-Biomembrane Interactions
and their Mathematical and Numerical Modeling. Nanomedicine, 2013. 8(6): p. 995-1011.
26. Roth, T.F. and K.R. Porter, Yolk Protein Uptake in the Oocyte of the Mosquito Aedes Aegyti. L.
The Journal Of Cell Biology, 1964. 20: p. 313-332.
27. Nel, A.E., et al., Understanding Biophysicochemical Interactions at the Nano–Bio Interface.
Nature Materials, 2009. 8(7): p. 543-557.
28. Yue, T. and X. Zhang, Molecular Understanding of Receptor-Mediated Membrane Responses
to Ligand-Coated Nanoparticles. Soft Matter, 2011. 7(19).
29. Yang, K. and Y.-q. Ma, Wrapping and Internalization of Nanoparticles by Lipid Bilayers: a
Computer Simulation Study. Australian Journal of Chemistry, 2011. 64(7): p. 894.
30. Groot, R.D. and T.J. Madden, On the Role of Hydrodynamic Interactions in Block Copolymer
Microphase Separation. Journal of Chemical Physics, 1999. 110(19): p. 9739.
31. Keaveny, E.E., et al., A Comparative Study between Dissipative Particle Dynamics and
Molecular Dynamics for Simple- and Complex-Geometry Flows. The Journal of Chemical
Physics, 2005. 123(10): p. 104107.
32. Moeendarbary, E., T.Y. Ng, and M. Zangeneh, Dissipative Particle Dynamics: Introduction,
Methodology and Complex Fluid Applications - A Review. International Journal of Applied
Mechanics, 2009. 01(04): p. 737-763.
33. Ingólfsson, H.I., et al., The Power of Coarse Graining in Biomolecular Simulations. WIREs:
Computational Molecular Science, 2014. 4(3): p. 225.
34. Groot, R.D., Mesoscopic Simulation of Polymer−Surfactant Aggregation. Langmuir, 2000.
16(19): p. 7493.
35. Rodgers, J.M., et al., Understanding the Phase Behavior of Coarse-Grained Model Lipid
Bilayers through Computational Calorimetry. The Journal of Physical Chemistry B, 2012.
116(5): p. 1551-1569.
36. Xie, L.-q., W.-d. Tian, and Y.-q. Ma, Computer Simulations of the Interactions of High-
Generation Polyamidoamine Dendrimers with Electronegative Membranes. Soft Matter,
2013. 9(39): p. 9319-9325.
37. Kelly, C.V., et al., Poly(amidoamine) Dendrimers on Lipid Bilayers I: Free Energy and
Conformation of Binding. The Journal of Physical Chemistry B, 2008. 112(31): p. 9337-9345.
38. Kelly, C.V., et al., Poly(amidoamine) Dendrimers on Lipid Bilayers II: Effects of Bilayer Phase
and Dendrimer Termination. The Journal of Physical Chemistry B, 2008. 112(31): p. 9346-
9353.
39. Guo, R., J. Mao, and L.-T. Yan, Unique Dynamical Approach of Fully Wrapping Dendrimer-like
Soft Nanoparticles by Lipid Bilayer Membrane. ACS Nano, 2013. 7(12): p. 10646-10653.
40. Deserno, M., Elastic Deformation of a Fluid Membrane upon Colloid Binding. Physical Review
E, 2004. 69(3): p. 031903.
41. Ding, H.-m., W.-d. Tian, and Y.-q. Ma, Designing Nanoparticle Translocation through
Membranes by Computer Simulations. ACS Nano, 2012. 6(2): p. 1230-1238.
42. Choi, Y., et al., DNA-Directed Synthesis of Generation 7 and 5 PAMAM Dendrimer
Nanoclusters. Nano Letters, 2004. 4(3): p. 391-397.
37
43. Venturoli, M., et al., Mesoscopic Models of Biological Membranes. Physics Reports, 2006.
437(1–2): p. 1-54.
44. Leroueil, P.R., et al., Wide varieties of cationic nanoparticles induce defects in supported lipid
bilayers. Nano Letters, 2008. 8(2): p. 420-424.
45. Perumal, O.P., et al., The Effect of Surface Functionality on Cellular Trafficking of Dendrimers.
Biomaterials, 2008. 29(24-25): p. 3469-76.
46. Mecke, A., et al., Lipid Bilayer Disruption by Polycationic Polymers: The Roles of Size and
Chemical Functional Group. Langmuir: The ACS Journal Of Surfaces And Colloids, 2005.
21(23): p. 10348-10354.
47. Goodsell, D. and G. Johnson, Molecule of the Month: Clathrin. RCSB PDB, 2007.
48. Alberts, B., et al., Molecular Biology of the Cell. 2015, New York: N.Y. Garland Science, Taylor
and Francis Group.
49. Mullen, D.G. and M.M. Banaszak Holl, Heterogeneous Ligand–Nanoparticle Distributions: A
Major Obstacle to Scientific Understanding and Commercial Translation. Accounts of
Chemical Research, 2011. 44(11): p. 1135-1145.
50. Gao, H., W. Shi, and L.B. Freund, Mechanics of Receptor-mediated Endocytosis. Proceedings
of the National Academy of Sciences of the United States of America, 2005. 102(27): p. 9469-
9474.
51. Doherty, G.J. and H.T. McMahon, Mechanisms of Endocytosis. Annual Review of
Biochemistry, 2009. 78(1): p. 857-902.
52. Leroueil, P.R., et al., Nanoparticle Interaction with Biological Membranes: Does
Nanotechnology Present a Janus Face? Accounts of Chemical Research, 2007. 40(5): p. 335-
342.
53. Wang, T., et al., Cellular Uptake of Nanoparticles by Membrane Penetration: A Study
Combining Confocal Microscopy with FTIR Spectroelectrochemistry. ACS Nano, 2012. 6(2): p.
1251-1259.
54. Akinc, A. and G. Battaglia, Exploiting Endocytosis for Nanomedicines. Cold Spring Harbor
Perspectives in Biology, 2013. 5(11).
38
10.0 Appendices
Appendix 1: Simulation Soft-core Repulsion Parameters
Table 4: Historic soft-core repulsion parameters.1
Lipid
Head
(+1)
Lipid
Head
(-1)
Lipid
Head
(0)
Lipid
Tail
(0)
Solvent
Bead
(0)
Surface
Dendrimer
Bead (+1)
Counterions
(-1)
Internal
Dendrimer
Bead (0)
Lipid Head (+1) 25 17 22 40 22 15 20 28
Lipid Head (-1) 17 25 22 40 22 15 20 28
Lipid Head (0) 22 22 25 40 25 15 22 28
Lipid Tail (0) 40 40 40 25 80 28 80 28
Solvent Bead (0) 22 22 25 80 25 20 20 80
Surface Dendrimer
Bead (+1)
15 15 15 28 20 25 20 28
Counterions (-1) 20 20 22 80 20 20 25 28
Internal Dendrimer
Bead (0)
28 28 28 28 80 28 28 25
Table 5: Soft-core repulsion parameters for bilayer self-assembly.
Lipid
Head
(+1)
Lipid
Head (-1)
Lipid
Head (0)
Lipid Tail
(0)
Solvent Bead
(0)
Receptor Head
(0)
Lipid Head (+1) 25 17 22 40 22 25
Lipid Head (-1) 17 25 22 40 22 25
Lipid Head (0) 22 22 25 40 25 25
Lipid Tail (0) 40 40 40 25 80 40
Solvent Bead (0) 22 22 25 80 25 15
Receptor Head (0) 25 25 25 40 15 25
39
Table 6: Soft-core repulsion parameters for interacting bilayer-dendrimer system. Starred parameters were introduced
not required by Yan and Yu’s receptorless system. These parameter were carefully chosen to model the receptor head
similarly to other lipid heads in an attempt to reproduce Yan and Yu’s results.1, 2
Lipid
Head
(+1)
Lipid
Head
(-1)
Lipid
Head
(0)
Lipid
Tail
(0)
Solvent
Bead
(0)
Receptor
Head (0)
Surface
Dendrimer
Bead (+1)
Counterions
(-1)
Internal
Dendrimer
Bead (0)
Lipid Head (+1) 25 17 22 40 22 25* 15 20 28
Lipid Head (-1) 17 25 22 40 22 25* 15 20 28
Lipid Head (0) 22 22 25 40 25 25* 15 22 28
Lipid Tail (0) 40 40 40 25 80 40* 28 80 28
Solvent Bead (0) 22 22 25 80 25 25* 15 20 80
Receptor Head (0) 25* 25* 25* 40* 25* 25* 15* 12* 28*
Surface Dendrimer
Bead (+1)
15 15 15 28 15 15* 25 20 28
Counterions (-1) 20 20 22 80 20 12* 20 25 28
Internal Dendrimer
Bead (0)
28 28 28 28 80 28* 28 28 25
Table 7: Soft-core repulsion parameters for interacting bilayer and dendrimer.
Lipid
Head
(+1)
Lipid
Head
(-1)
Lipid
Head
(0)
Lipid
Tail
(0)
Solvent
Bead
(0)
Receptor
Head (0)
Surface
Dendrimer
Bead (+1)
Counterions
(-1)
Internal
Dendrimer
Bead (0)
Lipid Head (+1) 25 17 22 40 22 25 15 20 28
Lipid Head (-1) 17 25 22 40 22 25 15 20 28
Lipid Head (0) 22 22 25 40 25 25 15 22 28
Lipid Tail (0) 40 40 40 25 80 40 28 80 28
Solvent Bead (0) 22 22 25 80 25 25 15 20 80
Receptor Head (0) 25 25 25 40 25 25 αRL 12 28
Surface Dendrimer
Bead (+1)
15 15 15 28 15 αRL 25 20 28
Counterions (-1) 20 20 22 80 20 12 20 25 28
Internal Dendrimer
Bead (0)
28 28 28 28 80 28 28 28 25
40
Table 8: Revised soft-core repulsion parameters for interacting bilayer and dendrimer.
Lipid
Head
(+1)
Lipid
Head
(-1)
Lipid
Head
(0)
Lipid
Tail
(0)
Solvent
Bead
(0)
Receptor
Head (0)
Surface
Dendrimer
Bead (+1)
Counterions
(-1)
Internal
Dendrimer
Bead (0)
Lipid Head (+1) 25 17 22 40 22 25 15 20 28
Lipid Head (-1) 17 25 22 40 22 25 15 20 28
Lipid Head (0) 22 22 25 40 25 25 15 22 28
Lipid Tail (0) 40 40 40 25 80 40 28 80 αTU
Solvent Bead (0) 22 22 25 80 25 25 15 20 80
Receptor Head (0) 25 25 25 40 25 25 αRL 12 28
Surface Dendrimer
Bead (+1)
15 15 15 28 15 αRL 25 20 28
Counterions (-1) 20 20 22 80 20 12 20 25 28
Internal Dendrimer
Bead (0)
28 28 28 αTU 80 28 28 28 25
41
Appendix 2: MATLAB Code for Bilayer Bead Density Profile
% Bilayer Bead Density by Displacement from Average z-coordinate
%create matrix with data from binary output file
[file,path]=uigetfile('SpaGe1.bin','Open the file');%to save N(r)_r data
if file
fid=fopen([path,file],'r');
a=fread(fid,[4 inf],'float32');
end
%calculate zavg
ztot=0;
for i=1:numel(a)/4;
z = a(3,i);
ztot = ztot+z;
end
zavg= ztot/(numel(a)/4);
%avg displacement from zavg
density= zeros(2,numel(a)/4);
for i=1:numel(a)/4;
density(1,i)=(a(3,i)-zavg);
density(2,i)=a(4,i);
end
%partition 0.05nm
x=linspace(-19.975, 20.025, 800);
for j=1:numel(a)/4
for i=1:numel(x)
if density(1,j)<= (x(i)+0.025) && density(1,j)>(x(i)-0.025)
density(1,j)=x(i);
end
end
end
%displacement histogram for relevant bilayer beads (types 1, 2, 3, 4, 6)
y=zeros(6,numel(x));
for i=1:6
for j=1:numel(a)/4
for m=1:numel(x)
if density(2,j)== i && density(1,j)== x(m)
y(i,m)=y(i,m)+1/88.2; %volume of xy-sliver 88.2nm
end
end
end
end
y=[y(1,:);y(2,:);y(3,:);y(4,:);y(6,:)];
42
y=y';
%histogram smoothing
smooth = sgolayfilt(y,3,29);
subplot(2,1,1)
plot(x, smooth,'linewidth', 2)
axis([-5 5 0 3.5])
title('Displacement Distribution of Bilayer Beads from Central z-
coordinate')
xlabel('Displacement from Bilayer Centre (nm)'),ylabel('Number Density
(beads/nm^3)');
legend('Lipid Head (+1)','Lipid Head (-1)','Lipid Head (0)','Lipid Tail
(0)','Receptor Head (0)')
subplot(2,1,2)
plot(x,smooth, 'linewidth', 2)
axis([0 5 0 1])
title('Magnified Displacement Distribution of Bilayer Beads from Central z-
coordinate')
xlabel('Displacement from Bilayer Centre (nm)'),ylabel('Number Density
(beads/nm^3)');
legend('Lipid Head (+1)','Lipid Head (-1)','Lipid Head (0)','Lipid Tail
(0)','Receptor Head (0)')
43
Appendix 3: Asphericity
Asphericity of the dendrimer can be calculated from the following equations:
(20)
Where I1 and I2 are the first two invariants:
(21)
(22)
𝜆1, 𝜆2, 𝜆3 are the eigenvalues of the gyration tensor:
𝑮 𝑻 =
𝟏
𝑵
∑ (𝒔𝒊⨂𝒔𝒊)𝑵
𝒊=𝟎 (23)
𝒔𝒊 = 𝒓𝒊 − 𝒓 𝑪𝑴 (24)
% Asphericity MATLAB
%create matrix with data from binary output file
[file,path]=uigetfile('SpaGe1.bin','Open the file');%to save N(r)_r data
if file
fid=fopen([path,file],'r');
a=fread(fid,[4 inf],'float32');
end
%remove counterions
b=a;
count=0;
for i=1:numel(a)/4;
if a(4,i)== 8
b(:, i-count) = [];
count=count+1;
else b(:,i-count)= a(:,i);
end
end
%calculate centre of mass (all beads have the same mass)
xtot=0;
ytot=0;
ztot=0;
for i=1:numel(b)/4;
x = b(1,i);
xtot = xtot+x;
end
xavg= xtot/(numel(b)/4);
for i=1:numel(b)/4;
y = b(2,i);
44
ytot = ytot+y;
end
yavg= ytot/(numel(b)/4);
for i=1:numel(b)/4;
z = b(3,i);
ztot = ztot+z;
end
zavg= ztot/(numel(b)/4);
rcm=[xavg;yavg;zavg]
%gyration tensor
dispoutertot=zeros(3);
disp = [0;0;0];
for i=1:numel(b)/4;
for j=1:3
disp(j,1)=b(j,i)-rcm(j,1);
end
dispouter=disp*disp';
dispoutertot=dispoutertot+dispouter;
end
Gt=dispoutertot/(numel(b)/4)
%calculate asphericity
eigvals=eig(Gt);
I1= sum(eigvals);
I2= eigvals(1)*eigvals(2)+eigvals(2)*eigvals(3)+eigvals(3)*eigvals(1);
As=1-(3*(I2)/(I1)^2)
45
Appendix 4: Radius of Gyration
The radius of gyration is determined by the following equation:
(25)
Where the summation runs over all N beads in the dendrimer at positions ri. rcm denotes the center
of mass of the dendrimer.
% Gyration Radius MATLAB
%create matrix with data from binary output file
[file,path]=uigetfile('SpaGe1.bin','Open the file');%to save N(r)_r data
if file
fid=fopen([path,file],'r');
a=fread(fid,[4 inf],'float32');
end
%remove counterions
b=a;
count=0;
for i=1:numel(a)/4;
if a(4,i)== 8
b(:, i-count) = [];
count=count+1;
else b(:,i-count)= a(:,i);
end
end
%calculate centre of mass (all beads have the same mass)
xtot=0;
ytot=0;
ztot=0;
for i=1:numel(b)/4;
x = b(1,i);
xtot = xtot+x;
end
xavg= xtot/(numel(b)/4);
for i=1:numel(b)/4;
y = b(2,i);
ytot = ytot+y;
end
yavg= ytot/(numel(b)/4);
for i=1:numel(b)/4;
z = b(3,i);
ztot = ztot+z;
end
zavg= ztot/(numel(b)/4);
rcm=[xavg;yavg;zavg]
46
%squared displacement from centre of gravity squared
dispsqutot=[0;0;0];
disp = [0;0;0];
for i=1:numel(b)/4;
for j=1:3
disp(j,1)=b(j,i)-rcm(j,1);
end
dispsqu = disp.^2;
dispsqutot=dispsqutot+dispsqu;
end
sum(dispsqutot)/((numel(b)/4));
Rg = sqrt(sum(dispsqutot)/((numel(b)/4)))
47
Appendix 5: Penetrability
Penetrability was defined as the fraction of dendrimers that were below the average z-coordinate of
all lipid beads enclosed by the maximum and minimum x,y coordinates of the dendrimer beads at
the fringes of the dendrimer structure.
% Penetrability MATLAB
%create matrix with data from binary output file
[file,path]=uigetfile('SpaGe1.bin','Open the file');%to save N(r)_r data
if file
fid=fopen([path,file],'r');
a=fread(fid,[4 inf],'float32');
end
[file,path]=uigetfile('SpaGe2.bin','Open the file');%to save N(r)_r data
if file
fid=fopen([path,file],'r');
c=fread(fid,[4 inf],'float32');
end
%remove counterions
b=c;
count=0;
for i=1:numel(c)/4;
if c(4,i)== 8
b(:, i-count) = [];
count=count+1;
else b(:,i-count)= c(:,i);
end
end
%box coordinates to determine relevant bilayer beads for penetrability
dimmax=max(b,[],2);
dimmin=min(b,[],2);
dimensions=[dimmin(1,1),dimmin(2,1); dimmax(1,1),dimmax(2,1)];
%remove irrelevant bilayer beads
d=a;
count=0;
for i=1:numel(a)/4;
if
a(1,i)<=dimensions(1,1)|a(2,i)<=dimensions(1,2)|a(1,i)>=dimensions(2,1)|a(2
,i)>=dimensions(2,2)
d(:, i-count) = [];
count=count+1;
else d(:,i-count)= a(:,i);
end
end
%calculate average z coordinate of remaining bilayer (all beads have the
same mass)
ztot=0;
for i=1:numel(d)/4;
48
z = (3,i);
ztot = ztot+z;
end
zavg= ztot/(numel(d)/4)
test=ones(1,numel(b)/4);
%test if dendrimer below middle of bilayer
for i=1:numel(b)/4
if b(3,i)<=zavg
test(1,i)=1;
else test(1,i)=0;
end
end
%fraction below middle of bilayer
fp=sum(test)/numel(test)
49
Appendix 6: Confidence Bands
Data file:
aRL fp
15 0.1109
15 0.1327
15 0.1426
15 0.1485
15 0.1089
0 0.1386
0 0.1762
0 0.1762
0 0.1802
0 0.2
-15 0.2693
-15 0.2178
-15 0.2495
-15 0.1881
-15 0.1861
R Code:
data = read.delim(file.choose(),header=T)
X=data$aRL
Y=data$fp
plot(X, Y, xlab="aRL", ylab="Penetrability fp", xlim=c(-20,
20), ylim=c(0, 0.4))
ml=lm(Y~X)
%abline(ml)
summary(ml)
confint(ml,level=0.95)
bands=predict(ml,newdata=data.frame(X),interval=c("prediction"
),level=0.95)
lines(X, bands[, "lwr"], lty = "dotted")
lines(X, bands[, "upr"], lty = "dotted")
50
Appendix 7: Intermediate Data
Table 9: Rg data for αTU=28
αRLTimestep 98000 99000 100000 101000 102000 μ σ
15 3.8E+00 3.7E+00 3.7E+00 3.7E+00 3.7E+00 3.7E+00 2.5E-02
0 3.9E+00 3.8E+00 3.9E+00 3.8E+00 3.9E+00 3.9E+00 5.2E-02
-15 4.0E+00 4.1E+00 4.0E+00 4.0E+00 4.0E+00 4.0E+00 3.1E-02
Table 10: As data for αTU=28
αRLTimestep 98000 99000 100000 101000 102000 μ σ
15 3.3E-02 1.9E-02 5.3E-03 1.1E-02 1.3E-02 1.6E-02 1.1E-02
0 3.8E-02 3.4E-02 3.2E-02 1.5E-02 2.7E-02 2.9E-02 9.1E-03
-15 2.4E-02 4.3E-02 3.1E-02 3.3E-02 5.6E-02 3.7E-02 1.3E-02
Table 11: fp data for αTU=28
αRLTimestep 98000 99000 100000 101000 102000 μ σ
15 1.1E-01 1.3E-01 1.4E-01 1.5E-01 1.1E-01 1.3E-01 1.8E-02
0 1.4E-01 1.8E-01 1.8E-01 1.8E-01 2.0E-01 1.7E-01 2.2E-02
-15 2.7E-01 2.2E-01 2.5E-01 1.9E-01 1.9E-01 2.2E-01 3.7E-02
Table 12: Rg data for αTU=40
αRLTimestep 238000 239000 240000 241000 242000 μ σ
0 3.2E+00 3.2E+00 3.2E+00 3.3E+00 3.2E+00 3.2E+00 4.2E-02
15 2.8E+00 2.8E+00 2.8E+00 2.9E+00 2.9E+00 2.9E+00 7.1E-02
-15 3.7E+00 3.7E+00 3.5E+00 3.6E+00 3.6E+00 3.6E+00 7.9E-02
-25 3.3E+00 3.3E+00 3.3E+00 3.3E+00 3.3E+00 3.3E+00 3.1E-02
Table 13: As data for αTU=40
αRLTimestep 238000 239000 240000 241000 242000 μ σ
0 3.2E-02 4.6E-02 3.6E-02 2.7E-02 2.6E-02 3.3E-02 8.5E-03
15 4.4E-02 3.8E-02 5.1E-02 1.0E-01 8.6E-02 6.4E-02 2.8E-02
-15 5.6E-02 7.6E-02 5.2E-02 3.5E-02 4.5E-02 5.3E-02 1.5E-02
-25 1.2E-01 1.3E-01 1.3E-01 1.2E-01 1.1E-01 1.2E-01 8.4E-03
Table 14: Rg data for αTU=40
αRLTimestep 238000 239000 240000 241000 242000 μ σ
0 1.1E-01 1.1E-01 1.1E-01 1.1E-01 8.5E-02 1.1E-01 1.2E-02
15 0.0E+00 0.0E+00 0.0E+00 0.0E+00 0.0E+00 0.0E+00 0.0E+00
-15 1.5E-01 1.7E-01 1.6E-01 1.3E-01 1.6E-01 1.5E-01 1.2E-02
-25 2.7E-01 2.6E-01 2.3E-01 2.6E-01 2.7E-01 2.6E-01 1.8E-02

Más contenido relacionado

Destacado

2014 presentation 2014_fr
2014 presentation 2014_fr2014 presentation 2014_fr
2014 presentation 2014_frOutLambda
 
Customer service culture
Customer service culture Customer service culture
Customer service culture Mahmoud Dargaly
 
Citations needed for the sum of all human knowledge: Wikidata as the missing ...
Citations needed for the sum of all human knowledge: Wikidata as the missing ...Citations needed for the sum of all human knowledge: Wikidata as the missing ...
Citations needed for the sum of all human knowledge: Wikidata as the missing ...Dario Taraborelli
 
The DigiMarketing Imperative
The DigiMarketing ImperativeThe DigiMarketing Imperative
The DigiMarketing ImperativeOgilvy
 
A journey in the world of animals. english
A journey in the world of animals. englishA journey in the world of animals. english
A journey in the world of animals. englishHarunyahyaEnglish
 
лекция № 1 ока опорно-двиг
лекция № 1 ока опорно-двиглекция № 1 ока опорно-двиг
лекция № 1 ока опорно-двигlali100226
 
Φάρμακα κοινωνικών δραστηριοτήτων ή φάρμακα σχετι
Φάρμακα κοινωνικών δραστηριοτήτωνή φάρμακα σχετιΦάρμακα κοινωνικών δραστηριοτήτωνή φάρμακα σχετι
Φάρμακα κοινωνικών δραστηριοτήτων ή φάρμακα σχετιcsdtesting
 
Ninong's video
Ninong's videoNinong's video
Ninong's videomtlacap818
 
A Journey to the Summit of Kilimanjaro
A Journey to the Summit of KilimanjaroA Journey to the Summit of Kilimanjaro
A Journey to the Summit of KilimanjaroNanosecond
 
Equity and Transparency in the New Province of Humanity
Equity and Transparency in the New Province of HumanityEquity and Transparency in the New Province of Humanity
Equity and Transparency in the New Province of HumanityValnora Leister
 
Schneider Electric offer in French
Schneider Electric offer in FrenchSchneider Electric offer in French
Schneider Electric offer in FrenchAoqi Liu
 
100%open innovation methods for charities toolkit session
100%open innovation methods for charities   toolkit session100%open innovation methods for charities   toolkit session
100%open innovation methods for charities toolkit sessionNesta
 
Article overview: Deep Neural Networks Reveal a Gradient in the Complexity of...
Article overview: Deep Neural Networks Reveal a Gradient in the Complexity of...Article overview: Deep Neural Networks Reveal a Gradient in the Complexity of...
Article overview: Deep Neural Networks Reveal a Gradient in the Complexity of...Ilya Kuzovkin
 
CINF 51: Analyzing success rates of supposedly 'easy' reactions
CINF 51: Analyzing success rates of supposedly 'easy' reactionsCINF 51: Analyzing success rates of supposedly 'easy' reactions
CINF 51: Analyzing success rates of supposedly 'easy' reactionsNextMove Software
 
Patrick Dodson on Airbnb Experience from scratch
Patrick Dodson on Airbnb Experience from scratchPatrick Dodson on Airbnb Experience from scratch
Patrick Dodson on Airbnb Experience from scratchDzmitry Bazhko
 

Destacado (18)

2014 presentation 2014_fr
2014 presentation 2014_fr2014 presentation 2014_fr
2014 presentation 2014_fr
 
Customer service culture
Customer service culture Customer service culture
Customer service culture
 
Citations needed for the sum of all human knowledge: Wikidata as the missing ...
Citations needed for the sum of all human knowledge: Wikidata as the missing ...Citations needed for the sum of all human knowledge: Wikidata as the missing ...
Citations needed for the sum of all human knowledge: Wikidata as the missing ...
 
The DigiMarketing Imperative
The DigiMarketing ImperativeThe DigiMarketing Imperative
The DigiMarketing Imperative
 
A journey in the world of animals. english
A journey in the world of animals. englishA journey in the world of animals. english
A journey in the world of animals. english
 
лекция № 1 ока опорно-двиг
лекция № 1 ока опорно-двиглекция № 1 ока опорно-двиг
лекция № 1 ока опорно-двиг
 
Φάρμακα κοινωνικών δραστηριοτήτων ή φάρμακα σχετι
Φάρμακα κοινωνικών δραστηριοτήτωνή φάρμακα σχετιΦάρμακα κοινωνικών δραστηριοτήτωνή φάρμακα σχετι
Φάρμακα κοινωνικών δραστηριοτήτων ή φάρμακα σχετι
 
Apple
Apple Apple
Apple
 
Ninong's video
Ninong's videoNinong's video
Ninong's video
 
A Journey to the Summit of Kilimanjaro
A Journey to the Summit of KilimanjaroA Journey to the Summit of Kilimanjaro
A Journey to the Summit of Kilimanjaro
 
Equity and Transparency in the New Province of Humanity
Equity and Transparency in the New Province of HumanityEquity and Transparency in the New Province of Humanity
Equity and Transparency in the New Province of Humanity
 
Schneider Electric offer in French
Schneider Electric offer in FrenchSchneider Electric offer in French
Schneider Electric offer in French
 
2 association rules
2 association rules2 association rules
2 association rules
 
100%open innovation methods for charities toolkit session
100%open innovation methods for charities   toolkit session100%open innovation methods for charities   toolkit session
100%open innovation methods for charities toolkit session
 
Article overview: Deep Neural Networks Reveal a Gradient in the Complexity of...
Article overview: Deep Neural Networks Reveal a Gradient in the Complexity of...Article overview: Deep Neural Networks Reveal a Gradient in the Complexity of...
Article overview: Deep Neural Networks Reveal a Gradient in the Complexity of...
 
CINF 51: Analyzing success rates of supposedly 'easy' reactions
CINF 51: Analyzing success rates of supposedly 'easy' reactionsCINF 51: Analyzing success rates of supposedly 'easy' reactions
CINF 51: Analyzing success rates of supposedly 'easy' reactions
 
Patrick Dodson on Airbnb Experience from scratch
Patrick Dodson on Airbnb Experience from scratchPatrick Dodson on Airbnb Experience from scratch
Patrick Dodson on Airbnb Experience from scratch
 
IPES Presentation
IPES PresentationIPES Presentation
IPES Presentation
 

Similar a CHEN90023 Lachlan Russell 389374

Jacob Kleine undergrad. Thesis
Jacob Kleine undergrad. ThesisJacob Kleine undergrad. Thesis
Jacob Kleine undergrad. ThesisJacob Kleine
 
Protdock - Aatu Kaapro
Protdock - Aatu KaaproProtdock - Aatu Kaapro
Protdock - Aatu KaaproSwapnesh Singh
 
43_EMIJ-06-00212.pdf
43_EMIJ-06-00212.pdf43_EMIJ-06-00212.pdf
43_EMIJ-06-00212.pdfUmeshYadava1
 
Molecular dynamics and Simulations
Molecular dynamics and SimulationsMolecular dynamics and Simulations
Molecular dynamics and SimulationsAbhilash Kannan
 
Structural Bioinformatics.pdf
Structural Bioinformatics.pdfStructural Bioinformatics.pdf
Structural Bioinformatics.pdfRahmatEkoSanjaya1
 
Thin Film Pressure Estimation of Argon and Water using LAMMPS
Thin Film Pressure Estimation of Argon and Water using LAMMPSThin Film Pressure Estimation of Argon and Water using LAMMPS
Thin Film Pressure Estimation of Argon and Water using LAMMPSCSCJournals
 
Granger.Deans.Research.Grant.Proposal
Granger.Deans.Research.Grant.ProposalGranger.Deans.Research.Grant.Proposal
Granger.Deans.Research.Grant.ProposalKathleen Granger
 
Monte Carlo Simulations & Membrane Simulation and Dynamics
Monte Carlo Simulations & Membrane Simulation and DynamicsMonte Carlo Simulations & Membrane Simulation and Dynamics
Monte Carlo Simulations & Membrane Simulation and DynamicsArindam Ghosh
 
利用分子動力學電腦模擬研究聚穀氨醯胺及胰
利用分子動力學電腦模擬研究聚穀氨醯胺及胰利用分子動力學電腦模擬研究聚穀氨醯胺及胰
利用分子動力學電腦模擬研究聚穀氨醯胺及胰Hsin-Lin Chiang
 
Final report - Adam Zienkiewicz
Final report - Adam ZienkiewiczFinal report - Adam Zienkiewicz
Final report - Adam ZienkiewiczAdam Zienkiewicz
 
GlucosePermeabilityofPDMS-PU-PEGMAInterpenetratingPolymerNetworks
GlucosePermeabilityofPDMS-PU-PEGMAInterpenetratingPolymerNetworksGlucosePermeabilityofPDMS-PU-PEGMAInterpenetratingPolymerNetworks
GlucosePermeabilityofPDMS-PU-PEGMAInterpenetratingPolymerNetworksMelissa Nguyen
 
Photophysics of dendrimers colombi
Photophysics of dendrimers   colombiPhotophysics of dendrimers   colombi
Photophysics of dendrimers colombiGiorgio Colombi
 
Cloud Pharmaceuticals white paper.LIE_2016
Cloud Pharmaceuticals white paper.LIE_2016Cloud Pharmaceuticals white paper.LIE_2016
Cloud Pharmaceuticals white paper.LIE_2016Shahar Keinan
 
Isogeometric_fluid-structure_interaction.pdf
Isogeometric_fluid-structure_interaction.pdfIsogeometric_fluid-structure_interaction.pdf
Isogeometric_fluid-structure_interaction.pdfesmaeelghafari2
 
Optimal Management of a Microgrid with Radiation and Wind-Speed Forecasting: ...
Optimal Management of a Microgrid with Radiation and Wind-Speed Forecasting: ...Optimal Management of a Microgrid with Radiation and Wind-Speed Forecasting: ...
Optimal Management of a Microgrid with Radiation and Wind-Speed Forecasting: ...VICTOR MAESTRE RAMIREZ
 
ME4474_independantrearch
ME4474_independantrearchME4474_independantrearch
ME4474_independantrearchSamuel Gates
 
Tutorial Mesodyn
Tutorial MesodynTutorial Mesodyn
Tutorial Mesodynlab13unisa
 

Similar a CHEN90023 Lachlan Russell 389374 (20)

Jacob Kleine undergrad. Thesis
Jacob Kleine undergrad. ThesisJacob Kleine undergrad. Thesis
Jacob Kleine undergrad. Thesis
 
Protdock - Aatu Kaapro
Protdock - Aatu KaaproProtdock - Aatu Kaapro
Protdock - Aatu Kaapro
 
THESIS FINAL
THESIS FINALTHESIS FINAL
THESIS FINAL
 
43_EMIJ-06-00212.pdf
43_EMIJ-06-00212.pdf43_EMIJ-06-00212.pdf
43_EMIJ-06-00212.pdf
 
Molecular dynamics and Simulations
Molecular dynamics and SimulationsMolecular dynamics and Simulations
Molecular dynamics and Simulations
 
Structural Bioinformatics.pdf
Structural Bioinformatics.pdfStructural Bioinformatics.pdf
Structural Bioinformatics.pdf
 
Research summary
Research summaryResearch summary
Research summary
 
Thin Film Pressure Estimation of Argon and Water using LAMMPS
Thin Film Pressure Estimation of Argon and Water using LAMMPSThin Film Pressure Estimation of Argon and Water using LAMMPS
Thin Film Pressure Estimation of Argon and Water using LAMMPS
 
Granger.Deans.Research.Grant.Proposal
Granger.Deans.Research.Grant.ProposalGranger.Deans.Research.Grant.Proposal
Granger.Deans.Research.Grant.Proposal
 
Monte Carlo Simulations & Membrane Simulation and Dynamics
Monte Carlo Simulations & Membrane Simulation and DynamicsMonte Carlo Simulations & Membrane Simulation and Dynamics
Monte Carlo Simulations & Membrane Simulation and Dynamics
 
利用分子動力學電腦模擬研究聚穀氨醯胺及胰
利用分子動力學電腦模擬研究聚穀氨醯胺及胰利用分子動力學電腦模擬研究聚穀氨醯胺及胰
利用分子動力學電腦模擬研究聚穀氨醯胺及胰
 
thesis.compressed
thesis.compressedthesis.compressed
thesis.compressed
 
Final report - Adam Zienkiewicz
Final report - Adam ZienkiewiczFinal report - Adam Zienkiewicz
Final report - Adam Zienkiewicz
 
GlucosePermeabilityofPDMS-PU-PEGMAInterpenetratingPolymerNetworks
GlucosePermeabilityofPDMS-PU-PEGMAInterpenetratingPolymerNetworksGlucosePermeabilityofPDMS-PU-PEGMAInterpenetratingPolymerNetworks
GlucosePermeabilityofPDMS-PU-PEGMAInterpenetratingPolymerNetworks
 
Photophysics of dendrimers colombi
Photophysics of dendrimers   colombiPhotophysics of dendrimers   colombi
Photophysics of dendrimers colombi
 
Cloud Pharmaceuticals white paper.LIE_2016
Cloud Pharmaceuticals white paper.LIE_2016Cloud Pharmaceuticals white paper.LIE_2016
Cloud Pharmaceuticals white paper.LIE_2016
 
Isogeometric_fluid-structure_interaction.pdf
Isogeometric_fluid-structure_interaction.pdfIsogeometric_fluid-structure_interaction.pdf
Isogeometric_fluid-structure_interaction.pdf
 
Optimal Management of a Microgrid with Radiation and Wind-Speed Forecasting: ...
Optimal Management of a Microgrid with Radiation and Wind-Speed Forecasting: ...Optimal Management of a Microgrid with Radiation and Wind-Speed Forecasting: ...
Optimal Management of a Microgrid with Radiation and Wind-Speed Forecasting: ...
 
ME4474_independantrearch
ME4474_independantrearchME4474_independantrearch
ME4474_independantrearch
 
Tutorial Mesodyn
Tutorial MesodynTutorial Mesodyn
Tutorial Mesodyn
 

CHEN90023 Lachlan Russell 389374

  • 1. Department of Chemical and Bimolecular Engineering CHEN90023 Chemical Engineering Research Project Simulation of the Transmembrane Transport of Poly(amidoamine) Lachlan Russell 389374 Supervisors: Prof. Scales, Assoc. Prof. Li-Tang Yan
  • 2. ii Summary The effect of modifications made to Yan and Yu’s existing dissipative particle dynamics (DPD) model of an interacting generation five (G5) Poly(amidoamine) (PAMAM) dendrimer and a lipid bilayer were examined.1, 2 The primitive bilayer was modified to include another double-chain lipid to simulate the effect of ligand receptors on the transmembrane transport of dendrimer structures. The structures formed by the interacting lipid bilayer and G5 dendrimer were analysed to see if they resembled morphologies likely to be involved in clathrin-mediated endocytosis (CME). The effect of changes to DPD soft-core repulsion forces acting between the dendrimer surface beads and the receptor beads, as well as the internal dendrimer structure and the hydrophobic lipid tails, were analysed. It was found that self-assembly of the bilayer structure was not compromised when converting a quarter of the simulated lipid molecules into receptor-simulating lipids. Binding and penetration were found to be largely affected by the inclusion of an attractive interaction between receptors and ligands coupled with changes to the magnitude of the repulsion between the bilayer lipid tails and the internal structure of the dendrimer. Encapsulation of the dendrimer by the bilayer was observed for a particular set of soft-core repulsion parameters. The morphology closely matched that of a phagocytic mechanism. On the simulated length scale such a result was unexpected. Further refinement of the model is required if CME, as it is currently understood, is to be simulated and the unobservable dynamics conjectured. Due to computational constraints, many of the results presented are preliminary in nature.
  • 3. iii Contents Summary....................................................................................................................................................................................ii 1.0 Introduction.........................................................................................................................................................................4 2.0 Relevant Theory...................................................................................................................................................................5 2.1 Dissipative Particle Dynamics (DPD) ...............................................................................................................................5 2.2 Dendrimers .....................................................................................................................................................................9 2.3 Lipid Bilayers ...................................................................................................................................................................9 2.4 Transmembrane Transport: ..........................................................................................................................................10 3.0 Literature Review...............................................................................................................................................................12 3.1 A Case for Dissipative Particle Dynamics (DPD) ............................................................................................................12 3.2 Using Dissipative Particle Dynamics to Model Transmembrane Transport ..................................................................13 3.3 Ligand-Receptor Modelling...........................................................................................................................................14 3.4 Aim of Research ............................................................................................................................................................14 4.0 Materials and General Methodology.................................................................................................................................15 5.0 Results and Discussion.......................................................................................................................................................17 5.1 Self-Assembly of the Modified Lipid Bilayer..................................................................................................................17 5.2 Comparison with Yan and Yu’s Results .........................................................................................................................19 5.3 Simulation of the System with Dendrimer Heads modelled as Ligands........................................................................21 5.3.1 Initial Attempts to Generate Clathrin-mediated Endocytic Morphology ..............................................................21 5.3.2 Further Attempts to Generate Clathrin-mediated Endocytic Morphology ...........................................................24 6.0 Further Discussion .............................................................................................................................................................28 6.1 Known Shortcomings of the Model...............................................................................................................................28 6.2 Relationship of CME to Simulations..............................................................................................................................30 6.3 The Notable Shortage of Empirical Data.......................................................................................................................30 7.0 Conclusions........................................................................................................................................................................32 8.0 Suggestions for Further Work............................................................................................................................................33 8.1 Simple Immediate Revisions to the Current DPD Model ..............................................................................................33 8.2 Extension of Results......................................................................................................................................................33 9.0 References .........................................................................................................................................................................35 10.0 Appendices ......................................................................................................................................................................38 Appendix 1: Simulation Soft-core Repulsion Parameters ...................................................................................................38 Appendix 2: MATLAB Code for Bilayer Bead Density Profile...............................................................................................41 Appendix 3: Asphericity ......................................................................................................................................................43 Appendix 4: Radius of Gyration ..........................................................................................................................................45 Appendix 5: Penetrability....................................................................................................................................................47 Appendix 6: Confidence Bands ...........................................................................................................................................49 Appendix 7: Intermediate Data...........................................................................................................................................50
  • 4. 4 1.0 Introduction Due to their complex nature, much of the science surrounding forces acting between nanoparticles is not well understood.3 Dissipative particle dynamics (DPD) can be used to effectively model the mesoscopic (10 nm - 100 mm) dynamics of interacting particles (refer to Figure 1).4, 5 DPD attempts to incorporate the essence of many forces believed to be significant on this scale. The approximations utilised by DPD methods allow for mesoscale simulation of many physical interactions that otherwise could not be run at acceptable speeds. Interactions between synthetic dendrimers and cellular membranes are significant due to the growing use of dendrimers in medicine.6 These mesoscale interactions need to be studied in order to synthesise drugs which allow for specific transmembrane transport whilst moderating cytotoxicity.7 DPD has been used to simulate dendritic particles interacting with lipid bilayers.1, 2, 8 Such simulations can model the behaviour of different types of dendrimers reducing the need for costly drug testing by synthesis and administration. Figure 1: Schematic showing coarse-graining used to model the mesoscale. 9 Interactions between lipid bilayers and dendrimers have been studied extensively. However, modern experimental techniques are currently unable to answer many fast-acting molecular-level mechanistic questions.10 It is hoped, given DPD’s success with modelling other mesoscopic phenomena, that simulations can provide additional insight until such a time when observational techniques are sufficiently advanced to provide a more detailed description of interaction mechanisms and morphology. This project aims to extend previous DPD models of interacting lipid bilayers and dendrimers developed by Yan and Yu.1, 2 Whilst amphiphilic lipid molecules comprise the bulk of cell membranes, transport mechanisms often involve receptor proteins which are also present in the structure of the membrane. The simulations performed incorporate primitive receptor-type molecules into Yan and Yu’s simple bilayer. Such modifications have been utilised in other simulations however the effect of such changes on the transmembrane transport of dendrimers such as poly(amidoamine) (PAMAM) have yet to be rigorously studied.11-13 The effects of this change, along with the variation of DPD soft- core repulsion parameters, was compared with Yan and Yu’s experiments, other simulation studies and mechanisms proposed by biologists. Specifically, it was assessed whether clathrin-mediated
  • 5. 5 endocytic morphology could be replicated by these modifications as it has been shown that PAMAM dendrimers and other particles of similar length scales are internalised via this mechanism.14, 15 The theory behind dissipative particle dynamics will first be introduced followed by a brief description of dendrimers, cell membranes and transmembrane transport. Subsequently, a review of literature concerning the use of computer simulation to model these transport processes will be presented. The details of the modified DPD model will be covered and the results of simulations examined. Lastly, conclusions and suggestions for further research will be presented. 2.0 Relevant Theory 2.1 Dissipative Particle Dynamics (DPD) Hoogerbrugge and Koelman developed a coarse-grained/beaded simulation method combining features of molecular dynamics (MD) and lattice-gas automata. Dissipative particle dynamics simulates more quickly than MD and is more flexible than lattice-gas automata schemes.16 A reformulation of DPD was developed and applied to the modelling of bilayers by Groot and his colleagues.17, 18, 19 The simulation equations adopted by Yan and Yu are largely consistent with the aforementioned theory developed by Groot; these equations were utilised for the sake of consistent simulation.1 Newton’s equations of motion are solved for interacting particles: (1) For simplicity, the mass of particles are set to equal 1 and hence force equals acceleration. Forces act within a dimensionless cut-off radius (rc) of 1. The bilayer-dendrimer DPD system is governed by five forces: Conservative, Dissipative, Random, Bond and Electrostatic. The forces acting on particle i are summed at each timestep: (2) Conservative force: (3) (4), (5), (6)
  • 6. 6 aij denotes the soft-core repulsion parameter between particles i and j and is related to the compressibility of interacting particles.9 The repulsion parameter can often be related to Flory-Huggins χ parameters in the instance where water is used as the solvent: 𝝌𝒊𝒋 ≈ 𝜶 𝒊𝒋−𝜶 𝒊𝒊 𝟑.𝟐𝟕 (7) • αii =25 is used in general DPD systems for all particle types. • αij will be larger than 25 for strong bead-bead repulsion interactions; it will be smaller than 25 where there exists reasons for attraction between two types of beads (e.g. due to electrostatic or entropic effects). For many of the cross terms, estimations are tested and refined based on qualitative knowledge and simulation results. A complete set of accurate Flory-Huggins χ-parameters often cannot be generated for complex systems with many types of beads present. Dissipative and Random forces: The dissipative force and the random force act as the heat sink and source, respectively. They are balanced to maintain a Boltzmann distribution and hence create a thermostat. (8) (9) • wD and wR are r-dependent weight functions vanishing for r>rc. For simplicity they are chosen to be: (10) • vij=vi-vj and vi denotes the velocity of bead i. • ξij is a random number with zero mean and unit variance. • The noise amplitude, σ, is fixed at σ =3. • The forces are connected by Eq. (11). (11) • This thermostat conserves linear and angular momentum, resulting in a correct description of hydrodynamics.
  • 7. 7 Bond forces: Dendrimers and lipids are constructed by tying beads together using Hookean springs with the potential: (12) For the purposes of simulation: • The spring constant is set to ks=64. • The unstretched length is set to l0=0.5rc. • Hence, the average bond length is fixed. A three-body potential acting between adjacent bead triples models hydrocarbon chain stiffness in the lipids and dendrimer: (13) • The angle φ is defined by Eq. (14) where A, B denote the two bonds connecting beads i-1, i, and i+1: 𝐜𝐨𝐬−𝟏 𝑨∙𝑩 ‖𝑨‖‖𝑩‖ = 𝝓 (14) • The bending constant ka=20 is used for both the lipid and dendrimer. • The preferred angle, φ0, is specified as 00 and 1200 for lipids and dendrimer respectively. From these potentials, corresponding forces can be calculated. Electrostatic force:18 A lattice grid is constructed, over which the electrostatic field is spread. There is a trade-off made between correct representation and fast implementation of the field. Each charged bead has charge proportional to: (15) This charge is calculated for every grid node within a radius Re =1.6rc of the bead and normalised so that the sum of all proximal charged nodes is equal to the charge of the bead.
  • 8. 8 The electric field is solved according to Eq. 16: (16) • ε, εr and εo, denote the dielectric permittivity of the medium, the vacuum and the water respectively. • 𝜌̅ 𝑒,𝑛 is the averaged charge density. • ∇𝜑 𝑛 is the electric gradient at node n. • Γ =13.87 is the coupling constant corresponding to the Bjerrum length of water at 300K. At each timestep the electrostatic force on an ion can finally be determined by Eq. (17): (17) • Where qi is the charge of the ion. Algorithm: The velocity-Verlet algorithm is iterated for each bead at each timestep: 𝜆 = 0.65 has been found to reduce computational error accumulating whilst allowing for a relatively large timestep to be chosen. In the simulations, the radius of interaction, the bead mass, and the temperature are set such that rc=m=kBT=1. A characteristic time scale can thus be defined: (18)
  • 9. 9 2.2 Dendrimers Dendrimers are snowflake-like macromolecules (refer to Figure 2). They have defined controllable structures of known physical, chemical and biological properties. Their associated monodispersity and reproducibility make them ideal for biological applications as structure-activity relationships can be more accurately studied.20 They are defined by their core structure and, as additional shells (generations) are added, approximately double in size and number of surface functional groups. The chemical backbone and surface terminal groups have a large effect on the biological properties of dendrimers. Figure 2: Poly(amidoamine) (PAMAM): Structure comprises of repetitively branched amide and amine functional groups.20 Dendrimers are increasingly being used for drug delivery as, when paired with pharmaceutically active compounds, their defined properties can be used to improve drug specificity.21 Dendrimers are capable of delivering drugs by many mechanisms. They can function as micelles which encapsulate hydrophobic drugs; alternatively, the terminal functional groups of the dendrimer can covalently attach to pharmaceutical agents.20 Dendrimers used for biomedical purposes usually have specific ligands attached to their often cationic surface which interact with receptors on the surface of cellular membranes. 2.3 Lipid Bilayers Phospholipids are surfactant molecules with a hydrophilic head and two hydrophobic tails which can self-assemble to form bilayers (refer to Figure 3). Complex lipid bilayers encapsulate cells and many important organelles.2 Figure 3: Bilayer sheet.22
  • 10. 10 2.4 Transmembrane Transport: The plasma membranes of cells serve to separate the intracellular cytoplasm from the extracellular environment by controlling the transport of molecules across the membrane.23 Endocytosis is one of many types of processes by which cells internalise molecules (refer to Figure 4). It involves the deformation of the cell membrane and, subsequently, smaller membrane-bound carriers are generated.24 Figure 4: Various methods of mass transport across cell membranes.25 Clathrin-mediated endocytosis (CME) is the chief receptor-mediated endocytic pathway.25 Hence, it is the most widely studied endocytic mechanism. First observed by using electron microscopy, CME is vital for such functions as nutrient uptake and receptor signalling.26 Mechanistically it involves proteins with receptor sites binding to specific ligands, triggering the formation of ‘coated pits’ which subsequently become fully enveloped by the cell (refer to Figure 5).24 Figure 5: Clathrin-mediated endocytosis requires specific ligand-receptor binding interactions.27
  • 11. 11 Different parameters affect the ability of nanoparticles to pass through the cellular membranes including, size, shape and surface chemistry.6 Surface chemistry strongly affects the interactions between nanoparticles and cells.24 Nonspecific binding forces caused by nanoparticle characteristics such as cationic charge and roughness can also promote cellular uptake in conjuction with ligand- receptor interactions.27 Consequently, it is believed that multiple transmembrane transport mechanisms can work simultaneously (refer to Figure 6).6 Figure 6: Receptor-mediated uptake with additional nonspecific binding forces.27
  • 12. 12 3.0 Literature Review In recent years, multiple studies have modelled nanoparticle lipid bilayer interactions using computer simulation.7 This is because, despite extensive research in this area, experimentalists have largely failed to elucidate the kinetic evolutions of these interactions. Dissipative particle dynamics (DPD) and other coarse-graining techniques have been used widely to attempt to model these processes. Many of these studies have concerned themselves with the effects of nanoparticle shape, size and surface chemistry on the internalisation process.7 Few have concerned themselves with seeking to replicate the true complexity of cell membranes instead modelling them simply as lipid bilayer sheets. Lately, attempts have been made to model these membranes in increasingly representative ways. The effect of the inclusion of ligand-receptor type molecules on a simulated bilayer-dendrimer system requires further study in the wake of recent developments.12, 13, 28, 29 3.1 A Case for Dissipative Particle Dynamics (DPD) At the nanoscale, interfacial forces dominate interactions. These include van der Waals, electrostatic, depletion, hydration and hydrophobic forces.24 Compared with atomistic models which seek to incorporate much of the complexity of modern physics, DPD utilises a different force potential between interacting beads. In atomistic models, hard-core potentials such as Lennard–Jones (L–J) potential, which incorporates van der Waals attraction and Pauli exclusion forces are used. In DPD models, a soft-core conservative force is used.9 This soft-core repulsion is important as hard-core modelling causes a ‘caging’ effect, where atoms undergo multiple collisions before any transport occurs.18 Hence, models using soft repulsive spheres allow for the use of larger timesteps and length scales. Figure 7: Schematic illustrating the hard-core nature of Lennard–Jones potential and the soft-core nature of DPD inter- particle potential.9 The DPD model attempts to compensate for missing physics by incorporating a dissipative force to account for viscous effects; the molecular level randomness, which usually manifests as Brownian motion, is modelled by the random force.9 Both DPD and MD methods can produce correct equilibrium distributions of polymer chains. However, MD simulations do not include hydrodynamic interactions. Conversely, DPD is able to correctly simulate compressible Navier-Stokes behaviour.30, 31 This result is important at the mesoscale. It is currently possible to simulate novel scenarios of small dendrimers interacting with simple lipid bilayers using atomistic MD models; however, in order to generate meaningful results at greater length and time scales, coarse-graining is required.7 Coarse-graining enables significant speeding up of simulations of interacting membranes and particles orders of magnitude larger.22, 32
  • 13. 13 The advantages and disadvantages of other coarse-graining techniques have been discussed at length.33 Groot and Madden argue that what is chiefly required of a mesoscale physics model is that it can reproduce thermodynamic properties on the relevant length scale. Namely, if the solubilities, liquid compressibilities and the shape of interfaces are correct, then the model correctly represents the physical system for the given length scale.4 Well-conditioned DPD simulations have been shown to pass such tests.4 3.2 Using Dissipative Particle Dynamics to Model Transmembrane Transport The fundamentals and technical subtleties of DPD are well established.4, 18, 19, 34 Traditionally DPD has been used to model polymer-colloid systems where robust experimental results abound.4, 30, 34 Shilcock and Lipowsky were able to formulate self-assembling planar bilayers with experimentally consistent density profiles and lateral stress distributions using the DPD model.22 These models have been shown to describe the phase behaviour of phospholipids with accuracy.33, 35 Following the successful development of thermodynamically robust self-assembling bilayers many have attempted to study their interactions with nanoscale particles. The modelling of the nanoparticle-bilayer interactions have typically involved the use of the polyamidoamine (PAMAM), which is the most commonly used dendrimer in biomedical applications, and a simple bilayer comprised entirely of dipalmitoylphosphatidylcholine (DPPC).21 In addition to DPD, a variety of coarse- grained and atomistic simulation models have been used for this purpose.10, 13, 36, 37 Yan and Yu used the bilayer developed by Shilcock and Lipowsky to model interactions with cationic G5 PAMAM dendrimers. The effect on morphology caused by changes to the soft-core repulsion parameters, especially those concerning the outer-dendrimer hydrophilic components and the inner-dendrimer hydrophobic components, was analysed extensively.2 Cationic dendrimers are known to be cytotoxic.37, 38 By introducing controlled surface tension into the bilayer model, Yan and Yu were able to observe holes forming on the bilayer surface on interaction with PAMAM dendrimers consistent with experimentally observed phenomena.1 They were subsequently able to control the soft-core interaction parameters of the lipids used in the bilayer to modify their shape factor ( 𝜈 𝛼𝑙 < 1) to instead form vesicles of various sizes and surface tensions and again interact these with PAMAM dendrimers.8 An attempt to better model the endocytic process was incorporated by Guo, Mao and Yan.39 Canham-Helfrich theory was applied to incorporate contact and membrane bending energies.40 Figure 8: DPD simulated endocytosis with incorporated membrane elasticity forces.39
  • 14. 14 3.3 Ligand-Receptor Modelling The use of DPD to model endocytosis is still in its infancy. Until recently, the simplicity of the simulated bilayers was such that meaningful comparisons to cell membranes were tenuous. Most transport across cell membranes occurs with the involvement of receptor proteins. Very little work has been attempted to model clathrin-mediated endocytosis using simulations as it can be difficult to incorporate changes to the bilayer that do not compromise its self-assembling properties. Attempts to circumvent this problem have been made by modifying the surface of interacting dendrimers to have a proportion of surface beads exhibit a greater affinity for the bilayer.13 More direct attempts modelled ligand-receptor interactions by embedding symmetrical plug-like receptors into a bilayer and interacting them with spherical molecules with uniformly anchored ligand sites. CME mechanisms resembling those proposed by biologists were generated.28 Another strategy involved the incorporation of another type of lipid into the bilayer structure with a ligand-receptor headgroup. This method preserved self-assembly.11-13 The addition of ligands to the model has been approached in a variety of ways. Yang and Ma model their interacting particle as a single large bead which interacts with ligand receptors according to a modified Lennard–Jones potential in order to incorporate the attractive force expected between ligands and receptors.29 Other studies have all used different models to incorporate ligands into simulations.11, 12, 29, 41 While many of these studies produce interesting findings they all fail to model the interacting particle as anything more than solid polyhedrons or spheres. 3.4 Aim of Research The aim of this study is to introduce a receptor-simulating double-chain lipid into the bilayer and interact this modified bilayer with more complex particles such as PAMAM dendrimers. It has been shown that PAMAM dendrimers are internalised primarily through clathrin-mediated endocytosis.14, 15 It is hypothesised that these structural changes to the bilayer combined with the balancing of soft- core repulsion parameters can induce clathrin-mediated endocytic (CME) morphology. In order to test this hypothesis a self-assembling bilayer must first be able to be generated. The stability of the DPD model will be evaluated on comparing with Yan and Yu’s previous simulations.1, 2 Only then can it be determined if structures resembling CME can be generated with or without the additional refinement of some soft-core repulsion parameters. Any such refinements should be justifiable either with empirical data or, less ideally, using qualitative arguments. Depending on the strength of these justifications the hypothesis may be able to be supported. Consequently, it could be argued that the simulation results might describe currently unobservable dynamics. If so, the model or its derivatives may be able to assist in the future development of new dendrimers for drug delivery.
  • 15. 15 4.0 Materials and General Methodology The DPD simulation was constructed using interacting coarse-grained molecules largely consistent with previous work by Yan and Yu.1 The lipid bilayer was constructed by creating a coarse-grained representation of dipalmitoylphosphatidylcholine, the phospholipid commonly used to study lipid bilayers. DPPC consists of a phosphate group, a simple organic molecule choline and two palmitic acid chains. Figure 9: Dipalmitoylphosphatidylcholine.17 The simulation model of DPPC is consists of a hydrophilic head with charge +1 (green), a hydrophilic head with charge -1 (purple) a head without charge (blue) and a hydrophobic tail group (cyan). The separation between the beads was set to 0.5rc. Figure 10: Bead model of DPPC.1 Another type of lipid was introduced into the bilayer structure.11, 12 Primitive receptor-simulating double-chain lipids were created using a similar structure consisting of two uncharged hydrophilic heads (orange), another uncharged head (blue) and a hydrophobic tail group (cyan). The separation between the beads was set to 0.5rc. Figure 11: Bead model of primitive receptor-simulating double-chain lipid. For each of the double-chained lipids introduced into the simulation box (approximately 2360 in total) a random number generator gave the lipid a 25% chance of being a receptor-simulating double-chain lipid (refer to Further Discussion). The DPD simulation was allowed to run in an aqueous environment until a steady bilayer structure self-assembled.
  • 16. 16 A single G5 PAMAM dendrimer was subsequently introduced into the simulation box above the bilayer. For illustrative purposes, the coarse-grained representation of a G1 PAMAM dendrimer is depicted in Figure 12. Figure 12: Bead model of G1 PAMAM dendrimer.1 The yellow beads represent the hydrophobic inner amine and amide functional groups and are uncharged. The red beads represent the hydrophilic surface amine functional groups and carry a charge of +1. The separation between the beads was set to 0.5rc. Consistent with Yan and Yu’s modelling, counterions were also included in the simulation.2 Each simulation was run for more than 105 timesteps of length Δt=0.02τ in order to ensure accurate temperature control. The approximate area per lipid in a tensionless DPPC membrane is 0.64 nm2 . From this, the cut-off radius (rc) can be estimated to be approximately 0.7 nm. The time unit (τ) can be related to physical time by known inplane lipid diffusion coefficients. Experiments have shown this value to be approximately 5 μm2 /s. Relating this to Yan and Yu’s previous work τ≈7.7ns and, hence, 105 time steps equates to approximately 15 μs.2 Similarly, most of the soft-core repulsion parameters used, (refer to Appendix 1: Table 4), were compiled by Yan and Yu from the literature and, in the absence of sources, refined through qualitative arguments and iterative simulation.1 The introduction of the new receptor-simulating lipids required additional interaction parameters to be generated. The effect of varying some of these parameters, both new and old, was studied in order to attempt to build a model that successfully simulated a bilayer-dendrimer system interacting in a way that mimicked proposed clathrin-mediated endocytic mechanisms (refer to Results and Discussion). The Fortran code used to perform these DPD simulations has been made available (refer also to Theory: Dissipative Particle Dynamics (DPD)). Similarly, the MATLAB programs written to perform subsequent analysis of binary output from these programs can be found in the appendices.
  • 17. 17 5.0 Results and Discussion 5.1 Self-Assembly of the Modified Lipid Bilayer In an attempt to ensure self-assembly and an even distribution of receptors on the surface of the bilayer; parameters for the new receptor beads were set such that their affinity for the other headgroup beads was equivalent to their affinity for one another (α=25). As per other headgroup beads, the receptor beads were designed to be similarly repelled by the hydrophobic lipid chains (α=40).1 So as to achieve the slight protrusion of these receptors above the bilayer (refer to Figures 5, 6) receptor beads were given smaller repulsion parameters than other headgroup beads to govern their interactions with water (α=15) (refer to Appendix 1: Table 5). Consistent with expectations, for these carefully selected soft-core repulsion parameters, the introduction of the new lipid molecule had no effect on the self-assembling property of the bilayer. It was observed that an even surface distribution of this new coarse-grained molecule was achieved (refer to Figure 13). The total lipid density was determined to be 1.47𝑟𝑐 2 =0.72 nm2 approximately 13% greater than Yan and Yu’s literature value of 0.64 nm2 .1 This minor difference was due to a change to the ‘target lipid density’ input variable in the code used to simulate the self-assembly of the bilayer. The simulation box is continuous in the sense that beads cannot leave the box and are simply transported to the other side if they move across a boundary. The targeted density is achieved by the coded insertion or removal of more lipids once the bilayer structure is relatively stable. Simulation of the steady-state self-assembly of a bilayer required significant computer processing. This minor change to the ‘target lipid density’ was performed at the request of a colleague due to limited computing resources in order to test the effect of such a change on the stability of the bilayer structure. This increase in density lead to slightly less exposure of the lipid tails when compared to the surfaces originally constructed by Yan and Yu.1, 2 This deviation was partially offset by the slight protrusion of receptor beads into the water phase. Figure 13: Self-assembling bilayer produced with 1:3 ratio of receptor-simulating double-chain lipids and DPPC lipids.
  • 18. 18 Figure 14: Bilayer bead density by displacement from bilayer beads’ average z-coordinate (refer to Appendix 2). As illustrated by Figure 14, consistent with the DPD model, total bead density (solvent included) was centered on 3 beads/nm3 .17 The density distribution for the bilayer beads was largely consistent with previous studies with expected deviation arising from the 1:3 ratio of receptor-simulating double- chain lipids and DPPC lipids.1, 22 The paired receptor beads protruded approximately 0.4 nm further into the water than the two topmost combined headbeads on the DPPC lipids (refer to Figures 10,11). Due to computational constraints it was not possible to revise the receptor-water interactions, lipid- receptor split and lipid density to be consistent with literature (refer to Suggestions for Further Work).
  • 19. 19 5.2 Comparison with Yan and Yu’s Results To further test the stability of the modified DPD model, attempts were made to replicate results from Yan and Yu’s previous studies.1, 2 The affinity of all lipid headgroup beads for the external dendrimer beads and the internal dendrimer beads were set to αRL=15 and αTU=28 respectively, consistent with Yan and Yu’s soft-core interaction parameters (refer to Appendix 1: Table 6). After 105 timesteps (approximately 15 μs), a contrasting structure was generated: Figure 15: Comparison of dendrimer penetration after 15 μs. Note that the counterions are not shown in both figures; however, their presence is incorporated into both models consistently. Due to the presence of the receptor-simulating lipid in the bilayer structure, this simulation is subtly different to Yan and Yu’s previous work. Figure 15 depicts a bilayer structure which is less planar, with lipids capable of migrating onto the surface of the dendrimer; this was unexpected. The reasons for these discrepancies require further investigation. However, it is worth noting that the bending of the bilayer reduced as the simulation progressed and the migration of lipids onto the dendrimer was observed by Yan and Yu in another simulation which used the same soft-repulsion parameters.2 This finding potentially highlights the effect of DPD’s random force on simulations. The simulated dendrimer was also slightly more spherical, with the asphericity calculated to be As=0.033 compared with Yan and Yu’s figure centred on As=0.055 (refer to Appendix 3); the magnitude of this variation was deemed not to be significant given the dynamic nature of the system. Importantly, the structure of the simulated dendrimer was considerably more open in the modified system with the radius of gyration equal to Rg=3.8 nm compared with Rg=2.7 nm (refer to Appendix 4). The magnitude of this deviance was unexpected. The most likely causes of this deviation are unintentional changes from historical soft-core repulsion parameters that govern the simulation environment of the dendrimer equilibrating in water prior to being introduced into the simulation box with the lipid bilayer. Table 1: Soft-core repulsion parameters governing isolated dendrimer-water system. Solvent Bead (0) Surface Dendrimer Bead (+1) Internal Dendrimer Bead (0) Solvent Bead (0) 25 25 60 Surface Dendrimer Bead (+1) 25 25 30 Internal Dendrimer Bead (0) 60 30 25 (Yan & Yu 2009)
  • 20. 20 Table 2: Historic soft-core repulsion parameters used in the combined simulation.1, 2 Past simulation of the G5 PAMAM dendrimers in water using the parameters defined in Table 2 yielded compact spherical structures with diameters consistent with empirical measurements.2, 42 Figure 16: Equilibrium configuration of a G5 PAMAM dendrimer in water 2. The effect of the soft-core repulsion parameters changes, especially the reduction of the hydrophobicity of internal dendrimer beads (from 80 to 60), caused the dendrimer to equilibrate to a size approximately 40% larger than desired. Figure 17: Dendrimer introduced into the combined simulations The simulation codes which modelled the dendrimer isolated from the bilayer were not made available and the artifacts were only detected after returning to Australia and hence could not be rectified. In the combined simulation environment the parameters were restored to their original values; however, the dendrimer did not have sufficient time to structurally equilibrate before interacting with the bilayer in its more open form (refer to Suggestions for Further Work). Minor structural variation was likely further compounded by other sources. As previously mentioned, due to the inclusion of random forces in the DPD model an exact morphological match was not expected. Furthermore, the incorporated hydrophilicity of the receptor beads caused the bilayer surface to be more open. The internal dendrimer beads (yellow) were modelled as hydrophobic. In addition the parameters defining the interaction between the hydrophobic lipid tails (cyan) and both types of dendrimer beads (red, yellow) were less repulsive than the parameters defining the interactions between the hydrophobic lipid tails (cyan) and the lipid headgroup beads (green, blue, orange, purple). As a result penetrability of the dendrimer increased from fp=0.02 to fp=0.1109 when interacting with the new bilayer with the more open surface (refer to Appendix 5). Solvent Bead (0) Surface Dendrimer Bead (+1) Internal Dendrimer Bead (0) Solvent Bead (0) 25 20 80 Surface Dendrimer Bead (+1) 20 25 28 Internal Dendrimer Bead (0) 80 28 25
  • 21. 21 5.3 Simulation of the System with Dendrimer Heads modelled as Ligands Despite the non-steady-state morphological discrepancies and discussed deficiencies of the model, it was tested whether the generation of structures resembling those expected if clathrin-mediated endocytosis were responsible for transmembrane transport was possible. Further refinement of soft- core repulsion parameters was required. It was assumed that the surface amine functional groups of the dendrimer could substitute for ligand complexes (refer to Further Discussion). The effect of this assumption was tested by holding constant all soft-core repulsion parameters for the surface dendrimer beads excluding the parameter governing the surface dendrimer bead-receptor bead interaction (αRL). This parameter was, in many instances, modelled to be attractive in order to simulate the strong binding relationship between ligands and receptors.11 5.3.1 Initial Attempts to Generate Clathrin-mediated Endocytic Morphology The following structures were achieved after 105 timesteps (approximately 15 μs). The surface dendrimer bead soft-core interaction parameter was varied from a repulsive αRL=15 through to an attractive αRL=-30 (refer to Appendix 1: Table 7): Figure 18: Penetration profiles of simulations. For the purposes of clear illustration of the counterions and part of the bilayer structures have been omitted. In these simulations, encapsulation of the dendrimer was not observed. The asphericity and radii of gyration were larger than would be expected if encapsulation were responsible for penetration. Figure 19: Radii of gyration of simulated dendrimer- bilayer systems after approximately 105 timesteps. Figure 20: Asphericity of simulated dendrimer-bilayer systems after approximately 105 timesteps. αRL=15 αRL=0 αRL=-15 3.6 3.7 3.8 3.9 4 4.1 -20 -10 0 10 20 Radiusof GyrationRg(nm) Soft-core Repulsion (αRL) 0 0.01 0.02 0.03 0.04 0.05 0.06 -20 -10 0 10 20 Asphericity(As) Soft-core Repulsion (αRL)
  • 22. 22 Further data is required if the observed trends of increasing radii of gyration and asphericity with the decrease in the soft-core interaction parameter αRL are to be substantiated. Due to computing constraints, an inadequate number of different values of αRL were tested. Furthermore, sampling at t=100,000 was not performed by calculating statistics from multiple simulation runs with the same αRL value. Only one simulation per tested value of αRL was able to be run and statistics were calculated from neighboring timesteps (98,000; 99,000; 100,000; 101,000; 102,000). As such, error bars of one standard deviation from the mean are depicted by these graphs. Repeated simulations would allow for the calculation of standard errors and greater confidence when interpreting results (refer to Suggestions for Further Work). Such trends would be expected in these non-steady-state snapshots as a reduction in αRL should allow the dendrimer structure to migrate through the bilayer more quickly. This is because the dendrimer heads are more strongly attracted to receptor beads on both sides of the membrane, stretching the dendrimer across the internal bilayer structure as encapsulation was not observed. Furthermore, the increasing attraction between the receptor heads and outer dendrimer beads led to a wider spread of dendrimer on the bilayer surfaces consistent with similar simulations without the receptor-simulating lipid inclusion.2 These expectations were further supported by penetration data. Penetration similarly appeared to increase with with a decrease in the soft-core interaction parameter αRL: Figure 21: Penetration of the dendrimer into the bilayer after approximately 105 timesteps. Again, due to variability and incompleteness of data, further simulation is required in order to test whether this observed and expected trend is statistically significant. This statistical insignificance of the data is further demonstrated by Figure 22 using confidence bands calculated to bound 95% of least-squares linear regressions that could be expected to be generated given the sampling methodology (refer to Appendix 6). Note that a linearly-increasing relationship could be plotted between αRL=-15 and αRL=15 within these confidence bands and that linearity is not expected necessarily. 0 0.05 0.1 0.15 0.2 0.25 -20 -10 0 10 20 Penetration(fp) Soft-core Repulsion (αRL)
  • 23. 23 Figure 22: Penetration data bounded by linear regression confidence bands. In all simulations the dendrimers were able to penetrate the bilayer to varying degrees. In these simulations, the penetration was possible due to the hydrophobicity of the internal dendrimer beads (yellow) combined with the observation that the parameters defining the interaction between the hydrophobic lipid tails (cyan) and both types of dendrimer beads (red, yellow) (α=28) were less repulsive than the parameters defining the interactions between the hydrophobic lipid tails (cyan) and the lipid headgroup beads (green, blue, orange, purple) (α=40). As a result, the dendrimer did not simply adhere to the top surface of the bilayer. Instead the internal dendrimer beads bridged the structure as they were only weakly repelled by the internal hydrophobic section of the bilayer. The dendrimer head beads aggregated near the receptor beads on both sides of the bilayer due to the introduced receptor-ligand interaction. The structures generated were consistent with Yan and Yu’s findings that attraction between lipid tails and inner-dendrimer beads has a large effect on penetrability.2 When the magnitude of the attractive force was too great (αRL=-30), the self-assembling bilayer structure deteriorated (refer to Further Discussion). Figure 23: Deterioration of bilayer structure with αRL=-30 With αRL<-25, the attractive force between receptors and dendrimer heads was greater than the repulsive forces keeping like-beads apart from one another. This had the effect of creating physically unrealistic and unstable regions of high bead density. The chaos of the system was further
  • 24. 24 compounded by these unstable regions crossing the boundaries of the simulation environment and transporting to the other side of the box. Subsequent simulations used values of αRL equal to and greater than -25. 5.3.2 Further Attempts to Generate Clathrin-mediated Endocytic Morphology In accordance with Yan and Yu’s previous findings, to inhibit the observed bridging the repulsive parameter governing the interaction between the internal dendrimer beads and the lipid tails was increased from αTU=28 to αTU=40 (refer to Appendix 1: Table 8).2 Various values of αRL were trialled from a repulsive αRL=15 through to an attractive αRL=-25. After longer simulations, 245 timesteps (approximately 36 μs), the following morphologies were observed: Figure 24: Penetration profiles of simulations. For the purposes of clear illustration of the counterions and part of the bilayer structures have been omitted. Very different structures were observed after this increase in αTU. For αRL=15 the repulsion was so great that the dendrimer did not even adhere to the surface of the bilayer. Hence, its profile is not shown. However, in the absence of direct interaction with the bilayer, the radius of gyration of the dendrimer was reduced significantly to 2.85 nm. Some axial stretching (As≈ 0.06) was caused by the migration of counterions which accumulated in the volume separating the dendrimer and bilayer. Nevertheless, this observation of significant shrinkage was welcome due to better agreement with empirical measurements and previously discussed concerns (refer to Figure 17). For αRL=0, the dendrimer adhered to the surface of the bilayer. Although not made clear by the cross-section exhibited in Figure 24, random forces permitted the jumping of some surface dendrimer beads, with internal dendrimer beads attached, across the bilayer as was even more strongly observed for the case where the receptor-ligand interaction was modelled as attractive with αRL=-15. Note for αRL=-15 this attractive force was large enough to cause lipids to migrate onto the dendrimer. For αRL=-25 this migration was pronounced. Additionally, the strength of the ligand-receptor attraction was much greater relative to the random forces and dendrimer beads could no longer jump across the membrane. Consequently, a morphology resembling endocytosis evolved. It was also observed that the ligand-simulating dendrimer beads became more locally concentrated as αRL decreased (refer to Further Discussion). αRL=0 αRL=-15 αRL=-25
  • 25. 25 Figure 25: Radii of gyration of simulated dendrimer-bilayer systems after approximately 245 timesteps. As previously discussed, more simulations are required to make better supported assertions. The gyration radius was smallest for the dendrimer that did not bind to the bilayer (αRL=15). It increased with decreasing αRL as per the previous results generated (refer to Section 5.3.1) until it decreased with the onset of encapsulation for αRL=-25 (refer to Figure 25). Figure 26: Penetration of the dendrimer into the bilayer after approximately 245 timesteps. Penetrability statistics were consistent with the profiles generated as well as the negative correlation described in the analysis of earlier simulations (refer to Figure 26). Penetration was achieved by the lipids moving to encapsulate the dendrimer for αRL=-25 by direct penetration for αRL=0 and likely by a combination of both mechanisms for αRL=-15. Figure 27: Asphericity of simulated dendrimer-bilayer systems after approximately 245 timesteps. 2 2.5 3 3.5 4 -30 -20 -10 0 10 20RadiusofGyration Rg(nm) Soft-core Repulsion (αRL) 0 0.05 0.1 0.15 0.2 0.25 0.3 -30 -20 -10 0 10 20 Penetration(fp) Soft-core Repulsion (αRL) 0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 -30 -20 -10 0 10 20 Asphericity(As) Soft-core Repulsion (αRL)
  • 26. 26 As discussed, the dendrimer that did not bind to the bilayer was axially stretched by the counterions. For αRL=0 most of the dendrimer was bound to the surface of the bilayer in a comparatively spherical shape. The asphericity increased for αRL=-15 as the dendrimer structure opened. Surprisingly, the structure most closely resembling encapsulation developed the most aspherical form. The large asphericity of the structure generated for αRL=-25 in Figure 24 can be explained by its morphological development depicted by Figure 27: Figure 28: Morphological progression for αRL=-25. Due to the strong attraction between the receptors and the ligand-simulating dendrimer beads the dendrimer initially flattened out over the surface of the bilayer. Subsequently, receptors migrated on top of the dendrimer. The flat shape of the dendrimer was maintained and after approximately 27 µs the dendrimer broke through to the other side of the bilayer due to a combination of random forces, receptor-ligand attraction and increased pressure exerted by the growing number of repulsive lipid tails on top of the dendrimer. Subsequently, more receptor beads migrated to cover this breach and structures resembling encapsulation were observed. It is believed that if the simulation were able to continue a steady-state would be reached with a more spherical dendrimer structure evolving and an approximately equal number of dendrimer beads concentrated on either side of the midline of the bilayer (fp≈0.5) (refer to Suggestions for Further Work). However, continued simulation was not possible due to computing constraints. The relationship between these observed morphologies to CME and other simulations will be discussed in Section 6. 0µs 9µs 18µs 27µs 36µs 45µs
  • 27. 27 A phase table describing the morphologies for different soft-core repulsion parameters was generated: Table 3: Phase table describing different morphologies as a result of changing soft-core repulsion parameters. αRL αTU -25 -15 0 15 28 - Strong Direct Penetration Direct Penetration Direct Penetration 40 Endocytosis Direct Penetration With Pronounced Receptor Migration Adhesion with Direct Penetration No Adhesion Scope exists to further develop this table in conjunction with improving an understanding for how physical properties relate to the soft-repulsion parameters (refer to Suggestions Further Work).
  • 28. 28 6.0 Further Discussion 6.1 Known Shortcomings of the Model Given the simplicity of the simulated membrane a true representation of the clathrin-mediated endocytosis was unlikely to have been simulated. Figure 29: Schematic illustrating more representative membrane complexity.43 This experiment used a tensionless membrane. In reality it has been shown that surface tension varies and that locally, negative surface tension can be generated by the cytoskeleton, dynamin or the actin filaments.39 Consequently, internalisation via invagination and cytotoxicity associated with hole generation (refer to Figure 30) were unlikely to be well-modelled by these simulations. Figure 30: Empirically observed hole inducing effects of cationic PAMAM dendrimers on tense lipid bilayer using atomic force microscopy; 44 see also 1, 36-38, 45, 46. Other simulations have regulated the tension of the bilayer. In order to maintain a constant surface tension, frequent transformations (refer to Equation 19) to the dimensions of the simulation box can be performed whilst keeping the simulation volume constant.2 (19)
  • 29. 29 Alternatively, the N-varied DPD method maintains a constant membrane tension by adding and removing water and lipids from the edges of the simulation environment as required.39 Additionally, the receptor beads were too simple to form complex clathrin cages to encapsulate the dendrimer: Figure 31: A clathrin cage with a single triskelion (composed of three interlocking clathrin proteins) highlighted.47 In biological fluids, nanoparticles may interact with other circulated proteins and not with cellular membranes. Hence, simulating in a pure solvent environment may not be representative of reality even on such small length scales.7 The 1:3 ratio of lipid types was chosen hastily due to time pressures. Other studies have incorporated 1:1 ratio in accordance with current knowledge of mammalian membranes (refer to Suggestions for Further Work).13, 48 Also, the uniform distribution of receptors and ligand-simulating dendrimer beads on the surfaces of the membrane and dendrimer was not representative of a real system. It was assumed that the surface amine functional groups could substitute for ligands yielding this uniform distribution. Chemical synthesis of dendrimers is not able to achieve such uniformity. Synthesis of ligand coated PAMAM results in a distribution of the number of attached ligands. The number of attached ligands is much fewer in number than the number of surface amine groups.49 Simulating the effects of ligand distributions on cellular uptake has been studied but only using polygons as interacting particles (refer to Suggestions for Further Work).12 Consistent with Yan and Yu’s modelling, a number of counterions were included in the simulation and behaved according to the electrostatics detailed in Section 2.1 However, this implementation lacks refinement, further compounding potential sources of inaccuracy. Similar concerns stemming from the reuse of other parameters such as spring constants governing restorative forces exist. Finally, the energies associated with receptor-ligand adhesion are largely unknown.50 αRL is related to these unspecified energies. In order to observe endocytosis, αRL had to be made attractive enough to induce locally dense regions populated by receptor and ligand head. This modification did not disturb the thermostat; however, this clumping of beads may signify further deviation of the model from reality. Yang and Ma were incorporated a classical Lennard–Jones potential to govern the receptor-ligand interaction (refer to Figure 7) in order to prevent this issue.29
  • 30. 30 6.2 Relationship of CME to Simulations Using this receptor model, encapsulation of the dendrimer was able to be simulated following changes to the soft-core repulsion parameters. This encapsulation occurred via morphologies varying from descriptions of CME and the hypothesis was consequently rejected. In the absence of invaginations and actin-induced regions of negative surface tension, internalisation of the vesicle was not observed via the ‘coated pit’ morphology depicted in Figure 5.51 Encapsulation was primarily observed to occur by the migration of receptors to the adhesive front.50 Consequently, encapsulation appeared to occur by a mechanism more closely resembling phagocytosis (refer to Figure 4). This mechanism is associated with larger length scales than that which was simulated.25 Further investigation is required to determine whether this model can be used for the purpose of simulating phagocytosis or revised to produce simulations that mimic CME morphological progression. Other simulations have been able to achieve similar structures resembling endocytosis. Macroscale physics have been incorporated into these mesoscale models in the form of Canham-Helfrich theory. Canham-Helfrich theory introduces membrane bending energies so as to limit direct penetration.39 It was possible to achieve this outcome without the need for this added complexity by balancing soft repulsion parameters. 6.3 The Notable Shortage of Empirical Data Experimentalists have not been able to accurately observe many proposed mechanisms for endocytosis due to technological limitations at these lengthscales and timescales (refer to Figure 32).52, 53 There is currently a pressing need to be able to effectively differentiate between pathways and perform quantitative analysis of transport mechanisms.54 Figure 32:Currently there is little overlap between what can experimentally be observed and what can be simulated when studying endocytosis.25 As mentioned previously, in order for the new model to be of value, it must reproduce solubilities, liquid compressibilities and the shape of interfaces on the length scale of the course-grained molecules.4 Such comparisons are currently limited by a deficiency in empirical data surrounding endocytic processes. As a consequence, simulations do little more than attempt to generate hypothesised morphological progression. Many attempts to incorporate the modest available data
  • 31. 31 into the DPD model have been made or discussed (e.g. isolated dendrimer size, lipid diffusion, hole generation, hole size and lipid density). Due to the complexity of the system, many of the soft-core repulsion parameters use in this simulation were estimated by qualitative arguments coupled with iterative refinement or quoted from other studies which use similar methods.1, 17, 34 Some interaction parameters are able to be generated from Flory-Huggins theory and empirical testing.19 Calculating a complete set of Flory-Huggins 𝜒 parameters for complex systems, including those with multiple nanoscale species with Janus properties would be extremely difficult given the number of interaction pairs. Furthermore, energies associated with receptor-ligand binding, encapsulated gyration radii, internalisation velocities are all currently unknown. If the key issues concerning this model are addressed by reformulation, its future acceptance or rejection should be determined based on agreement with available endocytic data and practical value however blind (refer to Suggestions for Further Work).
  • 32. 32 7.0 Conclusions Incorporation of a receptor-simulating lipid into the bilayer and subsequent tuning of soft-core repulsion parameters generated endocytic morphology. It was found that self-assembly of the bilayer structure was not compromised when converting a quarter of the simulated lipid molecules into a receptor-simulating lipid. These receptor-simulating lipids extended on average 0.4 nm further into the solvent than the DPPC simulating lipids. The lipid density of the bilayer increased by approximately 13% when compared with Yan and Yu’s previous simulations.2 The stability of the DPD model, on comparing with Yan and Yu’s previous simulations, was found to be compromised by changes made to the interaction parameters governing the generation of the G5 PAMAM dendrimer in solvent, causing it to be 40% more open than empirically observed (Rg=3.4 nm > Rg=2.4 nm). These changes were partially rectified for the simulations by reverting the interaction parameters to their historic values; however, these changes could not impose the required structural changes before interaction with the bilayer, increasing penetration. Despite these issues, (resulting in enhanced direct penetration,) further refinement of the soft-core repulsion parameters revealed that endocytosis could be modelled. Statistical analysis of changes to the soft-core interaction parameter governing the behaviour of proximal dendrimer surface beads and receptor beads suggested that penetrability could be increased if this repulsive force was instead modelled as attractive. The magnitude of this attraction was limited such that for αRL>-25 the bilayer structure became unstable due to the resultant creation of hyper-dense regions within the simulation box. The observed structures of dendrimers generated by these changes were open and distributed throughout the structure of the bilayer, again suggesting entry by direct penetration. Encapsulation of the dendrimer was promoted after subsequently increasing the repulsive force between the internal dendrimer beads and the lipid tails of the bilayer. From these results it was concluded that the binding energy between receptors and ligands as well as the hydrophobic interaction between the bilayer lipid tails and internal structure of the interacting particle has a large effect on binding and the type of penetration. Morphological changes in the simulation in which a form of endocytosis was observed may provide some insight as to how such encapsulation progresses. However, due to the discussed limitations of the model the morphology more closely matched a phagocytic mechanism. On the simulated length scale such a result was unexpected indicating that further refinement of the model may be needed (refer to Suggestions for Further Work) if CME, as it is currently understood, is to be simulated and the presently unobservable dynamics conjectured. As discussed, due to computational constraints, many of the results of the simulations run in this study were largely preliminary. Continued effort in this area of research will have future implications for drug development. When models can be used with confidence, molecular architectures will be able to be tested without costly synthesis. In order to generate robust models and build this confidence, continued interest in formulation and modification is required to better describe the intricacies of the very complex yet fundamental process of endocytosis.
  • 33. 33 8.0 Suggestions for Further Work Future experiments require access to appropriate computing resources similar to the cluster systems housed at VLSCI. 8.1 Simple Immediate Revisions to the Current DPD Model As uncovered by the results generated, revisions to enhance the representativeness of the DPD model need to be made:  The assembly of the bilayer structure needs to resimulated with the ‘target lipid density’ set to 0.64 nm2 .  The prevalence of receptor-simulating lipids needs to be increased to approximately 50% in accordance with current knowledge of mammalian membranes.13, 48  The protrusion of these receptor lipids into the solvent may also need to be revised in light of information from empirical studies of bilayer structure.  Reversion to the original parameters governing the dendrimer-water system need to be readopted and simulated so that on introducing the dendrimer to the bilayer simulation box the structure of the dendrimer agrees with empirical measurements.  The introduction of an additional type of ligand bead distributed on the surface of the dendrimer in accordance with Mullen and Banaszak Holl’s findings needs to be incorporated to remove the erroneous assumption that surface amine functional groups can substitute for ligand complexes.49  To prevent the dense clumping of ligand and receptor beads, in accordance with Yang and Ma, this attractive pairwise interaction should be governed by a classical Lennard–Jones potential.29  The diffusion constant for membrane receptors in a bilayer is estimated to be approximately 104 nm2 /s.50 If accurate tracking of these receptors can be performed, αRL can be tuned to achieve expected diffusion.  The concentration of counterions and spring coefficients may require revision following examination of relevant literature. 8.2 Extension of Results Subsequent to the introduction of these changes, extended simulation is required to see whether CME can be observed. Multiple simulations running with the same set of soft-core interaction parameters are required so that sampling techniques and data interpretation can be strengthened. Simulations should ideally be run until steady-states have been reached so as to gain more a complete understanding of morphological progression. From these results comprehensive phase diagrams could be generated. It may be that CME still cannot be simulated following these simple revisions. A new model incorporating a more complex bilayer structure may be required. The distribution, size and structure of the receptors may need to be revised so that they are able to form clathrin cage-like configurations (refer to Figure 31). Similarly, strategies to control the surface tension of the bilayer on macro and local scales may need to be incorporated to simulate the roles of actin and dynamin in promoting invagination and vesicle release (refer to Figure 4).
  • 34. 34 If CME is not observed following the simple revisions this model could still be extended to simulate a larger interacting particle instead of G5 PAMAM. It is known that phagocytosis occurs on larger length scales. Subsequent to further evaluation as to whether the endocytic morphology generated by this study resembles phagocytosis, it would be interesting to observe whether similar encapsulation could be induced without the need to incorporate energies associated with membrane deformation into the model, as is often done when simulating on larger length scales.
  • 35. 35 9.0 References 1. Yan, L.-T. and X. Yu, Enhanced Permeability of Charged Dendrimers across Tense Lipid Bilayer Membranes. ACS Nano, 2009. 3(8): p. 2171-2176. 2. Yan, L.-T. and X. Yu, Charged Dendrimers on Lipid Bilayer Membranes: Insight through Dissipative Particle Dynamics Simulations. Macromolecules, 2009. 42(16): p. 6277-6283. 3. Min, Y., et al., The role of interparticle and external forces in nanoparticle assembly. Nat Mater, 2008. 7(7): p. 527-538. 4. Groot, R.D. and T.J. Madden, Dynamic Simulation of Diblock Copolymer Microphase Separation. The Journal of Chemical Physics, 1998. 108(20): p. 8713-8724. 5. Qian, H.-J., et al., Computer Simulation of Cyclic Block Copolymer Microphase Separation. Macromolecules, 2005. 38(4): p. 1395-1401. 6. Beddoes, C.M., C.P. Case, and W.H. Briscoe, Understanding Nanoparticle Cellular Entry: A Physicochemical Perspective. Advances in Colloid and Interface Science, 2015. 218(0): p. 48- 68. 7. Tian, F., Computer Simulation Studies on the Interactions Between Nanoparticles and Cell Membrane. Science China Chemistry, 2014. 57(12): p. 1662. 8. Yan, L.T. and X. Yu, Complexes Comprised of a Dendrimer and a Vesicle: Role of Vesicle Size and the Surface Tension of the Vesicle Membrane. Nanoscale, 2011. 3(9): p. 3812-8. 9. Abraham, J., Dissipative Particle Dynamics, in Encyclopedia of Microfluidics and Nanofluidics, D. Li, Editor. 2008, Springer US. p. 385-391. 10. Lee, H. and R. Larson, Multiscale Modeling of Dendrimers and Their Interactions with Bilayers and Polyelectrolytes. Molecules, 2009. 14(1): p. 423. 11. Li, Y., et al., Molecular Modeling of the Relationship between Nanoparticle Shape Anisotropy and Endocytosis Kinetics. Biomaterials, 2012. 33: p. 4965-4973. 12. Schubertova, V., F.J. Martinez-Veracoechea, and R. Vacha, Influence of Ligand Distribution on Uptake Efficiency. Soft Matter, 2015. 11(14): p. 2726-2730. 13. Vácha, R., F.J. Martinez-Veracoechea, and D. Frenkel, Intracellular Release of Endocytosed Nanoparticles Upon a Change of Ligand–Receptor Interaction. ACS Nano, 2012. 6(12): p. 10598-10605. 14. Kitchens, K.M., et al., Endocytosis and Interaction of Poly(amidoamine) Dendrimers with Caco-2 Cells. Pharmaceutical Research, 2007. 24(11): p. 2138-2145. 15. Seib, F.P., A.T. Jones, and R. Duncan, Comparison of the Endocytic Properties of Linear and Branched PEIs, and Cationic PAMAM Dendrimers in B16f10 Melanoma Cells. Journal of Controlled Release, 2007. 117(3): p. 291-300. 16. Hoogerbrugge, P.J. and J.M.V.A. Koelman, Simulating Microscopic Hydrodynamic Phenomena with Dissipative Particle Dynamics. EPL (Europhysics Letters), 1992. 19(3): p. 155. 17. Groot, R.D. and K.L. Rabone, Mesoscopic Simulation of Cell Membrane Damage, Morphology Change and Rupture by Nonionic Surfactants. Biophysical Journal, 2001. 81(2): p. 725-736. 18. Groot, R.D., Electrostatic Interactions in Dissipative Particle Dynamics—Simulation of Polyelectrolytes and Anionic Surfactants. The Journal of Chemical Physics, 2003. 118(24): p. 11265-11277. 19. Groot, R.D. and P.B. Warren, Dissipative Particle Dynamics: Bridging the Gap between Atomistic and Mesoscopic Simulation. The Journal of Chemical Physics, 1997. 107(11): p. 4423-4435. 20. Mintzer, M.A. and M.W. Grinstaff, Biomedical Applications of Dendrimers: A Tutorial. Chemical Society Reviews, 2011. 40(1): p. 173-190.
  • 36. 36 21. McNerny, D.Q., P.R. Leroueil, and J.R. Baker, Understanding Specific and Nonspecific Toxicities: A Requirement for the Development of Dendrimer-Based Pharmaceuticals. Wiley Interdisciplinary Reviews: Nanomedicine and Nanobiotechnology, 2010. 2(3): p. 249-259. 22. Shillcock, J.C. and R. Lipowsky, Equilibrium Structure and Lateral Stress Distribution of Amphiphilic Bilayers from Dissipative Particle Dynamics Simulations. Journal of Chemical Physics, 2002. 117(10): p. 5048. 23. Conner, S.D. and S.L. Schmid, Regulated Portals of Entry into the Cell. Nature, 2003. 422(6927): p. 37. 24. Canton, I. and G. Battaglia, Endocytosis at the Nanoscale. Chemical Society Reviews, 2012. 41(7): p. 2718-2739. 25. Qu, Z.G., et al., Advances in the Understanding of Nanomaterial-Biomembrane Interactions and their Mathematical and Numerical Modeling. Nanomedicine, 2013. 8(6): p. 995-1011. 26. Roth, T.F. and K.R. Porter, Yolk Protein Uptake in the Oocyte of the Mosquito Aedes Aegyti. L. The Journal Of Cell Biology, 1964. 20: p. 313-332. 27. Nel, A.E., et al., Understanding Biophysicochemical Interactions at the Nano–Bio Interface. Nature Materials, 2009. 8(7): p. 543-557. 28. Yue, T. and X. Zhang, Molecular Understanding of Receptor-Mediated Membrane Responses to Ligand-Coated Nanoparticles. Soft Matter, 2011. 7(19). 29. Yang, K. and Y.-q. Ma, Wrapping and Internalization of Nanoparticles by Lipid Bilayers: a Computer Simulation Study. Australian Journal of Chemistry, 2011. 64(7): p. 894. 30. Groot, R.D. and T.J. Madden, On the Role of Hydrodynamic Interactions in Block Copolymer Microphase Separation. Journal of Chemical Physics, 1999. 110(19): p. 9739. 31. Keaveny, E.E., et al., A Comparative Study between Dissipative Particle Dynamics and Molecular Dynamics for Simple- and Complex-Geometry Flows. The Journal of Chemical Physics, 2005. 123(10): p. 104107. 32. Moeendarbary, E., T.Y. Ng, and M. Zangeneh, Dissipative Particle Dynamics: Introduction, Methodology and Complex Fluid Applications - A Review. International Journal of Applied Mechanics, 2009. 01(04): p. 737-763. 33. Ingólfsson, H.I., et al., The Power of Coarse Graining in Biomolecular Simulations. WIREs: Computational Molecular Science, 2014. 4(3): p. 225. 34. Groot, R.D., Mesoscopic Simulation of Polymer−Surfactant Aggregation. Langmuir, 2000. 16(19): p. 7493. 35. Rodgers, J.M., et al., Understanding the Phase Behavior of Coarse-Grained Model Lipid Bilayers through Computational Calorimetry. The Journal of Physical Chemistry B, 2012. 116(5): p. 1551-1569. 36. Xie, L.-q., W.-d. Tian, and Y.-q. Ma, Computer Simulations of the Interactions of High- Generation Polyamidoamine Dendrimers with Electronegative Membranes. Soft Matter, 2013. 9(39): p. 9319-9325. 37. Kelly, C.V., et al., Poly(amidoamine) Dendrimers on Lipid Bilayers I: Free Energy and Conformation of Binding. The Journal of Physical Chemistry B, 2008. 112(31): p. 9337-9345. 38. Kelly, C.V., et al., Poly(amidoamine) Dendrimers on Lipid Bilayers II: Effects of Bilayer Phase and Dendrimer Termination. The Journal of Physical Chemistry B, 2008. 112(31): p. 9346- 9353. 39. Guo, R., J. Mao, and L.-T. Yan, Unique Dynamical Approach of Fully Wrapping Dendrimer-like Soft Nanoparticles by Lipid Bilayer Membrane. ACS Nano, 2013. 7(12): p. 10646-10653. 40. Deserno, M., Elastic Deformation of a Fluid Membrane upon Colloid Binding. Physical Review E, 2004. 69(3): p. 031903. 41. Ding, H.-m., W.-d. Tian, and Y.-q. Ma, Designing Nanoparticle Translocation through Membranes by Computer Simulations. ACS Nano, 2012. 6(2): p. 1230-1238. 42. Choi, Y., et al., DNA-Directed Synthesis of Generation 7 and 5 PAMAM Dendrimer Nanoclusters. Nano Letters, 2004. 4(3): p. 391-397.
  • 37. 37 43. Venturoli, M., et al., Mesoscopic Models of Biological Membranes. Physics Reports, 2006. 437(1–2): p. 1-54. 44. Leroueil, P.R., et al., Wide varieties of cationic nanoparticles induce defects in supported lipid bilayers. Nano Letters, 2008. 8(2): p. 420-424. 45. Perumal, O.P., et al., The Effect of Surface Functionality on Cellular Trafficking of Dendrimers. Biomaterials, 2008. 29(24-25): p. 3469-76. 46. Mecke, A., et al., Lipid Bilayer Disruption by Polycationic Polymers: The Roles of Size and Chemical Functional Group. Langmuir: The ACS Journal Of Surfaces And Colloids, 2005. 21(23): p. 10348-10354. 47. Goodsell, D. and G. Johnson, Molecule of the Month: Clathrin. RCSB PDB, 2007. 48. Alberts, B., et al., Molecular Biology of the Cell. 2015, New York: N.Y. Garland Science, Taylor and Francis Group. 49. Mullen, D.G. and M.M. Banaszak Holl, Heterogeneous Ligand–Nanoparticle Distributions: A Major Obstacle to Scientific Understanding and Commercial Translation. Accounts of Chemical Research, 2011. 44(11): p. 1135-1145. 50. Gao, H., W. Shi, and L.B. Freund, Mechanics of Receptor-mediated Endocytosis. Proceedings of the National Academy of Sciences of the United States of America, 2005. 102(27): p. 9469- 9474. 51. Doherty, G.J. and H.T. McMahon, Mechanisms of Endocytosis. Annual Review of Biochemistry, 2009. 78(1): p. 857-902. 52. Leroueil, P.R., et al., Nanoparticle Interaction with Biological Membranes: Does Nanotechnology Present a Janus Face? Accounts of Chemical Research, 2007. 40(5): p. 335- 342. 53. Wang, T., et al., Cellular Uptake of Nanoparticles by Membrane Penetration: A Study Combining Confocal Microscopy with FTIR Spectroelectrochemistry. ACS Nano, 2012. 6(2): p. 1251-1259. 54. Akinc, A. and G. Battaglia, Exploiting Endocytosis for Nanomedicines. Cold Spring Harbor Perspectives in Biology, 2013. 5(11).
  • 38. 38 10.0 Appendices Appendix 1: Simulation Soft-core Repulsion Parameters Table 4: Historic soft-core repulsion parameters.1 Lipid Head (+1) Lipid Head (-1) Lipid Head (0) Lipid Tail (0) Solvent Bead (0) Surface Dendrimer Bead (+1) Counterions (-1) Internal Dendrimer Bead (0) Lipid Head (+1) 25 17 22 40 22 15 20 28 Lipid Head (-1) 17 25 22 40 22 15 20 28 Lipid Head (0) 22 22 25 40 25 15 22 28 Lipid Tail (0) 40 40 40 25 80 28 80 28 Solvent Bead (0) 22 22 25 80 25 20 20 80 Surface Dendrimer Bead (+1) 15 15 15 28 20 25 20 28 Counterions (-1) 20 20 22 80 20 20 25 28 Internal Dendrimer Bead (0) 28 28 28 28 80 28 28 25 Table 5: Soft-core repulsion parameters for bilayer self-assembly. Lipid Head (+1) Lipid Head (-1) Lipid Head (0) Lipid Tail (0) Solvent Bead (0) Receptor Head (0) Lipid Head (+1) 25 17 22 40 22 25 Lipid Head (-1) 17 25 22 40 22 25 Lipid Head (0) 22 22 25 40 25 25 Lipid Tail (0) 40 40 40 25 80 40 Solvent Bead (0) 22 22 25 80 25 15 Receptor Head (0) 25 25 25 40 15 25
  • 39. 39 Table 6: Soft-core repulsion parameters for interacting bilayer-dendrimer system. Starred parameters were introduced not required by Yan and Yu’s receptorless system. These parameter were carefully chosen to model the receptor head similarly to other lipid heads in an attempt to reproduce Yan and Yu’s results.1, 2 Lipid Head (+1) Lipid Head (-1) Lipid Head (0) Lipid Tail (0) Solvent Bead (0) Receptor Head (0) Surface Dendrimer Bead (+1) Counterions (-1) Internal Dendrimer Bead (0) Lipid Head (+1) 25 17 22 40 22 25* 15 20 28 Lipid Head (-1) 17 25 22 40 22 25* 15 20 28 Lipid Head (0) 22 22 25 40 25 25* 15 22 28 Lipid Tail (0) 40 40 40 25 80 40* 28 80 28 Solvent Bead (0) 22 22 25 80 25 25* 15 20 80 Receptor Head (0) 25* 25* 25* 40* 25* 25* 15* 12* 28* Surface Dendrimer Bead (+1) 15 15 15 28 15 15* 25 20 28 Counterions (-1) 20 20 22 80 20 12* 20 25 28 Internal Dendrimer Bead (0) 28 28 28 28 80 28* 28 28 25 Table 7: Soft-core repulsion parameters for interacting bilayer and dendrimer. Lipid Head (+1) Lipid Head (-1) Lipid Head (0) Lipid Tail (0) Solvent Bead (0) Receptor Head (0) Surface Dendrimer Bead (+1) Counterions (-1) Internal Dendrimer Bead (0) Lipid Head (+1) 25 17 22 40 22 25 15 20 28 Lipid Head (-1) 17 25 22 40 22 25 15 20 28 Lipid Head (0) 22 22 25 40 25 25 15 22 28 Lipid Tail (0) 40 40 40 25 80 40 28 80 28 Solvent Bead (0) 22 22 25 80 25 25 15 20 80 Receptor Head (0) 25 25 25 40 25 25 αRL 12 28 Surface Dendrimer Bead (+1) 15 15 15 28 15 αRL 25 20 28 Counterions (-1) 20 20 22 80 20 12 20 25 28 Internal Dendrimer Bead (0) 28 28 28 28 80 28 28 28 25
  • 40. 40 Table 8: Revised soft-core repulsion parameters for interacting bilayer and dendrimer. Lipid Head (+1) Lipid Head (-1) Lipid Head (0) Lipid Tail (0) Solvent Bead (0) Receptor Head (0) Surface Dendrimer Bead (+1) Counterions (-1) Internal Dendrimer Bead (0) Lipid Head (+1) 25 17 22 40 22 25 15 20 28 Lipid Head (-1) 17 25 22 40 22 25 15 20 28 Lipid Head (0) 22 22 25 40 25 25 15 22 28 Lipid Tail (0) 40 40 40 25 80 40 28 80 αTU Solvent Bead (0) 22 22 25 80 25 25 15 20 80 Receptor Head (0) 25 25 25 40 25 25 αRL 12 28 Surface Dendrimer Bead (+1) 15 15 15 28 15 αRL 25 20 28 Counterions (-1) 20 20 22 80 20 12 20 25 28 Internal Dendrimer Bead (0) 28 28 28 αTU 80 28 28 28 25
  • 41. 41 Appendix 2: MATLAB Code for Bilayer Bead Density Profile % Bilayer Bead Density by Displacement from Average z-coordinate %create matrix with data from binary output file [file,path]=uigetfile('SpaGe1.bin','Open the file');%to save N(r)_r data if file fid=fopen([path,file],'r'); a=fread(fid,[4 inf],'float32'); end %calculate zavg ztot=0; for i=1:numel(a)/4; z = a(3,i); ztot = ztot+z; end zavg= ztot/(numel(a)/4); %avg displacement from zavg density= zeros(2,numel(a)/4); for i=1:numel(a)/4; density(1,i)=(a(3,i)-zavg); density(2,i)=a(4,i); end %partition 0.05nm x=linspace(-19.975, 20.025, 800); for j=1:numel(a)/4 for i=1:numel(x) if density(1,j)<= (x(i)+0.025) && density(1,j)>(x(i)-0.025) density(1,j)=x(i); end end end %displacement histogram for relevant bilayer beads (types 1, 2, 3, 4, 6) y=zeros(6,numel(x)); for i=1:6 for j=1:numel(a)/4 for m=1:numel(x) if density(2,j)== i && density(1,j)== x(m) y(i,m)=y(i,m)+1/88.2; %volume of xy-sliver 88.2nm end end end end y=[y(1,:);y(2,:);y(3,:);y(4,:);y(6,:)];
  • 42. 42 y=y'; %histogram smoothing smooth = sgolayfilt(y,3,29); subplot(2,1,1) plot(x, smooth,'linewidth', 2) axis([-5 5 0 3.5]) title('Displacement Distribution of Bilayer Beads from Central z- coordinate') xlabel('Displacement from Bilayer Centre (nm)'),ylabel('Number Density (beads/nm^3)'); legend('Lipid Head (+1)','Lipid Head (-1)','Lipid Head (0)','Lipid Tail (0)','Receptor Head (0)') subplot(2,1,2) plot(x,smooth, 'linewidth', 2) axis([0 5 0 1]) title('Magnified Displacement Distribution of Bilayer Beads from Central z- coordinate') xlabel('Displacement from Bilayer Centre (nm)'),ylabel('Number Density (beads/nm^3)'); legend('Lipid Head (+1)','Lipid Head (-1)','Lipid Head (0)','Lipid Tail (0)','Receptor Head (0)')
  • 43. 43 Appendix 3: Asphericity Asphericity of the dendrimer can be calculated from the following equations: (20) Where I1 and I2 are the first two invariants: (21) (22) 𝜆1, 𝜆2, 𝜆3 are the eigenvalues of the gyration tensor: 𝑮 𝑻 = 𝟏 𝑵 ∑ (𝒔𝒊⨂𝒔𝒊)𝑵 𝒊=𝟎 (23) 𝒔𝒊 = 𝒓𝒊 − 𝒓 𝑪𝑴 (24) % Asphericity MATLAB %create matrix with data from binary output file [file,path]=uigetfile('SpaGe1.bin','Open the file');%to save N(r)_r data if file fid=fopen([path,file],'r'); a=fread(fid,[4 inf],'float32'); end %remove counterions b=a; count=0; for i=1:numel(a)/4; if a(4,i)== 8 b(:, i-count) = []; count=count+1; else b(:,i-count)= a(:,i); end end %calculate centre of mass (all beads have the same mass) xtot=0; ytot=0; ztot=0; for i=1:numel(b)/4; x = b(1,i); xtot = xtot+x; end xavg= xtot/(numel(b)/4); for i=1:numel(b)/4; y = b(2,i);
  • 44. 44 ytot = ytot+y; end yavg= ytot/(numel(b)/4); for i=1:numel(b)/4; z = b(3,i); ztot = ztot+z; end zavg= ztot/(numel(b)/4); rcm=[xavg;yavg;zavg] %gyration tensor dispoutertot=zeros(3); disp = [0;0;0]; for i=1:numel(b)/4; for j=1:3 disp(j,1)=b(j,i)-rcm(j,1); end dispouter=disp*disp'; dispoutertot=dispoutertot+dispouter; end Gt=dispoutertot/(numel(b)/4) %calculate asphericity eigvals=eig(Gt); I1= sum(eigvals); I2= eigvals(1)*eigvals(2)+eigvals(2)*eigvals(3)+eigvals(3)*eigvals(1); As=1-(3*(I2)/(I1)^2)
  • 45. 45 Appendix 4: Radius of Gyration The radius of gyration is determined by the following equation: (25) Where the summation runs over all N beads in the dendrimer at positions ri. rcm denotes the center of mass of the dendrimer. % Gyration Radius MATLAB %create matrix with data from binary output file [file,path]=uigetfile('SpaGe1.bin','Open the file');%to save N(r)_r data if file fid=fopen([path,file],'r'); a=fread(fid,[4 inf],'float32'); end %remove counterions b=a; count=0; for i=1:numel(a)/4; if a(4,i)== 8 b(:, i-count) = []; count=count+1; else b(:,i-count)= a(:,i); end end %calculate centre of mass (all beads have the same mass) xtot=0; ytot=0; ztot=0; for i=1:numel(b)/4; x = b(1,i); xtot = xtot+x; end xavg= xtot/(numel(b)/4); for i=1:numel(b)/4; y = b(2,i); ytot = ytot+y; end yavg= ytot/(numel(b)/4); for i=1:numel(b)/4; z = b(3,i); ztot = ztot+z; end zavg= ztot/(numel(b)/4); rcm=[xavg;yavg;zavg]
  • 46. 46 %squared displacement from centre of gravity squared dispsqutot=[0;0;0]; disp = [0;0;0]; for i=1:numel(b)/4; for j=1:3 disp(j,1)=b(j,i)-rcm(j,1); end dispsqu = disp.^2; dispsqutot=dispsqutot+dispsqu; end sum(dispsqutot)/((numel(b)/4)); Rg = sqrt(sum(dispsqutot)/((numel(b)/4)))
  • 47. 47 Appendix 5: Penetrability Penetrability was defined as the fraction of dendrimers that were below the average z-coordinate of all lipid beads enclosed by the maximum and minimum x,y coordinates of the dendrimer beads at the fringes of the dendrimer structure. % Penetrability MATLAB %create matrix with data from binary output file [file,path]=uigetfile('SpaGe1.bin','Open the file');%to save N(r)_r data if file fid=fopen([path,file],'r'); a=fread(fid,[4 inf],'float32'); end [file,path]=uigetfile('SpaGe2.bin','Open the file');%to save N(r)_r data if file fid=fopen([path,file],'r'); c=fread(fid,[4 inf],'float32'); end %remove counterions b=c; count=0; for i=1:numel(c)/4; if c(4,i)== 8 b(:, i-count) = []; count=count+1; else b(:,i-count)= c(:,i); end end %box coordinates to determine relevant bilayer beads for penetrability dimmax=max(b,[],2); dimmin=min(b,[],2); dimensions=[dimmin(1,1),dimmin(2,1); dimmax(1,1),dimmax(2,1)]; %remove irrelevant bilayer beads d=a; count=0; for i=1:numel(a)/4; if a(1,i)<=dimensions(1,1)|a(2,i)<=dimensions(1,2)|a(1,i)>=dimensions(2,1)|a(2 ,i)>=dimensions(2,2) d(:, i-count) = []; count=count+1; else d(:,i-count)= a(:,i); end end %calculate average z coordinate of remaining bilayer (all beads have the same mass) ztot=0; for i=1:numel(d)/4;
  • 48. 48 z = (3,i); ztot = ztot+z; end zavg= ztot/(numel(d)/4) test=ones(1,numel(b)/4); %test if dendrimer below middle of bilayer for i=1:numel(b)/4 if b(3,i)<=zavg test(1,i)=1; else test(1,i)=0; end end %fraction below middle of bilayer fp=sum(test)/numel(test)
  • 49. 49 Appendix 6: Confidence Bands Data file: aRL fp 15 0.1109 15 0.1327 15 0.1426 15 0.1485 15 0.1089 0 0.1386 0 0.1762 0 0.1762 0 0.1802 0 0.2 -15 0.2693 -15 0.2178 -15 0.2495 -15 0.1881 -15 0.1861 R Code: data = read.delim(file.choose(),header=T) X=data$aRL Y=data$fp plot(X, Y, xlab="aRL", ylab="Penetrability fp", xlim=c(-20, 20), ylim=c(0, 0.4)) ml=lm(Y~X) %abline(ml) summary(ml) confint(ml,level=0.95) bands=predict(ml,newdata=data.frame(X),interval=c("prediction" ),level=0.95) lines(X, bands[, "lwr"], lty = "dotted") lines(X, bands[, "upr"], lty = "dotted")
  • 50. 50 Appendix 7: Intermediate Data Table 9: Rg data for αTU=28 αRLTimestep 98000 99000 100000 101000 102000 μ σ 15 3.8E+00 3.7E+00 3.7E+00 3.7E+00 3.7E+00 3.7E+00 2.5E-02 0 3.9E+00 3.8E+00 3.9E+00 3.8E+00 3.9E+00 3.9E+00 5.2E-02 -15 4.0E+00 4.1E+00 4.0E+00 4.0E+00 4.0E+00 4.0E+00 3.1E-02 Table 10: As data for αTU=28 αRLTimestep 98000 99000 100000 101000 102000 μ σ 15 3.3E-02 1.9E-02 5.3E-03 1.1E-02 1.3E-02 1.6E-02 1.1E-02 0 3.8E-02 3.4E-02 3.2E-02 1.5E-02 2.7E-02 2.9E-02 9.1E-03 -15 2.4E-02 4.3E-02 3.1E-02 3.3E-02 5.6E-02 3.7E-02 1.3E-02 Table 11: fp data for αTU=28 αRLTimestep 98000 99000 100000 101000 102000 μ σ 15 1.1E-01 1.3E-01 1.4E-01 1.5E-01 1.1E-01 1.3E-01 1.8E-02 0 1.4E-01 1.8E-01 1.8E-01 1.8E-01 2.0E-01 1.7E-01 2.2E-02 -15 2.7E-01 2.2E-01 2.5E-01 1.9E-01 1.9E-01 2.2E-01 3.7E-02 Table 12: Rg data for αTU=40 αRLTimestep 238000 239000 240000 241000 242000 μ σ 0 3.2E+00 3.2E+00 3.2E+00 3.3E+00 3.2E+00 3.2E+00 4.2E-02 15 2.8E+00 2.8E+00 2.8E+00 2.9E+00 2.9E+00 2.9E+00 7.1E-02 -15 3.7E+00 3.7E+00 3.5E+00 3.6E+00 3.6E+00 3.6E+00 7.9E-02 -25 3.3E+00 3.3E+00 3.3E+00 3.3E+00 3.3E+00 3.3E+00 3.1E-02 Table 13: As data for αTU=40 αRLTimestep 238000 239000 240000 241000 242000 μ σ 0 3.2E-02 4.6E-02 3.6E-02 2.7E-02 2.6E-02 3.3E-02 8.5E-03 15 4.4E-02 3.8E-02 5.1E-02 1.0E-01 8.6E-02 6.4E-02 2.8E-02 -15 5.6E-02 7.6E-02 5.2E-02 3.5E-02 4.5E-02 5.3E-02 1.5E-02 -25 1.2E-01 1.3E-01 1.3E-01 1.2E-01 1.1E-01 1.2E-01 8.4E-03 Table 14: Rg data for αTU=40 αRLTimestep 238000 239000 240000 241000 242000 μ σ 0 1.1E-01 1.1E-01 1.1E-01 1.1E-01 8.5E-02 1.1E-01 1.2E-02 15 0.0E+00 0.0E+00 0.0E+00 0.0E+00 0.0E+00 0.0E+00 0.0E+00 -15 1.5E-01 1.7E-01 1.6E-01 1.3E-01 1.6E-01 1.5E-01 1.2E-02 -25 2.7E-01 2.6E-01 2.3E-01 2.6E-01 2.7E-01 2.6E-01 1.8E-02