Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2014 Feb 2.
Published in final edited form as: Nanomedicine (Lond). 2012 Dec 2;8(3):343–357. doi: 10.2217/nnm.12.124

in silico Vascular Modeling for Personalized Nanoparticle Delivery

Shaolie S Hossain 1,4,*, Yongjie Zhang 2, Xinghua Liang 2, Fazle Hussain 3, Mauro Ferrari 1, Thomas J R Hughes 4, Paolo Decuzzi 1,5,*
PMCID: PMC3909646  NIHMSID: NIHMS526842  PMID: 23199308

Abstract

Aims

To predict the deposition of nanoparticles in a patient-specific arterial tree as a function of the vascular architecture, flow conditions, receptor surface density, and nanoparticle properties.

Materials & methods

The patient-specific vascular geometry is reconstructed from CT Angiography images. The Isogeometric Analysis framework integrated with a special boundary condition for the firm wall adhesion of nanoparticles is implemented. A parallel plate flow chamber system is used to validate the computational model in vitro.

Results

Particle adhesion is dramatically affected by changes in patient-specific attributes, such as branching angle and receptor density. The adhesion pattern correlates well with the spatial and temporal distribution of the wall shear rates. For the case considered, the larger (2.0 μm) particles adhere ≈ 2 times more in the lower branches of the arterial tree, whereas the smaller (0.5 μm) particles deposit more in the upper branches.

Conclusion

Our computational framework in conjunction with patient specific attributes can be used to rationally select nanoparticle properties to personalize, thus optimize, therapeutic interventions.

Keywords: nanoparticles, rational design, mathematical modeling, vascular adhesion, personalized medicine, patient-specific, wall shear rate

Introduction

It is well accepted that the outcome and safety of a therapeutic intervention is often affected by patient-specific attributes, at the gene, cell and organ levels [1, 2]. It is also increasingly recognized that results from clinical trials do not necessarily apply to individual patients, not even to the patients that were directly enrolled in the trials [3]. Every individual is shaped differently and the distribution, metabolism and elimination of drugs as well as their biochemical effects on the target cells are influenced by the patient age, genetic background and anatomical features. Computational modeling and nanomedicine are playing a major role in supporting the development of personalized therapeutic approaches [4-8].

Computational modeling can capture the hierarchical complexity of biological systems and diseases over multiple scales – temporal and spatial – and include patient-specific information to personalize the outcome of the analysis. in silico modeling has been proven useful in orthopedic applications [9, 10], for the treatment of cardiovascular diseases [11-14] and cancer [15, 16], and in pharmacogenomic analysis [6, 17]. Segmental bone replacements can be optimally designed after a careful analysis of patient anatomical features and computation of the mechanical loads[9]. In cardiovascular diseases, the authentic geometry of the blood vessels and their mechanical properties have been incorporated in sophisticated computational tools to predict the distribution of wall shear stresses, risk of aneurysm rupture and to optimize the deployment of vascular stents [18-20]. Multi-physic models have been developed for predicting the response of tumors to molecular-based, radiation and thermal ablation therapies [15, 16].

Computational pharmacogenomics is used to predict in silico the efficacy, toxicity and possible resistance of drug molecules on different cell types [6, 16].

On the other hand, a plethora of nanoparticle-based delivery systems have been developed over the last two decades for enhancing the tissue-specific accumulation of therapeutic molecules and the contrast generated by imaging agents. Indeed, the biodistribution and bioavailability of both therapeutic and imaging agents are dramatically affected by the size, shape, surface properties and mechanical stiffness (the 4Ss) of their carriers: the nanoparticles. This has been firstly predicted using mathematical models and then supported by experimental evidence, in vitro and in vivo [21-23]. Different fabrication strategies have been proposed to finely tune the geometrical, mechanical and surface properties of the nanoparticles [24-27]. But how such four parameters (the 4Ss) can be rationally exploited in conjunction with patient-specific attributes to personalize and, thus optimize, therapeutic interventions is still largely unexplored.

Along this line, a patient-specific computational model is here presented for predicting the nanoparticle lodging within an authentic vascular network. Hexahedral solid NURBS (Non-Uniform Rational B-Splines) are used to accurately mesh the tridimensional architecture of an arterial tree, derived from the Computed Tomography (CT) scan of a patient [28]. The Finite Element Method (FEM), reformulated within the Isogeometric Analysis framework [29], is employed to solve for the fluid and particle transport in the authentic vasculature (see Materials and Methods section for further details on the numerical approach). Information at the micro and nanoscale are introduced in the computational model by developing a special boundary condition at the vessel wall, accounting for cell/nanoparticle adhesion [30]. The model directly integrates information from the macroscale (vessel geometry and permeability, blood flow condition) with data pertaining to the micro and nanoscale (particle geometry, receptor density and affinity), thus avoiding massive, computationally inefficient discretization over multiple spatial and temporal scales. A Newtonian rheological law is used for blood, which is a reasonable assumption for larger vessels [31, 32]. More complex laws, such as the Generalized Power Law [33] can be readily implemented. The proposed in silico model is first validated against in vitro parallel plate flow chamber experiments and then is applied for studying the effect of patient-specific attributes, such as the vascular geometry and receptor surface density, on the deposition of spherical particles with different sizes within an authentic arterial tree. Although we do not perform a theoretical analysis in this work because of the many complexities involved, we do carry out experimental validation of the particle adhesion model for a range of particle size and wall shear rates that defines a regime of applicability.

Materials & Methods

Reconstructing the patient-specific vascular geometry

The input CT Angiography imaging data are often of poor quality due to large motions of the heart, as it supplies blood to the circulatory system. This makes it difficult to construct analysis-suitable patient-specific coronary models. To circumvent this problem, the raw imaging data were passed through a preprocessing pipeline where the image quality is improved by enhancing the contrast, filtering noise, classifying, and segmenting regions of interest [28]. A small bifurcation portion of the coronary tree was considered here including the left coronary artery (LCA), the left anterior descending artery (LAD) and the left circumflex artery (LCX) (Figure 1). The surface model of this bifurcation structure was extracted from the processed imaging data, and the vessel path was obtained after skeletonizing the volume bounded by the local luminal surface using Voronoi and Delaunay diagrams. The generated path can also be edited according to simulation requirements, e.g., extending the included branch angles to study how geometry such as the bifurcation angle influences particle delivery processes in coronary arteries. A skeleton-based sweeping method [28] was then used to generate hexahedral control meshes by sweeping a templated quadrilateral mesh of a circle along the arterial path. A template for the bifurcation configuration was used to decompose the geometry into three mapped meshable patches using the extracted skeleton. Each patch can be meshed using one-to-one sweeping techniques. Some nodes in the control mesh lie on the surface, and some do not. We project nodes lying on the surface to the vascular surface. Finally, solid NURBS models were generated based on the constructed control meshes and they were employed in Isogeometric Analysis [34, 35] to simulate blood flow and particle delivery in the coronary arteries.

Figure 1. Reconstructing the patient-specific vascular geometry.

Figure 1

The image shows, from left to right, the isocontour of a human heart, path extraction and editing of a small bifurcation portion from the left coronary artery (LCA) and reconstruction of the geometry ready for Isogeometric Analysis. Also, a nanoparticle with its ligand molecules is shown interacting with the receptor molecules decorating the surface of the endothelial cells in the vasculature.

Governing equations for the fluid flow and particle transport

A continuum-based approach was adopted to simulate blood flow and particle transport within a patient-specific vascular network (Figure 1). Blood was modeled as an incompressible Newtonian fluid with a density (ρ) of 1060kg/m3 and a dynamic viscosity (μ) of 0.003 N-s/m2, and the governing equations were formulated accordingly. Indeed, for sufficiently larger vessels (macrocirculation), the corpuscular component of blood (red blood cells, white blood cells and platelets) can be neglected and a Newtonian model for blood provides sufficiently accurate results [31, 32]. The strong form of the continuity and momentum balance equations in the fluid domain Ω. with its boundary T divided into three non-overlapping parts, the inflow (Γin) and outflow (Γout) boundaries and the vascular wall (Γs), can be written as:

ρut+ρuu+pμ(2su)=finΩ×(0,T) (0)
u=0inΩ×(0,T) (0)
u=gonΓin×(0,T) (0)
u=0onΓs×(0,T) (1)
pn+μ(2su)n=0onΓout×(0,T) (2)
u(x,0)=u0(x)inΩ (3)

where x is a point in the spatial domain Ω and t is a point in the time domain [0,T]. In these equations, u (x;t) represents the fluid velocity vector, p(x;t) the pressure, f the external body force, n the unit outward normal to the surface, denoted Γ, and u0 (x;0) the initial velocity vector. An inflow velocity vector g(x;t) was specified at the inlet (Γin), a no-slip boundary condition (Eq.(1)) was prescribed at the rigid and impermeable vascular wall (Γs) and a stress free outflow boundary condition (Eq.(2)) was implemented at the branch outlets (Γout).

The mass transport of the particles was assumed to be governed by a scalar advection-diffusion-reaction equation. This is a reasonable assumption for sufficiently small particles navigating in large vessels [36, 37]. In the strong form, the transport problem can be stated as:

Ct+uC(KC)+σC=0inΩ×(0,T) (3)
C=C0onΓin×(0,T) (4)
(KC)n=0onΓout×(0,T) (5)
(KC)n+ΠC=0onΓs×(0,T) (6)
C(x,0)=0inΩ (7)

where C (x;t) is the particle concentration, K is the diffusivity tensor, and σ is the reaction coefficient. At the inlet (particle injection site) Γin, a Dirichlet boundary condition is prescribed (Eq. (4)) where C0 is the particle concentration given as:

C0(x,t)=1for0tti,xΓinC0(x,t)=0fort>ti,xΓin (8)

Here ti denotes the duration of particle injection. At the outflow Γout, a homogenous Neumann boundary condition was specified (Eq.(5)); and a Robin-type boundary condition (Eq.(6)), a combination of the Dirichlet and Neumann conditions, was prescribed at the rigid wall interface (Γs), where Π is defined in the sequel as the vascular deposition parameter.

Particle diffusivity was assumed to be isotropic and constant, and was determined from the Einstein-Stokes relation

D=kBT3μdp (9)

as a function of the particle size, where dp is the diameter of the particles. Here kBT represents the Boltzmann thermal energy. Then

K=DI (10)

where I is the 3×3 identity matrix. The velocity field u in Eq.(3) was obtained from the solution of the Navier–Stokes equations (Eqs.(0)-(3)) with the assumption that the flow physics affects the mass transport, not vice versa.

Boundary condition for particle adhesion to the vessel wall

The particle surface is assumed to be decorated with ligand molecules, uniformly distributed with a surface density ml, that can specifically interact with counter molecules (receptors) expressed on the vessel wall with a surface density mr (Figure 1). The molecular interaction between ligands and receptors is characterized by an affinity constant Ka0, at zero mechanical load. From [30], a probability of particle adhesion Pa can be derived as a function of its geometry (size and shape) and surface properties (ligand density, ligand type; surface electrostatic charge). The mathematical parameter Pa is defined as the probability of having at least one close ligand–receptor bond, and can be considered as a measurement of the strength of adhesion: the larger is Pa, the larger is the avidity and strength with which the particle firmly adheres to the wall. For a spherical particle of diameter dp, the probability of adhesion can be expressed as follows [30]:

PamrmlKa0πr02exp[1λKBT6FsμSmrdp24r02] (10)

where λ is a characteristic length of the ligand-receptor bond, generally of the order of 0.1 – 1 nm; kBT is the Boltzmann thermal energy (= 4.142×10-21 J); Fs is the coefficient of hydrodynamic drag force on the spherical particle; μS is the wall shear stress; and S is the wall shear rate. In Eq.(10), the parameter r0 is the radius of adhesion

r0=dp2[1(1Δdp/2)2] (10)

with Δ the separation distance between the particle and the substrate, at equilibrium (i.e., firm particle adhesion). Note that Δ (= 0.68 nm) is generally much smaller than the particle diameter dp [30]. An alternative representation of Eq.(10) is,

Paα1α2πr02exp[βμSα2] (11)

where α1=mlKa0 (=0.023) and α2=mr[1(1Δdp/2)2] (≈1.369×1010 #/m2 for dp = 0.5 μm and ≈ 3.400×109 #/m2 for dp = 2 μm) are two governing parameters, and β=λ6FskBT (=2.39×1011N-1) is a constant. In Eq.(11), α1, α2, the wall shear rate S and the particle diameter dp are the independent governing parameters. For convenience, λ has been fixed to be 0.1 nm.

The mass flux of particles (∂C/∂n) in the direction n normal to the wall can be related to the local increase in mass of particles adhering per unit surface ψ(x;t) as

(KC)n|s=DCn|s=ψt (12)

which simply derives from mass balance at the surface (accumulation of mass over time). Note that Eq.(12) is not a boundary condition yet, in that ψ(x;t) is an unknown function. To close the system, an additional relation needs to be defined between ψ(x;t) and C(x;t). The number of particles per unit surface ψ(x;t) is in turn related to the volume concentration of particles at the wall through the equation

DCn|s=ψt=ΠC|s=PaSdp2C|s (13)

Note that Sdp/2 is the fluid velocity in the center of a particle in proximity of the wall. The term Π = PaSdp/2 is called the vascular deposition parameter, in that the larger is Π, the larger is the number of particle adhering stably to the vessel walls under flow.

In the Supplementary Materials, the system of governing equations (0)-(13) is specialized to the case of a channel with a rectangular cross section simulating a parallel plate flow chamber, used for the in vitro validation of the proposed computational approach. Note that the model presented can be also integrated with other adhesion models [38] that could predict the number of adhering particles as a function of the local biophysical conditions (wall shear rate; surface density of receptors; binding affinity) and nanoparticle features (size, shape, surface properties such as ligand density, electrostatic charge).

Finite Element Solution Approach

Eqs.(1)-(11) were solved by applying Finite Element based Isogeometric Analysis [14, 34] that uses NURBS to describe the geometry as well as the solution space. It is a technology that constructs an exact geometry, which is preserved under mesh refinement allowing for higher–order (smoother) and higher continuity discretization of complex geometries, for example the arterial system. Quadratic NURBS was used for the spatial discretization. The spatial order of error is therefore of third order. A residual-based variational multiscale method [35] was implemented to solve the system of equations, utilizing a Newton-Raphson procedure with multi-stage predictor-corrector algorithm at each time step. The generalized–α method [39, 40] was used for time advancement, which is an implicit second-order time-accurate method that is also unconditionally stable. The readers are referred to the numerical procedures described in [29, 34, 41-43] for further details on the underlying methodology.

Results & Discussion

Predicting particle deposition in a parallel plate flow chamber apparatus

The implementation of the boundary condition for particle adhesion was first validated against in vitro experiments conducted in a parallel plate flow chamber apparatus. A schematic of the system and the geometry used for the computational analysis are presented in Supplementary Figure 1. The particles are injected in a 20 mm long channel with a rectangular cross section (10 × 0.274 mm) under controlled hydrodynamic conditions (fixed wall shear rate S). While the particles are transported downstream toward the waste, some of them can interact with and adhere firmly to the substrate. Details on the experimental conditions are provided in the Supplementary Materials.

The number of particles adhering nadh per unit area A, normalized by the total number of injected particles ninj, is presented in Figure 2 for different wall shear rates, namely S = 10, 75 and 200 s-1, and particle diameters. The experimental results (crosses with standard deviation bars) are compared with the numerical predictions (gray area), which are presented for different values of the parameters α1 and α2, as utilized in Eq.(11). The gray areas are obtained for a ± 5% variation of the mean value of α1 (= 0.023) (Figure 2 – left column) and α2 (≈ 9.44 ×109 for dp ≈ 0.7 μm; 1.388×109 for dp ≈ 5 μm; and 1.031×109 for dp ≈ 7 μm) (Figure 2 – right column).

Figure 2. Comparison between in silico and in vitro results.

Figure 2

The number of adhering particles nadh per unit surface area A normalized by the total number of injected particles ninj is plotted as a function of the particle diameter dp and for three different wall shear rates: (A) S = 10 s-1, (B) S = 75 s-1 and (C) S = 200 s-1. Black crosses with the standard deviation bars represent the experimental results obtained in a parallel plate flow chamber apparatus. The gray areas represent the in silico results obtained for values of the parameters α1 (left) and α2 (right) varying within ± 5% of their average values (α1 = 0.023) and (α2 ≈ 9.44 ×109 #/m2 for dp ≈ 0.7 μm; 1.388×109 #/m2 for dp ≈ 5 μm; and 1.031×109 #/m2 for dp ≈ 7 μm).

Three different adhesive behaviors are depicted as a function of the wall shear rate S: i) nadh grows steadily with dp, for S = 10 s-1 (Figure 2A); ii) nadh grows, reaches a maximum and then decreases with dp, for S = 75 s-1 (Figure 2B); iii) nadh decreases steadily with dp, S = 200 s-1 (Figure 2C). Particle adhesion is determined by the balance between interfacial adhesive interactions (specific and non-specific) and the dislodging hydrodynamic forces. At low S, (Figure 2A) the former dominates the latter, thus the steady increase in nadh up to dp ≈ 7 μm. Conversely, at high S, the adhesive forces cannot balance the dislodging forces, and nadh decreases steadily with dp (Figure 2C). At intermediate values of S, a maximum in adhesion appears for an optimal particle diameter (≈ 5 μm). For the conditions analyzed here, this happens for dp ≈ 5 μm, at S = 75 s-1.

However, this maximum dp, as well as its absolute value, changes with the adhesive properties of both the particle and the substrate [30]. Indeed, this maximum identifies the threshold in particle diameter below (above), which adhesive interactions prevail (do not prevail) over the dislodging hydrodynamic forces. Such a biphasic behavior was already predicted for the general case of oblate spheroidal particles as well [30], and correlates with the dependence of the vascular deposition parameter Π on dp (Supplementary Figure 2). Note that although the particles used in the flow chamber experiments were not decorated with any ligands, their adhesive behavior would follow the biphasic relationship of Eq.(11) due to the surface adsorption of non-specific molecules, mediating the interactions with the substrate.

Figure 2 demonstrates that the implemented adhesive boundary condition can accurately predict the complex, biphasic behavior for particle deposition under flow. Only at large shear rates, S = 200 s-1, and for large particle diameters (> 5 μm), the numerical predictions significantly underestimate the number of adhering particles. In this respect, however, it should be noted that for S = 200 s-1 and dp > 5 μm, the absolute number of particles adhering is extremely small (≈ 2.5/mm2). Thus, both numerical and experimental inaccuracies could explain the observed discrepancy.

It is also important to note that due to the highly convective nature of the particle transport (global Péclet number Pe can be 1000 and higher for all the particle sizes considered) diffusion plays a negligible role in particle mass transport within the core of the flow. As a result there appears to be little effect of particle size on near wall accumulation of particles in terms of C|s (i.e., particle availability near the wall for adhesion), as shown in Supplementary Figure 3. This implies that adhesion/deposition pattern for particles of different size is largely governed by the vascular deposition parameter Π, under the condition considered here. Also, the surface density of the adhering particles does not vary significantly over the channel length, as demonstrated in Supplementary Figure 4.

Predicting particle transport and deposition in a patient-specific vascular network

The Isogeometric Analysis formulation was used to simulate the transport and wall adhesion of particles injected through a catheter in a patient-specific arterial tree. From the CT scan imaging data of a healthy volunteer, a hexahedral solid NURBS model for a portion of the left coronary artery (LCA) tree was generated following the steps described in [28]. The geometry of the problem is presented in Figure 3A. A time-dependent pulsatile inflow condition [44, 45] with a period of 1 s (heart rate = 60 beats per minute) was imposed at the LCA inlet, where also a cylindrical catheter was located through which particles were injected, both radially and axially at a speed of 4 cm/s, for 5 cardiac cycles (5 s). The simulations were run on a computational mesh consisting of 55,100 quadratic NURBS elements for 14 cardiac cycles (14 s total) with a time step of 0.01 s employing the general solution strategies described in the Methods. Boundary layer meshes with the finest boundary element thickness of the order of 10-6 cm, are used for more accurate computation of wall quantities, such as the wall shear rate. The simulations were run with a time step of 0.05 s until all the particles left the fluid domain.

Figure 3. Nanoparticle transport in a patient-specific vascular tree.

Figure 3

(A) Schematic of the coronary artery (branches identified) with the inlet velocity profile (inset at the right) and applied boundary conditions. (B)-(F) Volumetric concentration C of 0.5 μm nanoparticles, normalized by the concentration at the catheter outlet C0, at various times t post injection: (B) t = 0.2 s, (C) t = 1 s, (D) t = 5 s, (E) t = 10 s, and (F) t = 14 s.

The time evolution of the particle volumetric concentration within the vascular network is presented in Figures 3B-F. With time, the particle distribution front moves from the LCA inlet, where the catheter is located, toward the downstream branch and eventually into the side branches of the left anterior descending (LAD – left) and circumflex (LCX – right) artery. At 1 s, the particle front has passed the bifurcation (Figure 3C); at 5 s, the particles are uniformly distributed within the vascular tree (Figure 3D); and at 14 s, they have left the LCX and LAD (Figure 3F). This distribution was obtained for the 0.5 μm particles. In the sequel, the particle transport and wall deposition are analyzed as a function of the particle size and patient-specific attributes, such as the vascular geometry and the receptor surface density.

First, two particles with different diameters, namely dp = 0.5 and 2 μm, were injected in the LCA. Figure 4 shows a side-by-side comparison of the time evolution for the wall surface concentration of the adhering particles, normalized by the injected dose. Three times are considered: 5, 10 and 14 s. For both particles, the surface concentration increases with time and becomes higher moving down the LCA, and approaching the LCX and LAD outlets. The larger particles (dp = 2 μm) exhibit about twice the adhesion of the smaller particles (dp = 0.5 μm), with a maximum normalized surface concentration of 7×10-8 and 2×10-8 cm-2, respectively. The smaller particles are observed to lodge more extensively in the upper branch (LCA) (Figure 4C).

Figure 4. Nanoparticle adhesion to the vessel walls: effect of nanoparticle size.

Figure 4

The normalized surface density of adhering nanoparticles is plotted along the arterial tree at various times t post injection, namely (A) t = 5 s, (B) t = 10 s and (C) t = 14 s. The left and right columns present in silico data for the 0.5 and 2.0 μm particle, respectively, in terms of particle number per unit area normalized by the injected dose [cm-2].

As discussed in the previous section, particle adhesion is mostly governed by the vascular deposition parameter Π and the shear rate at the wall S. Therefore, the variation of S along the vascular tree and the flow conditions were carefully quantified over time, as shown in Supplementary Figure 5. In the LCA, large shear rates are computed within the first 5 cardiac cycles (i.e., continuous injection), with values of the order of 500 – 1000 s-1, occurring in proximity of the branching point. Much lower values are computed in the LAD and LCX with S < 200 s-1. After 5 s, when the flow is no longer perturbed by the catheter injection, the wall shear rate decreases along the vascular tree with characteristic values ranging between 300 – 400 s-1 in the LCA and lower than 100 s-1 in the LAD and LCX. A comparison of Figure 4 and Supplementary Figure 5d reveals that the surface deposition patterns for the particles correlate well with the corresponding time averaged S distributions.

In the lower branches (LAD and LCX), before and after injection, the wall shear rate is generally equal to or lower than ≈ 100 s-1. Within this S range, the vascular deposition parameter Π for the 2 μm particles is, on the average, larger than that for the 0.5 μm particles (Supplementary Figure 2: at S = 50 s-1, Π ≈ 3.5×10-6 mm/s for the 2 μm particles, Π ≈ 1.75×10-6 mm/s for the 0.5 μm particles), thus explaining the larger adhesion (≈ 2 times) of the former compared to the latter. On the other hand, for larger values of S, as those experienced in the LAD, the 0.5 μm adhere more effectively than the 2 μm particles, in agreement with what is shown in Figure 4 (see also Supplementary Figure 2: at S = 500 s-1, Π ≈ 0 mm/s for the 2 μm particles, and Π ≈ 0.5×10-7 mm/s for the 0.5 μm particles). Incidentally, these observations are consistent with those reported in the literature where it has been shown that low and oscillating S zones are associated with enhanced deposition and uptake of lipoprotein (LDL), and correlate well with atherosclerotic regions [46-50]. Because of the pulsatile nature of blood flow in the coronaries, the associated complex flow features create recirculation zones near the bifurcation resulting in alternate areas of high and low S levels (Supplementary Figures 5c, d). It should also be noted that the injection flow rate from the catheter can significantly alter the wall shear rate distribution and thus affect the particle deposition rates and patterns.

Second, the branching angle between the LAD and LCX was increased from 76.8° to 106.8°, in the same patient-specific coronary geometry. As expected, the velocity magnitude and the wall shear rate distribution are significantly affected by such a change [51]. While the flow features appear similar in the LCA, the velocity magnitude and distribution become noticeably different as the branching point is approached, and moving downstream towards the two outlets (Supplementary Figure 5). First, a larger recirculation area is seen at the branching. Second, due to the sharper bend encountered as the blood flows from the LCA to the LAD, a higher S zone appears at the LCA-LAD junction. Finally, the flow patterns in the LAD and LCX are different, introducing an asymmetry in the vascular tree. This difference is more noticeable in the post catheter injection period. As a consequence of the change in the S distribution, particles were observed to adhere more at the walls of the LAD rather than in the LCX. For both particle sizes, this is shown in Figure 5 and additional details are provided in Supplementary Figure 6.

Figure 5. Nanoparticle adhesion to the vessel walls: effect of vascular geometry.

Figure 5

The normalized surface density of adhering nanoparticles is plotted along the arterial tree at t = 14 s, post injection. The left and right columns present in silico data for the 0.5 and 2.0 μm particles, respectively. The top and bottom rows present in silico data for the smaller (76.8°) and larger (106.8°) branching angles, respectively. Data are presented in particle number per unit area normalized by the injected dose [cm-2].

Third, the contribution of the over-expression of receptor molecules on particle vascular deposition is investigated. For this, the surface density mr in the LAD is increased by a factor of 10 compared to other regions in the vascular tree. A side-by-side comparison of the time evolution for the particle distribution is presented in Figure 6. Quite expectedly, particle concentration in the LAD is observed to be larger by almost an order of magnitude compared to the previous case, with no receptor over-expression. Note that the vascular deposition parameter Π grows almost linearly with mr. Particle adhesion is considerably enhanced because of the greater receptor availability that promotes specific interaction and increases the likelihood of ligand-receptor bond formation. It is also very interesting to observe that the increased receptor density tends to support a larger particle deposition within the recirculation zone as compared to other regions (Supplementary Figure 7). Note that the surface properties of the particles, such as density and type of ligand molecules, are accounted for in the model through the parameters ml (surface density of ligands on the particle surface); Ka0 (ligand-receptor affinity) and Δ (separation distance at equilibrium) in Eqs. (15-17). Therefore, the model can also be applied to explore different particle design scenarios such as multivalent targeting, PEGylation and the inclusion of polymer spacers in between ligand molecules and the particle surface.

Figure 6. Nanoparticle adhesion to the vessel walls: effect of surface density of vascular receptors.

Figure 6

The normalized surface density of 0.5 μm adhering nanoparticles is plotted along the arterial tree at various times t post injection: (A) t = 1 s, (B) t = 5 s and (C) t = 14 s. Data are presented for a uniform receptor density (mr = 109 #/cm2) along the vasculature (left column) and for a left anterior descending artery receptor density 10 times larger than that in the LCA and LCX (right column). Note that the color map scales are different for the two cases and give the particle number per unit area normalized by the injected dose [cm-2].

The results presented herein demonstrate that there are sites within the analyzed vascular network where larger particles (2.0 μm) would tend to adhere more avidly than smaller particles (0.5 μm). This is in qualitative agreement with the data reported in the recent work by the group of Dr. Eniola-Adefeso [52]. The vascular deposition of particles depends on the fine balance between hydrodynamic dislodging forces and interfacial adhesion forces. As a consequence, there is an optimal particle size that maximizes adhesion (see [30] and Suppl. Materials). The larger the adhesion forces (i.e., higher density of receptors/ligands and/or higher ligand-receptor affinity) and the smaller the shear rates, the larger the optimal particle size. This biphasic behavior is also evident in [52] for aVCAM-1 decorated particles, exposed to a peak wall shear rate of 200 s-1, presenting an optimal particle size of 2 μm. Indeed, additional information on the in vivo local hydrodynamic conditions; level of expression of vascular receptors and overall animal particle biodistribution would be required to provide a more accurate interpretation of the experimental results described in [52].

Conclusion

The computational framework described in this work allows us to quantify over time the distribution of the shear stresses at the wall and the volumetric concentration and wall surface density of nanoparticles injected in an authentic vascular network. This information is provided as a function of physiologically relevant patient-specific attributes, such as the architecture of the vasculature; the expression of endothelial receptor molecules; and the cardiac cycle, and depends on specific properties of the nanoparticles that can be finely controlled during the fabrication process, such as the geometry and surface properties (electrostatic charge, ligand density). While this work mainly focuses on the dependence of particle size on adhesion, information such as the vessel permeability and the nanoparticle shape, surface and stiffness can be readily implemented in the current formulation.

The presented results demonstrate that the adhesion pattern of intravascularly injected particles follows the distribution of the wall shear rates and is therefore significantly affected by the vascular geometry. Larger particles (2.0 μm in diameter) adhere more in the lower branches of the arterial tree, where the wall shear rates S are moderate (< 200 s-1); whereas smaller particles (0.5 μm in diameter) adhere more in the upper branches, where S is higher. It is also shown that as the branching angle in the arterial tree increases, the adhesion patterns become non-symmetric: patients with larger branching angles (≈ 106.8°) would receive more nanoparticles in the LAD than in the LCX. On the other hand, nanoparticles would distribute equally in the two arterial branches for patients with smaller branching angles (≈ 76.8°). Additionally, upregulation in the expression of vascular receptors is responsible for a non-symmetric and non-uniform vascular adhesion of the nanoparticles. These results clearly emphasize the importance of including patient-specific information for a proper selection of vascular delivery systems.

Future Perspective

In the near future, the following scenarios can be envisioned where computational modeling and nanoparticle engineering would be intimately integrated. Currently, MRI and CT are the most easily accessible, minimally invasive systems for whole human body imaging and, in conjunction with ultrasonic measurements, can provide information on the three-dimensional architecture and velocity field in a patient-specific vasculature. Nanoparticles can be engineered as potent MRI, or CT, contrast agents (imaging nanoconstructs) [53-56], molecularly targeted to a specific receptor family expressed over the diseased endothelial walls. This could be the integrin family on the tumor vasculature, inflammatory adhesion molecules in cardiovascular diseases and several others [57]. The number of adhering imaging nanoconstructs can be quantified by relating the contrast enhancement measured by the clinical scanner with the imaging properties of the single nanoconstruct. Knowing the number of adhering nanoparticles, the vascular geometry and mean hydrodynamic conditions, the computational model can be used to back calculate the mean receptor surface density following a reverse engineering approach. This would be extremely useful for the in vivo rapid screening of potential vascular targets in individual patients [57]. With all this information at hand, the optimal particle configuration, in terms of size, shape, surface and stiffness can be identified using the computational models with the objective of maximizing drug release at the target site while minimizing non-specific sequestration in healthy tissue. Indeed, these scenarios are within reach but can only be achieved with an orchestrated development of potent, safe nanoparticle-based imaging tools and accurate, efficient computational methods.

Executive Summary

Introduction

  • Drug biodistribution, bioavailability, efficacy and toxicity are unavoidably patient-specific.

  • Computational models are utilized to predict the deposition of nanoparticles in a patient-specific arterial tree as a function of the vascular architecture, flow conditions, receptor surface density and nanoparticle properties.

Materials & Methods

  • A 3D NURBS model of the patient-specific coronary artery tree was reconstructed from CT imaging data

  • A Navier-Stokes solver coupled to a scalar advection diffusion equation was employed within the Finite Element based Isogeometric Analysis framework to simulate blood flow and particle transport in the coronary artery. A special boundary condition for particle adhesion to the vessel wall was implemented.

  • The particle adhesion model was validated against in vitro parallel plate flow chamber experiments for a range of spherical particle size and wall shear rates. There is an optimal particle size that maximizes adhesion. The larger the adhesion forces (i.e., higher density of receptors/ligands and/or higher ligand-receptor affinity) and the smaller the shear rates, the larger the optimal particle size. The model accurately predicted this complex, biphasic behavior for particle deposition under flow.

Results & Discussion

  • Patient-specific attributes, such as the vascular architecture, blood flow and surface density of endothelial receptors, affect the deposition pattern of systemically injected nanoparticles.

  • The deposition pattern for nanoparticles correlates with the spatio-temporal distribution of the wall shear rates.

Future perspective

  • Multiscale computational modeling and novel nanoparticle-based contrast agents, enabling vascular imaging over multiple scales, could provide sufficient information for personalizing the delivery of therapeutic and imaging molecules.

Supplementary Material

Table 1.

Adhesion model parameters used.

Parameters Value
Surface density of ligand molecules ml = 1015 #/m2
Surface density of receptor molecules mr = 1013 #/m2
Ligand-receptor affinity constant at zero load Ka0 = 2.3×10-7 m2
Characteristic length of ligand-receptor bond λ = 1×10-10 m
Dynamic viscosity of water μ = 0.001 N-s/m2
Drag coefficient on the spherical particle Fs = 1.668

Acknowledgments

This work was supported by the Cancer Prevention Research Institute of Texas through the grant CPRIT RP110262, and through the grants from the National Institutes of Health (USA) (NIH) U54CA143837 and U54CA151668. The additional support for MF from DoD/BCRP (W81XWH-09-1-0212) and the Ernest Cockrell Jr. Distinguished Endowed Chair, and that for TJRH from the Moncrief Grand Challenge Award Program and Portuguese CoLab Grant No. 04A are also gratefully acknowledged. The authors thank John A. Evans for his input and suggestions regarding numerical stabilization and Benjamin Urick from Texas Advanced Computing Center for his help with the artwork.

Footnotes

Author contributions: PD and SSH drafted the manuscript. SSH performed all the numerical implementation and calculations, developed the wall boundary condition and generated the figures; YZ helped in modifying the vascular geometry and writing the manuscript; XL helped in modifying the vascular geometry; FH provided input on the original draft and revision; MF provided input on the original draft and its revisions; TJRH developed the computational methodology, supervised the numerical simulations, and provided input on the original draft and its revisions; PD designed the “experimental plan”, and developed the wall boundary condition. All the authors discussed the results and commented on the manuscript.

Contributor Information

Yongjie Zhang, Email: jessicaz@andrew.cmu.edu.

Xinghua Liang, Email: xliang1@andrew.cmu.edu.

Fazle Hussain, Email: fhussain@uh.edu.

Mauro Ferrari, Email: mferrari@tmhs.org.

Thomas J. R. Hughes, Email: hughes@ices.utexas.edu.

References

  • 1.Weinshilboum R. Genomic medicine - Inheritance and drug response. New England Journal of Medicine. 2003;348(6):529–537. doi: 10.1056/NEJMra020021. [DOI] [PubMed] [Google Scholar]
  • 2.Evans WE, Mcleod HL. Drug therapy - Pharmacogenomics - Drug disposition, drug targets, and side effects. New England Journal of Medicine. 2003;348(6):538–549. doi: 10.1056/NEJMra020526. [DOI] [PubMed] [Google Scholar]
  • 3.Kent DM, Hayward RA. Limitations of applying summary results of clinical trials to individual patients - The need for risk stratification. Jama-Journal of the American Medical Association. 2007;298(10):1209–1212. doi: 10.1001/jama.298.10.1209. [DOI] [PubMed] [Google Scholar]
  • 4.Ghosh S, Matsuoka Y, Asai Y, Hsin KY, Kitano H. Software for systems biology: from tools to integrated platforms. Nat Rev Genet. 2011;12(12):821–832. doi: 10.1038/nrg3096. [DOI] [PubMed] [Google Scholar]
  • 5.Vizirianakis IS. Nanomedicine and personalized medicine toward the application of pharmacotyping in clinical practice to improve drug-delivery outcomes'. Nanomedicine-Nanotechnology Biology and Medicine. 2011;7(1):11–17. doi: 10.1016/j.nano.2010.11.002. [DOI] [PubMed] [Google Scholar]
  • 6.Nie S. Nanotechnology for personalized and predictive medicine. Nanomedicine: nanotechnology, biology, and medicine. 2006;2(4):305. [Google Scholar]
  • 7.Michor F, Liphardt J, Ferrari M, Widom J. What does physics have to do with cancer? Nat Rev Cancer. 2011;11(9):657–670. doi: 10.1038/nrc3092. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8*.Peer D, Karp JM, Hong S, Farokhzad OC, Margalit R, Langer R. Nanocarriers as an emerging platform for cancer therapy. Nature Nanotechnology. 2007;2(12):751–760. doi: 10.1038/nnano.2007.387. Documents on the use of nanoparticles in drug delivery and biomedical imaging for cancer and other diseases, elucidating the advantages over systemically injected free molecules. [DOI] [PubMed] [Google Scholar]
  • 9*.Sutradhar A, Paulino GH, Miller MJ, Nguyen TH. Topological optimization for designing patient-specific large craniofacial segmental bone replacements. Proceedings of the National Academy of Sciences of the United States of America. 2010;107(30):13222–13227. doi: 10.1073/pnas.1001208107. Describes the successful integration of computational modeling with patient-specific data for solving clinically relevant problems in orthopedic applications. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Williams JM, Adewunmi A, Schek RM, et al. Bone tissue engineering using polycaprolactone scaffolds fabricated via selective laser sintering. Biomaterials. 2005;26(23):4817–4827. doi: 10.1016/j.biomaterials.2004.11.057. [DOI] [PubMed] [Google Scholar]
  • 11.Hwang CW, Wu D, Edelman ER. Physiological Transport Forces Govern Drug Distribution for Stent-Based Delivery. Circulation. 2001;104(5):600–605. doi: 10.1161/hc3101.092214. [DOI] [PubMed] [Google Scholar]
  • 12.Migliavacca F, Dubini G. Computational modeling of vascular anastomoses. Biomechanics and Modeling in Mechanobiology. 2005;3(4):235–250. doi: 10.1007/s10237-005-0070-2. [DOI] [PubMed] [Google Scholar]
  • 13.Taylor CA, Hughes TJR, Zarins CK. Finite Element Modeling of Three-Dimensional Pulsatile Flow in the Abdominal Aorta: Relevance to Atherosclerosis. Annals of Biomedical Engineering. 1998;26(6):975–987. doi: 10.1114/1.140. [DOI] [PubMed] [Google Scholar]
  • 14*.Hossain S, Hossainy S, Bazilevs Y, Calo V, Hughes T. Mathematical modeling of coupled drug and drug-encapsulated nanoparticle transport in patient-specific coronary artery walls. Computational Mechanics. 2012;49(2):213–242. Discusses the application of isogeometric analysis to simulate drug and nanoparticle transport within the artery wall. [Google Scholar]
  • 15.Deisboeck TS, Wang Z, Macklin P, Cristini V. Multiscale Cancer Modeling. Annual Review of Biomedical Engineering. 2011;13(1):127–155. doi: 10.1146/annurev-bioeng-071910-124729. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Swanson KR, Rockne RC, Claridge J, Chaplain MA, Alvord EC, Anderson AR. Quantifying the role of angiogenesis in malignant progression of gliomas: In silico modeling integrates imaging and histology. Cancer Research. 2011 doi: 10.1158/0008-5472.CAN-11-1399. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Schoeberl B, Eichler-Jonsson C, Gilles ED, Muller G. Computational modeling of the dynamics of the MAP kinase cascade activated by surface and internalized EGF receptors. Nature Biotechnology. 2002;20(4):370–375. doi: 10.1038/nbt0402-370. [DOI] [PubMed] [Google Scholar]
  • 18.Kim HJ, Vignon-Clementel IE, Coogan JS, Figueroa CA, Jansen KE, Taylor CA. Patient-Specific Modeling of Blood Flow and Pressure in Human Coronary Arteries. Annals of Biomedical Engineering. 2010;38(10):3195–3209. doi: 10.1007/s10439-010-0083-6. [DOI] [PubMed] [Google Scholar]
  • 19.Torii R, Oshima M, Kobayashi T, Takagi K, Tezduyar TE. Fluid-structure interaction modeling of aneurysmal conditions with high and normal blood pressures. Computational Mechanics. 2006;38(4-5):482–490. [Google Scholar]
  • 20.Ladisa JF, Guler I, Olson LE, et al. Three-dimensional computational fluid dynamics modeling of alterations in coronary wall shear stress produced by stent implantation. Annals of Biomedical Engineering. 2003;31(8):972–980. doi: 10.1114/1.1588654. [DOI] [PubMed] [Google Scholar]
  • 21.Decuzzi P, Pasqualini R, Arap W, Ferrari M. Intravascular Delivery of Particulate Systems: Does Geometry Really Matter? Pharmaceutical Research. 2009;26(1):235–243. doi: 10.1007/s11095-008-9697-x. [DOI] [PubMed] [Google Scholar]
  • 22.Lee SY, Ferrari M, Decuzzi P. Design of bio-mimetic particles with enhanced vascular interaction. Journal of Biomechanics. 2009;42(12):1885–1890. doi: 10.1016/j.jbiomech.2009.05.012. [DOI] [PubMed] [Google Scholar]
  • 23.Van De Ven AL, Kim P, Haley OH, et al. Rapid tumoritropic accumulation of systemically injected plateloid particles and their biodistribution. Journal of Controlled Release. 2012;158(1):148–155. doi: 10.1016/j.jconrel.2011.10.021. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Rolland JP, Maynor BW, Euliss LE, Exner AE, Denison GM, Desimone JM. Direct fabrication and harvesting of monodisperse, shape-specific nanobiomaterials. Journal of the American Chemical Society. 2005;127(28):10096–10100. doi: 10.1021/ja051977c. [DOI] [PubMed] [Google Scholar]
  • 25.Chiappini C, Tasciotti E, Fakhoury JR, et al. Tailored Porous Silicon Microparticles: Fabrication and Properties. Chemphyschem. 2010;11(5):1029–1035. doi: 10.1002/cphc.200900914. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Champion JA, Katare YK, Mitragotri S. Making polymeric micro- and nanoparticles of complex shapes. Proceedings of the National Academy of Sciences of the United States of America. 2007;104(29):11901–11904. doi: 10.1073/pnas.0705326104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Acharya G, Shin CS, Mcdermott M, et al. The hydrogel template method for fabrication of homogeneous nano/microparticles. Journal of Controlled Release. 2010;141(3):314–319. doi: 10.1016/j.jconrel.2009.09.032. [DOI] [PubMed] [Google Scholar]
  • 28.Zhang Y, Bazilevs Y, Goswami S, Bajaj CL, Hughes TJR. Patient-specific vascular NURBS modeling for isogeometric analysis of blood flow. Computer Methods in Applied Mechanics and Engineering. 2007;196(29-30):2943–2959. doi: 10.1016/j.cma.2007.02.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Hughes TJR, Cottrell JA, Bazilevs Y. Isogeometric analysis: CAD, finite elements, NURBS, exact geometry and mesh refinement. Computer Methods in Applied Mechanics and Engineering. 2005;194(39-41):4135–4195. [Google Scholar]
  • 30**.Decuzzi P, Ferrari M. The adhesive strength of non-spherical particles mediated by specific interactions. Biomaterials. 2006;27(30):5307–5314. doi: 10.1016/j.biomaterials.2006.05.024. Describes the mathematical model for the specific/non-specific particle adhesion and demonstrates the advantages of non-spherical particles in firmly adhering to the vessel walls as compared to conventional spherical beads. [DOI] [PubMed] [Google Scholar]
  • 31.Gijsen FJH, Van De Vosse FN, Janssen JD. The influence of the non-Newtonian properties of blood on the flow in large arteries: steady flow in a carotid bifurcation model. Journal of Biomechanics. 1999;32(6):601–608. doi: 10.1016/s0021-9290(99)00015-9. [DOI] [PubMed] [Google Scholar]
  • 32*.Taylor CA, Humphrey JD. Open problems in computational vascular biomechanics: Hemodynamics and arterial wall mechanics. Computer Methods in Applied Mechanics and Engineering. 2009;198(45-46):3514–3523. doi: 10.1016/j.cma.2009.02.004. Reviews on the multiscale modeling of blood flow along the vascular three, from the larger arteries to the capillary bed, emphasizing the physical, biological and computational challenges. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Ballyk PD, Steinman DA, Ethier CR. Simulation of non-Newtonian blood flow in an end-to-end anastomosis. Biorheology. 1994;31:565–586. doi: 10.3233/bir-1994-31505. [DOI] [PubMed] [Google Scholar]
  • 34**.Bazilevs Y, Calo VM, Zhang Y, Hughes TJR. Isogeometric Fluid–structure Interaction Analysis with Applications to Arterial Blood Flow. Computational Mechanics. 2006;38(4):310–322. Describes the mathematical formulation and numerical procedures associated with the application of NURBS based isogeometric analysis technology to simulate arterial blood flow. [Google Scholar]
  • 35.Bazilevs Y, Calo VM, Cottrell JA, Hughes TJR, Reali A, Scovazzi G. Variational multiscale residual-based turbulence modeling for large eddy simulation of incompressible flows. Computer Methods in Applied Mechanics and Engineering. 2007;197(1-4):173–201. [Google Scholar]
  • 36.Olgac U, Poulikakos D, Saur SC, Alkadhi H, Kurtcuoglu V. Patient-specific three-dimensional simulation of LDL accumulation in a human left coronary artery in its healthy and atherosclerotic states. American Journal of Physiology-Heart and Circulatory Physiology. 2009;296(6):H1969–H1982. doi: 10.1152/ajpheart.01182.2008. [DOI] [PubMed] [Google Scholar]
  • 37.Stangeby DK, Ethier CR. Computational Analysis of Coupled Blood-Wall Arterial LDL Transport. Journal of Biomechanical Engineering. 2002;124(1):1–8. doi: 10.1115/1.1427041. [DOI] [PubMed] [Google Scholar]
  • 38.Liu J, Weller GER, Zern B, et al. Computational model for nanocarrier binding to endothelium validated using in vivo, in vitro, and atomic force microscopy experiments. Proceedings of the National Academy of Sciences of the United States of America. 2010;107(38):16530–16535. doi: 10.1073/pnas.1006611107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Chung J, Hulbert GM. A time integration algorithm for structural dynamics with improved numerical dissipation: The generalized- method. Journal of Applied Mechanics. 1993;60:371–375. [Google Scholar]
  • 40.Jansen KE, Whiting CH, Hulbert GM. A generalized-[alpha] method for integrating the filtered Navier-Stokes equations with a stabilized finite element method. Computer Methods in Applied Mechanics and Engineering. 2000;190(3-4):305–319. [Google Scholar]
  • 41.Hossain SS. Mathematical Modeling of Coupled Drug and Drug-encapsulated Nanoparticle Transport in Patient-specific Coronary Artery Walls. Department of Mechanical Engineering Ph D. 2009 [Google Scholar]
  • 42.Bazilevs Y, Da Veiga LB, Cottrell JA, Hughes TJR, Sangalli G. Isogeometric analysis: Approximation, stability and error estimates for h-refined meshes. Mathematical Models & Methods in Applied Sciences. 2006;16(7):1031–1090. [Google Scholar]
  • 43.Cottrell JA, Hughes TJR, Reali A. Studies of refinement and continuity in isogeometric structural analysis. Computer Methods in Applied Mechanics and Engineering. 2007;196(41,Äì44):4160–4183. [Google Scholar]
  • 44.Matsuo S, Tsuruta M, Hayano M, et al. Phasic coronary-artery flow velocity determined by doppler flowmeter catheter in aortic-stenosis and aortic regulation. American Journal of Cardiology. 1988;62(13):917–922. doi: 10.1016/0002-9149(88)90893-4. [DOI] [PubMed] [Google Scholar]
  • 45.Johnston BM, Johnston PR, Corney S, Kilpatrick D. Non-Newtonian blood flow in human right coronary arteries: Transient simulations. Journal of Biomechanics. 2006;39(6):1116–1128. doi: 10.1016/j.jbiomech.2005.01.034. [DOI] [PubMed] [Google Scholar]
  • 46.Gawaz M, Langer H, May AE. Platelets in inflammation and atherogenesis. The Journal of Clinical Investigation. 2005;115(12):3378–3384. doi: 10.1172/JCI27196. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Zarins CK, Giddens DP, Bharadvaj BK, Sottiurai VS, Mabon RF, Glagov S. Carotid bifurcation atherosclerosis. Quantitative correlation of plaque localization with flow velocity profiles and wall shear stress. Circulation Research. 1983;53(4):502–514. doi: 10.1161/01.res.53.4.502. [DOI] [PubMed] [Google Scholar]
  • 48.Caro CG, Fitz-Gerald JM, Schroter RC. Atheroma and Arterial Wall Shear Observation, Correlation and Proposal of a Shear Dependent Mass Transfer Mechanism for Atherogenesis. Proceedings of the Royal Society of London Series B Biological Sciences. 1971;177(1046):109–133. doi: 10.1098/rspb.1971.0019. [DOI] [PubMed] [Google Scholar]
  • 49.Feldman CL, Ilegbusi OJ, Hu Z, Nesto R, Waxman S, Stone PH. Determination of in vivo velocity and endothelial shear stress patterns with phasic flow in human coronary arteries: A methodology to predict progression of coronary atherosclerosis. American Heart Journal. 2002;143(6):931–939. doi: 10.1067/mhj.2002.123118. [DOI] [PubMed] [Google Scholar]
  • 50.Ku D, Giddens D, Zarins C, Glagov S. Pulsatile flow and atherosclerosis in the human carotid bifurcation. Positive correlation between plaque location and low oscillating shear stress. Arteriosclerosis, Thrombosis, and Vascular Biology. 1985;5(3):293–302. doi: 10.1161/01.atv.5.3.293. [DOI] [PubMed] [Google Scholar]
  • 51.Perktold K, Peter RO, Resch M, Langs G. Pulsatile non-newtonian blood flow in three-dimensional carotid bifurcation models: a numerical study of flow phenomena under different bifurcation angles. Journal of Biomedical Engineering. 1991;13(6):507–515. doi: 10.1016/0141-5425(91)90100-l. [DOI] [PubMed] [Google Scholar]
  • 52.Charoenphol P, Mocherla S, Bouis D, Namdee K, Pinsky DJ, Eniola-Adefeso O. Targeting therapeutics to the vascular wall in atherosclerosis, ÄìCarrier size matters. Atherosclerosis. 2011;217(2):364–370. doi: 10.1016/j.atherosclerosis.2011.04.016. [DOI] [PubMed] [Google Scholar]
  • 53.Winter PM, Caruthers SD, Kassner A, et al. Molecular Imaging of angiogenesis in nascent vx-2 rabbit tumors using a novel alpha(v)beta(3)-targeted nanoparticle and 1.5 tesla magnetic resonance imaging. Cancer Research. 2003;63(18):5838–5843. [PubMed] [Google Scholar]
  • 54.Chen W, Jarzyna PA, Van Tilborg GaF, et al. RGD peptide functionalized and reconstituted high-density lipoprotein nanoparticles as a versatile and multimodal tumor targeting molecular imaging probe. Faseb Journal. 2010;24(6):1689–1699. doi: 10.1096/fj.09-139865. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Jokerst JV, Gambhir SS. Molecular Imaging with Theranostic Nanoparticles. Accounts of Chemical Research. 2011;44(10):1050–1060. doi: 10.1021/ar200106e. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Ananta JS, Godin B, Sethi R, et al. Geometrical confinement of gadolinium-based contrast agents in nanoporous particles enhances T(1) contrast. Nature Nanotechnology. 2010;5(11):815–821. doi: 10.1038/nnano.2010.203. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Staquicini FI, Cardo-Vila M, Kolonin MG, et al. Vascular ligand-receptor mapping by direct combinatorial selection in cancer patients. Proceedings of the National Academy of Sciences. 2011;108(46):18637–18642. doi: 10.1073/pnas.1114503108. [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

RESOURCES