nanomaterials Article

Characterization of Nanoparticle Dispersion in Red Blood Cell Suspension by the Lattice Boltzmann-Immersed Boundary Method Jifu Tan 1 , Wesley Keller 2 , Salman Sohrabi 1 , Jie Yang 3, * and Yaling Liu 1,4, * 1 2 3 4

*

Department of Mechanical Engineering and Mechanics, Lehigh University, Bethlehem, PA 18015, USA; [email protected] (J.T.); [email protected] (S.S.) Department of Civil and Environmental Engineering, Lehigh University, Bethlehem, PA 18015, USA; [email protected] School of Mechanics and Engineering, Southwest Jiaotong University, Chengdu 610031, China Bioengineering Program, Lehigh University, Bethlehem, PA 18015, USA Correspondence: [email protected] (J.Y.); [email protected] (Y.L.); Tel.: +86-28-8760-0797 (J.Y.); +1-610-758-5839 (Y.L.)

Academic Editor: Ming Su Received: 16 November 2015; Accepted: 25 January 2016; Published: 5 February 2016

Abstract: Nanodrug-carrier delivery in the blood stream is strongly influenced by nanoparticle (NP) dispersion. This paper presents a numerical study on NP transport and dispersion in red blood cell (RBC) suspensions under shear and channel flow conditions, utilizing an immersed boundary fluid-structure interaction model with a lattice Boltzmann fluid solver, an elastic cell membrane model and a particle motion model driven by both hydrodynamic loading and Brownian dynamics. The model can capture the multiphase features of the blood flow. Simulations were performed to obtain an empirical formula to predict NP dispersion rate for a range of shear rates and cell concentrations. NP dispersion rate predictions from the formula were then compared to observations from previous experimental and numerical studies. The proposed formula is shown to accurately predict the NP dispersion rate. The simulation results also confirm previous findings that the NP dispersion rate is strongly influenced by local disturbances in the flow due to RBC motion and deformation. The proposed formula provides an efficient method for estimating the NP dispersion rate in modeling NP transport in large-scale vascular networks without explicit RBC and NP models. Keywords: lattice Boltzmann method; immersed boundary method; cell suspension; nanoparticle delivery; dispersion rate

1. Introduction Accurately predicting drug delivery is a critical task in drug development research and clinical trials [1,2]. It requires careful consideration of physiological conditions, such as hematocrit level (the volume ratio between red blood cell and the whole blood) [3,4], vessel geometry and flow conditions [5–7], drug carrier size and shape [3,8], dissolution rate [9] and external stimuli [10,11]. For small particles in red blood cell (RBC) suspensions, such as nanoparticles (NP) and platelets, recent studies have demonstrated that local flow field disturbances caused by RBC translation and deformation can enhance particle dispersion [3,12–16]. The migration of particles in RBC suspensions under shear has been shown to behave like a random walk process [17,18], with a dispersion rate much larger than thermal diffusion. Therefore, accurate predictions of NP dispersion in RBC suspensions must consider fluid-structure interaction between the immersed solid bodies (particles and cells) and the surrounding fluid. Previously-developed models for predicting NP dispersion in RBC

Nanomaterials 2016, 6, 30; doi:10.3390/nano6020030

www.mdpi.com/journal/nanomaterials

Nanomaterials 2016, 6, 30

2 of 14

suspensions have relied primarily on empirical data fitting. Aarts et al. experimentally studied shear-induced platelet diffusivity (D), which was fitted with shear rate (η) as a power law D “ kη n , where k is a constant and n is a function of hematocrit [19]. However, the model parameters are obtained empirically rather than predicted from the underlying physics. Decuzzi et al. extended the Taylor–Aris theory to calculate an effective NP diffusion rate that considers wall permeability and blood rheology [20,21]. They also reported about a three-fold increase in dispersion rate of 1 µm compared to thermal diffusion [16]. Recently, Fedosov’s group systematically studied micro- and nano-particles in drug delivery, including particle size, shape effect and RBC influence on particle margination and adhesion probabilities [3]. However, there is no analytical formula or quantitative rule to directly predict the NP dispersion rate so far, which is much needed in large-scale drug delivery simulations [20–22]. In order to address the deficiencies in previously-developed models for predicting NP dispersion, this paper presents a numerical study on NP dispersion in RBC suspensions that considers the effects of local flow field disturbances due to RBC motion. This study provides insight into the underlying physics driving NP dispersion in these systems and develops simple, yet effective, formulae for predicting dispersion rate as a function of characteristic physiological parameters. These simple predictive formulae will provide an efficient approach for assessing NP dispersion under different flow conditions and hematocrit level, thereby facilitating practical modeling of NP transport and distribution in large-scale vascular systems [22]. The remainder of this paper is organized as follows. The fluid-structure interaction model is introduced in Section 2, including a description of the modeling approach for the fluid environment, RBCs and NPs, as well as the fluid-structure coupling scheme. Section 3 outlines the model setup and test parameters for a parametric study on NP dispersion. Simulation results from the parametric study are presented in Section 4, along with formulae derived from a regression analysis for predicting the NP dispersion rate. Predictions from the formulae are compared to data reported in the literature. Finally, conclusions and recommendations from the study are presented in Section 5. 2. Fluid-Structure Interaction Model The transport of particles in RBC suspensions is governed by hydrodynamic forces and fluid-structure interaction effects. In order to simulate this two-way coupling between the fluid environment and the immersed soft matter, a numerical fluid-structure interaction (FSI) code was developed utilizing an immersed boundary coupling scheme. Previous research has shown that the immersed boundary method (IBM) is an efficient approach to simulate soft matter and biological cells [23–30], flapping insect wings [31–33], harmonic oscillation of thin lamina in fluid [34] and other FSI problems, such as particle settling [35]. The advantage of this approach is two-fold. First, the lattice Boltzmann method (LBM) is very efficient at modeling fluid flow and ideal for parallel computing [36]. Second, the IBM uses an independent solid mesh moving on top of a fixed fluid mesh, thus removing the burden of re-meshing in conventional arbitrary Lagrangian–Eulerian methods [37]. In this study, the fluid was modeled using a lattice Boltzmann scheme, while the RBCs were modeled as elastic membranes. The fluids inside and outside the cell membrane are modeled with the same viscosity, following that in other works [38,39]. This approach greatly simplified the computation and without losing too much generality. It also can capture the multiphase feature of the blood flow. NPs were treated as rigid point elements with motions governed by both hydrodynamic loading and Brownian dynamics. Additional details regarding the fluid-structure interaction model are provided in the following sections. 2.1. Lattice Boltzmann Fluid Model The lattice Boltzmann method (LBM) is a mesoscale approach to modeling fluid dynamics that has been used extensively in blood flow modeling [24,38,40]. Reviews of the underlying theory for the LBM can be found in the literature [41–44]. LBM is usually considered as a second order accurate method in space and time [45]. The key concept of the LBM is the transport of particles, characterized

Nanomaterials 2016, 6, 30

3 of 14

Ñ

Ñ

by a local density distribution function f i (x,t) in phase space (x, c i ), where t denotes the time and c i denotes the lattice velocity. The dynamics of the LBM involves streaming and collision processes. ´ ¯ Ñ f i x ` ∆t c i , t ` ∆t ´ f i px, tq “ looΩ mooin ` Fi looooooooooooooooooomooooooooooooooooooon

(1)

Collision

Streaming

where Ωi is the collision operator and Fi is the external force term. The simplest and most computationally-efficient collision scheme is the Bhatnagar–Gross–Krook (BGK) scheme [46], Ωi “

1 eq p f px, tq ´ f i px, tqq τ i

(2)

eq

where τ is a relaxation constant and f i px, tq is the population distribution at equilibrium, which is related to the local macroscale fluid velocity and the speed of sound cs . ¨ ˛ ´Ñ ѯ2 Ñ Ñ Ñ Ñ ´Ñ ¯ c ¨ u i c ¨u u¨u‹ eq ˚ fi u , ρ “ wi ρ ˝1 ` i 2 ` ´ ‚ cs 2cs 2 2cs 4

(3)

where wi is the weight coefficients. An effective fluid viscosity is related to the relaxation time τ and the speed of sound cs , pτ ´ 0.5q (4) ν “ c2s pτ ´ 0.5q “ 3 and local macroscale density and velocity are calculated as: ρ px, tq “

ÿ

Ñ

f i px, tq, ρ px, tq u px, tq “

ÿ

Ñ

i f i px, tq c i

`

i

∆t Ñ f 2

(5)

Disturbances in the flow are introduced through a force density term Fi , which can be expressed Ñ

Ñ

in terms of an external body force density f and fluid macroscale velocity u , 1 Fi “ p1 ´ qw 2τ i

˜Ñ Ñ Ñ Ñ ¸ Ñ ci´ u c ¨uÑ ` i 4 ci ¨ f 2 cs cs

(6)

For simplicity, the simulations presented in this paper consider a regular 2D domain with a D2Q9 (2D 9 velocities) fluid lattice, the details of which can be found in [45]. However, it is noted that the modeling approach is readily adaptable to the 3D case (see Figure S9 in the Supplementary Materials) and with more complex geometric configurations. Velocity and non-slip wall boundaries were used. There are several ways to implement velocity boundary conditions, such as Zou/He boundaries [47] and regularized boundary conditions [48]. The basic idea behind Zou/He boundaries is assuming that the bounce back rule applies to the non-equilibrium part of the density population. The regularized boundary condition is to use the moment flux tensor to reconstruct the non-equilibrium part of the density distribution, thus it is more stable, but requires extra calculation. Interested readers are referred to [49] for a summary of different boundary conditions in LBM. In this paper, treatment of velocity boundary conditions followed the recommendations in [47] and is discussed in Section 3. 2.2. Spring Connected Network Cell Membrane Model Cell models are commonly based on a continuum approach utilizing a strain energy function to define membrane response [50–52]. Recently, however, a particle-based cell model governed by molecular dynamics has emerged as a popular alternative due to its simplicity in mathematical description [53–55]. These two models have been shown to provide consistent predictions [53]. For the present study, a particle-based RBC model is adopted. The model consists of a network of vertices,

Nanomaterials 2016, 6, 30

4 of 14

Xi , i P t1...NV u, connected with elastic springs, which are effective in axial extension/contraction and flexure, as shown in6,Figure 1. Nanomaterials 2016, 30 4 of 14 8 µm

A

B

C

θ

2 µm Cell membrane

Stretching

Bending

Figure 1. (A) Spring connected network cell membrane model. Kinematics for local stretching (B) and

Figure 1. (A) Spring connected network cell membrane model. Kinematics for local stretching (B) and bending (C) modes of response. bending (C) modes of response.

The potential energy of the membrane is defined as:

The potential energy of the membrane as: V { X i } =isVdefined stretch + V ben ding

+ V area

(7)

The stretch energy mimics V the spectrin`network and u“V tXelastic V ` Vis given by:

(7) area bending 2 1 Vstretchspectrin = ks network Lj − Land j0 (8) The stretch energy mimics the elastic is given by: 2 j∈1...N s ÿ ` ˘2 1 where k s is the stretching constant, of L the L j is L j0 and L j 0 is the equilibrium spring (8) Vstretch “theklength s j ´spring 2 i

stretch

(

)

jP1...Ns stretching, Fedosov investigated a wormlike length. To consider the nonlinear effect of membrane chain potential [56]. In a separate study, Wida developed an exponential relationship for mechanical wherestiffness ks is the stretching constant, L is the length of the spring and L is the equilibrium spring related to the bond stretch j ratio ( λ ) [57]. In this paper, the samej0exponential form of the length. To consider nonlinear effect of membrane stretching, Fedosov investigated a wormlike spring constant, the k s = k s 0 exp [ 2( λ − 1) ] , where λ is the bond stretch ratio, has been adopted.

chain potential [56]. In a separate study, Wida developed an exponential relationship for mechanical The bending energy represents the bending resistance of the lipid bilayer and is defined as: stiffness related to the bond stretch ratio (λ) [57]. In this paper, the same exponential form of the spring 2 1 constant, k s “ k s0 exp r2pλ ´ 1qs, where λ is the stretch has been adopted. Vbending = bond kb θ j −θ ratio, j0 (9) 2 The bending energy represents the bending resistance of the lipid bilayer and is defined as: j∈1...Nb

(

)

˘2 where k b is the bending constant and θ 1 andÿθ j 0 `are instantaneous and equilibrium angles j Vbending “

kb

θ j ´ θ j0

between adjacent spring, respectively. In this2model, we used the angles when cells are at rest as the jP1...Nb equilibrium angles θ j 0 ; each angle is also different from the others.

(9)

where k b is the the bending constant θ j andarea θ j0 are and equilibrium between When cell deforms, the and cell surface and instantaneous volume remain relatively constant. angles To maintain 2 adjacent spring, respectively. In this model, we used the angles when cells are at rest the equilibrium kas ( A − A ) 0 is a constant area in the 2D model, an artificial restoring force potential V area = a 2 A0 angles θ j0 ; each angle is also different from the others. A area When the cell deforms, cell surface volume remain constant.area To maintain introduced to conserve thethe area where is theand instantaneous area, Arelatively and 0 is the equilibrium k a pA ´ A0 q2 k a is a constant. a constant area in the 2D model, an artificial restoring force potential Varea “ is introduced The equilibrium shape of cells is followed by the analytical curve from Fung2A [58], 0 as shown in to conserve theThe area A is the area, A0 and is the equilibrium area andbond k a is length a constant. Figure 1. cellwhere membrane wasinstantaneous discretized to 52 nodes angles. The equilibrium The equilibrium shape of cells is followed by the analytical curve from Fung [58], as shown in was set to be equal, but the equilibrium angle varies among all of the angles. Figure 1. The cell membrane was discretized to 52 nodes and angles. The equilibrium bond length was Immersed Boundary Coupling Scheme set to2.3. be equal, but the equilibrium angle varies among all of the angles. The immersed boundary method (IBM) was selected to model the interaction between the fluid

2.3. Immersed Boundarysolids Coupling Scheme and the immersed due to the algorithm’s efficiency. The IBM was first proposed by Peskin to

study blood flow in the heart [59]. The fluidwas is solved on to a spatially-fixed Eulerian grid, whilethe thefluid The immersed boundary method (IBM) selected model the interaction between immersed solids are modeled using a moving Lagrangian mesh, which is not constrained to the and the immersed solids due to the algorithm’s efficiency. The IBM was first proposed by Peskin geometric layout of the Eulerian fluid grid. Data are exchanged between the two domains through to study blood flow in the heart [59]. The fluid is solved on a spatially-fixed Eulerian grid, while nodal interpolation. The coupling scheme enforces velocity continuity at the fluid-structure the immersed solids are modeled using a moving Lagrangian mesh, which is not constrained to boundary and transfers forces from the structure back into the fluid through an effective force the geometric layout of the Eulerian fluid grid. Dataimmersed are exchanged between the two solid domains density. This two-way coupling automatically handles body contact and prevents through nodal interpolation. The coupling schemeforces enforces velocity continuity fluid-structure penetration through the development of restoring in the fluid. The approach at hasthe been used for boundary andoftransfers forcesinteraction from theproblems, structureincluding back into fluid through an [60], effective a variety fluid-structure thethe simulation of jelly fish bloodforce flowThis [4,24,26–30,38] plateletautomatically migration [14]. Comprehensive reviews of thecontact IBM andand its applications density. two-wayand coupling handles immersed body prevents solid can be through found in the [25,61]. It is noticed the current approach works great softused penetration development of that restoring forcescoupling in the fluid. The approach hasfor been for a variety of fluid-structure interaction problems, including the simulation of jelly fish [60],

Nanomaterials 2016, 6, 30

5 of 14

blood flow [4,24,26–30,38] and platelet migration [14]. Comprehensive reviews of the IBM and its applications can be found in [25,61]. It is noticed that the current coupling approach works great for soft deformable solids, an alternative coupling approach, e.g., treating a solid as a moving boundary for the fluid and transferring the force back to the solid, should be used if the solid is relatively rigid, such as a particulate settling study [44,62–64]. The immersed structure can be viewed as a parametric surface X(s,t), where the force exerted by the structure on the fluid is interpolated as a source term in the momentum equation using: ż fpx, tq “

Fps, tqδpx ´ Xps, tqqds

(10)

where Fps, tq is the structural force at location s at time of t and ds is the discretized length of the immersed structure. The force is spread to the local fluid nodes through a delta function δprq, $ &

´ πr ¯¯ 1´ 1 ` cos ´2 ď r ď 2 δprq “ 4 2 % 0 otherwise

(11)

Other delta functions can also be used. Peskin listed a few delta functions in [25]. The structure velocity is interpreted from the local fluid field through: ż upX, tq “

upx, tqδpx ´ Xps, tqqdx

(12)

The same delta function Equation (11) is used to interpolate cell node velocities from the local fluid flow field. 2.4. Nanoparticle Model NPs were modeled as rigid bodies with motion governed by hydrodynamic forces and Brownian dynamics [65,66]. Langevin dynamics was used to simulate the motion of particles. m

du “ ´ςu ` Fc ` Fr dt

(13)

where ς is the friction coefficient, Fc is the conservative force and Fr is the random force that satisfies the fluctuation dissipation theorem. xFr ptqy “ 0 (14) @ D Fr ptqFr pt1 q “ 2 k B Tς δpt ´ t1 q I (15) where kB T is the thermal energy, δ(t ´ t1 ) is the Dirac delta function and I is the unit-second order tensor. F c ` Fr ςt The solution of the above equation gives uptq “ ` C expp´ q, where C is a constant. ς m When the time step dt (4.2 ˆ 10´8 s in the present study) is much larger than the relaxation m time τR “ (2.2 ˆ 10´9 s), the particle position can be updated with the terminal velocity as ς F c ` Fr uptq “ ` u f . A position check was performed to avoid the penetration of particles into ς the RBCs, e.g., the random displacement induced by thermal fluctuation will be reversed if it will jump inside cells at the next time step. In the diffusion coefficient calculation, periodic boundary conditions were applied to particles when they crossed the channel wall to remove the effect of geometric confinement. The flow chart of the algorithm for the program is shown in Figure 2, which is a second order accurate Runge–Kutta method based primarily on the midpoint rule [25]. The fluid-structure interaction model was validated for a sphere settling process [67] and a lateral migration process [68] in a viscous fluid (Figures S1–S5). Validation studies for fluid-structure interaction, cell tumbling and

Nanomaterials 2016, 6, 30

6 of 14

tank treading motions (Figure S6, parameters listed in Table S1), as well as NP diffusion (Figures S7–S8) are shown in the Supplementary Materials, which is provided in the supplementary document. Nanomaterials 2016, 6, 30

6 of 14

Figure 2. The algorithm flow chart of immersed boundary method.

Figure 2. The algorithm flow chart of immersed boundary method. 3. Model Setup and Parametric Study

3. Model Setup and Parametric Study Research has shown that particles in the core region of the vessel migrate toward the cell-free layer regions, wherethat the migration process can region be modeled diffusion [17,18].toward This migration is layer Research has shown particles in the core of theasvessel migrate the cell-free influenced by physical conditions, such as hematocrit level (Ht), cell membrane stiffness (ks), particle regions, where the migration process can be modeled as diffusion [17,18]. This migration is influenced size (r), shear rate ( η), fluid viscosity (ν) and cell size (dc). In this study, two parameters (hematocrit by physical conditions, such as hematocrit level (Ht ), cell membrane stiffness (ks ), particle size (r), shear level and shear rate) are considered, while the other parameters are kept constant [13]. NP dispersion rate (η),isfluid viscosity (ν)pure and shear cell size two parameters (hematocrit level and shear c ). In this first studied under flow(d conditions atstudy, different shear rates for a given hematocrit level. rate) areThis considered, the was other parameters areNP kept constant NP dispersion is first studied pure shear while flow setup used to explore the dispersion rate[13]. when cells undergo tumbling and tank treading [69–71]. Then, the study extended investigate NP dispersion channel under pure shear flow motion conditions at different shearis rates for to a given hematocrit level.inThis pure shear flow at different hematocrit levels, which was used to model the blood flow in capillaries with flow setup was used to explore the NP dispersion rate when cells undergo tumbling and tankatreading shear rate between 0 and 500 s−1 [72]. motion physiologically-relevant [69–71]. Then, the study is extended to investigate NP dispersion in channel flow at different In the study, RBCs were assumed to be healthy with typical physical parameters and with a size hematocrit levels, which was used to model the blood flow in capillaries with a physiologically-relevant of 8 µm. NPs were assumed to be spherical with a typical size of 100 nm. For simplicity, surface ´ 1 shear rate between 0 and 500sos that[72]. charges were neglected, NPs did not adhere to other NPs or to the RBCs. It is noted that the In NP theconcentration study, RBCswas were assumed beinhealthy typicalascertain physical parameters with a size kept relativelyto low order to with more readily the effect of RBCand motion −3 Pa·s. Through dimensional on NP dispersion. Dynamic viscosity of the fluid was fixed at 1 × 10 of 8 µm. NPs were assumed to be spherical with a typical size of 100 nm. For simplicity, surface an empiricalsofunction between diffusion coefficient, cellthe sizeRBCs. and hematocrit wasthat the chargesanalysis, were neglected, that NPs did not adhere to othershear NPsrate, or to It is noted D defined as: was , where D low is thein dispersion is the cellascertain size, η isthe theeffect shear rate and motion =kept f ( H t )relatively NP concentration order torate, moredcreadily of RBC d c2η ´ 3 on NP dispersion. Dynamic viscosity of the fluid was fixed at 1 ˆ 10 Pa¨ s. Through dimensional H is hematocrit. The same dimensionless dispersion rate was also suggested in [15]. This formula analysis, tan empirical function between diffusion coefficient, shear rate, cell size and hematocrit was was validated through simulations presented in later sections. D defined as: The “ case f pHconsisted is the dispersion rate,with dc isathe celllength, size, aη 25-µm is the width shearand ratea and Ht test ofD a rectangular fluid domain 50-µm t q, where 2 dc η lattice grid size of 0.5 µm, as shown in Figure 3A. Simulation with a finer grid of 0.25 µm showed a is hematocrit. The same dimensionless dispersion rate was also suggested in [15]. This formula was similar fluid velocity field. In the shear flow case, the top and bottom surfaces were defined as validated through simulations in right later sections. velocity boundaries, whilepresented the left and edges of the domain were modeled as periodic The test case In consisted of aflow rectangular fluid domain a 50-µm length, 25-µm width and boundaries. the channel case, a parabolic velocitywith profile was applied at athe left inlet the µm, right as outlet boundary was modeled as an openwith condition. a latticeboundary, grid sizeand of 0.5 shown in Figure 3A. Simulation a finerNon-slip grid ofboundaries 0.25 µm showed were defined along the upper Noticethe thattop the and physical systemsurfaces has to be were converted a similar fluid velocity field. In and thelower shearsurfaces. flow case, bottom defined as into LB units through reference dimensions, such as length (dx), time (dt) and mass (dm). Here, we velocity boundaries, while the left and right edges of the domain were modeled as periodic boundaries. In the channel flow case, a parabolic velocity profile was applied at the left inlet boundary, and the right outlet boundary was modeled as an open condition. Non-slip boundaries were defined along the upper and lower surfaces. Notice that the physical system has to be converted into LB units through reference

Nanomaterials 2016, 6, 30

7 of 14

dimensions, such as length (dx), time (dt) and mass (dm). Here, we select dx = 0.5 µm. The reference mass dm = ρdx3 = 1.28 ˆ 10´16 kg assuming the blood plasma density is 1025 kg/m3 . Depending on the Reynolds number, a relaxation time τ is typically recommended between 0.8 and 1.0; see [73,74]. As suggested in [44], a relaxation time τ of 1.0 was used for all simulations for computational efficiency and larger time steps. The time step dt can be determined through normalizing the fluid viscosity τ ´ 0.5 Nanomaterials dt 2016, 6, 30 7 of 14 “ ν 2 , where ν is the kinematic viscosity of the fluid. Thus, the time step can be calculated 3 dx 8 s. = 0.5 µm. The reference mass dm normalized, = ρdx3 = 1.28 × 10 kg assuming the blood density as 4.2 ˆselect 10´dx Other parameters can be as−16well, since the basic plasma reference values for 3 is 1025 kg/m . Depending on the Reynolds number, a relaxation time τ is typically recommended length, time and mass have been selected. As all of the simulations performed here have low Reynolds between 0.8 and 1.0; see [73,74]. As suggested in [44], a relaxation time τ of 1.0 was used for all numbers, the compressibility effect of the Mach number can be safely ignored. simulations for computational efficiency and larger time steps. The time step dt can be determined RBC membranes were as bi-concave τ − 0.5 dt curves with the dimensions shown in Figure 1A. through normalizing themodeled fluid viscosity = ν 2 , where ν is the kinematic viscosity of the fluid. 3 dx A single RBC was composed of 52 nodes. The cell parameters were selected based on recommended parameters canartificial be normalized, well, since kthe= 1 was Thus, the time step literature can be calculated as 4.2 10−8 s. Other values reported in the [53,75], as× listed in Table 1. The area as constraint a basic reference values for length, time and mass have been selected. As all of the simulations performed selected so that the area change was within 1%. Periodic boundary conditions (PBC) were applied to here have low Reynolds numbers, the compressibility effect of the Mach number can be safely ignored. the left and RBC rightmembranes boundaries of modeled the fluidasdomain for curves both RBCs and NPs. Theshown area ratio between were bi-concave with the dimensions in Figure 1A. RBCs and theAfluid domain defined the hematocrit level. were selected based on recommended single RBC waswas composed of as 52 nodes. The cell parameters values reported in the literature [53,75], as listed in Table 1. The artificial area constraint ka = 1 was Table 1. Cell membrane model parameters. selected so that the area change was within 1%. Periodic boundary conditions (PBC) were applied to the left and right boundaries of the fluid domain for both RBCs and NPs. The area ratio between RBCs andParameters the fluid domain was defined as the hematocritRecommended level. Specified Value Range References

Stretching coefficient ks0 5 µN/m 5~12 µN/m Table 1. Cell membrane model parameters. Bending coefficient kb 8 ˆ 10´19 J 2 ˆ 10´19 to 1 ˆ 10´17 J Parameters Stretching coefficient ks0 4. Results and Discussion Bending coefficient kb

Specified Value 5 µN/m 8 × 10−19 J

Recommended Range 5~12 µN/m 2 × 10−19 to 1 × 10−17 J

[53,75] [53,75] References [53,75] [53,75]

4.1. NP 4.Dispersion under Pure Shear Flow Results and Discussion The NP ratePure wasShear studied 4.1. NPdispersion Dispersion under Flow over a range of shear rates for a single layer of three cells. This setup was designed to eliminate the cell-cell interaction between different layers, so that we can The NP dispersion rate was studied over a range of shear rates for a single layer of three cells. focus onThis thesetup shear rate effect to oneliminate NP dispersions. ratesbetween rangingdifferent from 20 to 500 s´1 we were was designed the cell-cellShear interaction layers, so that canselected in orderfocus to cover and RBC tank treading of500 thes−1flow on theboth shearthe rateRBC effecttumbling on NP dispersions. Shear rates ranging regions from 20 to wereregime. selected For all in order to cover both tumbling and RBC tank treading regions of the flowat regime. For all shear rates investigated inthe theRBC study, the dimensionless number ηt was held 25. Therefore, the shear rates investigated in the study, the dimensionless number was held at 25. Therefore, the η t simulation time is longer for lower shear rate case. Three RBCs and 792 NPs were considered for each 1 and simulation time is of longer lower shear rate case.NPs Threeand RBCs and 792 NPs were considered each200 s´1 , simulation. Snapshots the for interaction between RBCs at shear rates of 40 s´for simulation. Snapshots of the interaction between NPs and RBCs at shear rates of 40 s−1 and 200 s−1, representative of RBC tumbling and RBC tank treading regions of the flow regime, respectively, are representative of RBC tumbling and RBC tank treading regions of the flow regime, respectively, are shown in Figure 3 (see3 the theSupplementary Supplementary Materials). shown in Figure (see corresponding the correspondingmovies movies in in the Materials). B

25 µm

A

50 µm Figure 3. Interaction between nanoparticle (NP) and red blood cell (RBC) at shear rates of (A) 40 s−1

Figure 3. Interaction between nanoparticle (NP) and red blood cell (RBC) at shear rates of (A) 40 s´1 (RBC tumbling) and (B) 200 s−1 (RBC tank treading). The bold red lines outline the RBC membranes, ´1 (RBC tumbling) and (B) 200 sdenote (RBC bold red lines outline the RBC membranes, while the green markers NPs.tank Flowtreading). streamlinesThe are shown in the background. while the green markers denote NPs. Flow streamlines are shown in the background. The mean square displacement over the y direction at different shear rates was calculated to obtain the NP dispersion rates, as shown in Figure 4. It shows that the dispersion rate is strongly The mean square displacement over the y direction −1at different shear rates was calculated to influenced by cell motion. In the RBC tumbling ( η < 40 s ) and RBC tank treading ( η > 200 s−1)

obtain the NP dispersion rates, as shown in Figure 4. It shows that the dispersion rate is strongly regions of the flow regime, the NP dispersion rate is approximately linear with the shear rate. 1 ) and RBC tank treading (η > 200 s´1 ) influenced by cell the isRBC tumbling (η 200 s´1 ,agrees the dispersion rate D can be written

should be close to thermal diffusion in the absence of shear flow. Therefore, for a given hematocrit level, and with η < 40 s−1 and η > 200 s−1,Dthe “dispersion kη ` D rate D can be written as: 0

D = kη + D0

(16)

(16)

where D0 is the thermal diffusion coefficient and k is a constant that depends on the hematocrit level. where D0 is the thermal diffusion coefficient and k is a constant that depends on the hematocrit It is noted that this formula is readily adaptable to different particle sizes, because the constant term D0 level. It is noted that this formula is readily adaptable to different particle sizes, because the constant already contains the particle size effect. The contribution of RBC motion is represented in the constant term D0 already contains the particle size effect. The contribution of RBC motion is represented in the k. It is noted that the influence of particle concentration on dispersion rate has been neglected. constant k . It is noted that the influence of particle concentration on dispersion rate has been neglected. -12

9

x 10

D (m2/s)

8 7 6 5 4

0

100

200 300 η (s-1)

400

500

Figure 4. NP dispersion rate as a function of shear rate. Error bars indicate the standard variance for

Figure 4. NP dispersion rate as a function of shear rate. Error bars indicate the standard variance for three simulations. RBCs undergo tumbling motion at a low shear rate ( η < 40 s−1) and tank treading three simulations. RBCs undergo tumbling motion at a low shear rate (η < 40 s´1 ) and tank treading motion at a high shear rate ( η > 200 s−1). In between, there is a transition region. Linear regression motion lines at a for high shear rate (η > 200 s´1 ). In between, there is a transition region. Linear regression the tumbling and tank treading regions are shown, as well. lines for the tumbling and tank treading regions are shown, as well. 4.2. NP Dispersion under Channel Flow

4.2. NP Dispersion under Channel Flow It should be noted that the pure shear flow case does not exist in physiological flow. While the shear case provides an understanding of NP dispersion enhancement under different shear flow

It should be noted that the pure shear flow case does not exist in physiological flow. While the rates, it is important to simulate NP dispersion in realistic channel flow cases. Figure 5A,B presents shear case provides andispersion understanding of NP dispersion under different shearrate flow rates, snapshots of NP in a channel flow simulationenhancement with a hematocrit of 23.5% and a shear it is important toatsimulate NP dispersion in realistic channel flow 5A,B presentsthe snapshots of 200 s−1 0.26 s and 0.46 s, respectively. The channel width is cases. 25 µm.Figure For these simulations, specified shear was measured as the shear rateaathematocrit the wall, unless noted and otherwise. Due to the of NP dispersion in arate channel flow simulation with of 23.5% a shear rate of 200 s´1 at 0.26 s and 0.46 s, respectively. The channel width is 25 µm. For these simulations, the specified

Nanomaterials 2016, 6, 30

9 of 14

shear rate was measured as the shear rate at the wall, unless noted otherwise. Due to the increased cell 2016, 30 pure shear flow simulations, the number of NPs was reduced to 378. 9 of 14The NPs volume,Nanomaterials compared to6,the were initially positioned in the core region of the channel. Since the shear rate is linearly changing increased cell volume, compared to the pure shear flow simulations, the number of NPs was reduced across the channel, RBCs not positioned exhibit distinctive such as tumbling tankrate treading, as to 378. The NPs weredid initially in the coremotions, region of the channel. Since theorshear is shown linearly in the previous pure shear flow. The higher hematocrit and cell-cell interaction also confined Nanomaterials 2016, 6, across 30 of 14 changing the channel, RBCs did not exhibit distinctive motions, such as tumbling or9tank in theThe previous pure of shear higher hematocrit andtreading cell-cell interaction also some the cell treading, motion as inshown the flow. majority theflow. cellsThe behaved like a tank motion, while increasedthe cellcell volume, to the pure shear flow simulations, the like number of NPs was reduced motion in the flow. The majority the cells due behaved a tank treading motion, cells in confined the core region werecompared bended or folded in theofchannel to the symmetry of the velocity near to 378. The NPs were initially positioned in the core region of the channel. Since the shear rate is while some cells in the core region were bended or folded in the channel due to the symmetry of the the center line of the channel. The RBC motion under other hematocrit levels and shear rates was linearly changing across line the channel, RBCs did not exhibit distinctive motions, such as tumbling tank velocity near the center of the channel. The RBC motion under other hematocrit levels andorshear similar treading, to Figure However, they are notshear shown as 5. shown in the previous pure flow.here. The higher hematocrit and cell-cell interaction also rates was similar to Figure 5. However, they are not shown here. confined the cell motion in the flow. The majority of the cells behaved like a tank treading motion, while some cells in the core region were bended or folded in the channel due to the symmetry of the velocity near the center line of the channel. The RBC motion under other hematocrit levels and shear rates was similar to Figure 5. However, they are not shown here.

(A)

(B)

Figure 5. Snapshots from a channel flow simulation for a cell-particle mixture with a hematocrit level

Figure of 5. 23.5% Snapshots channel simulation a cell-particle mixture with a hematocrit level −1 at 0.26 and a from shear arate of 200 sflow s (A) andfor 0.46 s (B). Fluid streamlines are shown in the of 23.5% and a shear rate 200 s´1 at 0.26 s (A) and 0.46nanoparticles. s (B). FluidFor streamlines shown in the background, while theofyellow markers represent 100-nm illustrationare purposes, (A) (B) background, while the yellow markers represent 100-nm illustration purposes, cells and particles are not shown to scale. A short movie nanoparticles. of the simulationFor is provided in the Figure 5. Snapshots fromshown a channel simulation for amovie cell-particle mixture with a hematocrit level in the Supplementary Materials. cells and particles are not toflow scale. A short of the simulation is provided of 23.5% and a shear rate of 200 s−1 at 0.26 s (A) and 0.46 s (B). Fluid streamlines are shown in the Supplementary Materials.

As illustrated in Figure 5A,B,markers the NPsrepresent tend to 100-nm migratenanoparticles. toward the wall. In order topurposes, characterize background, while the yellow For illustration cells and particles are the not channel, shown tothe scale. A short movie the simulation in number the the NP distribution across channel width wasofdivided into binsisofprovided 1 µm. The Supplementary Materials. Asofillustrated Figure 5A,B, theand NPs tend to toward In order characterize NPs within in each bin was counted divided by migrate the total number ofthe NPswall. to obtain the NP to fraction each bin. The NP fraction across channel width height is plotted in Figure at time of 0, number the NP within distribution across the channel, thethechannel was divided into6A bins of 1points µm. The As illustrated Figure 5A,B, shear the NPs to migrate toward the wall. InNP order to characterize −1. Figure and 0.52 s,bin forwas aincharacteristic ratetend of 200 presents the fraction of NPs 0.26 within each counted and divided bysthe total 6B number ofthe NPs tofraction obtainacross the NP the NP distribution channel, channel was into bins µm.NP The number −1. 1 channel height at t =across 0.52 sthe under shearthe rates of 100width s−1, 200 s−1divided , 300 s−1 and 500 sof The fractional within each The NPbinfraction across channel is plotted intoFigure 6ANP at time points of of NPsbin. within each wasare counted andthe divided bysample theheight total number NPs obtain the values shown in the figure the average of three usingofdifferent random seedsfraction for the ´1runs 0, 0.26 and 0.52 s, for a characteristic shear rate of 200 s . Figure 6B presents the NP fraction within each bin. The NP fraction across channel height is plotted in Figure 6A at time points of 0, across NP Brownian motion model. Figure 6Bthe clearly demonstrates that particle migration speed toward ´ 1 ´ 1 ´ 1 ´ 1 . The NP −1 0.26channel and 0.52walls s,atfortincreases a characteristic shear raterates of 200of s 100 . Figure presents across the channel height = 0.52 swith under shear s 6B , 200 s , the 300NP s fraction and 500 s the the shear rate. −1, 200 s−1, 300 s−1 and 500 s−1. The NP fractional channel height at t = 0.52 s under shear rates of 100 s fractional values shown in the figure are the average of three sample runs using different random 0.1 in the figure are the average of three sample runs using different random seeds for the values shown -1 seeds for the Figure 6B demonstrates particle migration A NP Brownian motion B clearly t = 0.52model. s 0.12 500 sthat NP Brownian motion model.t =Figure speed toward -1 0.26 s 6B clearly demonstrates that particle migration 300 s speed toward the 0.08channel walls increases with shear rate. 0.1 the channel walls increases with rate. t = 0 shear s -1 t = 0.52 s t = 0.26 s t=0s

0.04 0.08

0 0.04

5

10 15 20 Channel height(μm)

25

100 s-1

0.12 0.06

500 s-1

NP fraction

0.02 0.06

B

0.08

0.1 0.04

300 s-1 200 s-1

NP fraction

NP fraction

A

NP fraction

200 s

0.06 0.1

0.08 0.02

100 s-1

0.06 0 0.04

5

10 15 20 Channel height(μm)

25

0.02

0.02 Figure 6. The NP fraction across the channel height for a hematocrit level of 23.5%. (A) NP fraction at −1 t = 0, 0.26 0 s and 0.52 s for a shear rate of 200 s ; and (B) NP0fraction at t = 0.52 s for shear rates of 5 s−1 and 10 500 15 25 5 10 15 20 25 s−1. μ 20 100 s−1, 200 s−1, 300 Channel height( m)

Channel height(μm)

In order to characterize NP migration thea dispersion rate of was calculated from the Figure 6. The NP fraction across the channelspeed, height for hematocrit level 23.5%. (A) NP fraction at NP

Figure 6. The NP fraction across the channel height for a hematocrit level of 23.5%. (A) NP fraction and (B) NP fraction at t =levels, 0.52 s and for shear rates ofshear = 0, 0.26displacement. s and 0.52 s forThe a shear rate of 200 s−1;for meant square dispersion rates different hematocrit at various ´1 ; and (B) NP fraction at t = 0.52 s for shear rates of at t = 0,100 0.26 s and 0.52 s for a shear rate of 200 s s−1, 200 s−1, 300 s−1 and 500 s−1. 100 s´1 , 200 s´1 , 300 s´1 and 500 s´1 . In order to characterize NP migration speed, the dispersion rate was calculated from the NP mean square displacement. The dispersion rates for different hematocrit levels, and at various shear

In order to characterize NP migration speed, the dispersion rate was calculated from the NP mean square displacement. The dispersion rates for different hematocrit levels, and at various shear rates,

Nanomaterials 2016, 6, 30

10 of 14

Nanomaterials 2016, 6, 30

10 of 14

are shown in Figure 7A. From the pure shear simulation results shown in Equation (14), a modified rates, are shown in Figure 7A. developed, From the pure shear simulation results shown in Equation (14), a dimensionless dispersion rate was modified dimensionless dispersion rate was developed,

D ´ D0 f pH q Dr D“r = D −2 D0 =“f ( H t) t ddcc2ηη

(17)

(17)

The dimensionless dispersion rate rate Dr isDplotted in Figure 7B, where the error bars show standard The dimensionless dispersion r is plotted in Figure 7B, where the error bars show variancestandard for threevariance sample forruns. three sample runs. -11

6

Ht 23.5% Ht 20.2% Ht 13.4% Ht 10.1%

B

x 10

2

1.5

4

2

0 0

2.5

Dr

Dispersion rate (m2/s)

A

-3

x 10

Dimensionless

8

100

200

300 400 Shear (s-1)

500

600

1

0.5

0.1

0.15

0.2

0.25

Hematocrit (Ht)

Figure 7. Dispersion rate of NPs in blood flow. (A) NP dispersion rate at different hematocrit (Ht) and

Figure 7.shear Dispersion rate of NPs in blood flow. (A) NP dispersion rate at different hematocrit (Ht) and rates; (B) relationship between dimensionless dispersion rate (Dr) and hematocrit (Ht). Error shear rates; (B) relationship dimensionless dispersion rate (Dr ) and hematocrit (Ht ). Error bars bars show the standardbetween variance from three samples. show the standard variance from three samples. Figure 7 shows that the lateral dispersion of NPs (i.e., migration of NPs toward the vessel walls) is much larger than what is predicted by thermal diffusion alone. This migration is influenced by Figure 7 shows that the lateral dispersion of NPs (i.e., migration of NPs toward the vessel walls) is both the hematocrit level and the shear rate. While the relationship between dispersion rate and shear much larger what is predicted by 7A), thermal diffusion between alone. This migration is influenced rate isthan approximately linear (Figure the relationship dispersion and hematocrit is not by both the hematocrit level and 7B). the Nevertheless, shear rate. While the between dispersion andas: shear rate fully linear (Figure a best fit linerelationship with reasonable approximation can berate written

is approximately linear (Figure 7A), the between dispersion and hematocrit is not fully D −relationship D0 Dr = = 4.643 × 10 −3 H t +5.834 × 10 −4 (18) 2 linear (Figure 7B). Nevertheless, a best fit d c ηline with reasonable approximation can be written as: Particle dispersion rate predictions (18) were also compared to dispersion rates D ´ D0 from Equation ´4 “ 4.643 ˆ1-µm 10´3particles Ht ` 5.834 ˆis10 (18) Dr “ published in the literature for platelets [15,76] and [13]. It noted that the comparison 2 dc η of Equation (18) with platelets and microparticles was considered due to the lack of the NP dispersion rate in the literature. The predictions and measured dispersion rates are summarized in Table 2. The Particle dispersion rate predictions from Equation (18) were also compared to dispersion rates dimensionless dispersion rate is also plotted in Figure 8. As shown in both the table and the figure, published in the literature for platelets [15,76] and 1-µm particles [13]. It is noted that the comparison the dispersion rate predictions using Equation (18) are in good agreement with measured rates of Equation (18) in with microparticles considered to thedispersion lack of the NP dispersion reported the platelets literature. and Discrepancies betweenwas the predicted and due measured rates may be due to the linear correlation assumption for hematocrit and/or the effect particle concentration, rate in the literature. The predictions and measured dispersion ratesofare summarized in Table 2. which was very low for this study and was assumed to have a negligible effect on NP dispersion. The dimensionless dispersion rate is also plotted in Figure 8. As shown in both the table and the Another possible reason is that platelets generally have on average an aspect ratio of 1.4, and this may be figure, the dispersion rate predictions using Equation (18) are in good agreement with measured rates a confounding factor and contribute to the differences [77,78]. Nevertheless, the order of magnitude of the reporteddispersion in the literature. betweenare thecorrectly predicted andby measured dispersion rates may rate, as wellDiscrepancies as its trend with hematocrit predicted Equation (18).

be due to the linear correlation assumption for hematocrit and/or the effect of particle concentration, Tablelow 2. Comparison particleand dispersion predictions Equation (18) with dispersion which was very for thisofstudy was rate assumed to from have a negligible effect onrates NP dispersion. reported in the literature. Another possible reason is that platelets generally have on average an aspect ratio of 1.4, and this may Htfactor and Shear (s−1) Dispersion Rate (cm2/s)[77,78]. Prediction (cm2/s) theReference be a confounding contribute to the differences Nevertheless, order of magnitude −6 [0.2, rate, 0.4] as well400 0.68] × 10 [0.39, 0.63] predicted × 10−6 of the dispersion as its trend[0.5, with hematocrit are correctly by[76] Equation (18). [0.2, 0.4] 1100 [1.5, 2.1] × 10−6 [1.1, 1.7] × 10−6 [76] −9 −9 [0.1, 0.15, 0.2] 44.8 [8.2, 11.9, 17.2] × 10 [31.3, 37.9, 44.6] × 10 [13] Table 2. Comparison of particle dispersion rate predictions from Equation (18) with dispersion rates [0.1, 0.2] 804 [0.9, 1.4] × 10−7 [5.4, 7.8] × 10−7 [15]

reported in the literature. Ht

Shear (s´1 )

Dispersion Rate (cm2 /s)

Prediction (cm2 /s)

Reference

[0.2, 0.4] [0.2, 0.4] [0.1, 0.15, 0.2] [0.1, 0.2]

400 1100 44.8 804

[0.5, 0.68] ˆ 10´6 [1.5, 2.1] ˆ 10´6 [8.2, 11.9, 17.2] ˆ 10´9 [0.9, 1.4] ˆ 10´7

[0.39, 0.63] ˆ 10´6 [1.1, 1.7] ˆ 10´6 [31.3, 37.9, 44.6] ˆ 10´9 [5.4, 7.8] ˆ 10´7

[76] [76] [13] [15]

Nanomaterials 2016, 6, 30

11 of 14

Nanomaterials 2016, 6, 30

11 of 14 -3

Dimensionless

Dr

4

x 10

3

Prediction Saadatmand,2011 Zhao,2012 Crowl,2011 Simulation

2

1

0 0

0.1

0.2

0.3

0.4

0.5

Hematocrit (Ht) Figure 8. Comparison of the particle dispersion rate predicted from Equation (18) with the data

Figure 8. Comparison of the particle dispersion rate predicted from Equation (18) with the data reported in the literature. The dashed line is the prediction from Equation (18). reported in the literature. The dashed line is the prediction from Equation (18).

5. Conclusion and Future Work

5. Conclusion and Future Work

This paper presents a numerical study on NP dispersion in blood flow considering the influence of RBC motion and deformation. which did not exist from previous modeling This paper presents a numericalAnalytical study onformulae, NP dispersion in blood flow considering the influence results [3,16], were proposed to characterize NP migration to the vessel walls as a function of shear of RBC motion and deformation. Analytical formulae, which did not exist from previous modeling rate at different hematocrit levels. The formula’s predictions agree well with data reported the rate results [3,16], were proposed to characterize NP migration to the vessel walls as a function ofinshear literature, thereby facilitating practical modeling of NP transport and distribution in large-scale at different hematocrit levels. The formula’s predictions agree well with data reported in the literature, vascular systems. thereby facilitating practical modeling of NP transport and distribution in large-scale vascular systems. In the future, the model will be extended to 3D to explore NP binding and distribution in In the future, the The model will beformula extended 3Ddispersion to explorerate NP will binding distribution in capillary capillary vessels. proposed fortothe also and be used to evaluate NP vessels. The proposed formula for the dispersion rate will also be used to evaluate NP transport and transport and distribution in a large-scale vascular network.

distribution in a large-scale vascular network. Supplementary Materials: Figures S1–S9, Table S1 and three short movies are available online at

Supplementary Materials: Figures S1–S9, Table S1 and three short movies are available online at http://www.mdpi.com/2079-4991/6/2/30/s1. http://www.mdpi.com/2079-4991/6/2/30/s1. Acknowledgments: The study presented in this paper was supported by the National Science Foundation (NSF)

Acknowledgments: The study presented in this(CAREER) paper was supported by the National Science Foundation Faculty Early Career Development Program Grant CBET-1113040, NSF CBET-1067502, NSF (NSF)DMS-1516236 Faculty Early Career Development Program (CAREER) Grant CBET-1113040, NSF CBET-1067502, NSF and the National Institute of Health (NIH) Grant EB015105. DMS-1516236 and the National Institute of Health (NIH) Grant EB015105. Author Contributions: J.T. developed the code, performed the simulation and wrote the manuscript. W.K.

Author Contributions: J.T. developed performed the and simulation wrote manuscript. developed the code, analyzed the data.the S.S.code, contributed to the code analyzedand the data. J.Y.the helped design the W.K. developed the code, analyzed the data. S.S. contributed to the code and analyzed the data. J.Y. helped simulation. Y.L. conceived, designed the simulation and analyzed the data. All revised the manuscript. design the simulation. Y.L. conceived, designed the simulation and analyzed the data. All revised the manuscript. Conflicts of Interest: The authors declare no conflict of interest.

Conflicts of Interest: The authors declare no conflict of interest.

References

References 1.

1. 2.

3. 4. 5.

Sanhai, W.R.; Sakamoto, J.H.; Canady, R; Ferrari, M. Seven Challenges for Nanomedicine. Nat. Nano 2008,

Sanhai, W.R.; Sakamoto, J.H.; Canady, R.; Ferrari, M. Seven Challenges for Nanomedicine. Nat. Nano 2008, 3, 3, 242–244. 242–244. 2. Liu, [CrossRef] J.; Bradley,[PubMed] R.; Eckmann, D.M.; Ayyaswamy, P.S.; Radhakrishnan, R. Multiscale Modeling of Nanocarriers Drug Delivery.P.S.; Curr.Radhakrishnan, Nanosci. 2011, 7, 727–735. Liu, Functionalized J.; Bradley, R.; Eckmann,in Targeted D.M.; Ayyaswamy, R. Multiscale Modeling of 3. Muller, K.; Fedosov, D.A.; Gompper, G. Margination of microandNanosci. nano-particles flow and its Functionalized Nanocarriers in Targeted Drug Delivery. Curr. 2011,in7,blood 727–735. [CrossRef] effect on drug delivery. Sci. Rep. 2014, 4, doi:10.1038/srep04871. [PubMed] 4. Tan, J.; Thomas, A.; Liu, Y. Influence of red blood cells on nanoparticle targeted delivery in Muller, K.; Fedosov, D.A.; Gompper, G. Margination of micro- and nano-particles in blood flow and its effect microcirculation. Soft Matter 2012, 8, 1934–1946. on drug delivery. Sci. Rep. 2014, 4. [CrossRef] [PubMed] 5. Doshi, N.; Prabhakarpandian, B.; Rea-Ramsey, A.; Pant, K.; Sundaram, S.; Mitragotri, S. Flow and adhesion Tan, J.; Thomas, A.; Liu, Y. Influence of red blood cells on nanoparticle targeted delivery in microcirculation. of drug carriers in blood vessels depend on their shape: A study using model synthetic microvascular Soft Matter 2012, 8, 1934–1946. [CrossRef] [PubMed] networks. J. Control. Release 2010, 146, 196–200. Doshi, N.; Prabhakarpandian, B.; Rea-Ramsey, A.; Pant, K.; Sundaram, S.; Mitragotri, S. Flow and adhesion of drug carriers in blood vessels depend on their shape: A study using model synthetic microvascular networks. J. Control. Release 2010, 146, 196–200. [CrossRef] [PubMed]

Nanomaterials 2016, 6, 30

6. 7. 8. 9. 10. 11.

12. 13.

14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28.

29. 30.

12 of 14

Tan, J.; Shah, S.; Thomas, A.; Ou-Yang, H.D.; Liu, Y. The influence of size, shape and vessel geometry on nanoparticle distribution. Microfluid. Nanofluid. 2013, 14, 77–87. [CrossRef] [PubMed] Tan, J.; Wang, S.; Yang, J.; Liu, Y. Coupled particulate and continuum model for nanoparticle targeted delivery. Comput. Struct. 2013, 122, 128–134. [CrossRef] [PubMed] Geng, Y.; Dalhaimer, P.; Cai, S.; Tsai, R.; Tewari, M.; Minko, T.; Discher, D.E. Shape effects of filaments versus spherical particles in flow and drug delivery. Nat. Nanotechnol. 2007, 2, 249–255. [CrossRef] [PubMed] Liu, Y.L.; Shah, S.; Tan, J. Computational Modeling of Nanoparticle Targeted Drug Delivery. Rev. Nanosci. Nanotechnol. 2010, 1, 66–83. [CrossRef] Wang, S.; Zhou, Y.; Tan, J.; Xu, J.; Yang, J.; Liu, Y. Computational modeling of magnetic nanoparticle targeting to stent surface under high gradient field. Comput. Mech. 2014, 53, 403–412. [CrossRef] [PubMed] Guduru, R.; Liang, P.; Runowicz, C.; Nair, M.; Atluri, V.; Khizroev, S. Magneto-electric Nanoparticles to Enable Field-controlled High-Specificity Drug Delivery to Eradicate Ovarian Cancer Cells. Sci. Rep. 2013, 3. [CrossRef] [PubMed] Reasor, D.A.; Mehrabadi, M.; Ku, D.N.; Aidun, C.K. Determination of Critical Parameters in Platelet Margination. Ann. Biomed. Eng. 2013, 41, 238–249. [CrossRef] [PubMed] Saadatmand, M.; Ishikawa, T.; Matsuki, N.; Abdekhodaie, M.J.; Imai, Y.; Ueno, H.; Yamaguchi, T. Fluid particle diffusion through high-hematocrit blood flow within a capillary tube. J. Biomech. 2011, 44, 170–175. [CrossRef] [PubMed] Crowl, L.M.; Fogelson, A.L. Computational model of whole blood exhibiting lateral platelet motion induced by red blood cells. Int. J. Numer. Methods Biomed. Eng. 2010, 26, 471–487. [CrossRef] [PubMed] Zhao, H.; Shaqfeh, E.S.G.; Narsimhan, V. Shear-induced particle migration and margination in a cellular suspension. Phys. Fluids 2012, 24. [CrossRef] Lee, T.-R.; Choi, M.; Kopacz, A.M.; Yun, S.-H.; Liu, W.K.; Decuzzi, P. On the near-wall accumulation of injectable particles in the microcirculation: smaller is not better. Sci. Rep. 2013, 3. [CrossRef] [PubMed] Eckstein, E.C.; Bailey, D.G.; Shapiro, A.H. Self-Diffusion of Particles in Shear-Flow of a Suspension. J. Fluid Mech. 1977, 79, 191–208. [CrossRef] Leighton, D.; Acrivos, A. The Shear-Induced Migration of Particles in Concentrated Suspensions. J. Fluid Mech. 1987, 181, 415–439. [CrossRef] Aarts, P.A.M.M.; Steendijk, P.; Sixma, J.J.; Heethaar, R.M. Fluid shear as a possible mechanism for platelet diffusivity in flowing blood. J. Biomech. 1986, 19, 799–805. [CrossRef] Gentile, F.; Ferrari, M.; Decuzzi, P. The Transport of Nanoparticles in Blood Vessels: The Effect of Vessel Permeability and Blood Rheology. Ann. Biomed. Eng. 2008, 36, 254–261. [CrossRef] [PubMed] Gentile, F.; Decuzzi, P. Time dependent dispersion of nanoparticles in blood vessels. J. Biomed. Sci. Eng. 2010, 3, 517–524. [CrossRef] Hossain, S.S.; Zhang, Y.J.; Liang, X.H.; Hussain, F.; Ferrari, M.; Hughes, T.J.R.; Decuzzi, P. In silico vascular modeling for personalized nanoparticle delivery. Nanomedicine 2013, 8, 343–357. [CrossRef] [PubMed] Kruger, T.; Gross, M.; Raabe, D.; Varnik, F. Crossover from tumbling to tank-treading-like motion in dense simulated suspensions of red blood cells. Soft Matter 2013, 9, 9008–9015. [CrossRef] [PubMed] Zhang, J.F.; Johnson, P.C.; Popel, A.S. Red blood cell aggregation and dissociation in shear flows simulated by lattice Boltzmann method. J. Biomech. 2008, 41, 47–55. [CrossRef] [PubMed] Peskin, C.S. The immersed boundary method. Acta Numer. 2002, 11, 479–517. [CrossRef] Liu, Y.L.; Liu, W.K. Rheology of red blood cell aggregation by computer simulation. J. Comput. Phys. 2006, 220, 139–154. [CrossRef] Liu, Y.; Liu, W.K.; Belytschko, T.; Patankar, N.; To, A.C.; Kopacz, A.; Chung, J.H. Immersed electrokinetic finite element method. Int. J. Numer. Methods Eng. 2007, 71, 379–405. [CrossRef] Liu, Y.; Oh, K.; Bai, J.G.; Chang, C.; Chung, J.; Lee, K.; Liu, W.K. Manipulation of Nanoparticles and Biomolecules by Electric Field and Surface Tension. Comput. Methods Appl. Mech. Eng. 2008, 197, 2156–2172. [CrossRef] Liu, Y.; Zhang, L.T.; Wang, X.; Liu, W.K. Coupling of Navier-Stokes Equations with Protein Molecular Dynamics and Its Application to Hemodynamics. Int. J. Numer. Methods Fluids 2004, 46. [CrossRef] Liu, Y.L.; Chung, J.H.; Liu, W.K.; Ruoff, R.S. Dielectrophoretic assembly of nanowires. J. Phys. Chem. B 2006, 110, 14098–14106. [CrossRef] [PubMed]

Nanomaterials 2016, 6, 30

31. 32. 33. 34.

35. 36. 37. 38.

39. 40. 41. 42.

43. 44. 45. 46. 47. 48. 49. 50. 51. 52. 53. 54. 55.

13 of 14

Wu, J.; Shu, C.; Zhao, N.; Yan, W.W. Fluid Dynamics of Flapping Insect Wing in Ground Effect. J. Bionic Eng. 2014, 11, 52–60. [CrossRef] Wu, J.; Yang, S.C.; Shu, C.; Zhao, N.; Yan, W.W. Ground effect on the power extraction performance of a flapping wing biomimetic energy generator. J. Fluids Struct. 2015, 54, 247–262. [CrossRef] Suzuki, K.; Minami, K.; Inamuro, T. Lift and thrust generation by a butterfly-like flapping wing-body model: Immersed boundary-lattice Boltzmann simulations. J. Fluid Mech. 2015, 767, 659–695. [CrossRef] De Rosis, A.; Leveque, E. Harmonic oscillations of a thin lamina in a quiescent viscous fluid: A numerical investigation within the framework of the lattice Boltzmann method. Comput. Struct. 2015, 157, 209–217. [CrossRef] Wang, Y.; Shu, C.; Teo, C.J.; Wu, J. An immersed boundary-lattice Boltzmann flux solver and its applications to fluid-structure interaction problems. J. Fluids Struct. 2015, 54, 440–465. [CrossRef] Tölke, J. Implementation of a Lattice Boltzmann kernel using the Compute Unified Device Architecture developed by nVIDIA. Comput. Vis. Sci. 2010, 13, 29–39. [CrossRef] Takashi, N.; Hughes, T.J. An arbitrary Lagrangian-Eulerian finite element method for interaction of fluid and a rigid body. Comput. Methods Appl. Mech. Eng. 1992, 95, 115–138. [CrossRef] Kruger, T.; Varnik, F.; Raabe, D. Efficient and accurate simulations of deformable particles immersed in a fluid using a combined immersed boundary lattice Boltzmann finite element method. Comput. Math. Appl. 2011, 61, 3485–3505. [CrossRef] Sui, Y.; Chew, Y.T.; Roy, P.; Low, H.T. A hybrid method to study flow-induced deformation of three-dimensional capsules. J. Comput. Phys. 2008, 227, 6351–6371. [CrossRef] Reasor, D.A.; Clausen, J.R.; Aidun, C.K. Coupling the lattice-Boltzmann and spectrin-link methods for the direct numerical simulation of cellular blood flow. Int. J. Numer. Methods Fluids 2012, 68, 767–781. [CrossRef] Chen, S.; Doolen, G.D. Lattice boltzmann method for fluid flows. Annu. Rev. Fluid Mech. 1998, 30, 329–364. [CrossRef] Luo, L.S. The lattice-gas and lattice boltzmann methods: Past, present, and future. In Proceedings of the International Conference on Applied Computational Fluid Dynamics, Beijing, China, 17–20 October 2000; pp. 52–83. Succi, S. The Lattice Boltzmann Equation: For Fluid Dynamics and Beyond; Oxford University Press: New York, NY, USA, 2001. Ladd, A.J.C. Numerical Simulations of Particulate Suspensions via a Discretized Boltzmann-Equation. 1. Theoretical Foundation. J. Fluid Mech. 1994, 271, 285–309. [CrossRef] Latt, J. Hydrodynamic Limit of Lattice Boltzmann Equations; University of Geneva: Geneva, Switzerland, 2007. Qian, Y.; D’Humières, D.; Lallemand, P. Lattice BGK models for Navier-Stokes equation. EPL 1992, 17. [CrossRef] Zou, Q.S.; He, X.Y. On pressure and velocity boundary conditions for the lattice Boltzmann BGK model. Phys. Fluids 1997, 9, 1591–1598. [CrossRef] Latt, J.; Chopard, B. Lattice Boltzmann method with regularized pre-collision distribution functions. Math. Comput. Simul. 2006, 72, 165–168. [CrossRef] Latt, J.; Chopard, B.; Malaspinas, O.; Deville, M.; Michler, A. Straight velocity boundaries in the lattice Boltzmann method. Phys. Rev. E 2008, 77. [CrossRef] [PubMed] Pozrikidis, C. Effect of membrane bending stiffness on the deformation of capsules in simple shear flow. J. Fluid Mech. 2001, 440, 269–291. [CrossRef] Skalak, R.; Tozeren, A.; Zarda, R.P.; Chien, S. Strain Energy Function of Red Blood-Cell Membranes. Biophys. J. 1973, 13, 245–280. [CrossRef] Kruger, T. Computer Simulation Study of Collective Phenomena in Dense Suspensions of Red Blood Cells under Shear; Springer Science & Business Media: Berlin, Germany, 2012. Dao, M.; Li, J.; Suresh, S. Molecularly based analysis of deformation of spectrin network and human erythrocyte. Mater. Sci. Eng. C 2006, 26, 1232–1244. [CrossRef] Fedosov, D.A.; Lei, H.; Caswell, B.; Suresh, S.; Karniadakis, G.E. Multiscale Modeling of Red Blood Cell Mechanics and Blood Flow in Malaria. PLoS Comput. Biol. 2011, 7. [CrossRef] [PubMed] Nakamura, M.; Bessho, S.; Wada, S. Spring-network-based model of a red blood cell for simulating mesoscopic blood flow. Int. J. Numer. Methods Biomed. Eng. 2013, 29, 114–128. [CrossRef] [PubMed]

Nanomaterials 2016, 6, 30

56. 57. 58. 59. 60. 61. 62. 63. 64.

65. 66.

67. 68. 69. 70. 71. 72. 73. 74. 75. 76. 77. 78.

14 of 14

Fedosov, D.A.; Caswell, B.; Karniadakis, G.E. A Multiscale Red Blood Cell Model with Accurate Mechanics, Rheology, and Dynamics. Biophys. J. 2010, 98, 2215–2225. [CrossRef] [PubMed] Nakamura, M.; Bessho, S.; Wada, S. Analysis of Red Blood Cell Deformation under Fast Shear Flow for Better Estimation of Hemolysis. Int. J. Numer. Methods Biomed. Eng. 2014, 30, 42–54. [CrossRef] [PubMed] Evans, E.; Fung, Y.-C. Improved measurements of the erythrocyte geometry. Microvasc. Res. 1972, 4, 335–347. [CrossRef] Peskin, C.S.; McQueen, D.M. A three-dimensional computational method for blood flow in the heart I. Immersed elastic fibers in a viscous incompressible fluid. J. Comput. Phys. 1989, 81, 372–405. [CrossRef] Herschlag, G.; Miller, L. Reynolds number limits for jet propulsion: A numerical study of simplified jellyfish. J. Theor. Biol. 2011, 285, 84–95. [CrossRef] [PubMed] Mittal, R.; Iaccarino, G. Immersed Boundary Methods. Annu. Rev. Fluid Mech. 2005, 37, 239–261. [CrossRef] Ladd, A.J. Numerical simulations of particulate suspensions via a discretized Boltzmann equation. Part 2. Numerical results. J. Fluid Mech. 1994, 271, 311–339. [CrossRef] Aidun, C.K.; Lu, Y.N.; Ding, E.J. Direct analysis of particulate suspensions with inertia using the discrete Boltzmann equation. J. Fluid Mech. 1998, 373, 287–311. [CrossRef] Basagaoglu, H.; Allwein, S.; Succi, S.; Dixon, H.; Carrola, J.T.; Stothoff, S. Two- and three-dimensional lattice Boltzmann simulations of particle migration in microchannels. Microfluid. Nanofluid. 2013, 15, 785–796. [CrossRef] Branka, ´ A.; Heyes, D. Algorithms for Brownian dynamics simulation. Phys. Rev. E 1998, 58. [CrossRef] Radhakrishnan, R.; Uma, B.; Liu, J.; Ayyaswamy, P.S.; Eckmann, D.M. Temporal multiscale approach for nanocarrier motion with simultaneous adhesion and hydrodynamic interactions in targeted drug delivery. J. Comput. Phys. 2013, 244, 252–263. [CrossRef] [PubMed] Zhang, L.; Gerstenberger, A.; Wang, X.D.; Liu, W.K. Immersed finite element method. Comput. Methods Appl. Mech. Eng. 2004, 193, 2051–2067. [CrossRef] Feng, J.; Hu, H.H.; Joseph, D.D. Direct Simulation of Initial-Value Problems for the Motion of Solid Bodies in a Newtonian Fluid. 1. Sedimentation. J. Fluid Mech. 1994, 261, 95–134. [CrossRef] Fischer, T.; Stohr-Lissen, M.; Schmid-Schonbein, H. The red cell as a fluid droplet: tank tread-like motion of the human erythrocyte membrane in shear flow. Science 1978, 202, 894–896. [CrossRef] [PubMed] Sui, Y.; Low, H.T.; Chew, Y.T.; Roy, P. Tank-treading, swinging, and tumbling of liquid-filled elastic capsules in shear flow. Phys. Rev. E 2008, 77. [CrossRef] [PubMed] Pozrikidis, C. Numerical simulation of the flow-induced deformation of red blood cells. Ann. Biomed. Eng. 2003, 31, 1194–1205. [CrossRef] [PubMed] Decuzzi, P.; Ferrari, M. The adhesive strength of non-spherical particles mediated by specific interactions. Biomaterials 2006, 27, 5307–5314. [CrossRef] [PubMed] Kruger, T.; Varnik, F.; Raabe, D. Shear stress in lattice Boltzmann simulations. Phys. Rev. E 2009, 79. [CrossRef] [PubMed] Holdych, D.J.; Noble, D.R.; Georgiadis, J.G.; Buckius, R.O. Truncation error analysis of lattice Boltzmann methods. J. Comput. Phys. 2004, 193, 595–619. [CrossRef] Fedosov, D.A. Multiscale modeling of Blood flow and soft matter. In Applied Mathematics; Brown University: Providence, RI, USA, 2010. Crowl, L.; Fogelson, A.L. Analysis of mechanisms for platelet near-wall excess under arterial blood flow conditions. J. Fluid Mech. 2011, 676, 348–375. [CrossRef] Wang, W.W.; King, M.R. Multiscale Modeling of Platelet Adhesion and Thrombus Growth. Ann. Biomed. Eng. 2012, 40, 2345–2354. [CrossRef] [PubMed] Mody, N.A.; King, M.R. Platelet adhesive dynamics. Part I: Characterization of platelet hydrodynamic collisions and wall effects. Biophys. J. 2008, 95, 2539–2555. [CrossRef] [PubMed] © 2016 by the authors; licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons by Attribution (CC-BY) license (http://creativecommons.org/licenses/by/4.0/).

Characterization of Nanoparticle Dispersion in Red Blood Cell Suspension by the Lattice Boltzmann-Immersed Boundary Method.

Nanodrug-carrier delivery in the blood stream is strongly influenced by nanoparticle (NP) dispersion. This paper presents a numerical study on NP tran...
2MB Sizes 1 Downloads 21 Views