Modeling of Biomechanics and Biorheology of Red Blood Cells in Type 2 Diabetes Mellitus

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Article

Modeling of Biomechanics and Biorheology of


Red Blood Cells in Type 2 Diabetes Mellitus
Hung-Yu Chang,1 Xuejin Li,1,* and George Em Karniadakis1,*
1
Division of Applied Mathematics, Brown University, Providence, Rhode Island

ABSTRACT Erythrocytes in patients with type-2 diabetes mellitus (T2DM) are associated with reduced cell deformability and
elevated blood viscosity, which contribute to impaired blood flow and other pathophysiological aspects of diabetes-related
vascular complications. In this study, by using a two-component red blood cell (RBC) model and systematic parameter variation,
we perform detailed computational simulations to probe the alteration of the biomechanical, rheological, and dynamic behavior
of T2DM RBCs in response to morphological change and membrane stiffening. First, we examine the elastic response of T2DM
RBCs subject to static tensile forcing and their viscoelastic relaxation response upon release of the stretching force. Second, we
investigate the membrane fluctuations of T2DM RBCs and explore the effect of cell shape on the fluctuation amplitudes. Third,
we subject the T2DM RBCs to shear flow and probe the effects of cell shape and effective membrane viscosity on their
tank-treading movement. In addition, we model the cell dynamic behavior in a microfluidic channel with constriction and quantify
the biorheological properties of individual T2DM RBCs. Finally, we simulate T2DM RBC suspensions under shear and compare
the predicted viscosity with experimental measurements. Taken together, these simulation results and their comparison
with currently available experimental data are helpful in identifying a specific parametric model—the first of its kind, to our
knowledge—that best describes the main hallmarks of T2DM RBCs, which can be used in future simulation studies of hemato-
logic complications of T2DM patients.

INTRODUCTION
Diabetes mellitus (DM), the fastest growing chronic disease elastic nature of the cytoskeleton, the RBC is capable of
worldwide, is a metabolic dysfunction characterized by dramatic deformations when passing through narrow capil-
elevated blood glucose levels (hyperglycemia) (1). Type 2 laries as small as 3 mm in diameter without any damage.
DM (T2DM) is the most common form of diabetes. People For RBCs in pathological conditions, the alterations in
with T2DM demonstrate significantly higher mortality rates cell geometry and membrane properties of diseased RBCs
relative to nondiabetics due to an increased risk of developing could lead to impaired functionality including loss of de-
macro- and microvascular complications (2). Hemorheolog- formability (5,6). For example, the shape distortion and
ical abnormalities that emerge in diabetic patients play a key membrane stiffening of RBCs induced by parasitic infec-
role in the pathogenesis and progression of life-threatening tious diseases like malaria (7,8) and certain genetic blood
coronary and peripheral artery diseases. One of the hemor- disorders like sickle cell disease (9) cause increased cell
heological determinants is the impaired deformability of rigidity and decreased cell deformability. Impairment of
red blood cells (RBCs) involved in T2DM (3). A healthy RBC deformability has also been demonstrated in T2DM.
human RBC is a nucleus-free cell; it is primarily comprised For example, using micropipette aspiration and filtration
of a fluid-like lipid bilayer contributing to the bending resis- techniques, McMillan et al. (10) and Kowluru et al. (11)
tance, an attached spectrin network (cytoskeleton) maintain- showed that T2DM RBCs are less deformable and more
ing cell shape and facilitating its motion, and transmembrane fragile compared to nondiabetic subjects. Agrawal et al.
proteins bridging the connections between lipid and spectrin (12) found that the size of a T2DM RBC is larger than
domains (4). Owing to the fluid nature of the lipid bilayer and that of a normal RBC because of the possible metabolic
disturbances. Babu and Singh (13) showed that the develop-
ment of irregularity in the contour of the T2DM RBCs under
Submitted May 1, 2017, and accepted for publication June 12, 2017. hyperglycemia would cause a significant reduction in cell
*Correspondence: xuejin_li@brown.edu or george_karniadakis@brown. deformability. In addition, several studies using atomic
edu force microscopy (AFM) directly examined the biomechan-
Editor: Vivek Shenoy. ical properties of diabetic RBCs and demonstrated that they
http://dx.doi.org/10.1016/j.bpj.2017.06.015
 2017 Biophysical Society.

Biophysical Journal 113, 481–490, July 25, 2017 481


Chang et al.

are less deformable than normal RBCs (Fig. 1), which could spectrin network (38–44); two-component RBC models
be due to the oxidation and glycosylation of hemoglobin and treat the lipid bilayer and the cytoskeleton as two distinct
proteins on cell membrane under abnormal glycemic condi- components but also explicitly incorporate bilayer-cytoskel-
tions (14–16). etal interactions. Therefore, two-component RBC models
RBC deformability is an important hemorheological are capable of modeling the biophysics of RBCs arising
parameter in determining whole-blood viscosity and, hence, from the interactions between the lipid bilayer and the cyto-
blood-flow resistance in the microcirculation (17–19). skeletal network (37). Such models have been successfully
Increased blood viscosity has been demonstrated in patients applied to quantify molecular-level mechanical forces
with T2DM, inducing insufficient blood supply and vascular involved in bilayer-cytoskeletal dissociation (37,45), predict
damage, and eventually leading to diabetic microangiopathy mechanical properties of defective RBCs in hereditary
and other circulation problems. For example, Skovborg spherocytosis with different cytoskeleton connectivity
et al. (17) revealed that the blood viscosity of diabetic (46), and discover new mechanisms responsible for the
subjects is 20% higher than that of controls. Ercan et al. stiffening of malaria-infected RBCs (8,46).
(20) suggested that elevated plasma cholesterol contributes In this study, we extend the two-component whole-cell
to increased blood viscosity by an additional effect of hyper- model to T2DM RBCs and investigate the morphological
glycemia in T2DM patients. In long-standing DM with and biomechanical characteristics of T2DM RBCs and
nonproliferative retinopathy, Turczy nski et al. (21) showed hemorheological properties of T2DM blood. Specifically,
that blood viscosity is positively correlated with retinopathy we test and validate the whole-cell model through rigorous
severity, which is attributed to the decrease of RBC comparisons with experimental data from five different sets
deformability. of independent measurements that probe different aspects of
Along with the aforementioned experimental studies, biomechanical and rheological properties of T2DM RBCs,
recent advances in computational modeling and simulation including RBC stretching deformation and shape relaxation,
enable investigation of a broad range of biomechanical membrane fluctuations, RBC dynamics in shear flow, and
and rheological blood-related problems at different length blood viscosity in T2DM. The rest of the article includes
scales (22–29). In recent years, mesoscopic-particle-based a brief description of the development of diabetic RBC
RBC models, which treat both the fluid and the membrane models, detailed simulation results and discussion, and a
as particulate materials, are increasingly popular as a summary of major findings and conclusion.
promising tool for modeling the structural, mechanical,
and rheological properties of RBCs in normal and patho-
logical conditions (30–33). Two different particle-based METHODS
RBC models using dissipative particle dynamics (DPD) In this study, we probe the biomechanics, rheology, and dynamics of T2DM
(34), i.e., one-component RBC models (35,36) and two- RBCs with the help of the two-component RBC model based on the
component RBC models (37), have been developed. In DPD method. The intracellular and extracellular fluids are modeled by
collections of free DPD particles and their separation is enforced by
general, one-component RBC models depict the cell mem-
bounce-back reflections at the RBC membrane surface. The RBC mem-
brane as a single shell with effective properties that repre- brane interacts with the fluid and wall particles through DPD forces, and
sent the combined effects of the lipid bilayer and the the system temperature is maintained by the DPD thermostat. For complete-
ness, we review briefly the two-component RBC model below, whereas for
a detailed description of the DPD method and the RBC model, we refer to
(34–37).

Two-component RBC model


In the two-component whole-cell model, the cell membrane is modeled by
two distinct components, i.e., the lipid bilayer and the cytoskeleton, and
each component is constructed by a two-dimensional triangulated network
with Nv vertices. In general, the two-component RBC model takes into
account the elastic energy, bending energy, bilayer-cytoskeleton interaction
energy, and constraints of the fixed surface area and enclosed volume, given by

Vðxi Þ ¼ Vs þ Vb þ Va þ Vv þ Vint (1)

where Vs is the elastic energy that mimics the elastic spectrin network,
given by
FIGURE 1 (a) Young’s modulus of normal and diabetic RBCs measured
2   3
in experiments, with data as follows: crosses, Fornal et al. (14); triangles,
X kB Tlm 3xj2  2xj3 k
Ciasca et al. (15); squares, Zhang et al. (16); circles, Lekka et al. (90).
Vs ¼ 4   þ
p 5; (2)
(b) Sketch of the RBC models with equilibrium biconcave (S/V ¼ 1.44)
j˛1:::Ns 4p 1  xj ðn  1Þln1
j
and near-oblate (S/V ¼ 1.04) shapes. To see this figure in color, go online.

482 Biophysical Journal 113, 481–490, July 25, 2017


Modeling of T2DM RBCs

where lj is the length of the spring j, lm is the maximum spring extension, between neighboring triangles, and the edges of the triangles are modeled
xj ¼ lj =lm , p is the persistence length, kB T is the energy unit, kp is the spring as springs. Thus, when the angle qj is adjusted, the springs may be stretched
constant, and n is a specified exponent. The bending resistance from the as well, which could result in local in-plane strains (deformation). To eval-
lipid bilayer of the RBC membrane is modeled by uate such an effect, we estimate the ratio between the bending and shear
X    moduli of RBC membrane, kc =ðm0 R2 Þ ¼ Oð103 Þ, where R is the radius
Vb ¼ kb 1  cos qj  q0 ; (3) of a sphere having the same surface area as that of an RBC. Thus, bending
j˛1:::Ns contributes little to the RBC deformation.
In addition, membrane viscosity also plays an important role in the dy-
where kb is the bending constant, qj is the instantaneous angle between two namic behavior of RBCs in physiological and pathological conditions. To
adjacent triangles having the common edge j, and q0 is the spontaneous estimate the effective membrane viscosity, hm , we simply combine the
angle. viscous contributions from the lipid bilayer, hb , and cytoskeleton, hs .
Constraints on the area and volume conservation of RBCs are imposed to Following
pffiffiffi priorpwork
ffiffiffi (37), these two viscous terms can be calculated as
mimic the area-preserving lipid bilayer and the incompressible interior hi ¼ 3gTi þ ð 3gCi =4Þ ði ¼ s; bÞ, where gTi and gCi are dissipative pa-
fluid. The corresponding energy is given by rameters of the DPD model. It is worth noting that the lipid bilayer is
 2
  modeled as a two-dimensional triangulated network in this RBC model.
X kd Aj  A0 2 ka Acell  Atotcell;0 Therefore, it cannot capture the diffusion of lipid molecules. However, a
Va ¼ þ ; (4)
j˛1:::Nt
2A 0 2A tot
cell;0
higher-fidelity model that includes the two components of the RBC mem-
brane plus transmembrane proteins could address this issue (51,52). For
example, Li et al. (53) have shown that a decrease in lipid bilayer diffusivity
 2 (or an increase in lipid bilayer viscosity) does not change the shear modulus
kv Vcell  Vcell;0
tot
of the RBC membrane. We have recently developed a whole-cell RBC
Vv ¼ ; (5) model at protein resolution using OpenRBC (33), which could be poten-
2Vcell;0
tot
tially applied to investigate the diffusion of lipid molecules in the RBC
membrane at the whole-cell level.
where Nt is the number of triangles in the membrane network, A0 is the tri-
angle area, and kd , ka , and kv are the local area, global area, and volume
constraint coefficients, respectively. The terms Atot tot
cell;0 and Vcell;0 represent
the specified total area and volume, respectively.
Parameter estimation
The bilayer-cytoskeleton interaction potential, Vint , is expressed as a Under physiological conditions, a normal RBC has a distinctive biconcave
summation of harmonic potentials given by shape with a large surface area/volume ðS=VÞ ratio, which facilitates
 2 oxygen transport through the cell membrane and contributes to the remark-
X kbs djj0  djj0 ;0 able cell deformability. In this study, we model a normal RBC (N-RBC)
Vint ¼ ; (6)
j;j0 ˛1:::Nbs
2 with the following parameters: total number of vertices, Nv ¼ 500;
RBC membrane shear modulus, m0 ¼ 4.73 mN/m; bending modulus,
where kbs and Nbs are the spring constant and the number of bond connec- kc ¼ 2.4  1019 J; and effective membrane viscosity, hm ¼ 0.128 Pa$s.
tions between the lipid bilayer and the cytoskeleton, respectively. djj0 is the We note that hm here is the three-dimensional membrane viscosity derived
distance between vertex j of the cytoskeleton and the corresponding projec- from the geometric relationship hm ¼ h2D 2D
m =R (54), where hm ¼ 0.416 mN s/m

tion point j0 on the lipid bilayer, with the corresponding unit vector njj0 ; djj0 ;0 is the two-dimensional membrane viscosity approaching the measured
is the initial distance between the vertex j and the point j0 , which is set to value of 0.36 mN s/m in experiments (47). According to Berk et al. (55),
zero in the simulations presented here. the cytoskeleton viscosity is at least one or two orders of magnitude
greater than the bilayer viscosity, and therefore we set hb ¼ 0.0027 Pa$s
and hs ¼ 0.125 Pa$s, respectively. In addition, RBC surface area,
2 3
cell;0 ¼ 132.87 mm , cell volume Vcell;0 ¼ 92.45 mm , and surface
Atot tot
Mechanical properties of the RBC membrane
area/volume ratio, S=V ¼ 1.44, are given to our N-RBC model.
It is known that the membrane elasticity of RBCs characterizes their resis- Different from the N-RBC, the diabetic RBCs have decreased cell de-
tance to deformation, and membrane viscosity characterizes the viscous formability and increased cell volume. Here, we have developed three po-
resistance of the cell membrane to shear deformation (47). Following the tential T2DM RBC models (D-RBC1, D-RBC2, and D-RBC3) based on
linear analysis for a regular hexagonal network by Dao et al. (48), we corre- previous in vitro experiments. First, we selected experimentally determined
late the model parameters and the network macroscopic elastic properties, shear modulus data (56) and set the shear modulus of a diabetic RBC model
i.e., shear modulus (m), which is determined by (D-RBC1) to m ¼ 2.0m0 . Second, previous AFM measurements suggest that
pffiffiffi ! there is a visible change in cell shape from the normal biconcave shape to a
3k B T x0 1 1 near-oblate shape with a reduced S=V ratio in the pathophysiology of
m ¼  þ
2ð1  x0 Þ3 4ð1  x0 Þ2 4
T2DM RBCs (57). Considering this fact, here, we propose a modified
4plm x0
; (7) diabetic RBC model (D-RBC2) with an oblate shape (S=V  1.04)
pffiffiffi (Fig. 1 b). Third, lower membrane fluidity and increased membrane viscos-
3kp ðn þ 1Þ
þ ; ity for T2DM RBCs compared to controls have been reported and suggested
4l0nþ1 to be the consequences of nonenzymatic glycation-induced changes in the
RBC membrane (58). In addition, previous studies also suggest a compara-
where l0 is the equilibrium spring length and x0 ¼ l0 =lm . In addition, the ble shape recovery time ðtc Þ between normal and diabetic RBCs (10,59). It
corresponding area compression is given by K ¼ 2m þ ka þ kd (36,49). is known that the recovery time, tc ¼ hm =m, is primarily determined by the
The relation between the modeled bending constant, kb , and the macro- viscoelastic properties of the RBC membrane (47,60). In combination with
scopic bending
pffiffiffi rigidity, kc , of the Helfrich model can be derived as the above experimental results, we have developed a third diabetic RBC
kb ¼ ð2= 3Þkc for a spherical membrane (50). This expression describes model (D-RBC3) by considering the aberrant cell shape and impaired de-
the bending contribution of the energy in Eq. 3 but may not fully represent formability, as well as adjusted effective membrane viscosity. In summary,
actual bending resistance of the RBC membrane. In the triangulated mem- the D-RBC1 model holds a biconcave shape (S=V ¼ 1.44), as does that of
brane network, the bending elasticity is expressed in terms of the angle qj an N-RBC with an increased m; the D-RBC2 model takes a near-oblate

Biophysical Journal 113, 481–490, July 25, 2017 483


Chang et al.

shape with a reduced S=V ratio ( 1.04) accompanied by an increase in m;


and the D-RBC3 model has the same characteristics as the D-RBC2 model
but with the additional trait of enhanced hm . Parameters related to the mem-
brane properties among the different RBC models, including the shear
modulus, m, effective membrane viscosity, hm , and surface area/volume ra-
tio ðS=VÞ, are summarized in Table 1.
The simulations were performed using a modified version of the atom-
istic code LAMMPS. The time integration of the motion equations is
computed through a modified velocity-Verlet algorithm with l ¼ 0:50
and time step Dt ¼ 1.0  10–4 t z 0.66 ms. It takes (1.0–5.0)  106
time steps for a typical simulation performed in this work. Compared to
the one-component RBC model, the two-component RBC model takes
about the same computational time for the prediction of a single-cell
dynamic response. However, for a blood-flow system with a large number
of RBCs (for example, for a hematocrit level at 45.0%), we found that the
computational costs are increased by a factor of 3–5 for the two-component FIGURE 2 Stretching response of normal and T2DM RBC membrane at
RBC model compared to those for the one-component RBC model. different values of the stretching force. The error bars are obtained by
increasing or decreasing m by 20% from the default values (Table 1). The
experimental data are adopted from Suresh et al. (61), and the different
RESULTS AND DISCUSSION stretched RBCs at stretching force Fs ¼ 100 pN are presented on the right.
To see this figure in color, go online.
In this section, we employ the two-component RBC model
to investigate the mechanics, rheology, and dynamics of
tion in RBC deformability, which is consistent with our
T2DM RBCs. First, we probe the static and dynamic re-
recent computational results that the S=V ratio is one of
sponses of T2DM RBCs subject to tensile forcing and quan-
the main dominant factors in cell deformability (43). For
tify the cell deformation. Second, we simulate the dynamic
D-RBC3, we find that the DA values are almost exactly
behavior of T2DM RBCs in shear flow and investigate the
the same as those obtaiend for D-RBC2 (Fig. 2). This result
effect of membrane viscosity on the tank-treading fre-
demonstrates that the influence of the membrane viscosity
quency. Finally, we study the biorheological properties of
on RBC stretching deformation is negligible, because it is
individual T2DM RBCs and predict the shear viscosity of
performed under the equilibrium stretched state at a given
T2DM RBC suspensions.
Fs . Hence, we conclude that the membrane shear stiffness
and S=V ratio are the two most important parameters in
Mechanical properties of T2DM RBCs static RBC deformation.
In addition, instantaneous membrane fluctuations of
To investigate the elastic response of RBCs in normal and
RBCs, also called ‘‘membrane flickering’’ (62), are com-
T2DM conditions, we subject the RBC to stretching defor-
monly used to characterize the membrane stiffness.
mation by imposing an external tensile force at diametri-
Following the work of Fedosov et al. (63) and Peng et al.
cally opposite directions analogous to that in optical
(37), we then examine the membrane fluctuations of T2DM
tweezers experiments (61). In our simulations, the total
RBCs by computing instantaneous fluctuation height of the
stretching force, Fs , is applied in the opposite direction to
RBC membrane surface (Fig. 3). From this plot, we find a
eNv (e ¼ 0.05) vertices of the lipid bilayer component of
good agreement between the fluctuation distributions in ex-
the RBC membrane. The stretching response of the RBC
periments (red line) and simulations (black squares) for
is then characterized by the variation of axial ðDA Þ and
normal RBCs. For T2DM RBCs, we find narrower distribu-
transverse ðDT Þ diameters of the RBC (Fig. 2). For D-
tions or smaller full-width half-maximum (FWHM) values
RBC1, we find a significant decrease in DA when compared
compared to those for normal RBCs, which shows the influ-
with that of N-RBC, which is due to the increase in shear
ence of local membrane curvature and effective geometry on
stiffness of T2DM RBCs. For D-RBC2 with a near-oblate
the membrane fluctuations. These results are consistent with
shape, we find a further decrease in DA , indicating a further
previous simulations of malaria-infected RBCs at the
reduction in cell deformability. This result confirms that the
schizont stage (63): When simulations employed a biconcave
decrease in S=V ratio from  1.44 to  1.04 causes a reduc-
RBC shape, a wider distribution than that in experiment (62)
for the schizont stage was observed; however, by employing a
TABLE 1 Model Parameters Extracted from Available nearly spherical membrane (spherical RBCs at the schizont
Measured Data for RBCs in Normal and T2DM Conditions stage observed in experiments) in the RBC model, the distri-
N-RBC D-RBC1 D-RBC2 D-RBC3 bution became much narrower and presented a better agree-
m ðmN=mÞ 4.73 (47) 9.46 (56) 9.46 (56) 9.46 (56) ment with the experiments. These computational results
S=V ðmm2 =mm3 Þ 1.44 (89) 1.44 1.04 (57) 1.04 (57) show that both membrane shear modulus and shape alteration
hm ðPa,sÞ 0.128 (37) 0.128 0.128 0.256 (59) play important roles in RBC membrane fluctuations. More-
Numbers in parentheses indicate the literature source for the given parameter. over, the membrane distributions obtained from D-RBC2

484 Biophysical Journal 113, 481–490, July 25, 2017


Modeling of T2DM RBCs

In fact, the dynamic recovery of the RBC, RðtÞ ¼


f ðDA ; DT Þ, can be described by a time-dependent exponen-
tial decay,
       
DA
DT
 DDAT DA
DT
þ DDAT
RðtÞ ¼  t  N ,  0  N
DA
DT
þ DDAT DA
DT
 DDAT
t N 0 N

a
t
¼ exp  ; (8)
tc

where the subscripts 0 and N correspond to the ratios at


t ¼ 0 and N. The exponent a ¼ 1:0 is a generalization for
estimating tc (60,65,66), whereas a ¼ 0:7 has been used
by Fedosov et al. (36) for the purpose of better fitting their
computational results. For comparison, we obtained
FIGURE 3 Membrane fluctuation distributions of normal and T2DM different curves by fitting the simulation data to exponential
RBCs. The solid red line represents experimental data for membrane
fluctuation distribution of normal RBCs (62). The full-width, half-
functions with a ¼ 1:0 and a ¼ 0:7 (see Fig. S1). We found
maximum fluctuation value is 128 nm for N-RBC, 110 nm for D- that both fitting curves are in fairly good agreement with the
RBC1, 66 nm for D-RBC2, and 65 nm for D-RBC3. To see this figure in RBC recovery dynamics. In the following work, we adopted
color, go online. a ¼ 1:0 for estimating tc of different RBC models.
Fig. 4 c shows the corresponding best fit to the relaxation
and D-RBC3 remain almost the same, indicating a small ef- dynamics in Fig. 4 a. The shape recovery time is estimated
fect of membrane fluidity on the cell membrane fluctuations. to be tc ¼ 0.11, 0.08, 0.07, and 0.12 s for N-RBC, D-RBC1,
It is interesting to see that our simulation results are qualita- D-RBC2, and D-RBC3, respectively. The comparable tc
tively in accordance with an up-to-date experimental work by values for N-RBC and D-RBC3 show a correspondence
Lee et al. (64) showing that the membrane fluctuations of dia- with the experimental results (59), thus, we conclude that
betic RBCs are significantly lower than those of nondiabetic the effective membrane viscosity contributes largely to
RBCs. RBC relaxation. The underestimated tc value in D-RBC1
Upon external tensile forces, a normal RBC undergoes and D-RBC2 could attribute to the increased membrane
elastic deformation, and it restores to its original state shear stiffness and the reduced S/V ratio. In addition, there
when the external force is released. Therefore, the RBC de- is no significant difference in tc between D-RBC1 and D-
formability can also be reflected on the shape relaxation of RBC2, which is consistent with the experimental results
RBCs. It is well-accepted that the shape recovery time that the geometric structure of the RBC has little effect on
tc z0:1 s for normal RBCs (47,60,65). When the cell age the cell recovery process (60,67). Our computational results
was taken into account, Williamson et al. (59) found that on RBC mechanics from stretching behavior to recovery
tc ¼ 0:12750:011 and 0:13250:018 s for young RBCs in response demonstrate that the RBC membrane elasticity,
normal and diabetic conditions, respectively. It increases S/V ratio, and effective membrane viscosity are all essential
to tc ¼ 0:15250:010 and 0:14950:014 s for old RBCs in to probe RBC deformability in T2DM.
healthy and diabetic cells. Hence, the difference in tc be-
tween young and old cells is pronounced, whereas that
Dynamic behavior of individual T2DM RBC in flow
between normal and diabetic cells is not significant. We
then perform a stretching-relaxation test of normal and The RBCs in shear flow have been observed to exhibit two
T2DM RBCs with different model parameters (Fig. 4). primary types of dynamics: tank-treading (TT) motion and
We find that the recovery processes are different from tumbling (TB) motion, depending on cell geometry (S/V
each other even though all of these modeled RBCs are ratio), cell elasticity (capillary number Cam ¼ ho gR _ 0 =m,
able to recover their original shapes. In addition, assuming where R0 is the radius of a sphere with the same volume
a pathological RBC state where the bilayer-cytoskeletal as an RBC), and fluid properties such as viscosity ratio
interaction is significantly lower, we find an apparent uncou- (l ¼ hi =ho , where hi and ho are the cytoplasm and suspend-
pling between bilayer and cytoskeleton when RBCs are _ the
ing fluid viscosities) (23,68–71). At low shear rate, g,
subjected to tensile forcing (Fig. 4 b). Herein, we want to resistance to shear causes RBCs to tumble. As g_ increases,
emphasize that in most cases, the one-component and the dynamics of RBCs changes from TB to TT motion (72).
two-component RBC models do not differ too much; how- Physiologically, the TT motion of RBCs in flowing blood
ever, under extreme mechanical conditions or disease states, produces a lift force to prevent the RBCs from migration to-
the two-component RBC model is needed. ward the peripheral blood vessel, allowing high flow

Biophysical Journal 113, 481–490, July 25, 2017 485


Chang et al.

FIGURE 4 (a) Dynamic cell deformation and relaxation processes of different RBC models at the forces shown at top. (b) An apparent uncoupling be-
tween the lipid bilayer and cytoskeleton occurs by reducing the bilayer-cytoskeletal interaction of the D-RBC3 model in one or two orders of magnitude. (c)
Estimated shape recovery time, tc , of different RBC models. t ¼ 0 is the time when the external force is released and the cell starts to recover its original
shape. To see this figure in color, go online.

efficiency and sufficient oxygen transfer in blood circulation found a slight decrease in TT frequency of diabetic RBCs
(73). In this study, we simulate the TT motion of an RBC in compared to normal RBCs. To gain further insight into the
shear flow by placing a single RBC in linear shear flow (a RBC TT motion, we investigate the TT frequency of an in-
Couette flow) between two planar solid walls. To produce dividual RBC in normal and T2DM conditions. As sug-
shear flow in the fluidic channel, two solid walls move gested by previous experimental and computational
with the same speed but in opposite directions, and g_ can studies on the dynamics of individual RBC in shear flow,
be mediated by changing the speed of the solid walls.
According to previous experimental studies (74,75), we
set ho ¼ 0.0289 Pa$s (74) and hi ¼ 0.006 Pa$s (75).
Fig. 5 a shows the angular trajectory (q) as a function of
time for different RBC models at g_ ¼ 105 s1. It is evident
that both D-RBC1 and D-RBC2 have TT motion faster than
that of N-RBC, which is indicated by a shorter period for the
marked particle on the cell membrane to complete a TT
cycle, as shown in Fig. 5 b. The accelerated TT motion
for D-RBC1 compared to that of N-RBC is probably due
to the average elongation decrease with increasing cell
membrane stiffness (70). For an oblate shape of D-RBC2,
it rotates even faster than for D-RBC1, which may owe to
an increased cell thickness, as reported in previous compu-
tational work on higher inclination and TT speed of osmot-
ically swollen RBCs (76). In addition, D-RBC3 exhibits a
slower rotation and an extended TT period. This result
shows that TT motion is sensitive to the membrane viscous
dissipation, and although the swollen cell tends to speed up
the TT dynamics, the elevated effective membrane viscosity
slows it down even more.
TT frequency (f), i.e., the number of TT cycles per
second, is an important characteristic in the RBC TT mo-
tion. The TT frequency can be estimated from the time-
dependent angular trajectories of a marked particle based
on f ¼ 1/Ptt , where Ptt is the time for an oscillation cycle FIGURE 5 TT motion of an RBC in shear flow. (a) Angular trajectory (q)
of a marked particle in the RBC membrane during the TT motion for
of the markered particle. In previous studies, Tran-Son-
different RBC models at shear rate g_ ¼ 105 s1 . q is the inclination angle
Tay et al. (77) and Williamson et al. (59) investigated the between the position vector of the marked particle and the flow direction.
TT frequency of RBCs in normal and diabetic conditions. (b) Corresponding snapshots of different RBC models at t ¼ 0.20, 0.25,
They obtained a linear relationship between f and g_ and 0.30, 0.35, and 0.40 s. To see this figure in color, go online.

486 Biophysical Journal 113, 481–490, July 25, 2017


Modeling of T2DM RBCs

RBCs tumble at low shear rates (35,36,45,73). Therefore,


we collected and analyzed the RBC TT motion at relatively
high shear rates, herein g_ > 45 s1. Fig. 6 shows that f
increases linearly with g_ for all four RBC models. At the
same g,_ D-RBC2 has the largest f value, N-RBC and
D-RBC1 have moderate f value, whereas D-RBC3 has the
smallest TT frequency. Comparing with experimental
data, we find that our N-RBC and D-RBC3 have the corre-
sponding TT frequency. These simulation results show that
in addition to capturing the dynamic behavior of a normal
RBC under shear flow, D-RBC3 with explicit consideration
of increased effective membrane viscosity leads to an accu-
rate model of the impaired RBC dynamics in T2DM.

Rheological properties of T2DM RBCs in FIGURE 7 RBC traversal across a microfluidic channel of 6.0 mm width
shear flow at various values of pressure difference. The cell transit velocity, v, is
defined as the transit distance divided by the average transit time for an
The rheological behavior of blood samples from nondia- RBC traversing the narrow channel, and DP is a local pressure gradient
betic and diabetic patients have been studied in experiments in the channel. For better comparison, the normalized cell transit velocity
(v ¼ v/vs ) and local pressure difference (DP ¼ DP/DPs ) are adopted,
(17–19,78). Decreased cell deformability and increased cell where vs ¼ 1.6 mm/s and DPs ¼ 0.1 kPa. Experiments (open circles) and
volume in T2DM have shown significant implications for simulations (solid circles) of a normal RBC traversing a 6-mm -wide chan-
microcirculatory alterations. Computational modeling and nel from Quinn et al. (79) are shown. (Inset) Schematic of the microfluidic
simulations help in predicting how RBCs behave in shear channel with a symmetric converging-diverging geometry. To see this figure
flow and providing insights into how viscous flow trans- in color, go online.
forms RBC shapes, and vice versa, how deformed RBCs
distort the surrounding flow (22,39,40,79–83). Here, to and 2.7 mm high and has widths ranging from 4.0 to 6.0
address the effects of cell elasticity and shape on the mm. The narrow and wide domains are connected by walls
biorheological behavior of individual T2DM RBCs, we formed by stationary DPD particles.
investigate the dynamic behavior of T2DM RBCs in a Fig. 7 and Movies S1, S2, and S3, show the typical dy-
microfluidic channel with constriction. The microfluidic namic processes of normal and T2DM RBCs flowing
channel, as illustrated in the inset of Fig. 7, contains a sym- through the converging-diverging microfluidic channel.
metric converging and diverging nozzle-shaped section. At Our simulations indicate that the RBCs passing through
its narrowest domain, the microchannel is 30.0 mm long the microchannel are sensitive to the S/V ratio (or cell
volume). As shown in Fig. 7, a decrease in the S/V ratio
from 1.44 to 1.04 results in a decrease in cell transit velocity
when individual RBCs travel through a 6-mm-wide micro-
channel. Similar to previous computational studies of
RBCs and tumor cells passing through narrow slits
(43,84,85), the cell volume increase (S/V decrease) would
slow down the passing process. The average cell transit
velocity decreases more significantly, however, when the
individual T2DM RBCs pass through a 4-mm-wide micro-
channel see Movie S3. These simulations reveal that the
increase of flow resistance by T2DM RBCs is larger than
the resistance by normal RBCs, hence pointing to the impor-
tance of the S/V ratio as a determinant of T2DM RBCs
traversal across narrow capillaries.
Computational models have also been proven to be an
important tool in predicting the macroscopic flow properties
FIGURE 6 Functional dependence of RBC TT frequency with respect to of an RBC suspension or blood flow (e.g., yield stress and
_ The error bars were obtained by increasing or decreasing hm by
shear rate g. shear viscosity) from the mesoscopic properties of individ-
10% from their default values. Simulation results are compared with exper-
ual RBCs (e.g., membrane viscosity and cell deformability).
imental data by Fischer (red circle) (74), by Tran-Son-Tay et al. (red cross)
(77), and by Williamson et al. (green star) (59). The red and green lines Next, we employ our RBC models to examine the blood
show the linear fits for normal and diabetic RBCs, respectively, in experi- viscosity over a range of g_ from 1 to 500 s1. Fig. 8 and
ments. To see this figure in color, go online. Movies S4 and S5 show the relative viscosity of RBC

Biophysical Journal 113, 481–490, July 25, 2017 487


Chang et al.

tion from membrane shear resistance and a near-oblate


structural constraint. However, only the D-RBC3 model
can exhibit reasonable relaxation processes in comparison
with the corresponding experiments, indicating that mem-
brane viscosity plays an important role in RBC deformability.
Second, we quantitatively probed the linear relationship
between tank-treading frequency and shear-flow rate of
RBCs under normal and diabetic states by our N-RBC and
D-RBC3 models. The agreement of the simulation results
with the currently available experiments shows that the
RBC complex dynamics in shear flow can be predicted by
taking into account the membrane elasticity, cell geometry,
and viscous dissipation. Finally, we extended the simulation
from individual cell dynamics to the collective dynamics
of RBC suspensions. We predicted a higher viscosity of
FIGURE 8 Functional dependence of shear viscosity of T2DM RBC T2DM blood compared to that of normal subjects, and the
suspension on shear rate at hematocrit Ht ¼ 45:0%. Experimental data
results are also consistent with experimental data.
are as follows: circles are from Skovborg et al. (17), crosses from Zingg
et al. (18), and diamonds from Peduzzi et al. (78); red symbols are for It is known that the main effect of T2DM on blood prop-
normal RBC suspension and green symbols for diabetic RBC suspension. erties derives from the uncontrolled concentration of
To see this figure in color, go online. glucose. Glycated hemoglobin (HbA1c), which is formed
by a simple chemical reaction between blood glucose and
suspensions with the N-RBC and D-RBC3 models. The normal hemoglobin in RBCs, reflects the time-averaged
model predictions clearly capture the increased blood blood glucose level in an individual and can be used for
viscosity in T2DM, in good agreement with the experi- the diagnosis of T2DM. Experimental studies have shown
mental data. This abnormality in blood viscosity is attrib- that T2DM RBCs are less deformable and more fragile if
uted to the reduced cell deformability associated with there is an increased amount of HbA1c (87,88). However,
the alterations in cell shape. Our model does not include the precise effect of blood glucose level on cell deformabil-
cell-cell aggregation interactions. Hence, it fails to model ity and blood rheology remains an open question. The cur-
the formation of rouleaux structures and a tremendous rent RBC models, including the two-component models
viscosity increase at low shear rates (38,86); this can be a used here, still have limitations in facilitating the detailed
topic for future work. Nevertheless, our model with explicit exploration of diverse biophysical and biomechanical as-
description of the RBC structure and membrane properties pects involved in such cases. There is, hence, a compelling
could lead to useful insights into the cell mechanistic need to develop a hybrid model constructed by combining a
processes and guide future work for better understanding kinetic model for the formation of HbA1c at the subcelluar
the correlation between metabolic dysfunction and hemato- level with the particle-based RBC model. Such a model can
logical abnormalities in T2DM. be used to quantify the functional relationship between
glucose (or HbA1c) sensitivity and the biomechanical
properties of T2DM RBCs as well as the hemorheological
CONCLUSIONS
properties of T2DM blood. With the help of more extensive
We have presented a two-component whole-cell model that clinical tests and experimental studies, unsolved questions
seems to capture the biomechanical, rheological, and concerning the correlation between the degree of hypergly-
dynamic properties of RBCs in T2DM based on currently cemia and the alterations in diabetic hematology could be
available data. The RBC membrane is modeled by two possibly addressed in the near future.
distinct networks for representing the spectrin network
and the lipid bilayer. We have proposed and examined three
potential T2DM RBC models (D-RBC1, D-RBC2, and SUPPORTING MATERIAL
D-RBC3) based on the alterations in cell shape and mem- One figure and five movies are available at http://www.biophysj.org/
brane viscoelasticity measured experimentally, including biophysj/supplemental/S0006-3495(17)30665-3.
increased shear stiffness, altered cell shape, and elevated
membrane viscosity.
Using these RBC models, we first studied the static and AUTHOR CONTRIBUTIONS
dynamic deformation of T2DM RBCs subject to different
H.-Y.C., X.L., and G.E.K. conceived and designed research; H.-Y.C. and
tensile forcing. The significant reduction in stretching X.L. performed research; H.-Y.C., X.L., and G.E.K. analyzed data;
response of our diabetic RBC models compared to N-RBCs H.-Y.C., X.L., and G.E.K. contributed new reagents/analytic tools; and
at the same tensile force shows the large stiffness contribu- H.-Y.C., X.L., and G.E.K. wrote the article.

488 Biophysical Journal 113, 481–490, July 25, 2017


Modeling of T2DM RBCs

ACKNOWLEDGMENTS 21. Turczynski, B., K. Michalska-Ma1ecka, ., W. Romaniuk. 2003.


Correlations between the severity of retinopathy in diabetic patients
Part of this research was conducted using computational resources and ser- and whole blood and plasma viscosity. Clin. Hemorheol. Microcirc.
vices at the Center for Computation and Visualization (CCV) at Brown 29:129–137.
University. 22. Bagchi, P. 2007. Mesoscale simulation of blood flow in small vessels.
Biophys. J. 92:1858–1877.
The work described in this article was supported by National Institutes of
Health (NIH) grants U01HL114476 and U01HL116323. 23. Li, X., P. M. Vlahovska, and G. E. Karniadakis. 2013. Continuum- and
particle-based modeling of shapes and dynamics of red blood cells in
health and disease. Soft Matter. 9:28–37.

REFERENCES 24. Freund, J. B. 2014. Numerical simulation of flowing blood cells. Annu.
Rev. Fluid Mech. 46:67–95.
1. Mathers, C. D., and D. Loncar. 2006. Projections of global mortality 25. Yazdani, A., X. Li, and G. E. Karniadakis. 2016. Dynamic and
and burden of disease from 2002 to 2030. PLoS Med. 3:e442. rheological properties of soft biological cell suspensions. Rheol.
2. Fowler, M. J. 2008. Microvascular and macrovascular complications of Acta. 55:433–449.
diabetes. Clin. Diabetes. 26:77–82. 26. Balogh, P., and P. Bagchi. 2017. A computational approach to modeling
3. Caimi, G., and R. L. Presti. 2004. Techniques to evaluate erythrocyte cellular-scale blood flow in complex geometry. J. Comput. Phys.
deformability in diabetes mellitus. Acta Diabetol. 41:99–103. 334:280–307.

4. Li, J., M. Dao, ., S. Suresh. 2005. Spectrin-level modeling of the cyto- 27. Hosseini, S. M., and J. J. Feng. 2012. How malaria parasites reduce the
skeleton and optical tweezers stretching of the erythrocyte. Biophys. J. deformability of infected red blood cells. Biophys. J. 103:1–10.
88:3707–3719. 28. Li, X., H. Li, ., G. E. Karniadakis. 2017. Computational biome-
5. Chien, S. 1987. Red cell deformability and its relevance to blood flow. chanics of human red blood cells in hematological disorders.
Annu. Rev. Physiol. 49:177–192. J. Biomech. Eng. 139:021008.
6. Kim, J., H. Lee, and S. Shin. 2015. Advances in the measurement of red 29. Secomb, T. W. 1995. Mechanics of blood flow in the microcirculation.
blood cell deformability: a brief review. J Cell. Biotechnol. 1:63–79. Symp. Soc. Exp. Biol. 49:305–321.
7. Glenister, F. K., R. L. Coppel, ., B. M. Cooke. 2002. Contribution of 30. McWhirter, J. L., H. Noguchi, and G. Gompper. 2009. Flow-induced
parasite proteins to altered mechanical properties of malaria-infected clustering and alignment of vesicles and red blood cells in microcapil-
red blood cells. Blood. 99:1060–1063. laries. Proc. Natl. Acad. Sci. USA. 106:6039–6043.
8. Zhang, Y., C. Huang, ., S. Suresh. 2015. Multiple stiffening effects 31. Wu, T., and J. J. Feng. 2013. Simulation of malaria-infected red blood
of nanoscale knobs on human red blood cells infected with Plasmo- cells in microfluidic channels: passage and blockage. Biomicrofluidics.
dium falciparum malaria parasite. Proc. Natl. Acad. Sci. USA. 7:44115.
112:6068–6073. 32. Li, X., M. Dao, ., G. E. Karniadakis. 2017. Biomechanics and bio-
9. Barabino, G. A., M. O. Platt, and D. K. Kaul. 2010. Sickle cell biome- rheology of red blood cells in sickle cell anemia. J. Biomech. 50:34–41.
chanics. Annu. Rev. Biomed. Eng. 12:345–367. 33. Tang, Y.-H., L. Lu, ., G. E. Karniadakis. 2017. OpenRBC: a fast
10. McMillan, D. E., N. G. Utterback, and J. La Puma. 1978. Reduced simulator of red blood cells at protein resolution. Biophys. J.
erythrocyte deformability in diabetes. Diabetes. 27:895–901. 112:2030–2037.
11. Kowluru, R., M. W. Bitensky, ., T. Buican. 1989. Reversible sodium 34. Hoogerbrugge, P., and J. Koelman. 1992. Simulating microscopic
pump defect and swelling in the diabetic rat erythrocyte: effects on hydrodynamic phenomena with dissipative particle dynamics. Euro-
filterability and implications for microangiopathy. Proc. Natl. Acad. phys. Lett. 19:155.
Sci. USA. 86:3327–3331. 35. Pivkin, I. V., and G. E. Karniadakis. 2008. Accurate coarse-grained
12. Agrawal, R., T. Smart, ., C. Pavesio. 2016. Assessment of red blood modeling of red blood cells. Phys. Rev. Lett. 101:118105.
cell deformability in type 2 diabetes mellitus and diabetic retinopathy 36. Fedosov, D. A., B. Caswell, and G. E. Karniadakis. 2010. A multiscale
by dual optical tweezers stretching technique. Sci. Rep. 6:15873. red blood cell model with accurate mechanics, rheology, and dynamics.
13. Babu, N., and M. Singh. 2004. Influence of hyperglycemia on aggrega- Biophys. J. 98:2215–2225.
tion, deformability and shape parameters of erythrocytes. Clin. Hemor- 37. Peng, Z., X. Li, ., S. Suresh. 2013. Lipid bilayer and cytoskeletal
heol. Microcirc. 31:273–280. interactions in a red blood cell. Proc. Natl. Acad. Sci. USA.
14. Fornal, M., M. Lekka, ., J. Styczen. 2006. Erythrocyte stiffness in dia- 110:13356–13361.
betes mellitus studied with atomic force microscope. Clin. Hemorheol. 38. Fedosov, D. A., W. Pan, ., G. E. Karniadakis. 2011. Predicting human
Microcirc. 35:273–276. blood viscosity in silico. Proc. Natl. Acad. Sci. USA. 108:11772–
15. Ciasca, G., M. Papi, ., M. De Spirito. 2015. Mapping viscoelastic 11777.
properties of healthy and pathological red blood cells at the nanoscale 39. Lei, H., and G. E. Karniadakis. 2012. Quantifying the rheological and
level. Nanoscale. 7:17030–17037. hemodynamic characteristics of sickle cell anemia. Biophys. J.
16. Zhang, S., H. Bai, and P. Yang. 2015. Real-time monitoring of mechan- 102:185–194.
ical changes during dynamic adhesion of erythrocytes to endothelial 40. Lei, H., D. A. Fedosov, ., G. E. Karniadakis. 2013. Blood flow in
cells by QCM-D. Chem. Commun. (Camb.). 51:11449–11451. small tubes: quantifying the transition to the non-continuum regime.
17. Skovborg, F., A. V. Nielsen, ., J. Ditzel. 1966. Blood-viscosity in J. Fluid Mech. 722:214–239.
diabetic patients. Lancet. 1:129–131. 41. Lei, H., and G. E. Karniadakis. 2013. Probing vasoocclusion phenom-
18. Zingg, W., J. C. Sulev, ., R. M. Ehrlich. 1971. Blood viscosity in dia- ena in sickle cell anemia via mesoscopic simulations. Proc. Natl. Acad.
betic children. Diabetologia. 7:461–462. Sci. USA. 110:11326–11330.
19. Cho, Y. I., M. P. Mooney, and D. J. Cho. 2008. Hemorheological 42. Li, X., E. Du, ., G. E. Karniadakis. 2016. Patient-specific blood
disorders in diabetes mellitus. J. Diabetes Sci. Technol. 2:1130–1138. rheology in sickle-cell anaemia. Interface Focus. 6:20150065.
20. Ercan, M., D. Konukoglu, ., S. Önen. 2002. The effects of cholesterol 43. Pivkin, I. V., Z. Peng, ., S. Suresh. 2016. Biomechanics of red blood
levels on hemorheological parameters in diabetic patients. Clin. Hem- cells in human spleen and consequences for physiology and disease.
orheol. Microcirc. 26:257–263. Proc. Natl. Acad. Sci. USA. 113:7804–7809.

Biophysical Journal 113, 481–490, July 25, 2017 489


Chang et al.

44. Li, X., E. Du, ., G. E. Karniadakis. 2017. Patient-specific modeling of 66. Hochmuth, R. M., K. L. Buxbaum, and E. A. Evans. 1980. Temperature
individual sickle cell behavior under transient hypoxia. PLOS Comput. dependence of the viscoelastic recovery of red cell membrane.
Biol. 13:e1005426. Biophys. J. 29:177–182.
45. Li, X., Z. Peng, ., G. E. Karniadakis. 2014. Probing red blood cell me- 67. Ruef, P., and O. Linderkamp. 1999. Deformability and geometry of
chanics, rheology and dynamics with a two-component multi-scale neonatal erythrocytes with irregular shapes. Pediatr. Res. 45:114–119.
model. Philos. Trans. A Math. Phys. Eng. Sci. 372:20130389. 68. Fischer, T. M., M. Stöhr-Lissen, and H. Schmid-Schönbein. 1978. The
46. Chang, H.-Y., X. Li, ., G. E. Karniadakis. 2016. MD/DPD multiscale red cell as a fluid droplet: tank tread-like motion of the human erythro-
framework for predicting morphology and stresses of red blood cells in cyte membrane in shear flow. Science. 202:894–896.
health and disease. PLOS Comput. Biol. 12:e1005173. 69. Fischer, T. M., and R. Korzeniewski. 2015. Angle of inclination of
47. Hochmuth, R. M., and R. E. Waugh. 1987. Erythrocyte membrane elas- tank-treading red cells: dependence on shear rate and suspending me-
ticity and viscosity. Annu. Rev. Physiol. 49:209–219. dium. Biophys. J. 108:1352–1360.
48. Dao, M., J. Li, and S. Suresh. 2006. Molecularly based analysis of 70. Yazdani, A., R. M. Kalluri, and P. Bagchi. 2011. Tank-treading and
deformation of spectrin network and human erythrocyte. Mater. Sci. tumbling frequencies of capsules and red blood cells. Phys. Rev. E
Eng. C. 26:1232–1244. Stat. Nonlin. Soft Matter Phys. 83:046305.
49. Fedosov, D. A., B. Caswell, and G. E. Karniadakis. 2010. Systematic 71. Yazdani, A., and P. Bagchi. 2013. Influence of membrane viscosity on
coarse-graining of spectrin-level red blood cell models. Comput. capsule dynamics in shear flow. J. Fluid Mech. 718:569–595.
Methods Appl. Mech. Eng. 199:1937–1948. 72. Dupire, J., M. Socol, and A. Viallat. 2012. Full dynamics of a red blood
50. Helfrich, W. 1973. Elastic properties of lipid bilayers: theory and cell in shear flow. Proc. Natl. Acad. Sci. USA. 109:20808–20813.
possible experiments. Z. Naturforsch. C. 28:693–703. 73. Basu, H., A. K. Dharmadhikari, ., D. Mathur. 2011. Tank treading
51. Li, H., and G. Lykotrafitis. 2012. Two-component coarse-grained mo- of optically trapped red blood cells in shear flow. Biophys. J.
lecular-dynamics model for the human erythrocyte membrane. 101:1604–1612.
Biophys. J. 102:75–84. 74. Fischer, T. M. 2007. Tank-tread frequency of the red cell membrane:
52. Li, H., and G. Lykotrafitis. 2014. Erythrocyte membrane model with dependence on the viscosity of the suspending medium. Biophys. J.
explicit description of the lipid bilayer and the spectrin network. 93:2553–2561.
Biophys. J. 107:642–653. 75. Skalak, R., N. Ozkaya, and T. C. Skalak. 1989. Biofluid mechanics.
53. Li, H., Y. Zhang, ., G. Lykotrafitis. 2016. Modeling of band-3 protein Annu. Rev. Fluid Mech. 21:167–200.
diffusion in the normal and defective red blood cell membrane. Soft 76. Dodson, W. R., 3rd, and P. Dimitrakopoulos. 2012. Tank-treading of
Matter. 12:3643–3653. swollen erythrocytes in shear flows. Phys. Rev. E Stat. Nonlin. Soft
Matter Phys. 85:021922.
54. Prado, G., A. Farutin, ., L. Bureau. 2015. Viscoelastic transient of
confined red blood cells. Biophys. J. 108:2126–2136. 77. Tran-Son-Tay, R., S. P. Sutera, and P. R. Rao. 1984. Determination of
red blood cell membrane viscosity from rheoscopic observations of
55. Berk, D. A., R. M. Hochmuth, and R. E. Waugh. 1989. Viscoelastic
tank-treading motion. Biophys. J. 46:65–72.
properties and rheology. In Red Blood Cell Membranes: Structure:
Function: Clinical Implications. P. Agre and J. C. Parker, eds. (CRC 78. Peduzzi, M., M. Melli, ., F. Guerrieri. 1984. Comparative evaluation
Press), pp. 423–454. of blood viscosity in diabetic retinopathy. Int. Ophthalmol. 7:15–19.
56. Bokori-Brown, M., P. G. Petrov, ., C. P. Winlove. 2016. Red blood 79. Quinn, D. J., I. Pivkin, ., S. Suresh. 2011. Combined simulation and
cell susceptibility to pneumolysin: correlation with membrane experimental study of large deformation of red blood cells in microflui-
biochemical and physical properties. J. Biol. Chem. 291:10210–10227. dic systems. Ann. Biomed. Eng. 39:1041–1050.
57. Jin, H., X. Xing, ., J. Cai. 2010. Detection of erythrocytes influenced 80. Lykov, K., X. Li, ., G. E. Karniadakis. 2015. Inflow/outflow boundary
by aging and type 2 diabetes using atomic force microscope. Biochem. conditions for particle-based blood flow simulations: application to
Biophys. Res. Commun. 391:1698–1702. arterial bifurcations and trees. PLOS Comput. Biol. 11:e1004410.
58. Waczulı́kova, I., L. Sikurová, ., B. Krahulec. 2000. Decreased fluidity 81. Cordasco, D., and P. Bagchi. 2016. Dynamics of red blood cells in
of isolated erythrocyte membranes in type 1 and type 2 diabetes. The oscillating shear flow. J. Fluid Mech. 800:484–516.
effect of resorcylidene aminoguanidine. Gen. Physiol. Biophys. 82. Vahidkhah, K., P. Balogh, and P. Bagchi. 2016. Flow of red blood cells
19:381–392. in stenosed microvessels. Sci. Rep. 6:28194.
59. Williamson, J. R., R. A. Gardner, ., S. P. Sutera. 1985. Microrheo- 83. Xiao, L. L., Y. Liu, ., B. M. Fu. 2017. Effects of flowing RBCs on
logic investigation of erythrocyte deformability in diabetes mellitus. adhesion of a circulating tumor cell in microvessels. Biomech. Model.
Blood. 65:283–288. Mechanobiol. 16:597–610.
60. Hochmuth, R. M., P. R. Worthy, and E. A. Evans. 1979. Red cell exten- 84. Xiao, L. L., Y. Liu, ., B. M. Fu. 2016. Numerical simulation of a sin-
sional recovery and the determination of membrane viscosity. gle cell passing through a narrow slit. Biomech. Model. Mechanobiol.
Biophys. J. 26:101–114. 15:1655–1667.
61. Suresh, S., J. Spatz, ., T. Seufferlein. 2005. Connections between sin- 85. Salehyar, S., and Q. Zhu. 2017. Effects of stiffness and volume on the
gle-cell biomechanics and human disease states: gastrointestinal cancer transit time of an erythrocyte through a slit. Biomech. Model. Mecha-
and malaria. Acta Biomater. 1:15–30. nobiol. 16:921–931.
62. Park, Y., M. Diez-Silva, ., S. Suresh. 2008. Refractive index maps and 86. McMillan, D. E. 1983. The effect of diabetes on blood flow properties.
membrane dynamics of human red blood cells parasitized by Plasmo- Diabetes. 32 (Suppl 2):56–63.
dium falciparum. Proc. Natl. Acad. Sci. USA. 105:13730–13735. 87. Singh, M., and S. Shin. 2009. Changes in erythrocyte aggregation and
63. Fedosov, D. A., H. Lei, ., G. E. Karniadakis. 2011. Multiscale deformability in diabetes mellitus: a brief review. Indian J. Exp. Biol.
modeling of red blood cell mechanics and blood flow in malaria. 47:7–15.
PLOS Comput. Biol. 7:e1002270. 88. Kung, C.-M., Z.-L. Tseng, and H.-L. Wang. 2009. Erythrocyte fragility
64. Lee, S., H. Park, ., Y. Park. 2017. Refractive index tomograms and increases with level of glycosylated hemoglobin in type 2 diabetic pa-
dynamic membrane fluctuations of red blood cells from patients with tients. Clin. Hemorheol. Microcirc. 43:345–351.
diabetes mellitus. Sci. Rep. 7:1039. 89. Tomaiuolo, G. 2014. Biomechanical properties of red blood cells in
65. Mills, J. P., L. Qie, ., S. Suresh. 2004. Nonlinear elastic and visco- health and disease towards microfluidics. Biomicrofluidics. 8:051501.
elastic deformation of the human red blood cell with optical tweezers. 90. Lekka, M., M. Fornal, ., J. Styczen. 2005. Erythrocyte stiffness
Mech. Chem. Biosyst. 1:169–180. probed using atomic force microscope. Biorheology. 42:307–317.

490 Biophysical Journal 113, 481–490, July 25, 2017

You might also like