You are on page 1of 9

International Journal of Biological Macromolecules 227 (2023) 173–181

Contents lists available at ScienceDirect

International Journal of Biological Macromolecules


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

Single-polymer dynamics of starch-like branched ring polymers in steady


shear flow
Deyin Wang a, Xiaohui Wen a, *, Dong Zhang b, Xinguan Tan a, Jiajun Tang a
a
College of Mathematics and Physics, Chengdu University of Technology, Chengdu 610059, China
b
College of Life Sciences and Institute of Quantitative Biology, Zhejiang University, Hangzhou 310058, China

A R T I C L E I N F O A B S T R A C T

Keywords: The stretching dynamics and dynamical behaviors of individual branched ring polymer (BRP), a coarse-grained
Ring polymer model for some types of the starch, in steady shear flow are studied by using a hybrid mesoscale simulation
Branched ring polymer approach that combines multiparticle collision dynamics with standard molecular dynamics. By analyzing the
MPCD
stretched configuration of BRPs, we find the polymer size increases nonmonotonically with increasing branch
Tumbling motion
Tank-treading motion
length. Meanwhile, the decrease of the alignment angle of the stretched configuration of BRPs follows a universal
power law during the first downward phase as the shear rate increases. Constructing the three-dimensional
surface of the polymer's ring backbone and tracing the temporal fluctuations of the surface's normal vector
along the simulation trajectory, the tumbling and tank-treading motion are clearly reflected by periodic and non-
periodic changes of the normal vector. Interestingly, these temporal changes are much more regular than that of
the gyration tensor. Thus, a novel cross-correlation function, which is the correlation between fluctuations of the
normal vector along the flow direction and the velocity-gradient direction, is proposed to analyze the tumbling
motion that usually coexists with the tank-treading motion. This function can naturally address the fails of
traditional method that analyzing the tumbling motion by determining the correlation of temporal fluctuations of
the gyration tensor Gαα. By analyzing the dynamical behaviors of BRPs, diverse dependences of the tumbling
frequency ωTB and tank-treading frequency ωTT on the shear rate γ̇ are observed at a wide range of shear rates and
polymer sizes. Furthermore, our simulations also reveal that the tank-treading motion is more stable than the
tumbling motion for small-branch-size BRPs but the tumbling motion is more stable than the tank-treading
motion for large-branch-size BRPs.

1. Introduction they serve as a model system for biophysical problems, such as circular
plasmid DNA [22,23] and chromatin packing in nucleosomes [24,25].
In recent years, extensive investigations on dynamical and confor­ Moreover, circular macromolecules also play a critical role in emerging
mational properties of linear [1–11] and branched [12–18] polymers biotechnologies, such as macrocyclic peptides for drug delivery appli­
under shear flow have been conducted. The continuous tumbling motion cations [26]. Thus, much more researches have focus on dynamical and
of linear chains, which always accompany with large structural changes, structural properties of ring polymers under shear flow [27–36]. Unlike
has been revealed in theoretical studies [1–3], simulations [4–6,11] and the clear tumbling motion of linear chains, Chen et al. [29] found a
experiments [7–9]. The tank-treading motion of the star polymer in the coexistence of tumbling and tank-treading motions for ring polymers
strong shear flow regime, which represents the motion that the arms and proposed an angle autocorrelation function to analyze the tank-
rotate around the center of mass but the overall shape has little fluctu­ treading motion. Direct observation on dynamical behaviors and con­
ations, is observed by simulations [16,18]. formations of DNA ring polymers in shear flow shows similar average
The physical properties of polymers are always influenced by the fractional extensions, orientation angle, and gradient thickness but
chain end [19]. However, ring polymers (RPs) has no ends, which results different probability distributions of molecular extension between RPs
in different dynamics compared to linear and linear branched polymers and linear chains [33]. By investigating properties of RPs under ultra­
[20,21]. Ring polymers are directly relevant to biological systems, and high shear rate, Wang et al. [35] found a nonmonotonic increase in

* Corresponding author.
E-mail address: wenxiaohui13@cdut.edu.cn (X. Wen).

https://doi.org/10.1016/j.ijbiomac.2022.12.100
Received 7 October 2022; Received in revised form 28 November 2022; Accepted 10 December 2022
Available online 16 December 2022
0141-8130/© 2022 Elsevier B.V. All rights reserved.
D. Wang et al. International Journal of Biological Macromolecules 227 (2023) 173–181

polymer size when the hydrodynamic interaction (HI) is taken in ac­ monomers connected by a bond. The fully stretched bond length l0 and
count. Moreover, the shear-induced hydrodynamic inflation and the the spring constant K are chosen to be 1.5σ, which is chosen to avoid
strong suppression of tumbling of RPs under high shear rates [36] have bond crossing [57], and 30ε/σ2. ε and σ are taken as the units of energy
also been observed. These results show that ring polymers exhibit a and length.
richer variety of dynamical behaviors under wide range of shear rates.
More recently, branched ring polymers (BRPs), benefit from the 2.2. Multiparticle collision dynamics (MPCD)
possible novel functionality induced by fast and random motions of
arms, under shear flow has been investigated by many studies [37–40]. The solvents used to generate shear flow are modeled as point par­
Jeong et al. [38] found that the BRP with short arms generally has a ticles of mass m in MPCD [58–60]. In the streaming step, the solvent
more compact and less deformed structure, which is formed by the particles move ballistically with their respective velocities within the
intrinsically fast random motion of the short branches which constantly time interval h between two consecutive collision steps, and the position
disturbs the ring backbone's conformation. Roh et al. [39] analyzed of particle i is updated according to
several distinct torsional modes around branch points along the ring
ri (t + h) = ri (t) + vi (t)h, (3)
backbone of BRPs. They found noticeable differences among the
torsional modes with respect to the probability distributions of the In the collision step, all the particles are grouped into cubic cells of
gauche- and trans- states and the torsional dynamics, which is ascribed to edge length a and then rotate their relative velocities with respect to the
an extra torsional stiffness imposed by the short branches. However, the center-of-mass velocity of the selected collision cell around an random
knowledge on dynamical properties of individual BRP under shear flow axis by a fixed angle ϕ [61]. The velocity of ith particle in the collision
and effects of branches on the BRP's topology is still inadequate. Hence, step is updated according to
more studies on the variety of dynamical behaviors and conformational
vi (t + h) = vcm (t) + (R(ϕ) − I )[vi (t) − vcm (t) ], (4)
properties of BRP under a wide range of shear rates are needed. More­
over, insight into the dynamics of the BRP under shear flow is desirable
where R(ϕ) is the rotation matrix determined by ϕ, I is the unit matrix,
both from a fundamental and an applied viewpoint of polymer physics ∑Nc
and vcm = j=1vj/Nc is the center-of-mass velocity of the selected
and biophysics. Meanwhile, studies that using polymeric materials to
collision cell which has Nc particles. Mass, energy, and momentum are
remove metal ions [41–51], inorganic [52,53] and organic [54,55]
conserved in this process, which ensures that hydrodynamic behavior
compounds from water has been carried out in recent years. The com­
emerges on larger length scales [60,62].
plex flow field will appear during the industrial processing of the waste
The coupling of branched ring polymers and the fluid occurs in the
water, knowledge of polymeric materials under complex flow may be
collision step, here each cell contains Ncs solute particles and Ncm
helpful for designing, improving and industrial application of such
monomers of the polymer. The center-of-mass velocity of the selected
adsorbents.
cell at time t should be determined by [63,64].
In the present work, using the hybrid mesoscopic simulation method,
∑Ncs ∑Ncm
we investigate the dynamical behaviors, as well as conformational i=1 mvi (t) + j=1 Mvj (t)
properties, of BRPs in shear flow over a wide range of the polymer size vcm (t) = . (5)
mN sc + MN mc
and shear rates. We focus on the single-polymer dynamics and employ
several methods to analyze the motion of BRPs. We aim to elucidate the Then the velocities of both solute particles and monomers of the BRP
influence of branch size on dynamical behaviors of BRPs and clarify the in the selected collision cell are updated by Eq. (4). The mass, local
difference, similarity and connection between tumbling and tank- momentum, and energy are still conserved during the collision step. To
treading motions in response to different flow strength of the shear field. ensure the Galilean invariance, each collision grid is shifted with a
random vector at every collision step [65,66].
2. Model and simulation method Lees-Edwards boundary conditions are applied to generate desired
shear flow [67]. This method yields a linear fluid velocity profile v = γ̇ẑ
x
2.1. Branched ring polymers (BRPs) in the flow direction (x-axis). To maintain the temperature of the system
during the dynamic simulation, a local cell-level thermostat, the Max­
The branched ring polymer is composed of evenly distributed linear wellian thermostat (MBS) which is achieved by scaling the velocity of
arms and a ring backbone, and each monomer on backbone is linked by particles, is applied [68].
an arm. The number of monomers of backbone and each arm are
denoted by Lb and La respectively and the arms of polymer are Na, then 2.3. Simulation details
the total number of monomers of a BRP is N = Lb + NaLa. For simplicity,
we use relative branch size l*, which is defined as l* = La/Lb, to denote a For the shear flow, the mass of solvent particles is chosen to be m = 1,
BRP with backbone length Lb and arm length La. cubic cell length is set to a = σ , the rotation angle is set to ϕ = 130∘, the
Purely repulsive short-range interactions between beads are taken average number of fluid particles in a collision cell is set to 〈Nsc〉 = 10,
into account using the Lennard-Jones (LJ) potential and the collision period is set to h = 0.1τ, here the time unit satisfies τ =
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
⎧ [( )12 ( )6 ] ma2 /kB T . The choice of parameters for MPCD solvents yields a
⎪ σ σ √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
( ) ⎨ 4ε
⎪ − , rij ≤ 21/6 σ viscosityηs = 8.7 mkB T/a4 and a Schmidt number Sc ≈ 17 which have
ULJ rij = rij rij , (1) been shown to be reliable parameters for liquid-like system [69]. For the



0, 1/6
rij > 2 σ BRP, we set its mass of each monomer as M = 10m = 10 and use standard
velocity Verlet algorithm with a time step hMD = 0.005τ to integrate the
where rij denotes the distance between the centers of ith and jth Newton's equations of motion.
monomers. The bond that connects adjacent monomers is described by a Each simulation system is confined in a cubic box of length L = 100σ,
finitely extensible nonlinear elastic (FENE) potential [56]. where the periodic boundary conditions are applied. For more accurate
[ statistics, several independent runs were conducted for each set of
( )2 ]
( ) Kl20 rij parameters.
UFENE rij = − ln 1 − , rij < l0 (2)
2 l0

where rij denotes the distance between the centers of ith and jth

174
D. Wang et al. International Journal of Biological Macromolecules 227 (2023) 173–181

3. Results and discussions

It has been well confirmed that shear flow field extends the polymer
to a stretched state. The snapshots of fully stretched conformation of a
RP, BRP and the BRP's ring backbone under steady shear flow with shear
rate γ̇ = 0.3 at flow-gradient plane (the xoz plane) are shown in Fig. 1a,c
and e, respectively. The corresponding monomer density distributions of
these three conformations are shown in Fig. 1b,d and f, respectively.
Here, the monomers of RP, BRP's ring backbone and each arm of the BRP
are all chosen to be 20. As shown in Fig. 1, the ring backbone of the BRP
(Fig. 1e-f) reaches a much higher compressed structure than the pure
ring polymer (Fig. 1a-b) at the velocity-gradient direction, indicating
that the RP with and without arms (branches) experience different
strength of shear-induced strain force along the flow direction. How­
ever, as shown in Fig. S1, at the flow-vorticity plane (the xoy plane), the
conformational expansion of the ring backbone of the BRP (Fig. S1e-f) is
larger than that of the RP (Fig. S1a-b), because the arm of the BRP
(Fig. S1c-d) tend to move along the flow direction. Interestingly, for the
BRP with very short arms (l* = 0.1), the expansion of monomer density
distributions of the ring backbone is very close to that of a RP at the flow-
gradient plane (see Fig. S2), indicating that the polymer size plays a key
role in the formation of the compressed structure in response to applied
strong shear flow field. Moreover, as shown in Fig. S3 and S4, when
there is no any external field applied to MPC fluid (γ̇ = 0), both the
large-size (l* = 1.0) and small-size (l* = 0.1) BRP display wider ex­
pansions on configuration than the RP (l* = 0), because the self-avoid
random movements of arms play a critical role in the formation of
conformation of BRPs when shear flow is very weak. This result also can
be attribute to the self-diffusion behavior of BRPs, which is similar to
many other types of branched polymers [70–73].
The gyration tensor of a polymer is defined as

1 ∑N
〈( )( )〉
Gαβ = ri,α − rcm,α ri,β − rcm,β , (6)
N i=1 Fig. 2. The components of ensemble averaged gyration tensors as functions of
shear rate.
where i denotes the index of monomers, N is the total number of
monomers, rcm represents the center of mass of the polymer and α, β ∈ in Fig. 2a, for the BRP with very large size (l* = 2.0), <Gxx> increases
{x, y, z}. Thus the instantaneous configuration size of a BRP can be quickly in the weak-to-intermediate shear flow regime (region I, where
characterized by the radius of gyration tensor Rg, which is determined by γ̇ ≤ 0.02) but slowly in the intermediate-to-strong shear flow regime
Rg 2 = Gxx + Gyy + Gzz . The ensemble averaged gyration tensors of fully (region II, where γ̇ > 0.02) as the shear rate increases. However, <Gyy>
stretched BRPs at various shear flow rates are shown in Fig. 2. As shown and <Gzz> for the BRP decrease slowly in region I but quickly in region

Fig. 1. The snapshot and monomer density distributions of a (a,b) RP, (c,d) BRP and (e,f) the ring backbone of the BRP. Here the horizontal axis for (b), (d) and (f)
has been adjusted by the configuration size.

175
D. Wang et al. International Journal of Biological Macromolecules 227 (2023) 173–181

II and the similar trends for the RP (l* = 0) are shown in Fig. 2b, structure the BRP has and the response of the BRP to applied shear flow
indicating that the BRPs' and RPs' responses to applied shear field field is nonmonotonic. To explore the nonmonotonic increase of relative
dramatically change as the shear rate increases into a strong shear flow extension on the BRP's structure, we fit the results with the power law
regime. Notably, in Fig. 2a, the increasing speed of <Gxx> in region I < Rg 2 > / < Rg 2 >0 ∼ γ̇α . The results as a function of the BRP's relative
(< Gxx >∼ γ̇1/2 ) and region II (< Gxx >∼ γ̇ 1/3 ) are similar to the branch size l* are shown in Fig. 3b, confirming that there is a non­
decreasing speed of <Gzz> in region II (< Gzz >∼ γ̇ − 1/3 ) and region I monotonic regime between l* = 0.5 and 1.5, indicating the stable
(< Gzz >∼ γ̇− 1/2 ), respectively. Meanwhile, <Gyy> decreases slower structure exists for the BRP when the relative branch size l* is close to
than <Gzz> in region I but faster than <Gzz> in region II, indicating that, 1.0. Furthermore, the nonmonotonic region exists between l* = 0 and
in the flow-vorticity plane, the response of the BRP to the applied shear 0.3, implying there are different dynamical and conformational prop­
flow is on a new stage in the intermediate-to-strong shear flow regime. erties for the RP and the BRP. Meanwhile, in this region, the exponent α
However, for the BRP that has short arms (see Fig. S5), this trend van­ of the BRP is smaller than that of the RP, confirming that ring polymers
ishes. Furthermore, the power law dependence of <Gzz> on shear rate that have short branches are more compact and less deformable which
(< Gzz >∼ γ̇− 1/3 ) for the BRP in region I and the RP in region II are equal has been revealed by Roh et al. [39].
to the scaling results for isolated polymer chains [5], confirming that the The alignment of configuration of a BRP is characterized by the
polymer size plays a key role in the polymer's dynamical and confor­ orientation angle χ G, which is determined by
mation properties. 2Gxz
tan(2χ G ) = . (7)
The changes on the squared-rescaled radius of the gyration tensor < Gxx − Gzz
2 2
Rg > / < Rg >0 for various BRPs and the RP as varying shear rate are We plot in Fig. 4 the alignment parameter tan(2χ G) as a function of
shown in Fig. 3a. As presented in Fig. 3a, for both the RP and BRPs, the the shear rate. For BRPs that have small size (l* ≤ 0.4), the amplitude of
squared-rescaled radius < Rg 2 > / < Rg 2 >0 increases fast in the weak- tan(2χ G) follows a power law with tan(2χ G ) ∼ γ̇ − 1/3 in the strong flow
to-intermediate flow regime but very slow in the intermediate-to- regime (see Fig. 4a), which is similar to the result for star polymers [16]
strong flow regime, confirming that the structure of the polymer and ring polymers [35]. However, as the shear rate increases further, the
under shear flow is sensitive to changes on shear rate in the weak-to- nonmonotonic part exits for RPs (red-circle plot in Fig. 4a). A similar
intermediate flow regime but insensitive in the intermediate-to-strong behavior has been reported by [35]. For BRPs that have intermediate-to-
flow regime. However, as the shear rate increases further from a me­ large branch size (l* ≥ 0.5), Fig. 4b shows that the amplitude of tan(2χ G)
dium value, the squared-rescaled radius < Rg 2 > / < Rg 2 >0 for the BRP also follows the power law tan(2χ G ) ∼ γ̇− 1/3 in the intermediate shear
that has very large size (l* = 2.0) keeps a fast linear increasing trend,
suggesting that the longer arm the BRP has, the more deformable

Fig. 3. (a) The squared and scaled radius of gyration tensor < Rg 2 > / < Rg 2 >0 Fig. 4. Alignment angles tan(2χ G) of BRPs as a function of shear rate γ̇ for
as a function of shear rates for several BRPs, and (b) the exponent of power-law relatively (a) small-size BRPs (l* ≤ 0.4) and (b) large-size BRPs (l* ≥ 0.5). The
dependence of the squared gyration radius < Rg 2 > / < Rg 2 >0 , on the shear dashed lines indicate the power-law dependence of alignment angles on the
rate γ̇ as a function of the relative branch size l*. shear rate.

176
D. Wang et al. International Journal of Biological Macromolecules 227 (2023) 173–181

flow regime, indicating a universal response of BRPs as the shear rate the closed shape of the ring backbone makes it possible to construct a
grows. The result also implying that, rather than the polymer size, the three-dimension surface. We construct the surface of the ring backbone
random motion of arms plays the key role in conformational properties and calculate its normal vector by the method proposed by Jeong et al.
of the BRP when the polymer starts to respond to shear flow field as the [28], the details of surface constructing are presented in Fig. S6. The
shear rate grows. However, in the strong shear flow regime, the expo­ components of normal vector, Nβ, where β ∈ {x, y, z}, of the surface of the
nent for power-law dependence of tan(2χ G) on γ̇ increases as the relative ring backbone of a BRP, which is the same BRP in Fig. 5a, as a function of
branch size l* grows, indicating that, as the shear rate increases further, time are plotted in Fig. 5b. As shown in Fig. 5b, in regime I, the regular
the structure and configure of the BRP is dramatically shifted by the periodic fluctuations of Nz (the component of normal vector along the
strong shear flow field, resulting in a very small alignment angle. velocity-gradient direction) coupled with the periodic changes of Nx (the
Meanwhile, the random movement of arms is highly restricted by the component of normal vector along the flow direction) clearly show the
highly compressed structure, leading to more stable dynamical motions tumbling motions of the BRP. The decreasing amplitude of Nz in regime
of the BRP. II implies the emerging of tank-treading motions and the non-periodic
Adding to the BRP's complex dynamical behaviors, are the BRP's fluctuations of Nz and Nx in regime III implies the tumbling motion no
random shape fluctuations formed by random movement of arms and longer exits. Moreover, the random changes between negative and
two possible motions: (1) periodic stretched-folded-stretched state positive value for Nz imply that the occasional tumbling motion exists in
transition on the polymer's shape (the polymer always exhibits obvious regime III, where the steady tank-treading motion exists.
shape fluctuations), which is known as the tumbling motion; (2) the It has been shown in Fig. 5 that the normal vector of the surface,
polymer's rotational motion around the center of mass as the compresses which is constructed by the closed shape of the ring backbone, exhibits
structure keep stable, which is known as the tank-treading motion. stronger periodic fluctuations than the gyration tensor as the BRP ex­
Especially, for BRPs, the tank-treading motion exists on the ring back­ hibits steady tumbling motions. Thus, rather than using the traditional
bone can be described by the motion of monomers that circulate along method, which considers the auto-correlation of conformational fluc­
the contour of the ring [29]. The temporal fluctuations on the structure, tuations reflected by the gyration tensor [11], we obtain the charac­
which are reflected by the changes on gyration tensor as time goes on, of teristic time for tumbling dynamics by determining the cross-correlation
a BRP under shear flow are shown in Fig. 5a. The steady periodic between the periodic fluctuations of the normal vector in the flow and
changes of Gxx/ < Gxx > − 1, here <⋯> means the time average of Gxx, the velocity-gradient direction, which is defined by
in regime I are caused by the BRP's tumbling motions, the obscure pe­
〈δNx (t0 ) 〉〈δNz (t0 + t) 〉
riodic changes on Gxx/ < Gxx > − 1 and the decreasing of Gyy − Gzz in Cxz (t) = √〈̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
〉〈 〉̅. (8)
regime II are caused by the coexistence of tumbling and tank-treading δNx 2 (t0 ) δNz 2 (t0 )
motions, and the little fluctuations of Gxx/ < Gxx > − 1 and Gyy − Gzz
imply the stable tank-treading motions. Fig. 6 shows the cross-correlation functions of a BRP and RP for
When investigating tumbling and tank-treading motions of BRPs, it is several various shear rates. In Fig. 6, the tumbling motion of the BRP is
convenient to only focus on the motions of the ring backbone, because not strictly periodic due to the auto-correlation function in Fig. 6 always
decays to zero at large time-lags [29], similar to linear polymer chains
[11]. The characteristic time of the tumbling motion is defined as TTB =

Fig. 5. (a) Gxx/ < Gxx > − 1 (red) and Gyy − Gzz (blue) for a BRP under shear
flow as a function of time. (b) Temporal fluctuations of normal vector of three-
dimensional surface, which is constructed from the ring backbone of the BRP. Fig. 6. Cross-correlation function of (a) a BRP with branch size l* = 1.0 and (b)
Here the branch size of the BRP is l* = 1.25 and the shear strength is γ̇ = 0.5. a RP at various shear rates.

177
D. Wang et al. International Journal of Biological Macromolecules 227 (2023) 173–181

2(t+ − t− ), where t+ represents the deep minimum of auto-correlation due to the very large conformation. Furthermore, for l* = 1.5, the
function at time t > 0 and t− is the location of pronounced peak at exponent increases to 1.05, which is much further from the result of RPs,
time t < 0 [11]. It has been found that traditional method fails to in the strong flow regime. Fig. S8 clearly shows the decreasing trend and
characterize the tumbling time at high shear rate [35]. Interestingly, the increasing trend on the exponent of power-law dependence in the in­
velocity-gradient-direction component of normal vector always keeps termediate and strong flow regime, the results are in consistent with the
periodic fluctuations when the BRP or RP exhibits tumbling motions, results for alignment angle (see Fig. 4). In Fig. 7b, the plateau appears
implying this novel method can naturally address this problem when and quickly moves toward the lower shear rate as the arm length grows
RPs or BRPs are under very strong shear flow. The existence of charac­ further (l* > 1.5), indicating the tumbling frequency becomes inde­
teristic time of periodic tumbling motion of linear polymers and ring pendent of γ̇. Similar behaviors for capsules [74] and star polymers [16]
polymers can be determined by the presence of a peak in the power have been observed in the strong flow regime.
spectral density (PSD) of orientation angle [4,33]. Thus, another way to Fig. 8 shows the tumbling frequency ωTB plotted as a function of
find the characteristic time TTB is determining the PSD of the normal relative branch size l* for several shear rates. As shown in Fig. 8, for
vector's z-component, Nz. The representatives of PSDs of RPs and BRPs at BRPs that have very small relative branch size (l* ≤ 0.1), the tumbling
various shear rates are shown in Fig. S7a and b, the existence and shifts frequency increases as the arm length increases, which corresponds to
of clear peaks are observed for PSDs. Meanwhile, the characteristic time the decrease of α. It confirms that the short branches have a function of
determined by the PSD (the reciprocal of peak frequency) and cross- increasing the resistance on structural deformation, which is similar to
correlation as a function of shear rate are plotted in Fig. S7c and d. branched linear polymers [30,38,40]. Correspondingly, it results in a
Clearly, for both RP (l* = 0.0) and BRP (l* = 1.75), the results deter­ more compact structure (see Fig. 3b) that makes the polymer easier to
mined by two methods are consistent in the strong shear flow regime. perform fast tumbling motions. The decreasing trend on tumbling fre­
The characteristic frequency of tumbling motions, ωTB, which can be quency as l* increases further from 0.1 corresponds to the trend on
determined by cross-correlation of components of the normal vector scaling results for squared-rescaled radius of the gyration tensor (see
along the flow and the velocity-gradient direction, for several BRPs are Fig. 3b), indicating the strong correlation between conformational size
plotted in Fig. 7. As shown in Fig. 7a, in the intermediate flow regime, and the tumbling behaviors for BRPs. Notably, the frequencies at
the exponent of power-law dependence of the frequency ωTB on γ̇, different shear rates always keep a steady gap as the relative branch size
ωTB ∼ γ̇0.68 , for BRPs is close to that of the RP. The result is also l* smaller than 1.5, implying that BRPs have consistent tumbling mo­
consistent with the result, ωTB ~ Wi2/3, where Wi is the Weissenberg tions under various intermediate-to-strong shear flow while the arm
number which is proportion to γ̇, obtained from linear, ring and star length is short-to-intermediate and the tumbling frequency is mainly
polymers [1,4,7,11,29]. Shifting into the strong flow regime, the expo­ determined by the polymer size. However, the overlaps on frequencies
nent of power law for the BRP is much larger than the RP and gradually exist as l* increases further from 1.5 because the tumbling frequency
grows further as the arm length increases, indicating the dynamics of reaches the upper limit as the shear rate increases in the intermediate
BRPs are more complicated than RPs. As shown in Fig. 7b, the exponent shear flow regime. More details on the changes of the tumbling fre­
decreases to about 0.5 in the intermediate flow regime, implying the quency in the weak-to-intermediate shear flow regime are shown in
BRP that has long arms (l* ≥ 1.5) is hard to perform tumbling motions Fig. S9.
It is well-known that not only soft matter objects [33,75], but also
various types of polymers [3,16,29,34,37] present stable tank-treading
motions when applying steady shear flow field. For BRPs, the steady
tank-treading motion is also observed in the strong shear flow regime.
The frequency of tank-treading motion of the BRP can be obtained by
determining the autocorrelation of the angle, θ, between the longitudi­
nal axis of the instantaneous conformation and the vector connecting
marked monomer and the center of mass, a method proposed by Chen
et al. [29]. The angular autocorrelation function is defined as
〈θ(t0 ) 〉〈θ(t0 + t) 〉
Cθ (t) = 〈 2 〉 (9)
θ (t0 )

and then the characteristic tank-treading time TTT is determined by

Fig. 7. Tumbling frequency ωTB as a function of shear rate γ̇ for relatively (a)
small-size BRPs (l* ≤ 0.4) and (b) large-size BRPs (l* ≥ 1.5). The dashed lines Fig. 8. Tumbling frequency ωTB as a function of the BRP's relative branch size
indicate the power-law dependence of tumbling frequency on shear rate. l* at various shear rates.

178
D. Wang et al. International Journal of Biological Macromolecules 227 (2023) 173–181

the twice the time when the secondary peak of autocorrelation function
Cθ(t) appears [29]. In the strong shear flow regime, the tank-treading
frequency ωTT as a function of shear rate γ̇ for BRPs are plotted in
Fig. 9. Clearly, the peak and trend of ωTT are close to that of the tumbling
frequency ωTB (see Fig. 7), because the characteristic time of the tank-
treading motion is consistent with that of the tumbling motion when
the combination of these two motions exists. In fact, as the combined
motion which contains the tumbling and tank-treading motion exists,
each monomer's one complete circular motion relative to the center of
mass of the polymer not only represents a complete cycle of the tumbling
motion but also implies a cycle of the tank-treading motion. In the weak-
to-intermediate shear flow regime, the tank-treading motion of the BRP
is not obvious enough to observe. We ascribe this result to the symmetric
distribution of arms along the ring backbone which makes the tank-
treading motion of monomers that circulate along the contour of the
ring backbone much more difficult. The dependence of the tank-treading
frequency ωTT on the shear rate γ̇ is ωTT ∼ γ̇ 0.93 for small-size BRPs and Fig. 10. Tank-treading frequency ωTT as a function of the BRP's relative branch
the exponent of the power-law dependence increases to about 1.0, which size l* at various shear rates.
represents a linear dependence, as the relative branch size l* grows.
Interestingly, an almost linear dependence of the tank-treading fre­ to Fig. 8, indicating that the tumbling and tank-treading behaviors are
quency on the shear rate was observed for red cells in shear flow [76]. mainly determined by the polymer size and these motions have an upper
This similarity indicates that the large-size BRP behaves like vesicles in limit on frequency which decreases as the relative branch size l* in­
the strong shear flow regime. Similar to the tumbling motion (see Fig. 7), creases. Meanwhile, for l* ≤ 0.5, compared plots with Fig. 8, the smaller
the frequency of the tank-treading motion for BRPs that have very long amplitudes of error bars in Fig. 10 imply that the tank-treading motion is
branches (l* ≥ 1.75) reaches a plateau as the shear rate increases. more stable than tumbling motion for BRPs that have small size.
Fig. 10 shows the tank-treading frequency ωTT plotted as a function Furthermore, comparing Fig. 9b with Fig. 7b, the smaller amplitudes of
of relative branch size l* for several shear rates. The increasing trend of error bars in Fig. 7b indicate that tumbling motions are more stable than
ωTT for small BRPs (l* ≤ 0.1) is similar to that of ωTB in Fig. 8, confirming tank-treading motions for large-size BRPs (l* ≥ 1.5).
that BRPs have less deformable structures in response to the applied
shear flow field, which can be ascribed to the random and fast move­ 4. Conclusions
ments of arms [38], leading to a more fast and stable tank-treading
motion like the tumbling. Clearly, trends of plots in Fig. 10 are similar In summary, the nonequilibrium single-polymer dynamics of BRPs
over a wide range of relative branch sizes and shear rates have been
systematically studied by using a hybrid mesoscale simulation approach.
As the branch size of the BRP increases, the increasing trend of the gy­
ration tensor along flow direction and decreasing trend along the
velocity-gradient direction gradually displays two phases, indicating a
more intensive compression on BRPs in the strong shear flow regime.
Notably, the decreasing of gyration tensor along the velocity-gradient
direction always follows the power law < Gzz >∼ γ̇ − 1/3 observed for
RPs in first phase and this result corresponds to the decreasing trend of
orientation angle, indicating that the random motion of arms plays the
key role in the stretching dynamics when the BRP starts to be affected by
shear flow field as the shear rate grows. Moreover, by analyzing the size
fluctuations of BRPs as the branch size grows, we find that the polymer
size increases monotonically with increasing shear rates but non­
monotonically with increasing branch length when the relative branch
size near 0 and 1.0, which means that the arm length equals the ring
backbone length.
Furthermore, to find more representative information reflecting the
tumbling motion and overcome the fail that analyzing tumbling fre­
quency of RPs by tracing correlation of fluctuations on the gyration
tensor in the strong shear flow regime [35], we construct three-
dimensional surface of ring backbone and analyze the temporal fluctu­
ations on the surface's normal vector along trajectory. A novel cross-
correlation function, which is the correlation between fluctuations of
the normal vector along the flow direction and the velocity-gradient
direction, is proposed to analyze the tumbling motion that usually co­
exists with tank-treading motion in the strong shear flow regime. For
BRPs with small-to-intermediate size, the increasing trend of tumbling
frequency in the intermediate flow regime, which follows the power law
ωTB ∼ γ̇0.68 , is similar to that of linear, ring and star polymers
Fig. 9. Tank-treading frequency ωTT as a function of shear rate γ̇ for relatively [1,4,7,11,29]. Moreover, as the relative branch size increases, the
(a) small-size BRPs (l* ≤ 0.4) and (b) large-size BRPs (l* ≥ 1.5). The dashed increasing speed of ωTB decreases in the intermediate flow regime but
lines indicate the power-law dependence of the tumbling frequency on the
increases in the strong flow regime. However, for BRPs with very long
shear rate.

179
D. Wang et al. International Journal of Biological Macromolecules 227 (2023) 173–181

branches, the tumbling frequency will quickly reach the upper limit as [10] P. LeDuc, C. Haber, G. Bao, D. Wirtz, Dynamics of individual flexible polymers in a
shear flow, Nature 399 (1999) 564–566.
the shear rate increases. Such behaviors are very similar to capsules [74]
[11] C.-C. Huang, G. Sutmann, G. Gompper, R. Winkler, Tumbling of polymers in
and star polymers [16]. Furthermore, the tank-treading frequency is semidilute solution under shear flow, Europhys. Lett. 93 (2011) 54004.
obtained from the autocorrelation function of the orientation angle and [12] S. Jeong, J.M. Kim, S. Cho, C. Baig, Effect of short-chain branching on interfacial
shows similar increasing trend to the tumbling motion as varying the polymer structure and dynamics under shear flow, Soft Matter 13 (2017)
8644–8650.
shear rate and the relative branch size. By comparing the statistical re­ [13] T. Yamamoto, N. Masaoka, Numerical simulation of star polymers under shear flow
sults of frequencies of the tumbling and tank-treading motion for various using a coupling method of multi-particle collision dynamics and molecular
BRPs, we find that the tank-treading motion is more stable than tum­ dynamics, Rheol. Acta 54 (2014) 139–147.
[14] A. Nikoubashman, C.N. Likos, Branched polymers under shear, Macromolecules 43
bling motion for small-branch-size BRPs but the tumbling motion is (2010) 1610–1620.
more stable than tank-treading motion for large-branch-size BRPs. [15] M. Ripoll, R. Winkler, G. Gompper, Hydrodynamic screening of star polymers in
The molecular characteristic of BRPs revealed in this paper may help shear flow, Eur. Phys. J. E: Soft Matter Biol. Phys. 23 (2007) 349–354.
[16] M. Ripoll, R. Winkler, G. Gompper, Star polymers in shear flow, Phys. Rev. Lett. 96
for analyzing the dynamical behaviors and mechanisms of complex flow (2006), 188302.
induced structural changes of RPs and BRPs in shear flow. Such [17] L.J. Kasehagen, C.W. Macosko, Nonlinear shear and extensional rheology of long-
knowledge could be useful for guiding drug designing and trans­ chain randomly branched polybutadiene, J. Rheol. 42 (1998) 1303–1327.
[18] S.P. Singh, A. Chatterji, G. Gompper, R.G. Winkler, Dynamical and rheological
portation in body fluid and polymeric material designing for industrial properties of ultrasoft colloids under shear flow, Macromolecules 46 (2013)
processing. 8026–8036.
[19] P.C. Hiemenz, T.P. Lodge, Polymer Chemistry, CRC Press, 2007.
[20] P.-G.De Gennes, P.-G. Gennes, Scaling Concepts in Polymer Physics, Cornell
CRediT authorship contribution statement university press, 1979.
[21] T.C.B. McLeish, Tube theory of entangled polymer dynamics, Adv. Phys. 51 (2002)
Deyin Wang: Conceptualization, Formal analysis, Investigation, 1379–1527.
[22] M. Trabi, D.J. Craik, Circular proteins—no end in sight, Trends Biochem. Sci. 27
Methodology, Software, Writing – original draft. Xiaohui Wen: (2002) 132–138.
Conceptualization, Formal analysis, Funding acquisition, Methodology, [23] T. Sanchez, I. Kulic, Z. Dogic, Circularization, photomechanical switching, and a
Resources, Supervision. Dong Zhang: Formal analysis, Writing – review supercoiling transition of actin filaments, Phys. Rev. Lett. 104 (2010), 098103.
[24] T. Cremer, C. Cremer, Chromosome territories, nuclear architecture and gene
& editing. Xinguan Tan: Formal analysis, Writing – review & editing.
regulation in mammalian cells, Nat. Rev. Genet. 2 (2001) 292–301.
Jiajun Tang: Formal analysis, Writing – review & editing. [25] J.D. Halverson, J. Smrek, K. Kremer, A.Y. Grosberg, From a melt of rings to
chromosome territories: the role of topological constraints in genome folding, Rep.
Prog. Phys. 77 (2014), 022601.
Declaration of competing interest [26] A.A. Vinogradov, Y. Yin, H. Suga, Macrocyclic peptides as drug candidates: recent
progress and remaining challenges, J. Am. Chem. Soc. 141 (2019) 4167–4181.
[27] D. Wang, F.-Q. Li, X.-H. Wang, S.-B. Li, L.-L. He, Effects of chain stiffness and shear
The authors declare that they have no known competing financial flow on nanoparticle dispersion and distribution in ring polymer melts, J. Zheijang
interests or personal relationships that could have appeared to influence Univ. Sci. A 21 (2020) 229–239.
the work reported in this paper. [28] S.H. Jeong, S. Cho, E.J. Roh, T.Y. Ha, J.M. Kim, C. Baig, Intrinsic surface
characteristics and dynamic mechanisms of ring polymers in solution and melt
under shear flow, Macromolecules 53 (2020) 10051–10060.
Acknowledgements [29] W. Chen, J. Chen, L. An, Tumbling and tank-treading dynamics of individual ring
polymers in shear flow, Soft Matter 9 (2013) 4312–4318.
[30] S. Jeong, S. Cho, J.M. Kim, C. Baig, Interfacial molecular structure and dynamics of
This study was supported by the National Natural Science Founda­ confined ring polymer melts under shear flow, Macromolecules 51 (2018)
tion of China (Grant No. 11504033). 4670–4677.
[31] J. Yoon, J. Kim, C. Baig, Nonequilibrium molecular dynamics study of ring polymer
melts under shear and elongation flows: a comparison with their linear analogs,
Appendix A. Supplementary data J. Rheol. 60 (2016) 673–685.
[32] W. Chen, Y. Li, H. Zhao, L. Liu, J. Chen, L. An, Conformations and dynamics of
Snapshots and monomer density distributions for BRPs (Fig. S1-S4); single flexible ring polymers in simple shear flow, Polymer 64 (2015) 93–99.
[33] M.Q. Tu, M. Lee, R.M. Robertson-Anderson, C.M. Schroeder, Direct observation of
the ensemble averaged gyration tensor of BRPs (Fig. S5); the construc­ ring polymer dynamics in the flow-gradient plane of shear flow, Macromolecules
tion process of intrinsic surface of a ring backbone (Fig. S6); the power 53 (2020) 9406–9419.
spectral densities (PSDs) and the tumbling frequency of BRPs (Fig. S7); [34] W. Chen, H. Zhao, L. Liu, J. Chen, Y. Li, L. An, Effects of excluded volume and
hydrodynamic interaction on the deformation, orientation and motion of ring
the tumbling frequency of for various BRPs (Fig. S8); the tumbling fre­ polymers in shear flow, Soft Matter 11 (2015) 5265–5273.
quency of BRPs at various shear rates (Fig. S9). Supplementary data to [35] Z. Wang, Q. Zhai, W. Chen, X. Wang, Y. Lu, L. An, Mechanism of nonmonotonic
this article can be found online at doi:https://doi.org/10.1016/j.ijbi increase in polymer size: comparison between linear and ring chains at high shear
rates, Macromolecules 52 (2019) 8144–8154.
omac.2022.12.100
[36] M. Liebetreu, C.N. Likos, Hydrodynamic inflation of ring polymers under shear,
Commun. Mater. 1 (2020).
References [37] S.H. Jeong, T.Y. Ha, S. Cho, E.J. Roh, J.M. Kim, C. Baig, Melt rheology of short-
chain branched ring polymers in shear flow, Macromolecules 54 (2021)
10350–10359.
[1] R.G. Winkler, Semiflexible polymers in shear flow, Phys. Rev. Lett. 97 (2006),
[38] S.H. Jeong, S. Cho, T.Y. Ha, E.J. Roh, C. Baig, Structural and dynamical
128301.
characteristics of short-chain branched ring polymer melts at interface under shear
[2] A. Celani, A. Puliafito, K. Turitsyn, Polymers in linear shear flow: a numerical
flow, Polymers 12 (2020) 3068.
study, Europhys. Lett. 70 (2005) 464.
[39] E.J. Roh, J.M. Kim, C. Baig, Molecular dynamics study on the structure and
[3] P.S. Lang, B. Obermayer, E. Frey, Dynamics of a semiflexible polymer or polymer
relaxation of short-chain branched ring polymer melts, Polymer 175 (2019)
ring in shear flow, Phys. Rev. E 89 (2014), 022606.
107–117.
[4] C.M. Schroeder, R.E. Teixeira, E.S. Shaqfeh, S. Chu, Characteristic periodic motion
[40] J.M. Kim, C. Baig, Communication: role of short chain branching in polymer
of polymers in shear flow, Phys. Rev. Lett. 95 (2005), 018301.
structure and dynamics, J. Chem. Phys. 144 (2016), 081101.
[5] I. Saha Dalal, A. Albaugh, N. Hoda, R.G. Larson, Tumbling and deformation of
[41] M.R. Awual, A novel facial composite adsorbent for enhanced copper (ii) detection
isolated polymer chains in shearing flow, Macromolecules 45 (2012) 9493–9499.
and removal from wastewater, Chem. Eng. J. 266 (2015) 368–375.
[6] C. Aust, S. Hess, M. Kröger, Rotation and deformation of a finitely extendable
[42] M.R. Awual, Assessing of lead (iii) capturing from contaminated wastewater using
flexible polymer molecule in a steady shear flow, Macromolecules 35 (2002)
ligand doped conjugate adsorbent, Chem. Eng. J. 289 (2016) 65–73.
8621–8630.
[43] M.R. Awual, Solid phase sensitive palladium (ii) ions detection and recovery using
[7] R.E. Teixeira, H.P. Babcock, E.S. Shaqfeh, S. Chu, Shear thinning and tumbling
ligand based efficient conjugate nanomaterials, Chem. Eng. J. 300 (2016)
dynamics of single polymers in the flow-gradient plane, Macromolecules 38 (2005)
264–272.
581–592.
[44] M.R. Awual, Ring size dependent crown ether based mesoporous adsorbent for
[8] S. Gerashchenko, V. Steinberg, Statistics of tumbling of a single polymer molecule
high cesium adsorption from wastewater, Chem. Eng. J. 303 (2016) 539–546.
in shear flow, Phys. Rev. Lett. 96 (2006), 038304.
[45] M.R. Awual, Novel nanocomposite materials for efficient and selective mercury
[9] D.E. Smith, H.P. Babcock, S. Chu, Single-polymer dynamics in steady shear flow,
ions capturing from wastewater, Chem. Eng. J. 307 (2017) 456–465.
Science 283 (1999) 1724–1727.

180
D. Wang et al. International Journal of Biological Macromolecules 227 (2023) 173–181

[46] M.R. Awual, A. Islam, M.M. Hasan, M.M. Rahman, A.M. Asiri, M.A. Khaleque, M. [61] T. Ihle, D. Kroll, Stochastic rotation dynamics. Ii. Transport coefficients, numerics,
C. Sheikh, Introducing an alternate conjugated material for enhanced lead (ii) and long-time tails, Phys. Rev. E 67 (2003), 066706.
capturing from wastewater, J. Clean. Prod. 224 (2019) 920–929. [62] C.-C. Huang, G. Gompper, R.G. Winkler, Hydrodynamic correlations in
[47] M.R. Awual, A facile composite material for enhanced cadmium (ii) ion capturing multiparticle collision dynamics fluids, Phys. Rev. E 86 (2012), 056711.
from wastewater, J. Environ. Chem. Eng. 7 (2019), 103378. [63] A. Malevanets, J. Yeomans, Dynamics of short polymer chains in solution,
[48] M.R. Awual, M.M. Hasan, A ligand based innovative composite material for Europhys. Lett. 52 (2000) 231.
selective lead (ii) capturing from wastewater, J. Mol. Liq. 294 (2019), 111679. [64] M. Ripoll, K. Mussawisade, R. Winkler, G. Gompper, Low-Reynolds-number
[49] M.R. Awual, M.M. Hasan, J. Iqbal, M.A. Islam, A. Islam, S. Khandaker, A.M. Asiri, hydrodynamics of complex fluids by multi-particle-collision dynamics, Europhys.
M.M. Rahman, Ligand based sustainable composite material for sensitive nickel (ii) Lett. 68 (2004) 106.
capturing in aqueous media, J. Environ. Chem. Eng. 8 (2020), 103591. [65] T. Ihle, D. Kroll, Stochastic rotation dynamics: a galilean-invariant mesoscopic
[50] M.A. Islam, M.J. Angove, D.W. Morton, B.K. Pramanik, M.R. Awual, A mechanistic model for fluid flow, Phys. Rev. E 63 (2001), 020201.
approach of chromium (vi) adsorption onto manganese oxides and boehmite, [66] T. Ihle, D.M. Kroll, Stochastic rotation dynamics. I. Formalism, galilean invariance,
J. Environ. Chem. Eng. 8 (2020), 103515. and green-kubo relations, Phys. Rev. E 67 (2003), 066705.
[51] K.T. Kubra, M.S. Salman, M.N. Hasan, A. Islam, M.M. Hasan, M.R. Awual, Utilizing [67] A. Lees, S. Edwards, The computer study of transport processes under extreme
an alternative composite material for effective copper (ii) ion capturing from conditions, J. Phys. C Solid State Phys. 5 (1972) 1921.
wastewater, J. Mol. Liq. 336 (2021), 116325. [68] C.-C. Huang, A. Chatterji, G. Sutmann, G. Gompper, R.G. Winkler, Cell-level
[52] M.R. Awual, M.M. Hasan, A. Islam, M.M. Rahman, A.M. Asiri, M.A. Khaleque, M. canonical sampling by velocity scaling for multiparticle collision dynamics
C. Sheikh, Introducing an amine functionalized novel conjugate material for toxic simulations, J. Comput. Phys. 229 (2010) 168–177.
nitrite detection and adsorption from wastewater, J. Clean. Prod. 228 (2019) [69] J. Padding, A. Louis, Hydrodynamic interactions and brownian forces in colloidal
778–785. suspensions: coarse-graining over time and length scales, Phys. Rev. E 74 (2006),
[53] M.R. Awual, M.M. Hasan, A.M. Asiri, M.M. Rahman, Cleaning the arsenic (v) 031402.
contaminated water for safe-guarding the public health using novel composite [70] C.R. Bartels, B. Crist Jr., L.J. Fetters, W.W. Graessley, Self-diffusion in branched
material, Compos. Part B 171 (2019) 294–301. polymer melts, Macromolecules 19 (1986) 785–793.
[54] H.M. Munjur, M.N. Hasan, M.R. Awual, M.M. Islam, M. Shenashen, J. Iqbal, [71] K.R. Shull, E.J. Kramer, L.J. Fetters, Effect of number of arms on diffusion of star
Biodegradable natural carbohydrate polymeric sustainable adsorbents for efficient polymers, Nature 345 (1990) 790–791.
toxic dye removal from wastewater, J. Mol. Liq. 319 (2020), 114356. [72] J.E. Martin, D. Adolf, Diffusion in branched polymer melts, Macromolecules 22
[55] M. Naushad, A.A. Alqadami, A.A. Al-Kahtani, T. Ahamad, M.R. Awual, (1989) 4309–4311.
T. Tatarchuk, Adsorption of textile dye using Para-aminobenzoic acid modified [73] E. Von Meerwall, D. Tomich, N. Hadjichristidis, L.J. Fetters, Phenomenology of
activated carbon: kinetic and equilibrium studies, J. Mol. Liq. 296 (2019), 112075. self-diffusion in star-branched polyisoprenes in solution, Macromolecules 15
[56] M. Bishop, M. Kalos, H. Frisch, Molecular dynamics of polymeric systems, J. Chem. (1982) 1157–1163.
Phys. 70 (1979) 1299–1304. [74] Y. Navot, Elastic membranes in viscous shear flow, Phys. Fluids 10 (1998)
[57] K. Kremer, G.S. Grest, Dynamics of entangled linear polymer melts: a molecular- 1819–1833.
dynamics simulation, J. Chem. Phys. 92 (1990) 5057–5086. [75] K.-I. Tsubota, S. Wada, H. Liu, Elastic behavior of a red blood cell with the
[58] A. Malevanets, R. Kapral, Solute molecular dynamics in a mesoscale solvent, membrane’s nonuniform natural state: equilibrium shape, motion transition under
J. Chem. Phys. 112 (2000) 7260–7269. shear flow, and elongation during tank-treading motion, Biomech. Model.
[59] K. Mussawisade, M. Ripoll, R. Winkler, G. Gompper, Dynamics of polymers in a Mechanobiol. 13 (2014) 735–746.
particle-based mesoscopic solvent, J. Chem. Phys. 123 (2005), 144905. [76] T.M. Fischer, M. Stöhr-Liesen, H. Schmid-Schönbein, The red cell as a fluid droplet:
[60] A. Malevanets, R. Kapral, Mesoscopic model for solvent dynamics, J. Chem. Phys. tank tread-like motion of the human erythrocyte membrane in shear flow, Science
110 (1999) 8605–8613. 202 (1978) 894–896.

181

You might also like