ArticlePDF Available

Dissipative particle dynamics simulations for biological systems: From protein structures to cell mechanics (In Chinese)

Authors:

Abstract and Figures

Dissipative particle dynamics (DPD) is a mesoscopic coarse-grained simulation method developed in recent years, which is an important method for studying the dynamic behaviors of soft matter and complex fluids. In this method, each DPD particle represents a coarse-grained virtual cluster of a set of atoms or groups of matter. The position and momentum of the DPD particle are updated in a continuous phase but spaced at discrete time steps. The coarse-grained DPD particles are subject to simplified pairwise interacting conservative, dissipative, and random forces. Especially, the dissipative force and random force in the DPD method act as heat sink and source, respectively, and the combined effect of these two forces act as a thermostat, which conserves momentum and thus provides the correct description of hydrodynamic interactions for the model system. In addition, a common choice of the soft repulsion for the conservative force allows using larger integration time steps in DPD simulations than that is usually allowed in classical molecular dynamics (MD) simulation. Hence, compared with the MD method, the computational cost of the DPD simulation is significantly reduced due to the smaller number of modeled particles and the larger computational time step, enabling the simulations of the static and dynamic behaviors of complex fluids and soft matter systems at physically attractive length scale and time scale. Moreover, the particle-based framework of the DPD method enables people to easily incorporate additional physical features into the model systems and extend its application to complex systems. For these reasons, the DPD method and its extensions have been successfully applied to numerous soft matter and complex fluid systems such as oil/water/surfactant systems, chemical morphology, microscopic morphology, phase separation, as well as dynamics and rheological properties of polymer solutions and colloidal suspensions. In this paper, we first introduce the theoretical formulation and parameterization of the DPD method. Then, we review recent advances in DPD modeling of biological systems, focusing on its applications at the molecular and cellular scales. At the molecular scale, we highlight examples of successful simulations of the protein structures and their interactions, the structure and dynamics of amphiphilic lipid molecule membranes (e.g., the self-assembly of lipid molecules, the structure and properties of lipid membranes, the fusion of lipid membrane, and the budding and fission of lipid membrane), the interaction of lipid membranes with protein molecules, and the interaction of nanoparticles with lipid membranes; at the cellular scale, we focus on the DPD modeling of blood cell flow and blood rheological behavior in the blood microcirculatory systems, including the shape deformation and flow dynamics of red blood cells, the margination and adhesion dynamics of white blood cells, the margination and aggregation behaviors of platelets, the hemorheological behavior of blood flow, as well as the separation of circulating tumor cells from blood flow using microfluidic devices. Additionally, we compare the advantages and disadvantages of the microscale blood flow simulations between the continuum-based methods and particle-based methods, including the DPD method. Finally, we briefly investigate the development trends and application prospects of DPD modeling in biological systems.
Content may be subject to copyright.
耗散粒子动力学方法在生物学领域的应用与研究进展:
从蛋白质结构到细胞力学
唐梓涵,李学进*,李德昌*
浙江大学工程力学系,杭州 310012
*联系人, E-mail: xuejin_li@zju.edu.cn;dcli@zju.edu.cn
2022-09-05 收稿, 2022-11-30 修回, 2022-11-30 接受, 2022-12-01 网络版发表
国家自然科学基金(12122212, 12072318, 11932017)资助
摘要 耗散粒子动力学(dissipative particle dynamics, DPD)是近年发展起来的一种介观尺度的数值模拟方法,是研
究软物质和复杂流体动力学行为的一种重要手段.这种新型介观模拟方法采用粗粒化粒子模型描述具有关联性的
原子团或物质团,并通过简单的软排斥作用力描述粗粒化粒子间的相互作用,从而实现更大时间和空间尺度的复
杂系统模拟计算,如油//表面活性剂体系聚合物和胶体溶液的化学形态微观形貌相分离以及复杂流体流
变特性的模拟等.本文首先介绍了DPD方法的理论框架,继而详细综述了DPD方法在生物系统中的应用.具体地,
在分子尺度,我们重点介绍了该方法在蛋白质结构及其相互作用两亲性脂质分子膜的结构与动力学脂质膜与
蛋白分子相互作用纳米颗粒与脂质膜相互作用等方面的研究现状和研究热点.在细胞尺度,我们归纳了DPD
法在模拟血液微循环系统中血细胞的流动和血液流变学行为等方面的应用进展,包括红细胞的变形及流动,白细
胞边聚及黏附行为,血小板边聚黏附及聚集行为,健康与疾病状态下血液流变学特征,循环肿瘤细胞迁移黏附
及分选富集等.此外,我们总结了用于模拟血细胞变形及血液流动的其他数值模型并进行了简单比较.最后,我们
简单展望了DPD方法在生物学领域的发展趋势和应用前景.
关键词 耗散粒子动力学,蛋白质,脂质膜,血细胞,血液流动
耗散粒子动力学(dissipative particle dynamics,
DPD)是一种介观的粗粒化模拟方法[1~3],最初主要用
于描述复杂流体的动力学行为.近年来, DPD方法被广
泛应用于理解蛋白质脂质膜,甚至细胞以及复杂生
物流体等的行为.在该方法中,具有关联性的原子团被
粗粒化表示成DPD粒子,通过赋予该粒子不同的亲疏
水性质和相互作用,可有效模拟复杂流体系统的平衡
态和非平衡态行为及形态学特征.
DPD方法作为一种粗粒化模拟方法,其粒子运动
由简单的运动学方程描述:
rfmt
d
d= , (1)
i
i
i
2
2
其中,mi为第i个粒子的质量,
fi
为粒子i所受的力,
ri
粒子i的位置.粒子i受到的作用力
fi
由保守力
Fij
C
散力
Fij
D
和随机力
Fij
R
组成[1~3]:
f F F F= ( + + ), (2)
ii j ij ij ij
C D R
其中,
F r
F v r r
F r
a r
r
r t
= ( ) ,
= ( )( ) ,
= ( ) ( ) ,
(3)
ij ij ij ij
ij ij ij ij ij
ij ij ij ij
C
D2
R1/ 2
引用格式:唐梓涵,李学进,李德昌.耗散粒子动力学方法在生物学领域的应用与研究进展:从蛋白质结构到细胞力学.科学通报, 2023, 68: 741–761
Tang Z H, Li X J, Li D C. Dissipative particle dynamics simulations for biological systems: From protein structures to cell mechanics (in Chinese). Chin Sci
Bull, 2023, 68: 741–761, doi: 10.1360/TB-2022-0913
© 2022《中国科学》杂志社 www.scichina.com csb.scichina.com
2023 68 7: 741 ~ 761
https://engine.scichina.com/doi/10.1360/TB-2022-0913
其中,aij为粒子ij之间的排斥作用参数,描述粒子ij
之间软排斥作用的强度.一般而言,aij可以定性或定量
给出.比如,一般设定DPD水粒子与水粒子间awater-water
25, 若其他粒子为疏水性粒子,则取aij大于25, 其取
值越大,疏水性越强;反之,若其他粒子为亲水性粒子,
则取aij小于或等于25, 其取值越小,亲水性越强.定性给
aij的方法在常见的粒子模型中得到了广泛应用,
如常见的脂质分子聚合物等系统,可以给出一些定
性的结果[4].另一方面,为了定量确定DPD粒子间的aij,
可根据DPD粒子的Florry-Huggins(χij)参数给定排斥作
用参数aij=aii+3.27χij
[5].一般而言,可以使用分子动力
学模拟计算DPD粒子的溶解度参数δi
[6],χij=V(δi–δj)2/
RT;或者使用蒙特卡洛模拟计算DPD粒子的混合能[7],
从而得到Florry-Huggins参数,χijEij/RT.(3),rij
粒子ij之间的距离,
rij
是相应的单位向量,
v v v=
ij i j
是粒子ij的相对速度,ξij表示均值为0
差为1的随机数.γσ分别为耗散和噪声强度系数,
取值满足关系σ2=2γkBT,其中kB为玻尔兹曼常数,T为绝
对温度.ω(rij)为权函数:
rr r r r
r r
( ) = 1 / , < ,
0, , (4)
ij
ij ij
ij
C C
C
其中,rC为截断半径.为了描述更复杂的分子结构和相
互作用,系统中还可以引入粒子间的成键相互作用(
括键伸缩作用键角弯折作用及二面角扭转作用)
短程吸引(Lennard-Jones相互作用)以及长程的静电
相互作用等.
DPD方法由于采用了相对于传统分子动力学(mo-
lecular dynamics, MD)更进一步的粗粒模型和更软的粒
子间相互作用势及动量守恒的热浴形式,可实现更大
时间与空间尺度的模拟计算,已被广泛应用于生物颗
(包括液滴胶粒细胞外泌体和生物大分子
)两亲性聚合物表面活性剂简单和复杂流体
等领域[8,9].本文专注于介绍DPD方法在生物系统中的
应用及研究进展,包括蛋白质结构及其相互作用
质分子及脂质膜系统,以及血液和血细胞流动系统中
的单细胞与多细胞行为等,1所示.
1蛋白质结构及其相互作用
蛋白质是构成生命的重要基础物质之一,由氨基
酸通过脱水缩合形成肽键构成链状结构.蛋白质氨基
酸具有相同的骨架(backbone)原子,根据侧链基团不同
而分为20.氨基酸脱水缩合形成的链状结构通过相
互作用进一步形成蛋白质特定的二级三级和四级结
,从而行使其特定功能.蛋白质的功能与其结构和动
力学密切相关, MD模拟方法可在原子尺度有效模拟蛋
白质分子的构象转变及其相互作用[11,12].但是,目前原
1(网络版彩色)耗散粒子动力学方法在生物系统中的应用.右上角插图修改自文献[10]
Figure 1 (Color online) Dissipative particle dynamics simulations of biological systems. The inset in the upper-right corner is adapted from Ref. [10]
2023 3 68 7
742 https://engine.scichina.com/doi/10.1360/TB-2022-0913
子尺度的MD模拟所能达到的时间和空间尺度与生理
条件下蛋白质分子功能实现的尺度尚有差距[13]. DPD
方法由于其固有的粗粒化表示方式以及简单的相互作
,可有效降低模拟系统的自由度,提高计算效率,
而实现蛋白质系统较大空间尺度和较长时间尺度的
模拟.
与粗粒化分子动力学模型类似[14,15],蛋白质分子
DPD模拟大多以氨基酸为基本单位进行模型构建,
2所示.粗粒化分子动力学模拟和DPD模拟在原理
上很相似.两者都是将拓扑关系相近的原子(或分子)
不同的粗粒化程度合并成一个粒子来描述,平均化处
理粗粒化粒子间相互作用.所不同的是,粗粒化分子动
力学模拟采用势函数描述粒子间相互作用,通过势函
数对坐标的导数求解作用力.DPD方法则直接给定
粒子间相互作用的保守力耗散力和随机力. DPD
法通过保守力将粒子描述成软球粒子,其粗粒化程度
更高,采用的作用力描述相对更简单,可使用更大的时
间步长,因此可模拟更大空间尺度和更长时间尺度的
系统.例如,为了精细描述蛋白质分子的二级结构,
Vishnyakov等人[16]用一个DPD粒子描述氨基酸的骨架
基团,并根据氨基酸侧链基团的不同性质,把氨基酸侧
链基团划分成1~2个具有不同亲疏水属性的DPD粒子,
2(a)所示.在该模型中, Vishnyakov等人[16]通过引
1-3(氨基酸nn+2)1-5(氨基酸nn+4)粒子间
Morse势来模拟肽链形成螺旋结构时氨基酸形成氢键
的相互作用.更进一步,通过控制骨架nn+1n+2
子间的弯曲刚度,实现了肽链α螺旋和β片等二级结构
的描述,并通过不同pH下氨基酸侧链的不同质子化状
态引入粒子电量[17],成功模拟了不同pH条件下α突触
核蛋白(α-synuclein)的构象转变[16].另一方面,根据蛋
白质分子的晶体结构和全原子分子动力学模拟结果获
取结构参数,是构建蛋白质分子DPD模型的有效方
[18~20].例如, ChoudhuryKuksenok[19]以全原子MD
模拟的聚丙氨酸肽链(poly-Ala)α螺旋结构为模板,
建了可描述α螺旋折叠的DPD模型,2(b)所示.
们通过全原子MD模拟获得了聚丙氨酸肽链α螺旋结构
氨基酸Cα原子间的键长键角二面角等参数,用于
构建DPD模型.除此之外,通过调整DPD模型α螺旋结
构中形成接触对(contact-pair)粒子间的排斥参数aij
角势能刚度Kθ二面角势能刚度Kφ,使得DPD模拟
α螺旋结构与全原子结果相匹配,从而实现DPD模型
的参数优化.根据该模型, ChoudhuryKuksenok[19]
功模拟了聚丙氨酸肽链从无序结构到α螺旋结构折叠
的过程,并尝试模拟了具有6α螺旋结构的蛋清溶菌
(hen egg white lysozyme)的稳定性.更具一般性地,
VaiwalaAyappa[20]利用包含20种氨基酸且具有典型
α螺旋ββ发夹(β-hairpin)等二级结构短肽的全
原子分子动力学模拟结果作为模板,获取DPD模型中
的键长键角二面角等参数,并引入一个具有双势
阱结构的势函数来表征序列中相邻3个氨基酸骨架粒
子间的弯曲作用.该双势阱势能函数的平衡角度分别
对应α螺旋和β片结构的状态,从而可实现α螺旋和β
之间的自发变构.此外, VaiwalaAyappa[20]还根据
Flory-Huggins参数[3,21]给出了20种氨基酸各个DPD
子间的排斥参数aij,使得该模型可方便地与脂质分子
DPD模型联合使用,用于研究蛋白-脂质膜系统.
2(网络版彩色)蛋白质DPD模型示意图.以氨基酸为基本单位. (a) 多个DPD粒子描述1个氨基酸[16], Copyright © 2012 American Chemical
Society. (b) 1DPD粒子描述1个氨基酸[19], Copyright © 2020 American Chemical Society
Figure 2 (Color online) Illustration of the DPD models for proteins. Take amino acids as the basic unit. (a) One residue is represented by multiple
coarse-grained particles[16], Copyright © 2012 American Chemical Society. (b) One residue is represented by one coarse-grained particle[19], Copyright
© 2020 American Chemical Society
743
https://engine.scichina.com/doi/10.1360/TB-2022-0913
模型可有效稳定蛋白质的二级结构,表征氨基酸的局
部接触,以及模拟蛋白质的折叠过程. DingMa[22,23]
根据不同氨基酸侧链具有的不同亲疏水和电荷性质,
将每个氨基酸采用一个DPD粒子来描述.20种氨基
酸划分为4种不同的DPD粒子:亲水粒子(GlySer
AsnGlnThr)疏水粒子(LeuMetAlaPhe
ProTrpValIle)带正电的粒子(LysArgHis)
和带负电的粒子(TyrCysAspGlu), 研究了血清蛋
白与纳米颗粒相互作用形成纳米颗粒-蛋白质冠(pro-
tein corona)结构的动力学过程,分析了纳米颗粒被血清
蛋白包裹后与细胞膜相互作用的过程,并与裸露的纳
米颗粒与细胞膜相互作用进行了详细比较.Ding
Ma[22,23]的研究中,仅需考虑血清蛋白的亲疏水性与纳
米颗粒和细胞膜的相互作用.因此,模型中表示氨基酸
DPD粒子仅用简单的弹簧键连接,未考虑血清蛋白
的二级结构等性质,可以大大降低模型构建的复杂度
和计算效率.
上述模型中,仅用1~3DPD粒子描述1个氨基酸,
极大地降低了模拟系统的自由度,因此可以模拟蛋白
质分子长时间尺度的构象变化(α螺旋与β片之间的
变构), 以及蛋白质分子与纳米颗粒的相互作用.为了模
拟更大空间和时间尺度的蛋白质动力学过程, Lin
[24]Rim等人[25]进行了更粗粒化的描述,他们采用1
DPD粒子描述3个氨基酸,构建了丝蛋白的DPD
,研究了丝蛋白通过纺丝作用形成丝纤维的过程.
该模型中,丝蛋白的肽链用多嵌段共聚物链描述,链上
1DPD粒子描述肽链上连续的3个氨基酸.为了描述
丝蛋白β片晶体(β-sheet crystalline)与无定形结构域
(amorphous domain)交联形成的网络结构,形成β片晶
体的氨基酸被定义为疏水粒子,无定形结构域的氨基
酸则为亲水粒子.形成β片晶体的疏水粒子在疏水作用
下倾向于形成团聚状态的节点”, 用于描述β片晶体内
氨基酸因形成大量氢键而形成的团聚结构.无定形结
构域由于用亲水粒子描述,在体系中倾向于伸展结构,
从而形成连接节点的”. 通过该模型, Lin等人[24]
Rim等人[25]模拟了丝蛋白自组装形成以β片晶体为节
无定形域为连接桥的网络结构过程,并研究了丝
蛋白序列长度,以及流体剪切作用对丝纤维形成及
其力学性能的影响. Shen等人[26]基于上述丝蛋白DPD
模型自组装形成的丝纤维网络结构,构建了包含多个β
片晶体和无定形结构域相互交联的丝蛋白网络结构全
原子模型,研究了预拉伸对丝纤维力学性能的影响及
其机理. Pan等人[27]在该DPD模型基础上引入碳纳米
,研究了碳纳米管对丝蛋白纤维的增强增韧机理.
据以氨基酸为基本单元的不同粗粒化表示,我们把典
型的蛋白分子DPD模型总结于1.
2两亲性脂质分子膜的结构与动力学
脂质分子是组成细胞膜内质网膜细胞核膜等
生物膜结构的主要成分,具有重要的生理功能[29].脂质
分子通常具有两亲性结构,包括亲水的头部基团(如磷
脂分子的头部基团–PO4–)和疏水的尾部基团(如磷脂分
子的尾部基团–CH2–). 与粗粒化分子动力学模型类似,
脂质分子通常以基团的不同亲疏水性质被划分成少量
的粗粒化粒子[30~35],这些粒子一般通过简单的弹簧键
连接,3(a)所示.
相比于传统的粗粒化分子动力学模拟方法, DPD
方法具有更简单的相互作用势函数,因此被广泛应用
于脂质分子膜的研究,包括脂质分子的自组装行为
脂质膜的结构与性质脂质膜的融合脂质膜的分裂
和出芽等.
2.1 脂质分子自组装
两亲性的脂质分子自组装在生物系统中具有重要
作用.脂质分子的自组装过程涉及大量的两亲性脂质
分子在溶液中长时间的动力学过程.因此, DPD模拟适
合用于模拟研究该过程[32~34].例如, Yamamoto等人[32]
采用DPD模拟研究了两亲性分子的自组装行为,发现
两亲性分子首先形成扁平的胶束结构,进而自发形成
囊泡结构,并且其结果显示,双尾的脂质分子由于比单
尾的相遇概率更高,其自组装形成囊泡的速度更快.
Qiang等人[33]考虑不同浓度下两亲性脂质分子的自组
装过程,结果显示,在不同脂质分子浓度下,通过调整
脂质分子亲水头部基团长度和疏水尾部基团长度可调
控自组装形成的平衡态构象,包括球形胶束脂质双
分子层膜带孔的脂质双分子层膜囊泡和圆柱体等
结构.更有意思的是, Arai等人[34]利用DPD模拟研究了
油酸分子(oleic acid)和二棕榈酰磷脂酰胆碱(dipalmi-
toylphosphatidylcholine, DPPC)磷脂分子在温度和浓度
共同调控下自组装形成不同尺寸和形状的囊泡等生物
膜结构的机理,用于理解脂质分子形成类细胞结构过
.类似于两亲性的脂质分子自组装过程, Wang
[36]研究了端部基团的性质和形状对两亲性聚合物自
组装结构的影响,其模拟结果显示,可通过引入端部基
2023 3 68 7
744 https://engine.scichina.com/doi/10.1360/TB-2022-0913
团调控自组装囊泡的壁厚和内腔尺寸. Tan等人[37,38]
研究了树枝状两亲性聚合物的自组装行为,发现调整
聚合物分子的分支度分枝长度和不同溶剂环境等因
素可获得球状胶束蠕虫状胶束螺旋状胶束囊泡
等孤立结构及其聚集体.上述结果显示, DPD方法可简
单表示两亲性脂质分子的结构和相互作用,可有效模
拟大规模两亲性分子的自组装行为,为理解自组装结
构的形成过程及其机理提供帮助,并进一步为设计具
有特定结构的生物膜结构提供理论指导.
2.2 脂质膜的结构与性质
利用DPD模拟可有效研究脂质分子的类型
温度溶剂等条件对自组装动力学过程的影响.
除此之外, DPD方法还被用于研究脂质分子构型
成等对生物膜的结构和性质的影响[30,31,39~41].例如,
ShillcockLipowsky[39]采用DPD模拟研究了具有单尾
链和双尾链的脂质双分子层的结构与力学性质,计算
了脂质双分子层的表面张力,估算了典型组分脂质双
分子层的面积模量和弯曲刚度,该数值与实验结果一
.另外,该模拟结果还得到了与粗粒化分子动力学模
拟一致的脂质膜侧向应力分布,DPD模拟所需计算
量要小得多.该结果显示, DPD模拟可有效研究比分子
动力学模拟更大空间尺度和时间尺度的两亲性脂质分
子的平衡动力学行为,为研究复杂的亚细胞系统开辟
了道路[39].基于类似的模型, Illya等人[40]进一步研究了
脂质分子尾链长度和脂质分子两根尾链长度对称性对
膜结构和性质的影响.为了模拟脂质双分子层面积模
量与脂质分子尾链长度不相关的实验结果[42], Illya
[40]在模型中设置脂质分子头部DPD粒子相互排斥的
作用强度随着尾链长度增加而增加,从而得到与实验
一致的结果.模拟结果显示,两根尾链长度不相同的脂
质分子形成的双分子层具有较小的面积模量和弯曲模
[40].更进一步, Illya等人[41]通过调整脂质分子DPD
子间排斥作用参数aij来区分两种不同的脂质分子,
究不同脂质分子混合形成的双分子层和囊泡的结构.
在典型磷脂DPD模型的基础上, de Meyer等人[30,31]构建
了胆固醇的DPD模型,3(a)所示.利用该模型, de
Meyer等人[30,31]系统研究了不同温度下胆固醇浓度对
二肉豆蔻酰磷脂酰胆碱(dimyristoylphosphatidylcholine,
DMPC)形成的磷脂膜中的凝聚效应的影响.结果显示,
胆固醇浓度和温度共同调控了磷脂双分子层的结构形
(3(b)). 为了验证该模拟的可靠性, de Meyer
[30,31]定量地研究了脂质分子膜的结构和力学性能,
如膜的厚度弯曲刚度单位磷脂面积磷脂的序参
数等,并且与全原子模拟和实验数据进行了多方面的
对比,用简单的模型展示了DPD模拟方法的有效性和
优势.
2.3 脂质膜结构的融合
脂质膜融合是细胞的重要生理过程,如神经递质
的传递细胞间的通讯脂质体包裹药物进入目标细
胞等过程.这些过程通常涉及脂质囊泡与细胞膜的融
.细胞内脂质囊泡一般在几十纳米的尺寸,有的甚至
达到微米量级.融合过程一般在几百纳秒到微秒,甚至
秒的时间尺度.由于目前计算能力的限制,利用全原子
模型进行脂质膜融合模拟所能达到的空间尺度和时间
尺度均有限.因此, DPD方法常被用于研究大规模和大
尺度下的脂质膜融合机制[43~45].例如, ShillcockLi-
powsky[43]通过在等温等容(NVT)系综下减少囊泡和平
面膜上磷脂个数来改变膜张力,研究了膜张力调控下
囊泡和平面膜的融合过程.得益于DPD模拟的高效性,
ShillcockLipowsky[43]进行了93次独立的模拟.模拟
1典型的蛋白分子DPD模型
Table 1 The summary of typical DPD models for proteins
氨基酸的粗粒化模型 文献 研究问题
主链和侧链分别用多个DPD粒子描述
[16] 蛋白质α螺旋与β片的构象转变
[18] 开发极化的蛋白模型和力场参数
[20] 开发20种天然氨基酸的DPD结构和通用力场
单个DPD粒子描述1个氨基酸 [19] 蛋白的α螺旋折叠
[22,23] 人血清蛋白介导下的纳米颗粒与生物膜作用
单个DPD粒子描述3个氨基酸 [24] 丝蛋白通过纺丝作用形成丝纤维的过程
非以氨基酸为单位设置DPD粒子 [28] 细胞膜介导下蛋白的疏水错配引起的蛋白倾斜和对膜的扰动
745
https://engine.scichina.com/doi/10.1360/TB-2022-0913
结果显示,在膜张力较小的情况下,没有观察到囊泡和
平面膜的融合,囊泡总是黏附在平面膜上.当膜张力增
,囊泡和平面膜在接触区域形成半融合结构,使得膜
张力释放,从而稳定构型,未能成功融合.继续增加膜
张力,囊泡和平面膜在100 ns时间尺度下成功融合.
膜张力增加到囊泡和平面膜保持完整性的临界值,
泡和平面膜将发生破裂而释放膜张力,不发生融合过
.基于类似的模型, Gao等人[44]分析了不同张力下的
融合路径. Grafmüller等人[45]详细分析了膜融合的3
子过程:囊泡和平面膜接触区域单个磷脂的翻转和插
入过程半融合结构形成过程半融合区域融合孔打
开过程.由于DPD模拟的高效性,可通过多次独立模拟
统计上述3个过程完成需要的平均时间,从而估算该过
程需要克服的能量势垒[45].
与两亲性脂质分子膜融合过程类似, DPD模拟还
被广泛应用于两亲性嵌段聚合物形成的囊泡和膜融合
过程[46~48].Li等人[46]GuoYang[47]以及Liu等人[48]
分别模拟了三嵌段聚合物支状嵌段聚合物和星型三
嵌段聚合物形成的囊泡膜融合过程.
2.4 脂质膜的出芽和分裂
出芽和分裂是与脂质膜融合相反的一种生理行为,
也广泛存在于细胞生理过程,如神经细胞中突触囊泡
的形成细胞外泌体囊泡的形成细胞器如高尔基体
形成囊泡输运蛋白质等.在脂质膜出芽和分裂的DPD
模拟中,一般考虑多种不同脂质组分的相互作用差异
导致脂质膜形态改变,该形态改变将进一步导致脂质
膜出芽和分裂[49~53].例如, YamamotoHyodo[49]采用
相同的脂质分子结构不同的排斥参数aij研究了两种
不同脂质分子AB形成的囊泡出芽与分裂过程.在该
研究中,通过调整囊泡内外膜脂质分子AB的含量,
展示了囊泡双分子层内外膜脂质分子分布不对称对囊
泡形态的影响.首先,两种不同的脂质分子在膜内出现
相分离现象,形成了两种脂质分子聚集的区域.内外层
脂质分子分布不对称降低了脂质膜的弯曲刚度,从而
促进出芽形成,并进一步导致出芽子囊泡和母囊泡连
接处颈部的分裂,完成子囊泡释放,3(c)左侧图所
.另一方面, YamamotoHyodo[49]通过增加AB
3(网络版彩色)脂质分子的DPD模型及其典型的应用. (a) 典型的脂质分子DPD模型[31].从左到右依次为DMPC脂质分子的化学结构示意
DMPC脂质分子相应的DPD模型胆固醇分子的化学结构示意图胆固醇分子相应的DPD模型表示, Copyright © 2010 American Chemical
Society. (b) 胆固醇浓度和温度调控的DMPC脂质双分子层结构[31]. LαLoPβ′Lβ′Lc′分别表示无序结构有序结构波纹结构脂质分子
倾斜的有序结构和非倾斜的有序结构. (c) 基于DPD方法模拟两种典型的膜出芽和分裂方式[49]
Figure 3 (Color online) The DPD models for lipids and their typical applications. (a) Illustrations of the typical DPD models for lipids[31]. The images
from left to right are the schematic diagram of the chemical structure of a DMPC lipid molecule, the corresponding DPD model of the DMPC lipid
molecule, the schematic diagram of the chemical structure of a cholesterol molecule, and the corresponding DPD model of the cholesterol molecule,
respectively. Copyright © 2010 American Chemical Society. (b) Structures of DMPC bilayers regulated by the concentration of cholesterol and
temperature[31]. The symbols Lα, Lo, Pβ′, Lβ′, and Lc′ represent the disordered structure, ordered structure, rippled structure, tilted and ordered structure,
and nontilted ordered structure, respectively. (c) Two typical pathways of membrane budding and fission based on DPD simulations[49]
2023 3 68 7
746 https://engine.scichina.com/doi/10.1360/TB-2022-0913
质分子间的排斥参数aij来模拟脂质分子AB聚集相之
间界面能的增加,或者升高系统温度,发现了不同的出
芽分裂路径:出芽后在AB界面处发生脂质膜分裂,
而后分裂的脂质膜A聚集相和B聚集相分别愈合,从而
实现子囊泡的释放,3(c)右侧图所示.基于类似的
模型, LaradjiKumar[50~52]重点研究了两种不同脂质分
子在膜上的相分离现象,以及出芽和分裂过程.由于脂
质膜限制了水分子的渗透, LaradjiKumar[50~52]通过调
整囊泡内水分子的数量来调控囊泡的体积,从而改变
脂质囊泡的面积-体积比,发现较大的面积-体积比更有
利于子囊泡的出芽和分裂.更进一步, LaradjiKu-
mar[50~52]通过改变两种脂质分子间的排斥参数aij来调
整脂质膜的线张力,研究了线张力对出芽的影响.与脂
质分子相类似, Li等人[46]在研究三嵌段聚合物形成的
囊泡分裂过程中,采用逐渐增大两种聚合物之间排斥
参数aij的方式成功模拟了一个囊泡自发分裂成两个近
乎等大子囊泡的过程,并分析了膜张力等因素对该过
程的影响.
在生理情况及实验中,脂质膜的出芽过程一般是
在较大尺寸膜上发生,如细胞膜内质网膜等.由于模
拟尺寸的限制,上述研究主要关注小尺寸囊泡上的出
芽过程,囊泡自身的曲率会对出芽过程有一定影响.
一方面,由于有限的脂质分子个数,出芽过程中膜张力
会发生较大变化.为了解决上述问题, Hong等人[53]重点
研究了平面膜的出芽过程.为了解决出芽过程中膜张
力变化问题, Hong等人[53]引入了粒子可变的DPD模拟
方法(N-varied DPD). 在该方法中, Hong等人[53]监测模
拟体系边界脂质膜的分子密度,当边界局部分子密度
下降到临界值,则在该处增加脂质分子;反之,当该处
分子密度超过临界值,则删除脂质分子.如此, Hong
[53]成功实现了在平面膜上恒定膜张力的出芽过程
模拟.
3脂质膜与蛋白相互作用
细胞膜通常分布着各种蛋白,如转运蛋白融合
蛋白黏附蛋白等.各种蛋白在膜上的生理功能也各
不相同,蛋白与细胞膜的相互作用调控着蛋白在细胞
膜上的形态分布等状态,显著影响蛋白的功能.如脂
质膜通过膜的局部曲率脂阀膜形态的热涨落等影
响膜上蛋白-蛋白间相互作用[54],脂质膜与蛋白相互作
用可调控蛋白分子在膜上的分布,一些蛋白的病理性
聚集有可能侵入脂质膜造成脂质膜破坏引起细胞毒性
[55].科研人员基于DPD方法对细胞膜介导下的蛋白
行为进行了一系列的研究[28,56~60].
Venturoli等人[28]研究了跨膜蛋白由于其疏水区域
长度和脂质膜厚度的失配(hydrophobic mismatch, 疏水
长度失配)而导致脂质双分子层变形,以及该失配效应
引起的蛋白在膜内的姿态问题.由于跨膜蛋白一般呈
现螺旋结构,在该模型中, Venturoli等人[28]用简单的圆
柱状结构表示跨膜蛋白,蛋白的两端为亲水基团,中间
段为疏水区域.利用该模型, Venturoli等人[28]研究了蛋
白尺寸(包括圆柱形蛋白结构的直径和疏水区域的长
)对膜结构的影响.模拟结果发现,当蛋白疏水区域长
度大于脂质双分子层疏水厚度时,由于蛋白与脂质分
子尾部的疏水相互作用,可导致脂质双分子层局部增
;反之,当蛋白疏水区域长度小于脂质双分子层疏水
厚度时,脂质双分子层局部变薄.另一方面,当蛋白直
径较小时,可通过倾斜补偿疏水长度失配导致的能量
损失,即蛋白通过倾斜一定角度,使更多的疏水区域埋
入脂质双分子层的疏水区;当蛋白直径较大,则通过改
变脂质膜局部的厚度来使更多的脂质分子尾部与蛋白
疏水区域相互作用.该模型被广泛用于研究多个蛋白
插入膜内由于疏水尺寸失配介导的聚集行为[56~58]. de
Meyer等人[59]在该模型基础上,采用伞状采样方法计算
了两个蛋白在不同直径和不同疏水区长度下相互靠近
的平均力势曲线,发现疏水长度失配是跨膜蛋白聚集
的驱动力. MorozovaWeiss[60]通过在圆柱形蛋白模型
的疏水区域端部添加一段疏水分支来模拟跨膜蛋白氨
基酸乙酰化对该蛋白在膜上行为的影响.模拟结果显
,乙酰化导致蛋白在膜中的倾斜度增加,从而减弱了
由于蛋白和膜疏水长度失配介导的蛋白质聚集行为.
另一方面,由于高效性, DPD方法被广泛应用于研
究蛋白或表面活性剂等对磷脂膜结构的影响[61~67].
GrootRabone[61]首先尝试采用DPD方法研究不带电
的表面活性剂对磷脂膜的穿孔破坏效应,讨论其抗菌
机理.与表面活性剂类似,抗菌肽等短肽蛋白具有类似
的两亲性结构, DPD方法能有效模拟这些小分子在大
空间尺度和长时间尺度下的动力学行为[63~67].Chen
等人[66]对比研究了8种具有α螺旋β片和无定形结构
短肽在不同浓度下对膜结构的影响.在该研究中, Chen
等人[66]采用Vishnyakov等人[16]稳定蛋白二级结构的方
,并引入带正电氨基酸与带负电磷脂的静电相互作
[68],详细分析了短肽的带电性两亲性二级结构
等性质对磷脂膜穿孔褶皱,以及短肽跨膜运动等行
747
https://engine.scichina.com/doi/10.1360/TB-2022-0913
为的影响.
4纳米颗粒与脂质膜相互作用
纳米颗粒常被用作药物输运的载体[69].为了精确
和高效地把药物输运到细胞内部,纳米颗粒与细胞膜
的相互作用至关重要.对纳米颗粒与细胞膜相互作用
的理解是药物设计的重要一环.纳米颗粒与细胞膜作
用一般包括吸附包裹和穿透等几个方面,主要由纳
米颗粒与细胞膜的亲疏水相互作用和静电相互作用等
主导. DPD方法由于其描述粒子间亲疏水相互作用固
有的优势和高效的计算效率,是理解纳米颗粒与脂质
膜相互作用的重要方法之一.
DPD方法首先被广泛应用于研究纳米颗粒形状对
其与脂质膜相互作用的影响.例如, YangMa[70]系统
研究了球状椭球状半椭球状棒状盘状和图钉
状等不同形状的纳米颗粒与脂质膜的相互作用.在该
研究中,他们采用典型的脂质分子DPD模型,纳米颗粒
则由亲水粒子组成相应的形状.为了描述纳米颗粒与
细胞膜的相互作用,在模型中还采用Lennard-Jones
函数来描述纳米颗粒粒子与脂质头部DPD粒子间的相
互作用[70].在该模型中,通过拉力使纳米颗粒穿过脂质
,以纳米颗粒能穿透脂质膜最小的拉力来评价该纳
米颗粒的穿透能力.其模拟结果发现,纳米颗粒形状的
各向异性和初始与膜接触角度对纳米颗粒与脂质膜相
互作用有关键影响,纳米颗粒侵入脂质膜的能力由纳
米颗粒与脂质膜的接触面积和纳米颗粒在接触点处的
局部曲率来主导.基于YangMa[70]的研究结果, Wang
等人[71]通过实验的手段,制备了具有肿瘤靶向和pH
应的棒状混合纳米材料,提高了药物的穿透能力.在另
外的模拟中, DingMa[72]同样用DPD模拟的方法研究
pH敏感聚合物修饰下的纳米颗粒与细胞膜的相互作
,发现pH敏感聚合物修饰下的纳米颗粒可以增强对
细胞膜的穿透能力.类似地, Li等人[73]通过改变纳米颗
粒表面部分粒子和部分脂质分子的相互作用,来模拟
纳米颗粒表面修饰有配体而脂质膜表面存在受体的情
,研究了受体-配体作用介导下不同形状纳米颗粒与
脂质膜的相互作用. YueZhang[74]通过在纳米颗粒表
面显式布置棒状结构用于描述配体分子,在脂质膜中
用圆柱状结构模拟配体分子,研究了受体-配体作用介
导下纳米颗粒进入脂质膜的过程.结果发现,根据膜张
配体面积密度和纳米颗粒大小的不同,可以观察
4种情况:膜破裂纳米颗粒黏附纳米颗粒穿透和
受体-配体介导的内吞.有意思的是, Chen等人[75]通过
在纳米颗粒表面引入电量,发现纳米颗粒表面电荷随
机分布可导致纳米颗粒在脂质膜表面定向扩散.更进
一步, DPD方法还被广泛用于研究Janus颗粒与脂质膜
相互作用[76~80]. Janus颗粒是表面具有两种或两种以上
不同性质的纳米颗粒,如典型的Janus颗粒,一半表面
是亲水属性,另一半表面是疏水属性.由于Janus颗粒
表面具有可控的不同属性,其与脂质膜的相互作用可
通过控制纳米颗粒表面属性来调控,使其在药物输运
方面备受关注[81]. DingMa[76]研究了具有球状和椭球
一半表面为亲水属性另一半表面为疏水属性的
Janus颗粒在受体-配体作用介导下与磷脂膜的相互作
.他们发现, Janus颗粒初始与膜接触表面的属性对颗
粒进入膜的方式具有重要影响.另一方面,通过调整磷
DPD粒子间的排斥参数aij, DingMa[76]模拟了Janus
颗粒与脂阀(lipid raft)的作用,探讨了Janus颗粒完全穿
透磷脂膜的途径.更进一步, Li等人[80]模拟了三嵌段球
Janus纳米颗粒与脂质膜相互作用,考虑了三嵌段表
面不同亲疏水分布和面积对颗粒穿过磷脂膜的影响.
Wang等人[79]则通过二维的三角形网格包裹一个球状
的颗粒构造了核-壳结构的纳米颗粒,通过调整三角形
网格的二面角弯曲刚度和面内剪切刚度可调控核-
结构纳米颗粒的刚度,从而研究了不同刚度纯亲水
纯疏水和Janus纳米颗粒穿过磷脂膜的力学过程.其结
果显示,对于纯亲水纳米颗粒,较大的刚度有利于其侵
入磷脂膜;而对于纯疏水的纳米颗粒,较大的刚度则抑
制其侵入磷脂膜. Yan等人[82~85]还采用DPD方法系统研
究了一种更软的纳米颗粒——树状大分子与脂质膜的
相互作用.
上述研究展示了DPD方法在研究纳米颗粒与脂质
膜相互作用方面的广泛应用.研究结果显示,不同的纳
米颗粒形状尺寸刚度表面性质等因素对纳米颗
粒与脂质膜相互作用具有显著影响.对纳米颗粒与脂
质膜相互作用机理的理解将有助于指导纳米颗粒的改
,促进更高效和精准的药物输运.系统研究纳米颗粒
改性对药物输运的影响受限于人力物力资源实验技
实验条件等因素.由于DPD模拟方法的灵活性,
可高效应用于模拟各种条件的改性对纳米颗粒与脂质
膜作用的影响,指导高效精准药物输运的纳米颗粒设
[23,72,86~88].马余强团队[23,72,86~88]采用DPD模拟提出用
具有不同特性的聚合物与纳米颗粒结合,优化纳米颗
-聚合物复合物与细胞膜作用,从而实现纳米颗粒可
2023 3 68 7
748 https://engine.scichina.com/doi/10.1360/TB-2022-0913
控穿膜的同时减弱其对细胞膜的破坏,并有可能实现
针对肿瘤细胞的靶向输运.
纳米颗粒对细胞膜的吸附侵入和穿透可能会改
变细胞膜结构,从而造成细胞膜损伤而具有细胞毒性.
DPD方法被广泛应用于纳米颗粒对脂质膜破坏机理的
研究. Li等人[89]构建了石墨烯的DPD模型,联合实验和
全原子分子动力学模拟,发现石墨烯片可通过边缘的
尖角侵入脂质膜. Dallavalle等人[90,91]则研究了不同尺
寸的六边形石墨烯对脂质膜形态的改变,其模拟结果
显示,尺寸较小的石墨烯片可顺利侵入脂质膜,尺寸较
大的石墨烯片则以与膜平行的姿态吸附在膜表面.
于石墨烯片与脂质分子疏水端具有较强的疏水作用,
其平行吸附在膜表面导致局部大量脂质分子发生转向,
造成脂质膜结构的破坏.另一方面,只有当石墨烯片氧
化区域是随机分布时才能侵入脂质膜,而当石墨烯边
缘氧化后,石墨烯片将很难侵入.燕立唐团队[92~94]详细
报道了采用DPD模拟方法研究石墨烯片侵入脂质膜的
过程,以及形成脂质分子-石墨烯片-脂质分子三明治结
半球囊泡穿透脂质膜表面吸附等结构的分子
机理,4(a)所示. Wang等人[95]发现,石墨烯片的氧
化程度氧化位点和膜张力共同调节石墨烯基纳米片
的细胞毒性,4(b)所示.DPD模拟结果显示,未氧
化的石墨烯片由于其疏水性可快速插入脂质膜,但不
会造成脂质囊泡破坏;而氧化石墨烯片由一些未氧化
的边缘尖角区域插入脂质膜,把其他氧化区域带入脂
质膜疏水区域造成膜结构紊乱,在较大的膜张力下可
发生脂质囊泡破裂[95]. Wang等人[96]根据实验发现的碳
纳米锥结构构建模型,研究了不同尺寸不同锥度和
不同边缘氧化的碳纳米锥对脂质膜的破坏机理.此外,
根据石墨烯片插入磷脂膜后因其疏水性可抽取脂质分
子而造成细胞膜损伤的假说[97], Zhang等人[98]采用DPD
方法模拟了石墨烯片对脂质膜中胆固醇的抽取过程.
根据对脂质分子头部粒子相互作用描述的区别,我们
把典型的脂质分子DPD模型总结于2.
5血液及血细胞的流动与流变行为
从蛋白到脂质分子,再到脂质膜及脂质囊泡系统,
DPD方法展示了其在微纳尺度的广泛应用.近年来,
DPD方法被推广到更大的空间尺度应用,包括单细胞
和多细胞系统,其典型的代表是血液及血细胞的流动
与流变行为.血液由血浆和血细胞两部分组成,是一种
具有相当黏性的复杂流体.血细胞是存在于血液中的
细胞(包括红细胞白细胞血小板), 能随血液的流动
遍及全身[99].血细胞在流场中的运动同时包含了弹性
力学和流体力学问题的求解,由于细胞的运动和变形
比较大,采用传统的动网格技术求解十分复杂.因此,
有研究人员提出用固定的Euler网格求解流体力学问
,用可自由运动的Lagrange网格求解细胞的运动,
4(网络版彩色)纳米颗粒与脂质膜的相互作用. (a) 不同石墨烯尺寸膜张力受体密度情况下,石墨烯与平面脂质膜的相互作用[94].示意
HSGSLAAMMRFVIFV分别表示为半球形囊泡石墨烯三明治结构石墨烯横跨膜石墨烯吸附在膜表面膜破裂扁平
囊泡不完全扁平囊泡等结构, Copyright © 2016 American Chemical Society. (b) 不同氧化点位的石墨烯与囊泡的相互作用[95]
Figure 4 (Color online) Interactions between nanoparticles and lipid membranes. (a) Interaction of graphene with planar lipid membranes for different
graphene sizes, membrane tensions, and receptor densities[94]. The symbols HS, GS, LA, AM, MR, FV, and IFV represent the hemisphere vesicle,
graphene-sandwiched structure, graphene lying across the membrane, graphene adhering to the membrane surface, membrane rupture, flat vesicle, and
incomplete flat vesicle, respectively. Copyright © 2016 American Chemical Society. (b) Interaction of graphene at various oxidation sites with lipid
vesicles[95]
749
https://engine.scichina.com/doi/10.1360/TB-2022-0913
通过浸没边界法进行耦合.同时,该方法可以与多种模
拟方法(如格子-玻尔兹曼方法和有限体积法等)相结合,
在血细胞流动的研究中具有广泛的适用性[100].另一类
求解的思路是将细胞膜内外的流体视为可自由运动的
粒子,细胞膜用弹性网格模型来描述,流体粒子之间存
在相互作用,且运动到细胞膜处时会有反弹作用[101].
根据这种思路,人们基于DPD模拟方法发展了血细胞
多尺度弹性网络力学模型,用于研究血细胞的流动和
流变行为.例如,在红细胞弹性网络力学模型中,可将
红细胞膜离散为一系列质点(DPD粗粒化粒子), 并采用
弹性网络结构连接形成三角形网格[102].弹性网络结构
作用能采用:
Vk T x x
p x
k
n l
=(3 2 )
4 (1 ) +( 1) , (5)
j N
j j
jj
m
s
1,...,
B
2 3
p
1
s
其中,Ns是质点的个数.为了计算Vs,需要给出红细胞膜
的剪切刚度μ0:
µk T
pl x
x
x x
k n
l
=3
42(1 )
1
4(1 ) +1
4
+3 ( + 1)
4. (6)
n
0
B
m 0
0
0
3
0
2
p
0+1
红细胞膜的抗弯特性体现为弹簧网络之间夹角θj
的变化:
V k= [1 cos( )], (7)
j N
jb
1,...,
b 0
s
其中,θ0为相邻两个三角形网格之间的平衡夹角,kb
红细胞膜的弯曲常数,与红细胞膜的弯曲刚度kc之间满
k k=(2/ 3 )
b c
的关系.此外,红细胞弹性网络模型还
包含了红细胞膜表面积和体积守恒约束,用以反映红
细胞膜磷脂分子层的近似不可压缩性和细胞浆的不可
压缩流体特性,表示为
( ) ( ) ( )
V
k A A
A
k A A
A
k V V
V
=
2+2+2, (8)
j N
j
a+v
1,...,
d 0
2
0
a0
tot 2
0
tot
v0
tot 2
0
tot
t
其中,
V0
tot
分别表示红细胞膜的表面积和红细胞
体积.同时,在用DPD方法模拟研究血细胞流动及聚集
的相关工作中,人们常用Morse势函数来描述红细胞-
红细胞的聚集行为,表达式为
V r D( ) = [e 2e ], (9)
r r r r
Mors e 0
2 ( ) ( )
0 0
其中,r0是平衡作用距离,D0Morse势的阱深,常用来
表示细胞与细胞以及细胞与颗粒之间相互作用的强度.
此外,红细胞与血管内皮细胞之间的黏附作用研究对
揭示心血管疾病的发病机制具有重要意义.一般用动
力学黏附模型来描述受体-配体(receptor-ligand)在时间
域的随机键合及分离过程,其中,键合几率与断键几率
的具体表达形式为
k k l l
k T
= exp ( )
2, (10)
on on
0on 0
2
B
k k l l
k T
= exp ( )
2, (11)
off off
0off 0
2
B
其中,σonσoff是指有效的键合和断键强度.
2典型的脂质分子DPD模型
Table 2 The summary of typical DPD models for lipids
磷脂模型 文献 研究问题
磷脂由亲水的头部和疏水的尾部粒子
构成,不考虑粒子的极性
[30,31] 胆固醇与磷脂膜的二元模型
[34] 磷脂和油酸的自组装
[43] 张力介导下的脂质双层和囊泡融合
[50,51] 通过排斥参数调节线张力研究囊泡出芽
[70] 不同形状的纳米颗粒与脂质膜的作用
[76] Janus纳米颗粒与脂质膜的作用
磷脂由亲水的头部和疏水的尾部粒子
构成,考虑粒子的极性
[66,67] 抗菌肽人胰岛淀素多肽与脂质膜作用
[72] pH敏感聚合物修饰的纳米颗粒与脂质膜作用
[82,83] 带电纳米颗粒与脂质膜作用
[85] 树状大分子与脂质膜作用
2023 3 68 7
750 https://engine.scichina.com/doi/10.1360/TB-2022-0913
红细胞弹性网络力学模型的参数一般由实验测量
的细胞力学特性来确定.例如,对于健康的红细胞,
常取细胞膜的剪切刚度μ0=4.7 μN/m, 细胞膜的抗弯刚
kc=2.4×10–19 J, 细胞膜面积
=135 μm2和体积
V0
tot
=
94.5 μm3[102,103].基于DPD模拟方法建立的红细胞弹性
网络力学模型具有多尺度特征.例如,在模拟单个红细
胞时,可以用Ns=23867个质点连接而成红细胞膜的网
络结构,这样质点与质点之间的距离大约在75 nm[102,103].
另一方面,为了提高计算效率,也可以用更高粗粒化程
度的弹性网络来描述,Ns=500个质点,此时质点与质
点之间的距离大概为550 nm[103].采用类似的方法,
究人员构建了其他两类血细胞(白细胞和血小板)以及
循环肿瘤细胞多尺度弹性网络力学模型[104~106].
利用DPD方法模拟血液及其有形成分的运动行为
,它可同时模拟血细胞血浆和血管及其壁面,因此
各个组分可在模型模拟中被无缝衔接在一起.鉴于这
些事实,基于DPD方法的血细胞弹性网络力学模型已
被成功应用于健康与疾病下血细胞的流动变形模拟以
及循环肿瘤细胞的分选富集研究,如管道流中红细胞
的变形镰状细胞贫血症红细胞和白细胞的黏附动力
,以及糖尿病患者血小板的迁移和边聚行为等.
5.1 健康与疾病下红细胞变形及流动特性的DPD
模拟
红细胞也称红血球,是血液中数量最多的一种血
细胞.它具有一定的弹性和可塑性,主要功能是运输氧
二氧化碳电解质以及葡萄糖等物质.在正常的
生理状态下,红细胞因具有良好的变形性,可以顺利通
过狭窄尺寸的毛细血管网和血窦孔隙.然而,在病理状
态下,红细胞变形能力降低,高切黏度增高,从而增加
了血液的外周阻力,影响到血液循环以及组织和器官
的血液供应.因此,研究毛细血管微循环中红细胞的运
动与变形,揭示红细胞运动行为的机理,对理解基本的
血液循环及相关血液疾病的临床表现具有重要的理论
价值.
事实证明,基于DPD方法的红细胞弹性网络力学
模型可以有效地模拟红细胞在毛细血管内的运动与变
形过程.例如, Quinn等人[107]结合微流控实验和DPD
值模拟研究了红细胞在毛细血管内的变形和流动规律.
他们发现,当红细胞挤入狭窄的毛细管道时,细胞会发
生很大的变形,但一旦穿过管道,则很快恢复其最初的
圆盘状结构.通过调节梯度压差改变流体流速,他们分
析了在不同压降下红细胞通过毛细管道时的流动速度,
发现细胞流速随梯度压差近乎呈线性变化. Yazdani
[100]借助DPD方法研究了红细胞在圆形管道中的变
形行为和平衡状态,发现红细胞在流体剪切力作用下,
可存在对称的降落伞(5(a))或非对称的拖鞋
两种平衡状态,且其平衡状态会受到毛细管数管径
及细胞膜内外黏度比的影响.
红细胞血液病是血液系统发病率较高的疾病,
理状态下红细胞变形能力受损会增大微血管流动阻力,
导致血流量大幅度下降,从而引起组织缺氧及微血管
梗塞.研究人员借助DPD粒子方法也模拟研究了病理
状态下红细胞膜微观结构缺陷对细胞流动和变形性能
的影响机制.例如, Peng等人[108]发展了一套基于红细
胞膜微结构的两组分红细胞模型,模拟了红细胞膜磷
-细胞骨架相互作用变化对红细胞变形和流动行为
的影响.他们发现,病理状态下红细胞膜微观结构缺陷
会导致两组分间相互作用减弱,进而增加红细胞在毛
细血管内的流动阻力. Bow等人[109]采用红细胞弹性网
络力学模型模拟了疟疾感染下红细胞的异常流动特性,
发现疟疾感染的红细胞膜表面结节对细胞膜有多重刚
化效应,导致红细胞可变形性降低,难以通过毛细血管
并引起微血管栓塞[109].类似地, Ye等人[110]基于DPD
法模拟了疟疾感染红细胞在剪切流中的流动行为,
现疟原虫感染的红细胞可以扰乱血流,并导致其向血
管壁边聚并黏附在血管壁表面. Papageorgiou等人[111]
结合微流控实验和DPD数值模拟定量研究了镰状细胞
贫血症下红细胞异常的变形和黏附行为,发现缺氧会
显著增加镰状红细胞的黏附特性,造成镰状细胞无法
顺利通过毛细血管或毛细血管后微静脉,进而引发血
管闭塞危象. Li等人[112]模拟了球形红细胞增多症下病
变红细胞在脾微循环结构时的微流动特征,发现病变
红细胞在流动过程中会伴随细胞膜表面膜脂质的丢失,
从而降低红细胞的变形能力,致使其穿越脾脏毛细血
管时变得困难.
大量的临床实例表明,血管分叉处是血管病变的
高发区,其中一个原因是分叉处血流动力学特性的变
.因此,从流体力学角度分析病变血管中血液流动行
为可以帮助人们理解某些疾病发生发展的过程,
由于入出流边界条件不对称,基于离散粒子模型研究
分叉血管中血流动面临很大的挑战.最近,研究人员基
DPD数值方法发展了新的开放式边界条件控制方法,
751
https://engine.scichina.com/doi/10.1360/TB-2022-0913
并运用这种方法模拟研究了红细胞在分叉血管内的三
维运动及变形情况,分析了在流体力作用下红细胞在
分叉管道处流向主次子干道的分布规律[113].研究发现,
红细胞变形性变差聚集性增强有助于微循环中血细
胞和血浆相分离;同时,当红细胞流经毛细血管的分支
区域时,它们更趋向于流入流速相对较大的支管道,
两个分支管道中的流量比超过6:1,可实现100%的细
-血浆分离并获得血液中的无细胞血浆.
5.2 白细胞边聚及黏附行为的DPD数值模拟
白细胞又称白血球,是无色球形有细胞核的
血细胞.它一般有活跃的移动能力,具有机体防御和免
疫调节功能.白细胞通过血管壁外渗到血管外并到达
炎症灶是个复杂的连续过程,一般包括白细胞边聚和
滚动黏附和游出等阶段.因此,白细胞的边聚及其在
血管壁上的黏附是机体发生免疫反应时最初的环节.
近些年,人们在红细胞弹性网络力学模型基础上,基于
DPD数值方法拓展了白细胞弹性网络力学模型,并模
拟分析了不同流体力作用下白细胞的边聚和黏附过程
及其引发的血管堵塞现象.
在白细胞弹性网络力学模型中,其细胞膜表面也
是由三角形网格构成,通过施加黏弹性键相互作用
细胞膜抗弯曲作用抵抗细胞膜剪切和弯曲变形,并控
制细胞膜表面积和细胞内部体积使其维持一定的形
[104].在正常的生理状态下,可取白细胞膜的剪切刚
μ0=105.0 μN/m, 抗弯刚度kc=3.0×10–18 J[104].
为了促进黏附,白细胞随血流流动时会向血管边
缘部迁移(边聚). 白细胞的边聚取决于一些条件,包括
血流流速红细胞压积红细胞的聚集以及红细胞和
白细胞的变形能力. FedosovGompper[105]借助DPD
法模拟了白细胞的边聚过程,并系统性研究了红细胞
压积和血流流速对边聚过程的影响.他们通过模拟证
:血细胞在血管内流动时,移动较快且变形能力较强
的红细胞会逐渐把体积较大移动较慢的白细胞推离
血管中心部并迁移至血管边缘部,5(b)所示.他们
发现,当血流流速相对较小(大致在微循环静脉血流范
)且红细胞压积处于中间范围(20%~40%),白细胞
更容易边聚到血管的边缘.此外,红细胞-红细胞之间
的聚集作用也可以加快白细胞的边聚过程.
随着白细胞进入边流,白细胞会出现附壁现象. Lei
Karniadakis[104]借助DPD数值方法模拟了炎症刺激下
白细胞的黏附过程及其与病理状态下红细胞共同作用
引起的微血管栓塞现象.他们发现,整个血管闭塞过程
可以分为3个阶段:首先,当白细胞通过边聚迁移至血
管边缘部位后,会继续附壁滚动导致血流速度下降,
逐渐黏附在血管壁表面(第一阶段); 其次,黏附在血管
壁面的白细胞会与病理状态下的红细胞发生聚集作用,
进而导致血流速度进一步下降(第二阶段); 最后,也是
炎症反应的后期阶段,白细胞与红细胞之间的聚集作
用进一步增强,多个血细胞被堵在黏附白细胞周围,
发微血管栓塞(第三阶段). 总之,基于DPD数值方法的
白细胞弹性网络力学模型已被证实可成功用于研究相
5血细胞变形及流动行为的DPD模拟. (a) 基于DPD数值模拟分析红细胞在毛细管道内的流动变形行为[100].从左至右描述的是红细胞从圆
盘形状到降落伞形状的变形过程. (b) 基于DPD数值方法模拟白细胞在毛细管道内的边聚行为[105].红细胞和白细胞分别用红色和白色表示.
(c) 利用DPD方法模拟循环肿瘤细胞在微血管网络中的迁移行为[10].循环肿瘤细胞和红细胞分别用蓝色和红色表示,数字1~6描述模拟体系中的
6个循环肿瘤细胞的序号.(c)箭头表示循环肿瘤细胞与血管壁面黏附作用力
Figure 5 DPD simulations of shape deformation and flow behaviors of blood cells. (a) DPD simulation of a healthy red blood cell (RBC) in shear flow
in microtubes[100]. Sequential snapshots as time progress from left to right show the RBC shape deformation from biconcave to parachute shapes. (b)
DPD simulation of the margination dynamics of a white blood cell (WBC) in microtubes[105]. The RBCs and WBC are rendered in red and white,
respectively. (c) DPD simulation of circulating tumor cells (CTCs) transport in the microvascular network[10]. The CTCs and RBCs are rendered in blue
and red, respectively, and the numbers from 1 to 6 represent the sequence number of the six CTCs in the model system. The arrows in (c) show the
adhesion forces between the CTCs and the blood vessel wall
2023 3 68 7
752 https://engine.scichina.com/doi/10.1360/TB-2022-0913
关血液疾病微血管栓塞过程涉及的多血细胞相互作用
及多阶段流动特征,有助于人们对不同类型微血管栓
塞的认识,并为临床诊断和干预治疗提供新的思路和
方案.
5.3 血小板边聚黏附及聚集行为的DPD模拟
血小板是血液中最小的血细胞,呈两面微凸的圆
盘形,当受到机械或化学刺激时,则伸出突起,呈不规
则形,具有维持血管内皮完整性以及黏附聚集
收缩和促凝等功能.一般情况下,由流动着的血液
对血管内皮细胞产生的剪应力作用可以用连续介质力
学的方法来描述;然而,血小板血栓形成过程中,由纤
维蛋白原等黏附蛋白介导的血小板黏附和聚集作用机
制是微观尺度的.而在微观细胞尺度和宏观运输现象
之间涉及多尺度多层次的耦合.为了弥补宏观流动
尺度和微观细胞尺度之间的差距,模拟并预测血流流
动引起的血栓形成初期血小板黏附和聚集,基于介观
尺度的DPD离散方法是一种合适的模拟手段.
在由三角形网格构成的血小板弹性网络力学模型
,同样通过施加黏弹性成键作用细胞膜抗弯曲作
用以及控制细胞膜表面积和细胞体积来模拟血小板的
力学性质和形状.DPD模拟时,通常取血小板膜的剪
切刚度μ0=470.0 μN/m, 抗弯刚度kc=3.0×10–18 J[114].
,相比红细胞和白细胞尺寸,血小板要小很多(一般
直径只有2~4 μm), 在模拟时通常取血小板表面积
=16.9 μm2和体积
V0
tot
=4.2 μm3[114].
在正常血液循环过程中,血小板处于静息状态,
在某些生理或病理状态下,血小板可被活化,发生变
黏附聚集以及参与凝血功能. Chang等人[115]
DPD方法模拟研究了II型糖尿病状态下红细胞和血
小板在圆形微管中的流动与迁移行为,发现增加血流
流速或红细胞压积也会增强血小板的边聚行为.同时,
DPD模拟时,他们还考虑了正常和病理状态下血细
胞在形状大小和可变形性等方面的不同,发现病理
状态下红细胞变形能力的下降会减缓血小板的边聚过
,而血小板平均体积的增加则又会增强其边聚效应.
此外,他们通过模拟还发现,白细胞的贴壁滚动或贴附
会影响周边的血流流动,进而减缓血小板的边聚过程.
因此,II型糖尿病而言,除了凝血酶和纤维蛋白原水
平升高等众多凝血因子的变化外,随血液一同流动的
红细胞和血小板在力学性质和形状尺寸方面的异常也
对血小板边聚和血栓形成的初期过程有一定的影响.
同时,在病理状态下,血管狭窄弯曲及分叉处常
是血栓形成的好发部位,这与该部位形成特殊流场有
. Soares等人[116]基于DPD方法模拟研究了血小板在
剪切流场下通过血管狭窄通道的过程. YazdaniKar-
niadakis[114]利用血细胞弹性网络力学模型模拟了红细
胞和血小板在收缩性微通道中的流动行为,并系统性
分析了微通道狭窄处的收缩水平红细胞压积和血流
流速对血小板流动及迁移行为的影响.他们发现,较高
的狭窄收缩水平和管壁剪切率会明显增强血小板的边
聚行为,这可以用来解释实验观察到的血管狭窄后血
小板聚集增强的现象.在后续的工作中,他们整合了血
细胞力学血小板黏附动力学和凝血级联来系统模拟
病理状态下血小板血栓的形成过程[117].这些定量的模
拟结果可以为相关的生理病理过程,ATP释放
叉血管处血浆撇取效应和血栓形成提供新的理解.
5.4 健康与疾病状态下血液流变学特征
血液是一种具有黏滞性的非牛顿流体,血黏度是
反映血液流动性质的一项重要指标,对其测定可为临
床许多疾病(如球形红细胞增多症镰状细胞贫血
深静脉血栓), 特别为血栓前状态与血栓性疾病的
诊治和预防提供一定的参考依据.近年来,基于介观尺
度的DPD离散方法已被成功应用于模拟血液及其有形
成分的流动性与变形性规律,以及探讨造成血黏度变
化或异常的微观影响机制.
我们知道,血液的黏滞性主要取决于红细胞的压
积及其变形和聚集性质.借助DPD模拟方法, Fedosov
等人[118]从血细胞悬浊液的模拟中准确预测了血液的
非牛顿特征:血黏度随着剪切率的增加而逐渐下降.
,他们还发现,红细胞之间的聚集以及单个红细胞的
变形性能会显著影响血液的这种非牛顿特性.特别地,
红细胞之间的强聚集作用会导致低切血黏度水平的显
著增加.
血黏度水平高于或低于健康人的参考值,都是某
些疾病发生和发展的病理性反映.因此,测定血黏度并
分析血黏度变化规律,对揭示血液流变学的改变与某
些疾病的发生和发展关系具有重要意义. II型糖尿病是
一种常见的内分泌代谢疾病,其严重的并发症之一是
血液的高黏高凝状态血栓形成进而引起血管病
[119]. Chang等人[120]基于DPD模拟研究了II型糖尿病
病理状态下的血黏度特征,发现糖化血红蛋白水平较
753
https://engine.scichina.com/doi/10.1360/TB-2022-0913
高的II型糖尿病患者血黏度平均值也较高.同时,他们
还分析了红细胞聚集性和变形性对血黏度异常的影响,
发现低切时血浆纤维蛋白原水平升高引起的红细胞聚
集程度增加以及高切下红细胞变形及定向程度增加是
导致II型糖尿病患者血黏度异常的主要因素.高黏滞综
合征是以血黏度增高为主要表现的病理综合征,致使
组织血液灌注量减少,产生缺血缺氧的症状[121]. Javadi
等人[122]基于DPD模拟分析了高黏滞综合征下血黏度
异常的影响因素,发现红细胞压积增加和聚集性增强,
以及纤维蛋白原浓度增加引起的血浆黏度增高是引起
低切血黏度异常增高的主要决定因素.镰状细胞贫血
症是一种比较严重的红细胞疾病,临床表现为慢性溶
血性贫血易感染和再发性疼痛危象引起慢性局部缺
,从而导致器官组织损害[123].基于患者临床数据
微流控实验和DPD数值模拟, Li等人[124]建立了一套能
适应多种实验条件的血液及其有形成分流变性能预测
分析平台,对不同症状(轻症/重症)患者的血黏度水平
进行了模型预测.他们发现,正常生理状态下血液呈现
非牛顿流体特征,而病理状态下血黏度则由于红细胞
变形能力的急剧下降呈现类似牛顿流体的特征.此外,
他们通过模拟还揭示了已获批准的临床药物(如羟基
)对病理状态下血黏度的调节机制.
5.5 循环肿瘤细胞迁移黏附及分选富集的DPD
模拟
循环肿瘤细胞(circulating tumor cells, CTCs)泛指
存在于外周血中的各类肿瘤细胞,对其进行捕获及分
选有助于肿瘤转移风险的预测和判断,在癌症早期诊
用药指导预后监测中具有潜在的应用价值[125].
但是,由于外周血液中循环肿瘤细胞的数量稀少且具
有异质性,目前循环肿瘤细胞的捕获和分选在技术上
存在巨大的挑战,成为制约其应用于临床的主要瓶颈.
近年来,微流控芯片技术因具有微型化集成化等特
征在诸多技术中异军突起,正在成为循环肿瘤细胞高
效分离与检测的关键技术之一.与此同时,人们也借助
DPD离散方法模拟研究循环肿瘤细胞在微流控通道内
的迁移黏附以及分选等过程,试图揭示其深层次的
细胞迁移和分离机制.
循环肿瘤细胞以不同形态存在于外周血中,既有
游离的单个细胞,也有聚集成群的细胞团.当单个循环
肿瘤细胞在微血管内边聚并贴附在血管内壁表面后,
将发生血管外渗现象.因此,研究循环肿瘤细胞在微血
管内的迁移及黏附过程有助于理解循环肿瘤细胞的血
道转移过程.基于此, Wang等人[10]基于DPD模拟方法
建立了循环肿瘤细胞力学模型,分析了循环肿瘤细胞在
复杂毛细血管网络中的边聚及黏附动力学行为(5(c)),
发现当血流比较缓慢且红细胞压积适中时,循环肿瘤
细胞更容易在微血管分叉处外渗. Xiao等人[106]采用类
似的循环肿瘤细胞力学模型,模拟研究了其在复杂微
血管内随血液流动时的边聚及黏附行为.他们发现,
血管内血液的颗粒特性对循环肿瘤细胞的迁移及黏附
过程都有显著的影响.此外,他们通过对比平直和弯曲
微血管的模拟结果,发现循环肿瘤细胞更易牢固黏附
在弯曲微血管的转折处.
针对循环肿瘤细胞的富集分离和检测的研究也十
分具有吸引力.例如, Rossinelli等人[126]基于DPD模拟
方法构建了一种具有确定性侧向位移阵列(determinis-
tic lateral displacement, DLD)CTC-iChip微流控芯片.
基于这种微流控平台,他们模拟了循环肿瘤细胞从红
细胞等微颗粒中分离的动力学过程,证实了在惯性升
力作用下,尺寸大的循环肿瘤细胞会发生显著的侧向
位移,而尺寸小的红细胞则易与阵列微柱碰撞,进而导
致侧向位移不显著.两种不同类型的细胞因此在侧向
位移上出现差值并最终实现分离.吴晨冰[127]采用DPD
方法构建了一种类似的DLD微流控芯片,分析了阵列
微柱截面形状对循环肿瘤细胞侧向迁移及分选效率的
影响,发现相比三角形截面的微柱DLD阵列,正方形截
面与圆形截面的微柱DLD阵列可容易实现循环肿瘤细
胞的高效分离.这些模拟研究结果为人们设计新型的
微流控操控实验平台,更好地利用细胞的物理和流动
特性实现循环肿瘤细胞的高效分选富集提供了一定的
理论指导.
5.6 模拟血细胞变形及血液流动的其他数值模型
简介
近年来,随着计算能力的提升和超级计算应用领
域的拓展,通过仿真模拟手段对血细胞建模并模拟研
究血细胞随血液流动时的变形迁移聚集黏附等
问题,越来越受到科技工作者的关注.除了DPD血细胞
力学模型外,其他一些血细胞力学模型,包括基于弹性
理论和流体力学的连续体力学模型与基于粒子模拟方
法的离散弹簧网络模型[100,101,128],相继被建立并用于血
细胞力学与血流动力学方面的研究.一方面,连续体力
学模型一般基于欧拉网格求解方法(如浸入式边界
2023 3 68 7
754 https://engine.scichina.com/doi/10.1360/TB-2022-0913
[129,130]和边界积分法[131]), 将细胞膜和细胞液视为均
质材料.这类方法的优点是容易与现有的计算流体力
学求解器(如有限体积法有限差分法)耦合,可同时处
理大量血细胞的变形及交互作用问题,因此常用于研
究较大空间长度和较长时间尺度上的血流动力学问题.
另一方面,基于粒子模拟方法的血细胞模型,如多粒子
碰撞动力学方法[132]光滑耗散粒子动力学方法[133]
光滑粒子动力学方法[134],通常将细胞膜细胞液以及
血浆都视为粗粒化粒子.这种无网格方法在处理细胞
膜复杂结构问题方面具有一定的优势,如病理状态下
细胞膜磷脂与细胞骨架的分离现象血细胞膜热涨落
等问题.但该类方法在模拟流体运动时并非严格遵循
Navier-Stokes方程,同时相比连续体力学模型,其计
算精度较低,一般需要模拟一定规模的离散粒子保证
计算精度,因此使用离散粒子法的计算消耗通常高于
基于连续体力学模型模拟.如何把离散弹簧网络细胞
力学模型和连续介质力学模型结合起来,发展并完善
连续体-离散粒子混合血流动力学模型,同时兼顾血
细胞膜微结构和计算效率,将是未来工作的一个
重点.
6总结与展望
耗散粒子动力学(DPD)方法是一种介观尺度的无
网格粒子类模拟算法.根据粗粒化程度的不同, DPD
法可有效处理不同空间和时间尺度的生物体系动力学
问题.本文重点介绍了DPD方法在蛋白质分子脂质
分子及其膜系统和细胞等多个尺度的应用.可以发现,
在蛋白质系统中,目前DPD方法通过附加各种约束,
经能描述其二级结构的稳定性和变构行为.但目前的
DPD蛋白质模型大多基于特定的蛋白质分子,其扩展
性有待提高.此外,如何不显著增加模型复杂度而能定
量精确描述蛋白质分子精细的相互作用,如氢键
,制约了DPD方法对蛋白质分子三级和四级结构描
述与对多数量多种类蛋白质协作行为的研究.为了
解决该问题,或许可以借鉴量子力学/分子力学组合模
拟方法的思路[135],采用DPD模拟方法还原诸如蛋白质
二级结构折叠变构转换脂质膜-蛋白分子相互作
脂质膜-纳米颗粒相互作用脂质膜融合脂质
膜出芽分裂等过程,发挥其介观时间尺度与空间尺度
优势.针对其中涉及的关键分子则采用传统分子动力
学计算,对局部小系统引入全原子力场,仅增加局部系
统的复杂度自由度,对局部系统进行精细原子层面
相互作用的描述,有可能规避全局系统复杂度增加引
起的模拟尺度制约,从而实现原子-亚细胞层面生理病
理过程的同步还原.例如,最近Pezeshkian等人[136]开展
了一项有趣的工作,他们将三角形网格的模型映射为
粗粒化的模型,提出了一个跨尺度模拟概念框架.他们
提出由实验获取细胞器的真实三维结构,进行三角形
网络的建模,映射到粗粒化模型中,进行粗粒化分子动
力学模拟,以实现微秒尺度下的粗粒化模拟.在局部纳
米尺度,他们提出可以将粗粒化模型重新映射到原子
级别的分辨率进行全原子的分子动力学模拟.根据该
概念框架, Pezeshkian等人[136]模拟了具有真实大小和
脂质组成的整个线粒体的膜.虽然仅进行了2 ns的模
,没有看到线粒体的变形,但该概念框架是一个有趣
的尝试.但是,目前此类跨尺度的模拟方法尚缺少处理
不同空间尺度过渡区间信息传递的统一理论框架,
作用力的传递能量的传递物质的传递等.这是目
前制约跨尺度模拟的主要屏障之一.总体而言, DPD
法对脂质分子及其膜系统的定量描述,是理解亚细胞
尺度细胞生理病理过程的有效手段.联合多数量
多种类蛋白质分子协作及其与脂质膜系统的相互作用,
实现对亚细胞结构(如细胞骨架细胞器)的定量
定量模拟,将具有重要意义.
在细胞尺度, DPD方法已在血液及血细胞的复杂
流动中得以应用并展现一定的应用潜力.血液及其血
细胞的运动属于典型的多相流动问题,对其流动和流
变行为的研究是探讨病理状态下相关疾病发生机理的
重要手段.目前,基于DPD方法的血细胞弹性网络力学
模型已被成功应用于健康与疾病下血细胞的流动性与
变形性规律研究.接下来,如何利用DPD方法更好地模
拟生理和病理状态下血细胞的物理与力学特性,把血
细胞形状的动态变化与各种疾病的病理变化及临床表
现联系起来,将是未来工作的一个重点和难点;同时,
基于微流控芯片技术的器官芯片近几年来发展迅速,
已经实现了体外模拟多种活体细胞组织及器官微环
[137,138].如何结合介观尺度的DPD离散方法和宏观尺
度的连续介质力学模型以及微观尺度的分子动力学模
,发展并完善多尺度多物理场耦合力学模型,模拟
更加复杂的血管网络和更逼真的血流微环境并建立患
者个体化的血管及器官模型,也有很大的开拓空间和
研究潜力.这部分研究将有助于从亚细胞-细胞-组织等
不同层面多尺度贯通探索相关疾病发生的基本机制及
干预措施.
755
https://engine.scichina.com/doi/10.1360/TB-2022-0913
参考文献
1 Hoogerbrugge P J, Koelman J M V A. Simulating microscopic hydrodynamic phenomena with dissipative particle dynamics. Europhys Lett, 1992,
19: 155–160
2 Español P, Warren P. Statistical mechanics of dissipative particle dynamics. Europhys Lett, 1995, 30: 191–196
3 Groot R D, Warren P B. Dissipative particle dynamics: Bridging the gap between atomistic and mesoscopic simulation. J Chem Phys, 1997, 107:
4423–4435
4 Zhang L, Becton M, Wang X. Designing nanoparticle translocation through cell membranes by varying amphiphilic polymer coatings. J Phys
Chem B, 2015, 119: 3786–3794
5 Groot R D, Madden T J. Dynamic simulation of diblock copolymer microphase separation. J Chem Phys, 1998, 108: 8713–8724
6 Wen X, Lan J, Cai Z, et al. Dissipative particle dynamics simulation on drug loading/release in polyester-PEG dendrimer. J Nanopart Res, 2014,
16: 2403
7 Xie X, Xu S, Pi P, et al. Dissipative particle dynamic simulation on the assembly and release of siRNA/polymer/gold nanoparticles based
polyplex. AIChE J, 2018, 64: 810–821
8 Wang J, Han Y, Xu Z, et al. Dissipative particle dynamics simulation: A review on investigating mesoscale properties of polymer systems.
Macromol Mater Eng, 2021, 306: 2000724
9 Liu M B, Liu G R, Zhou L W, et al. Dissipative particle dynamics (DPD): An overview and recent developments. Arch Computat Methods Eng,
2015, 22: 529–556
10 Wang S, Ye T, Li G, et al. Margination and adhesion dynamics of tumor cells in a real microvascular network. PLoS Comput Biol, 2021, 17:
e1008746
11 Hollingsworth S A, Dror R O. Molecular dynamics simulation for all. Neuron, 2018, 99: 1129–1143
12 Li D, Ji B. Protein conformational transitions coupling with ligand interactions: Simulations from molecules to medicine. Med Novel Technol
Devices, 2019, 3: 100026
13 Dror R O, Dirks R M, Grossman J P, et al. Biomolecular simulation: A computational microscope for molecular biology. Annu Rev Biophys,
2012, 41: 429–452
14 Tozzini V. Coarse-grained models for proteins. Curr Opin Struct Biol, 2005, 15: 144–150
15 Li D, Ji B, Hwang K, et al. Coarse grained modeling of biopolymers and proteins: Methods and applications. Int J Appl Mech, 2009, 01: 113–136
16 Vishnyakov A, Talaga D S, Neimark AV. DPD simulation of protein conformations: From α-helices to β-structures. J Phys Chem Lett, 2012, 3:
3081–3087
17 Nozax Y, Tanford C. Examination of titration behavior. Meth Enzymol, 1967, 11: 715–734
18 Peter E K, Lykov K, Pivkin I V. A polarizable coarse-grained protein model for dissipative particle dynamics. Phys Chem Chem Phys, 2015, 17:
24452–24461
19 Choudhury C K, Kuksenok O. Native-based dissipative particle dynamics approach for α-helical folding. J Phys Chem B, 2020, 124: 11379–
11386
20 Vaiwala R, Ayappa K G. A generic force field for simulating native protein structures using dissipative particle dynamics. Soft Matter, 2021, 17:
9772–9785
21 Truszkowski A, van den Broek K, Kuhn H, et al. Mesoscopic simulation of phospholipid membranes, peptides, and proteins with molecular
fragment dynamics. J Chem Inf Model, 2015, 55: 983–997
22 Ding H M, Ma Y Q. Computer simulation of the role of protein corona in cellular delivery of nanoparticles. Biomaterials, 2014, 35: 8703–8710
23 Ding H M, Ma Y Q. Design strategy of surface decoration for efficient delivery of nanoparticles by computer simulation. Sci Rep, 2016, 6: 26783
24 Lin S, Ryu S, Tokareva O, et al. Predictive modelling-based design and experiments for synthesis and spinning of bioinspired silk fibres. Nat
Commun, 2015, 6: 6892
25 Rim N G, Roberts E G, Ebrahimi D, et al. Predicting silk fiber mechanical properties through multiscale simulation and protein design. ACS
Biomater Sci Eng, 2017, 3: 1542–1556
26 Shen W H, Tang Z H, Wu X W, et al. An atomistic model of silk protein network for studying the effect of pre-stretching on the mechanical
performances of silks. Acta Mech Sin, 2022, 38: 222013
27 Pan L, Wang F, Cheng Y, et al. A supertough electro-tendon based on spider silk composites. Nat Commun, 2020, 11: 1332
28 Venturoli M, Smit B, Sperotto M M. Simulation studies of protein-induced bilayer deformations, and lipid-induced protein tilting, on a
mesoscopic model for lipid bilayers with embedded proteins. Biophys J, 2005, 88: 1778–1798
29 Lipowsky R. The morphology of lipid membranes. Curr Opin Struct Biol, 1995, 5: 531–540
2023 3 68 7
756 https://engine.scichina.com/doi/10.1360/TB-2022-0913
30 de Meyer F, Smit B. Effect of cholesterol on the structure of a phospholipid bilayer. Proc Natl Acad Sci USA, 2009, 106: 3654–3658
31 de Meyer F J M, Benjamini A, Rodgers J M, et al. Molecular simulation of the DMPC-cholesterol phase diagram. J Phys Chem B, 2010, 114:
10451–10461
32 Yamamoto S, Maruyama Y, Hyodo S. Dissipative particle dynamics study of spontaneous vesicle formation of amphiphilic molecules. J Chem
Phys, 2002, 116: 5842–5849
33 Qiang X, Wang X, Ji Y, et al. Liquid-crystal self-assembly of lipid membranes on solutions: A dissipative particle dynamic simulation study.
Polymer, 2017, 115: 1–11
34 Arai N, Yoshimoto Y, Yasuoka K, et al. Self-assembly behaviours of primitive and modern lipid membrane solutions: A coarse-grained molecular
simulation study. Phys Chem Chem Phys, 2016, 18: 19426–19432
35 Li X, Gao L, Fang W. Dissipative particle dynamics simulations for phospholipid membranes based on a four-to-one coarse-grained mapping
scheme. PLoS One, 2016, 11: e0154568
36 Wang M, Pei S, Fang T, et al. Dissipative particle dynamics simulation on vesicles self-assembly controlled by terminal groups. J Phys Chem B,
2018, 122: 10607–10614
37 Tan H, Wang W, Yu C, et al. Dissipative particle dynamics simulation study on self-assembly of amphiphilic hyperbranched multiarm copolymers
with different degrees of branching. Soft Matter, 2015, 11: 8460–8470
38 Tan H, Yu C, Lu Z, et al. A dissipative particle dynamics simulation study on phase diagrams for the self-assembly of amphiphilic hyperbranched
multiarm copolymers in various solvents. Soft Matter, 2017, 13: 6178–6188
39 Shillcock J C, Lipowsky R. Equilibrium structure and lateral stress distribution of amphiphilic bilayers from dissipative particle dynamics
simulations. J Chem Phys, 2002, 117: 5048–5061
40 Illya G, Lipowsky R, Shillcock J C. Effect of chain length and asymmetry on material properties of bilayer membranes. J Chem Phys, 2005, 122:
244901
41 Illya G, Lipowsky R, Shillcock J C. Two-component membrane material properties and domain formation from dissipative particle dynamics. J
Chem Phys, 2006, 125: 114710
42 Rawicz W, Olbrich K C, McIntosh T, et al. Effect of chain length and unsaturation on elasticity of lipid bilayers. Biophys J, 2000, 79: 328–339
43 Shillcock J C, Lipowsky R. Tension-induced fusion of bilayer membranes and vesicles. Nat Mater, 2005, 4: 225–228
44 Gao L, Lipowsky R, Shillcock J. Tension-induced vesicle fusion: Pathways and pore dynamics. Soft Matter, 2008, 4: 1208–1214
45 Grafmüller A, Shillcock J, Lipowsky R. Dissipative particle dynamics of tension-induced membrane fusion. Mol Simul, 2009, 35: 554–560
46 Li X, Liu Y, Wang L, et al. Fusion and fission pathways of vesicles from amphiphilic triblock copolymers: A dissipative particle dynamics
simulation study. Phys Chem Chem Phys, 2009, 11: 4051–4059
47 Guo Y, Yang S. Spontaneous formation and fusion of raspberry vesicle self-assembled from star block terpolymers in aqueous solution. Materials,
2021, 14: 7690
48 Liu Y T, Zhao Y, Liu H, et al. Spontaneous fusion between the vesicles formed by A2n(B2)ntype comb-like block copolymers with a semiflexible
hydrophobic backbone. J Phys Chem B, 2009, 113: 15256–15262
49 Yamamoto S, Hyodo S. Budding and fission dynamics of two-component vesicles. J Chem Phys, 2003, 118: 7937–7943
50 Laradji M, Kumar P B S. Dynamics of domain growth in self-assembled fluid vesicles. Phys Rev Lett, 2004, 93: 198105
51 Laradji M, Kumar P B S. Domain growth, budding, and fission in phase-separating self-assembled fluid bilayers. J Chem Phys, 2005, 123: 224902
52 Laradji M, Kumar P B S. Anomalously slow domain growth in fluid membranes with asymmetric transbilayer lipid distribution. Phys Rev E,
2006, 73: 040901
53 Hong B, Qiu F, Zhang H, et al. Budding dynamics of individual domains in multicomponent membranes simulated by N-varied dissipative
particle dynamics. J Phys Chem B, 2007, 111: 5837–5849
54 Li L, Ji J, Song F, et al. Intercellular receptor-ligand binding: Effect of protein-membrane interaction. J Mol Biol, 2022, 435: 167787
55 Mukherjee S, Sakunthala A, Gadhe L, et al. Liquid-liquid phase separation of α-Synuclein: A new mechanistic insight for α-Synuclein
aggregation associated with Parkinson’s disease pathogenesis. J Mol Biol, 2022, 435: 167713
56 Schmidt U, Guigas G, Weiss M. Cluster formation of transmembrane proteins due to hydrophobic mismatching. Phys Rev Lett, 2008, 101:
128104
57 Li S, Zhang X, Wang W. Cluster formation of anchored proteins induced by membrane-mediated interaction. Biophys J, 2010, 98: 2554–2563
58 Li S Y, Zhang X R, Wang W C. Selective aggregation of membrane proteins by membrane-mediated interactions. Sci China Chem, 2014, 57:
1683–1689
59 de Meyer F J M, Venturoli M, Smit B. Molecular simulations of lipid-mediated protein-protein interactions. Biophys J, 2008, 95: 1851–1865
60 Morozova D, Weiss M. On the role of acylation of transmembrane proteins. Biophys J, 2010, 98: 800–804
61 Groot R D, Rabone K L. Mesoscopic simulation of cell membrane damage, morphology change and rupture by nonionic surfactants. Biophys J,
757
https://engine.scichina.com/doi/10.1360/TB-2022-0913
2001, 81: 725–736
62 Huang K C, Lin C M, Tsao H K, et al. The interactions between surfactants and vesicles: Dissipative particle dynamics. J Chem Phys, 2009, 130:
245101
63 Gao L, Fang W. Effects of induced tension and electrostatic interactions on the mechanisms of antimicrobial peptide translocation across lipid
bilayer. Soft Matter, 2009, 5: 3312–3318
64 Chen L, Gao L, Fang W, et al. How the antimicrobial peptides kill bacteria: Computational physics insights. Commun Comput Phys, 2012, 11:
709–725
65 Chen L, Jia N, Gao L, et al. Effects of antimicrobial peptide revealed by simulations: Translocation, pore formation, membrane corrugation and
Euler buckling. Int J Mol Sci, 2013, 14: 7932–7958
66 Chen L, Li X, Gao L, et al. Theoretical insight into the relationship between the structures of antimicrobial peptides and their actions on bacterial
membranes. J Phys Chem B, 2015, 119: 850–860
67 Li X X, Wan M W, Gao L H, et al. Mechanism of inhibition of human islet amyloid polypeptide-induced membrane damage by a small organic
fluorogen. Sci Rep, 2016, 6: 21614
68 Gao L, Fang W. Self-energy and corresponding virial contribution of electrostatic interactions in dissipative particle dynamics: Simulations of
cationic lipid bilayers. J Chem Phys, 2010, 132: 031102
69 Patra J K, Das G, Fraceto L F, et al. Nano based drug delivery systems: Recent developments and future prospects. J Nanobiotechnol, 2018, 16: 71
70 Yang K, Ma Y Q. Computer simulation of the translocation of nanoparticles with different shapes across a lipid bilayer. Nat Nanotechnol, 2010, 5:
579–583
71 Wang S, Zhou Q, Yu S, et al. Rod-like hybrid nanomaterial with tumor targeting and pH-responsive for cancer chemo/photothermal synergistic
therapy. J Nanobiotechnol, 2022, 20: 332
72 Ding H M, Ma Y Q. Controlling cellular uptake of nanoparticles with pH-sensitive polymers. Sci Rep, 2013, 3: 2804
73 Li Y, Yue T, Yang K, et al. Molecular modeling of the relationship between nanoparticle shape anisotropy and endocytosis kinetics. Biomaterials,
2012, 33: 4965–4973
74 Yue T, Zhang X. Molecular understanding of receptor-mediated membrane responses to ligand-coated nanoparticles. Soft Matter, 2011, 7: 9104–
9112
75 Chen P, Huang Z, Liang J, et al. Diffusion and directionality of charged nanoparticles on lipid bilayer membrane. ACS Nano, 2016, 10: 11541–
11547
76 Ding H M, Ma Y Q. Interactions between Janus particles and membranes. Nanoscale, 2012, 4: 1116–1122
77 Li Y, Li X, Li Z, et al. Surface-structure-regulated penetration of nanoparticles across a cell membrane. Nanoscale, 2012, 4: 3768–3775
78 Arai N, Yasuoka K, Zeng X C. A vesicle cell under collision with a Janus or homogeneous nanoparticle: Translocation dynamics and late-stage
morphology. Nanoscale, 2013, 5: 9089–9100
79 Wang S, Guo H, Li Y, et al. Penetration of nanoparticles across a lipid bilayer: Effects of particle stiffness and surface hydrophobicity. Nanoscale,
2019, 11: 4025–4034
80 Li J, Wang J, Yan Y, et al. Manipulating the interactions between the lipid bilayer and triblock Janus nanoparticles: Insight from dissipative
particle dynamics. Mol Syst Des Eng, 2021, 6: 156–162
81 Su H, Price C A H, Jing L, et al. Janus particles: Design, preparation, and biomedical applications. Mater Today Bio, 2019, 4: 100033
82 Yan L T, Yu X. Enhanced permeability of charged dendrimers across tense lipid bilayer membranes. ACS Nano, 2009, 3: 2171–2176
83 Yan L T, Yu X. Charged dendrimers on lipid bilayer membranes: Insight through dissipative particle dynamics simulations. Macromolecules,
2009, 42: 6277–6283
84 Yan L T, Yu X. Complexes comprised of a dendrimer and a vesicle: Role of vesicle size and the surface tension of the vesicle membrane.
Nanoscale, 2011, 3: 3812–3818
85 Guo R, Mao J, Yan L T. Unique dynamical approach of fully wrapping dendrimer-like soft nanoparticles by lipid bilayer membrane. ACS Nano,
2013, 7: 10646–10653
86 Ding H, Tian W, Ma Y. Designing nanoparticle translocation through membranes by computer simulations. ACS Nano, 2012, 6: 1230–1238
87 Huang L, Yu Y, Lu X, et al. Designing a nanoparticle-containing polymeric substrate for detecting cancer cells by computer simulations.
Nanoscale, 2019, 11: 2170–2178
88 Xia Q, Zhu T, Jiang Z, et al. Enhancing the targeting ability of nanoparticles via protected copolymers. Nanoscale, 2020, 12: 7804–7813
89 Li Y, Yuan H, von dem Bussche A, et al. Graphene microsheets enter cells through spontaneous membrane penetration at edge asperities and
corner sites. Proc Natl Acad Sci USA, 2013, 110: 12295–12300
90 Dallavalle M, Calvaresi M, Bottoni A, et al. Graphene can wreak havoc with cell membranes. ACS Appl Mater Interfaces, 2015, 7: 4406–4414
91 Dallavalle M, Bottoni A, Calvaresi M, et al. Functionalization pattern of graphene oxide sheets controls entry or produces lipid turmoil in
2023 3 68 7
758 https://engine.scichina.com/doi/10.1360/TB-2022-0913
phospholipid membranes. ACS Appl Mater Interfaces, 2018, 10: 15487–15493
92 Guo R, Mao J, Yan L T. Computer simulation of cell entry of graphene nanosheet. Biomaterials, 2013, 34: 4296–4301
93 Mao J, Guo R, Yan L T. Simulation and analysis of cellular internalization pathways and membrane perturbation for graphene nanosheets.
Biomaterials, 2014, 35: 6069–6077
94 Mao J, Chen P, Liang J, et al. Receptor-mediated endocytosis of two-dimensional nanomaterials undergoes flat vesiculation and occurs by
revolution and self-rotation. ACS Nano, 2016, 10: 1493–1502
95 Wang Q, Zhai X, Crowe M, et al. Heterogeneous oxidization of graphene nanosheets damages membrane. Sci China-Phys Mech Astron, 2019, 62:
64611
96 Wang S, Li X, Gong X, et al. Mechanistic modeling of spontaneous penetration of carbon nanocones into membrane vesicles. Nanoscale, 2020,
12: 2686–2694
97 Tu Y, M, Xiu P, et al. Destructive extraction of phospholipids from Escherichia coli membranes by graphene nanosheets. Nat Nanotechnol,
2013, 8: 594–601
98 Zhang L, Xu B, Wang X. Cholesterol extraction from cell membrane by graphene nanosheets: A computational study. J Phys Chem B, 2016, 120:
957–964
99 Sebastian B, Dittrich P S. Microfluidics to mimic blood flow in health and disease. Annu Rev Fluid Mech, 2018, 50: 483–504
100 Yazdani A, Li X, Karniadakis G E. Dynamic and rheological properties of soft biological cell suspensions. Rheol Acta, 2016, 55: 433–449
101 Li X, Vlahovska P M, Karniadakis G E. Continuum- and particle-based modeling of shapes and dynamics of red blood cells in health and disease.
Soft Matter, 2013, 9: 28–37
102 Pivkin I V, Karniadakis G E. Accurate coarse-grained modeling of red blood cells. Phys Rev Lett, 2008, 101: 118105
103 Fedosov D A, Caswell B, Karniadakis G E. A multiscale red blood cell model with accurate mechanics, rheology, and dynamics. Biophys J, 2010,
98: 2215–2225
104 Lei H, Karniadakis G E. Probing vasoocclusion phenomena in sickle cell anemia via mesoscopic simulations. Proc Natl Acad Sci USA, 2013, 110:
11326–11330
105 Fedosov D A, Gompper G. White blood cell margination in microcirculation. Soft Matter, 2014, 10: 2961–2970
106 Xiao L L, Liu Y, Chen S, et al. Effects of flowing RBCs on adhesion of a circulating tumor cell in microvessels. Biomech Model Mechanobiol,
2017, 16: 597–610
107 Quinn D J, Pivkin I, Wong S Y, et al. Combined simulation and experimental study of large deformation of red blood cells in microfluidic
systems. Ann Biomed Eng, 2011, 39: 1041–1050
108 Peng Z, Li X, Pivkin I V, et al. Lipid bilayer and cytoskeletal interactions in a red blood cell. Proc Natl Acad Sci USA, 2013, 110: 13356–13361
109 Bow H, Pivkin I V, Diez-Silva M, et al. A microfabricated deformability-based flow cytometer with application to malaria. Lab Chip, 2011, 11:
1065–1073
110 Ye T, Phan-Thien N, Cheong Khoo B, et al. Numerical modelling of a healthy/malaria-infected erythrocyte in shear flow using dissipative particle
dynamics method. J Appl Phys, 2014, 115: 224701
111 Papageorgiou D P, Abidi S Z, Chang H Y, et al. Simultaneous polymerization and adhesion under hypoxia in sickle cell disease. Proc Natl Acad
Sci USA, 2018, 115: 9473–9478
112 Li H, Lu L, Li X, et al. Mechanics of diseased red blood cells in human spleen and consequences for hereditary blood disorders. Proc Natl Acad
Sci USA, 2018, 115: 9574–9579
113 Lykov K, Li X, Lei H, et al. Inflow/outflow boundary conditions for particle-based blood flow simulations: Application to arterial bifurcations and
trees. PLoS Comput Biol, 2015, 11: e1004410
114 Yazdani A, Karniadakis G E. Sub-cellular modeling of platelet transport in blood flow through microchannels with constriction. Soft Matter, 2016,
12: 4339–4351
115 Chang H Y, Yazdani A, Li X, et al. Quantifying platelet margination in diabetic blood flow. Biophys J, 2018, 115: 1371–1382
116 Soares J S, Gao C, Alemu Y, et al. Simulation of platelets suspension flowing through a stenosis model using a dissipative particle dynamics
approach. Ann Biomed Eng, 2013, 41: 2318–2333
117 Yazdani A, Deng Y, Li H, et al. Integrating blood cell mechanics, platelet adhesive dynamics and coagulation cascade for modelling thrombus
formation in normal and diabetic blood. J R Soc Interface, 2021, 18: 20200834
118 Fedosov D A, Pan W, Caswell B, et al. Predicting human blood viscosity in silico. Proc Natl Acad Sci USA, 2011, 108: 11772–11777
119 DeFronzo R A, Ferrannini E, Groop L, et al. Type 2 diabetes mellitus. Nat Rev Dis Primers, 2015, 1: 15019
120 Chang H Y, Li X, Karniadakis G E. Modeling of biomechanics and biorheology of red blood cells in type 2 diabetes mellitus. Biophys J, 2017,
113: 481–490
121 Gertz M A. Acute hyperviscosity: Syndromes and management. Blood, 2018, 132: 1379–1385
759
https://engine.scichina.com/doi/10.1360/TB-2022-0913
122 Javadi E, Deng Y, Karniadakis G E, et al. In silico biophysics and hemorheology of blood hyperviscosity syndrome. Biophys J, 2021, 120: 2723–
2733
123 Barabino G A, Platt M O, Kaul D K. Sickle cell biomechanics. Annu Rev Biomed Eng, 2010, 12: 345–367
124 Li X, Du E, Lei H, et al. Patient-specific blood rheology in sickle-cell anaemia. Interface Focus, 2016, 6: 20150065
125 Alix-Panabières C, Schwarzenbach H, Pantel K. Circulating tumor cells and circulating tumor DNA. Annu Rev Med, 2012, 63: 199–215
126 Rossinelli D, Tang Y H, Lykov K, et al. The in-silico lab-on-a-chip: Petascale and high-throughput simulations of microfluidics at cell resolution.
In: Proceedings of the International Conference for High Performance Computing, Networking, Storage and Analysis. New York: Association for
Computing Machinery, 2015. 1–12
127 Wu C B. Computational simulations of circulating tumor cell separation using microfluidic-based approaches (in Chinese). Master Dissertation.
Hangzhou: China Jiliang University, 2022 [吴晨冰.基于微流控芯片技术分选循环肿瘤细胞的模拟与机理研究.硕士学位论文.杭州:中国计
量大学, 2022]
128 Ye T, Phan-Thien N, Lim C T. Particle-based simulations of red blood cells—A review. J Biomech, 2016, 49: 2255–2266
129 Mittal R, Iaccarino G. Immersed boundary methods. Annu Rev Fluid Mech, 2005, 37: 239–261
130 Wang X, Gong X, Sugiyama K, et al. An immersed boundary method for mass transfer through porous biomembranes under large deformations. J
Comput Phys, 2020, 413: 109444
131 Pozrikidis C. The axisymmetric deformation of a red blood cell in uniaxial straining Stokes flow. J Fluid Mech, 1990, 216: 231–254
132 McWhirter J L, Noguchi H, Gompper G. Flow-induced clustering and alignment of vesicles and red blood cells in microcapillaries. Proc Natl
Acad Sci USA, 2009, 106: 6039–6043
133 Ye T, Phan-Thien N, Lim C T, et al. Hybrid smoothed dissipative particle dynamics and immersed boundary method for simulation of red blood
cells in flows. Phys Rev E, 2017, 95: 063314
134 Ye T, Pan D, Huang C, et al. Smoothed particle hydrodynamics (SPH) for complex fluid flows: Recent developments in methodology and
applications. Phys Fluids, 2019, 31: 011301
135 Friesner R A, Guallar V. Ab initio quantum chemical and mixed quantum mechanics/molecular mechanics (QM/MM) methods for studying
enzymatic catalysis. Annu Rev Phys Chem, 2005, 56: 389–427
136 Pezeshkian W, König M, Wassenaar T A, et al. Backmapping triangulated surfaces to coarse-grained membrane models. Nat Commun, 2020, 11:
2296
137 Lin B C. Laboratory on a Microfluidic Chip (in Chinese). Beijing: Science Press, 2013 [林炳承.微纳流控芯片实验室.北京:科学出版社, 2013]
138 Whitesides G M. The origins and the future of microfluidics. Nature, 2006, 442: 368–373
2023 3 68 7
760 https://engine.scichina.com/doi/10.1360/TB-2022-0913
Summary for 耗散粒子动力学方法在生物学领域的应用与研究进展:从蛋白质结构到细胞力学
Dissipative particle dynamics simulations for biological systems:
From protein structures to cell mechanics
Zihan Tang, Xuejin Li*& Dechang Li*
Department of Engineering Mechanics, Zhejiang University, Hangzhou 310012, China
* Corresponding authors, E-mail: xuejin_li@zju.edu.cn;dcli@zju.edu.cn
Dissipative particle dynamics (DPD) is a mesoscopic coarse-grained simulation method developed in recent years, which is
an important method for studying the dynamic behaviors of soft matter and complex fluids. In this method, each DPD
particle represents a coarse-grained virtual cluster of a set of atoms or groups of matter. The position and momentum of the
DPD particle are updated in a continuous phase but spaced at discrete time steps. The coarse-grained DPD particles are
subject to simplified pairwise interacting conservative, dissipative, and random forces. Especially, the dissipative force and
random force in the DPD method act as a heat sink and a source, respectively, and the combined effect of these two forces
act as a thermostat, which conserves momentum and thus provides the correct description of hydrodynamic interactions for
the model system. In addition, a common choice of the soft repulsion for the conservative force allows using larger
integration time steps in DPD simulations than that usually allowed in classical molecular dynamics (MD) simulations.
Hence, compared with the MD method, the computational cost of the DPD simulation is significantly reduced due to the
smaller number of modeled particles and the larger computational time step, enabling the simulations of the static and
dynamic behaviors of complex fluids and soft matter systems at physically attractive length scale and time scale. Moreover,
the particle-based framework of the DPD method enables people to easily incorporate additional physical features into the
model systems and extend its application to complex systems. For these reasons, the DPD method and its extension have
been successfully applied to numerous soft matter and complex fluid systems such as oil/water/surfactant systems,
chemical morphology, microscopic morphology, phase separation, as well as dynamics and rheological properties of
polymer solutions and colloidal suspensions.
In this paper, we first introduce the theoretical formulation and parameterization of the DPD method. Then, we review
recent advances in DPD modeling of biological systems, focusing on its applications at the molecular and cellular scales.
At the molecular scale, we highlight examples of successful simulations of the protein structures and their interactions, the
structure and dynamics of amphiphilic lipid molecule membranes (e.g., the self-assembly of lipid molecules, the structure
and properties of lipid membranes, the fusion of lipid membranes, and the budding and fission of lipid membranes), the
interaction of lipid membranes with protein molecules, and the interaction of nanoparticles with lipid membranes; at the
cellular scale, we focus on the DPD modeling of blood cell flow and blood rheological behaviors in the blood
microcirculatory systems, including the shape deformation and flow dynamics of red blood cells, the margination and
adhesion dynamics of white blood cells, the margination and aggregation behaviors of platelets, the hemorheological
behavior of blood flow, as well as the separation of circulating tumor cells from blood flow using microfluidic devices.
Additionally, we compare the advantages and disadvantages of the microscale blood flow simulations between the
continuum-based methods and particle-based methods, including the DPD method. Finally, we briefly present the
development trends and application prospects of DPD modeling in biological systems.
dissipative particle dynamics, protein, lipid membrane, blood cell, blood flow
doi: 10.1360/TB-2022-0913
761
https://engine.scichina.com/doi/10.1360/TB-2022-0913
Article
Full-text available
The development of chemo/photothermal nanotherapeutic systems with excellent photothermal performance, stable drug loading, tumor targeting and strong membrane penetration still remains a challenge. To address this problem, herein a rod-like nanocomposite system (AuNR@FA-PR/PEG) forming from folic acid (FA) terminated carboxylated cyclodextrin (CD) pseudopolyrotaxane (FA-PR) and polyethylene glycol (PEG) modifying gold nanorods (AuNR) was reported. Cisplatin (CDDP) was loaded in AuNR@FA-PR/PEG via coordination bonds to prepare a rod-like pH-responsive nanosystem (AuNR@FA-PR/PEG/CDDP) with chemotherapy/photothermal therapy. The rod-like morphology of AuNR@FA-PR/PEG was characterized by transmission electron microscope. In vitro drug release experiments showed the pH-responsive of AuNR@FA-PR/PEG/CDDP. In vivo real-time imaging assays proved AuNR@FA-PR/PEG/CDDP could rapidly enrich in the tumor area and stay for a long time because of folate targeting and their rod-like morphology. In vivo photothermal imaging assays showed AuNR@FA-PR/PEG/CDDP excellent photothermal performance, the average temperature of tumor region could reach 63.5 °C after 10 min irradiation. In vitro and in vivo experiments also demonstrated that the combined therapy of chemotherapy and photothermal therapy had an outstandingly synergistic effect and improved the therapeutic efficacy comparing with chemotherapy and photothermal therapy alone. Therefore, the prepared rod-like AuNR@FA-PR/PEG/CDDP will provide a new strategy for the effective treatment of cancer.
Article
Full-text available
Silk protein builds one of the strongest natural fibers based on its complex nanocomposite structures. However, the mechanical performance of silk protein, related to its molecular structure and packing is still elusive. In this study, we constructed an atomistic silk protein network model, which reproduces the extensive connection topology of silk protein with structure details of the β-sheet crystallites and amorphous domains. With the silk protein network model, we investigated the structure evolution and stress distribution of silk protein under external loading. We found a pre-stretching treatment during the spinning process can improve the strength of silk protein. This treatment improves the properties of silk protein network, i.e., increases the number of nodes and bridges, makes the nodes distributed homogeneously, and induces the bridges in the network well aligned to the loading direction, which is of great benefit to the mechanical performances of silk protein. Our study not only provides a realized atomistic model for silk protein network that well represents the structures and deformations of silk proteins under loading, but also gains deep insights into the mechanism how the pre-loading on silk proteins during spinning improves the mechanical properties of silk fibers.
Article
Full-text available
The spontaneous formation and fusion of raspberry vesicles was studied using the dissipative particle dynamics (DPD) method. The vesicles were formed through the self-assembly of amphiphilic E12O6F2 star terpolymers in selective solvent. E and F blocks are solvophobic and the O block is solvophilic. The shortest F block plays a major role in the formation of raspberry vesicles. Distinct vesicle formation mechanisms were observed at different polymer concentrations. At higher concentrations, vesicles form via the bending and closure of an oblate F-bump-E bilayer. At lower concentrations, the formation pathway contains: the initial formation of a vesicle with a core, the combination of such vesicles into cylindrical micelles, and the bending of the cylindrical micelles to form a hollow vesicle. In addition, raspberry vesicle fusion is regulated by F bumps through the continuous coalescence of them from apposed vesicle membranes. The contact area bends, followed by the formation of a fusion pore and a tilted inner layer. As the pore sealed, the hemifusion structure appears, which further restructures to form a vesicle. Our results provide guidance on understanding the dynamic processes of complex vesicles and biological membrane fusion.
Article
Full-text available
Hyperviscosity syndrome (HVS) is characterized by an increase of the blood viscosity by up to seven times the normal blood viscosity, resulting in disturbances to the circulation in the vasculature system. HVS is commonly associated with an increase of large plasma proteins, and abnormalities in the properties of red blood cells (RBCs) such as cell interactions, cell stiffness, and increased hematocrit. Here, we perform a systematic study of the effect of each biophysical factor on the viscosity of blood by employing the Dissipative Particle Dynamic (DPD) method. Our in silico platform enables manipulation of each parameter in isolation, providing a unique scheme to quantify and accurately investigate the role of each factor in increasing the blood viscosity. To study the effect of these four factors independently, each factor was elevated more than its values for a healthy blood while the other factors remained constant and viscosity measurement was performed for different hematocrits and flow rates. While all four factors were found to increase the overall blood viscosity, these increases were highly dependent on the hematocrit and the flow rates imposed. The effect of cell aggregation and cell concentration on blood viscosity were predominantly observed at low shear rates, in contrast to the more magnified role of cell rigidity and plasma viscosity at high shear rates. Additionally, cell-related factors increase the whole blood viscosity at high hematocrits compared to the relative role of plasma-related factors at lower hematocrits. Our results, mapped onto the flow rates and hematocrits along the circulatory system provide a correlation to underpinning mechanisms for HVS findings in different blood vessels.
Article
Full-text available
In tumor metastasis, the margination and adhesion of tumor cells are two critical and closely related steps, which may determine the destination where the tumor cells extravasate to. We performed a direct three-dimensional simulation on the behaviors of the tumor cells in a real microvascular network, by a hybrid method of the smoothed dissipative particle dynamics and immersed boundary method (SDPD-IBM). The tumor cells are found to adhere at the microvascular bifurcations more frequently, and there is a positive correlation between the adhesion of the tumor cells and the wall-directed force from the surrounding red blood cells (RBCs). The larger the wall-directed force is, the closer the tumor cells are marginated towards the wall, and the higher the probability of adhesion behavior happen is. A relatively low or high hematocrit can help to prevent the adhesion of tumor cells, and similarly, increasing the shear rate of blood flow can serve the same purpose. These results suggest that the tumor cells may be more likely to extravasate at the microvascular bifurcations if the blood flow is slow and the hematocrit is moderate.
Article
Full-text available
Normal haemostasis is an important physiological mechanism that prevents excessive bleeding during trauma, whereas the pathological thrombosis especially in diabetics leads to increased incidence of heart attacks and strokes as well as peripheral vascular events. In this work, we propose a new multiscale framework that integrates seamlessly four key components of blood clotting, namely transport of coagulation factors, coagulation kinetics, blood cell mechanics and platelet adhesive dynamics, to model the development of thrombi under physiological and pathological conditions. We implement this framework to simulate platelet adhesion due to the exposure of tissue factor in a three-dimensional microchannel. Our results show that our model can simulate thrombin-mediated platelet activation in the flowing blood, resulting in platelet adhesion to the injury site of the channel wall. Furthermore, we simulate platelet adhesion in diabetic blood, and our results show that both the pathological alterations in the biomechanics of blood cells and changes in the amount of coagulation factors contribute to the excessive platelet adhesion and aggregation in diabetic blood. Taken together, this new framework can be used to probe synergistic mechanisms of thrombus formation under physiological and pathological conditions, and open new directions in modelling complex biological problems that involve several multiscale processes.
Article
Gaining insights into the intercellular receptor-ligand binding is of great importance for understanding numerous physiological and pathological processes, and stimulating new strategies in drug design and discovery. In contrast to the in vitro protein interaction in solution, the anchored receptor and ligand molecules interact with membrane in situ, which affects the intercellular receptor-ligand binding. Here, we review theoretical, simulation and experimental works regarding the regulatory effects of protein-membrane interactions on intercellular receptor-ligand binding mainly from the following aspects: membrane fluctuations, membrane curvature, glycocalyx, and lipid raft. In addition, we discuss biomedical significances and possible research directions to advance the field and highlight the importance of understanding of coupling effects of these factors in pharmaceutical development.
Article
Aberrant aggregation of the misfolded presynaptic protein, α-Synuclein (α-Syn) into Lewy body (LB) and Lewy neuritis (LN) is a major pathological hallmark of Parkinson’s disease (PD) and other synucleinopathies. Numerous studies have suggested that prefibrillar and fibrillar species of the misfolded α-Syn aggregates are responsible for cell death in PD pathogenesis. However, the precise molecular events during α-Syn aggregation, especially in the early stages, remain elusive. Emerging evidence has demonstrated that the liquid-liquid phase separation (LLPS) of α-Syn occurs in the nucleation step of α-Syn aggregation, which offers an alternate non-canonical aggregation pathway in the crowded microenvironment. The liquid-like α-Syn droplets gradually undergo an irreversible liquid-to-solid phase transition into amyloid-like hydrogel entrapping oligomers and fibrils. This new mechanism of α-Syn LLPS and gel formation might represent the molecular basis of cellular toxicity associated with PD. This review aims to demonstrate the recent development of α-Syn LLPS, the underlying mechanism along with the microscopic events of aberrant phase transition. This review further discusses how several intrinsic and extrinsic factors regulate the thermodynamics and kinetics of α-Syn LLPS and co-LLPS with other proteins, which might explain the pathophysiology of α-Syn in various neurodegenerative diseases.
Article
A coarse-grained force field for molecular dynamics simulations of the native structures of proteins in a dissipative particle dynamics (DPD) framework is developed. The parameters for bonded interactions are derived by mapping the bonds and angles for 20 amino acids onto target distributions obtained from fully atomistic simulations in explicit solvent. A dual-basin potential is introduced for stabilizing backbone angles, to cover a wide spectrum of protein secondary structures. The backbone dihedral potential enables folding of the protein from an unfolded initial state to the folded native structure. The proposed force field is validated by evaluating the structural properties of several model peptides and proteins including the SARS-CoV-2 fusion peptide, consisting of α-helices, β-sheets, loops and turns. Detailed comparisons with fully atomistic simulations are carried out to assess the ability of the proposed force field to stabilize the different secondary structures present in proteins. The compact conformations of the native states were evident from the radius of gyration and the high intensity peaks of the root mean square deviation histograms, which were found to be within 0.4 nm. The Ramachandran-like energy landscape on the phase space of backbone angles (θ) and dihedrals (ϕ) effectively captured the conformational phase space of α-helices at ∼(ϕ = 50°,θ = 90°) and β-strands at ∼(ϕ = ±180°,θ = 90-120°). Furthermore, the residue-residue native contacts were also well reproduced by the proposed DPD model. The applicability of the model to multidomain complexes was assessed using lysozyme and a large α-helical bacterial pore-forming toxin, cytolysin A. Our study illustrates that the proposed force field is generic, and can potentially be extended for efficient in silico investigations of membrane bound polypeptides and proteins using DPD simulations.
Article
Polymer systems have typical multiscale characteristics, both in space and time. The mesoscopic properties of polymers are difficult to describe through traditional experimental approaches. Dissipative particle dynamics (DPD) is a simulation method used for solving mesoscale problems of complex fluids and soft matter. The mesoscopic properties of polymer systems, such as conformation, dynamics, and transport properties, have been studied extensively using DPD. This paper briefly summarizes the application of DPD to research involving microchannel flow, electrospinning, free‐radical polymerization, polymer self‐assembly processes, polymer electrolyte fuel cells, and biomedical materials. The main features and possible development avenues of DPD are described as well. The article describes the application of dissipative particle dynamics (DPD) in the polymer field. The basic principles of DPD are proposed. The advantages of DPD from six aspects are described: polymer flow in microchannels, electrospinning, free radical polymerization, polymer self‐assembly, polyelectrolyte fuel cell and biomedical materials. Finally, the future application of DPD is proposed.