Next Article in Journal
Sound-Absorbing Composites with Rubber Crumb from Used Tires
Next Article in Special Issue
Research on the Inertial Migration Characteristics of Bi-Disperse Particles in Channel Flow
Previous Article in Journal
Advancements and Research Trends in Microgrids Cybersecurity
Previous Article in Special Issue
Fibers Effects on Contract Turbulence Using a Coupling Euler Model
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review on the Some Issues of Multiphase Flow with Self-Driven Particles

State Key Laboratory of Fluid Power Transmission and Control, Zhejiang University, Hangzhou 310027, China
*
Author to whom correspondence should be addressed.
Appl. Sci. 2021, 11(16), 7361; https://doi.org/10.3390/app11167361
Submission received: 31 May 2021 / Revised: 17 July 2021 / Accepted: 9 August 2021 / Published: 10 August 2021
(This article belongs to the Special Issue Multiphase Flows in Microfluidics: Fundamentals and Applications)

Abstract

:
Multiphase flow with self-driven particles is ubiquitous and complex. Exploring the flow properties has both important academic meaning and engineering value. This review emphasizes some recent studies on multiphase flow with self-driven particles: the hydrodynamic interactions between self-propelled/self-rotary particles and passive particles; the aggregation, phase separation and sedimentation of squirmers; the influence of rheological properties on its motion; and the kinematic characteristics of axisymmetric squirmers. Finally, some open problems, challenges, and future directions are highlighted.

1. Introduction

Previous investigations on multiphase flows were mainly focused on the passive particles which are just driven by fluid and external force. However, in the multiphase flow with self-driven particles, the particles are mainly driven by internal force together with some influence of fluid and external force. Meanwhile, self-driven particles inject energy into fluid so as to change the motion of the fluid and then change its own motion in turn.
Multiphase flows with self-driven particles exist in natural and engineering applications. In nature, cells (e.g., Spermatozoa, Leucocyte), bacteria (e.g., Coccus, bacillus, Spirillum, Vibrio), fungus (e.g., Ascomycota), algae (e.g., Diatom, Dinoflagellates), protozoa (e.g., Entamoeba histolytica, Paramecium), movement proteins, etc., are all self-driven particles (as shown in Figure 1). Situations including when spermatozoa pass through the mucus to fuse with an egg [1], when Escherichia coli moves to high nutrient areas inside the human body, etc., all belong to the multiphase flow with self-driven particles [2].
In practice, synthesized self-driven particles contain artificial cell tissues, phosphorus-containing colloidal particles, soft field-responsive gel, catalytic particles for biodegradation, biomarkers and contrast media, miniature swimming devices and robots, etc. (as shown in Figure 2). In the presence of external fields, artificial particles become self-driven and obtain the abilities of targeted drug delivery, precise surgery, self-assembly, environmental modification, and water treatment; such situations also belong to the multiphase flow with self-driven particles.
When compared with multiphase flow with passive particles, the multiphase flow with self-driven particles involves more complex properties [3]. Firstly, the system is in a non-equilibrium state. Brownian motion of self-driven particles is caused by the interaction between random collision of fluid molecules and particles, which cannot satisfy the fluctuation–dissipation theorem that relies on the assumption that the response of a system in equilibrium to a small applied force is the same as its response to a spontaneous fluctuation, while the system consisting of fluid molecules and particles does not satisfy this assumption because of the interaction between small applied force and self-driving force. Secondly, there is a great diversity of self-driving modes which include gravity, torque, flagella driving, cilia shaking, surface deformation, chemical reaction, bubble release, electromagnetic light energy and actin aggregation and so on. Different driving modes lead to different stresses and patterns of flow, which will affect the particle motion in turn. Thirdly, there exist some abnormal characteristics such as turbulence with small Reynolds number [4], abnormal shear viscosity, weak shear hydrodynamic diffusion, biological convection, cluster structure, non-equilibrium phase transition between disordered and ordered phases, orientational ordering of quasi-liquid crystal and so on [5]. Moreover, the translational and rotational velocities of particles will deviate from the Maxwell distribution [6].
The study of multiphase flow with self-driven particles plays an important role in exploring the laws of nature and practical applications. Firstly, research on the moving mode and efficiency of a natural organism as well as the effect of environment on its kinematic characteristics can deepen the understanding of the tropism and rationality of natural selection in organisms. For example, how does the hydrodynamic interaction between organisms affect swimming efficiency? How do the different characteristics between the flows near the wall and far away from the wall influence the distribution and swimming properties of organisms? Secondly, exploring the energy conversion mechanism of self-driven particles in the flow is beneficial for the development of artificial active materials, such as cell tissue and phosphorus-containing colloid. Moreover, research on the influence of particle shapes on the flow properties contributes to the design of the miniature swimming devices and improvement of self-organization technology of rotating body in a water environment. Understanding the interaction between self-driven particles and passive particles also contributes to the development of catalytic active particles and achieving of the biodegradation of pollutants and environmental remediation in water treatment and water protection. Thirdly, exploring the influence of fluid physical property on particle motion contributes to the realization of the propagation and diffusion mechanism of active organisms in the human system, e.g., manipulating drugs to kill germs and control infections. Finding out the influence of wall surface on the properties of particle motion can ensure that particles give full play to the function of being biomarkers and contrast media, and contribute to particle localization and identification, targeted drug delivery, accuracy improvement of non-invasive surgery, screening of diseased cells and development of detection devices.
The energy injected into the fluid by self-driven particles is dependent on the particle driving modes. The self-driven particles described in this paper are the self-propelled and self-rotary particles and squirmers, which, as the most typical particles in natural and synthetic particles, directly interact with fluid by oscillating the cilia around the body. Based on the different stress applied to the fluid, squirmers can be divided into three kinds of modes, i.e., pushers (e.g., Escherichia coli, spermatozoa), pullers (e.g., Chlamydomonas) and neutral squirmers (e.g., paramecium), respectively.
In past research, although the research on the motion of self-propelled/rotary particles and squirmers has made some achievements, there are still some problems that have not been clarified. For example, the effect of number density ratio of self-driven particles to passive particles on the motion of passive particles was not taken into account; the settlement characteristics of squirmers under the combined influence of driving force, gravity, hydrodynamic force, wall effect and Reynolds number are not clear; the combined effect of Reynolds number, power law index, and driving modes on the swimming velocities and hydrodynamic efficiency of squirmers in shear-thickening fluids are still unknown; there is still a lack of studies taking particle volume fraction, shape factor, self-driving mode, and fluid viscosity into account to explore the formation mechanisms of self-organization of squirmers. Therefore, the following part of the paper summarizes the research progress of these related contents.

2. Interactions between Self-Propelled/Rotary Particles and Passive Particles

Self-driven particles frequently collide with passive particles in flow. For example, spermatozoa, algae, bacteria and movement proteins collide with intracellular vesicles or dead bacteria; artificial active particles collide with pollutants. There are some pieces of research on interactions between self-driven particles and passive particles. Ouyang et al. [7] studied the interaction between one self-rotary rotator and two passive particles, and found that when the rotator acted on passive particles, the Saffman force was larger than the Magnus force, which caused the passive particles to simultaneously rotate about rotator and revolve on its axis. In a small Reynolds number range, velocities of passive particles were the superposition of large amplitude and small amplitude pulsations, but the former would disappear if the Reynolds number was large enough. However, studies of interactions between a few self-driven and passive particles are insufficient to understand the properties of particle diffusion and aggregation. In addition, self-driven particles usually translate and rotate simultaneously, rather than only one of them doing so. Based on the experimental results, Kümmel et al. [8] pointed out that increasing the number of self-driven particles would change the structural and dynamic characteristics of passive particles. As the number density of passive particles increased, the particles would aggregate with a self-diffusing function (as shown in Figure 3). In their study, self-driven particles carried out translational self-diffusion under the function of electrophoresis.
Natural and synthetic self-driven particles usually move near the wall. Some papers have concluded that cells would move along the circumferential direction, Chlamydomonas cells would disperse or oscillate along the wall [9] (as shown in Figure 4 and Figure 5), tumble of Escherichia coli would be restrained, Volvox would become double beat, spermatozoa would change to backward swimming [10], and there existed three kinds of motion states for Chlamydomonas cells after touching the wall surface, i.e., escaping from the wall and swimming parallel with the wall [11].
Research on the influence of wall surface on the swimming properties of self-driven particles can provide benefits for controlling the natural biological and synthetic particles such as biofilm formation [12], guiding spermatozoa pass through the fallopian tube [13], biochemical sensing [14], targeting drug delivery [15], environmental modification [16], and improving the velocities of self-diffusion electrophoresis particles [17]. Breke et al. [18] discovered that hydrodynamic attraction led to Escherichia coli aggregation near the wall. However, Molaei et al. [9] believed that despite hydrodynamic attraction enhancing particle aggregation, it also appeared without hydrodynamic attraction. Near the wall, particle movements were related to the flow regime caused by the interaction between fluid and the wall, which depended on the wettability of the wall surface [19] and could be described by slip length ls (as shown in Figure 6). In addition, the slip length ls was close to 0 on the hydrophilic surface but a few tens of nanometers away on the smooth hydrophobic surface. Moreover, ls could be increased to the micron scale on a surface coated with hydrophobic molecular film or a surface on which bubble formation took place on the inner concave place. Poddar et al. [4] found that under the condition of ls ≠ 0, the trajectories of particles would change greatly with the increase in ls, and the results obtained from such cases were totally different from the case of ls ≈ 0 [20]. The values of ls changed along the wall for the rough surface [21].

3. Aggregation, Phase Separation and Sedimentation of Squirmers

Natural and synthetic self-driven particles, such as Escherichia coli or Brevibacterium with honeycomb or rafted structure (as shown in Figure 7), artificial Janus granules and synthetic photoactivated colloidal particles, often show the phenomena of aggregation, phase separation and sedimentation [5,22]. Clarifying and controlling such mechanisms can help improve the capability of screening diseased cells, developing detection devices together with particle localization and identification. There are several explanations for the aggregation of squirmers in previous studies. Some researchers believed it was caused by Brownian motion [23,24]; some accepted that high concentration and strong driving properties led to the increase in particle collision rate [25]; some supported that it was caused by the comprehensive influence of geometric dimensions of particles, self-driven properties and collision between particles [26]; and others attributed to the hydrodynamic interactions between particles and fluid [27,28].
For the particles with different phases, particle aggregation may result in phase separation. Matas-Navarro et al. [29] discovered that phase separation would be enhanced when the hydrodynamic interaction between fluid and particles was weak and would disappear when the interaction was strong. Yoshinaga et al. [30] concluded that hydrodynamic force suppressed phase separation. Phase separation is also related to the driving modes of particles. Theers et al. [31] found that pushers had the largest degree of phase separation, neutral squirmers followed and pullers were the last. However, Zottl et al. [32] and Blaschke et al. [33] discovered that neutral squirmers had the largest degree of phase separation. The reason for the different conclusions could be attributed to the different treatment of change in fluid density. If the compressibility of the flow was not considered, the fluid transport induced by squirmers would lead to the uneven distribution of fluid density. The driving characteristics of squirmers would be weakened in the areas with large fluid density, resulting in squirmers being stable within the group without phase separation. The research of [32,33] belonged to such cases, but the effect of fluid compressibility was taken into account in the research of [31]. The effect of fluid compressibility is related to the number density of particles. The above conclusions were based on the assumption that number density of particles is constant.
Near the wall, particles would present different settlement patterns, such as a stable state together with a fixed position and orientation [34], upright state together with floating above the wall [35], etc. The appearance of such states is correlated with the influence of gravity. The inertia of particles could not be ignored if settling velocities of particles were large enough [36]. Accordingly, it is necessary to explore the settlement characteristics of squirmers under the combined influence of driving force, gravity, hydrodynamic force, wall effect and inertia of particles.

4. Influence of Rheological Properties on the Motion of Squirmers

The fluid in which the microorganism lives frequently has non-Newtonian properties, e.g., Escherichia coli swims in the gastric juice, bacterium passes through the viscous-elastic barriers which are used to protect epithelia. The non-Newtonian properties of fluid contain the dependence of viscosity on shear strain rate, viscoelasticity and yield stress, etc. [37]. These characteristics make the motion of self-driven particles in non-Newtonian fluid different from that in Newtonian fluid. For example, the viscoelasticity will change the swimming stoke of particles, and elastic stress may affect the swimming velocities of particles. The existence of high molecular weight and macromolecules will increase the viscosity and further reduce the velocities of particles [38], while the interaction between driving property of self-driven particles and rheology of fluid may improve the velocities of particles [39,40,41]. Understanding the influence of rheological properties on the motion of self-driven particles is helpful to manipulate drugs to wipe out germs and control infection.
When squirmers swim in a fluid with viscosity varying with shear rate, the velocity and hydrodynamic efficiency, which is defined as the value of the useful to the total work, are related to the change of viscosity and driving mode caused by fluid shear. In the shear-thinning fluid, the non-local effect caused by the change of flow is more important than that of local viscosity reduction on its swimming [42]. The change of squirmer velocity depends on its starting velocity [43]. Compared with swimming in Newtonian fluid, although squirmers swim slowly, they swim more efficiently [44]. The power consumption of squirmer swimming decreases with the increase in Reynolds number, and the hydrodynamic efficiency is the same at a small Reynolds number for pullers and pushers [45]. The squirmers near the wall collide periodically with the wall with a constant period [46] and will change its trajectory due to rotation in parallel and opposite motion collisions [47].
Another property of non-Newtonian fluid is viscoelasticity. In the viscoelastic fluid, self-driven particles can not only paddle, but also move by the elasticity of fluid [48]. The change of squirmer motion caused by viscoelasticity is related to the driving mode. Whether the pullers and pushers move towards the wall depends on their initial position, while the neutral squirmer does not depend on the initial position [49]. The residence time of a squirmer in the near-wall region depends on the fluid elasticity and the initial position [50]. The elasticity of the fluid makes it easier for a squirmer to move towards the wall [51]. The degree of drift caused by viscoelasticity depends on the Deborah number, and ratio of driving velocity to flow velocity and driving mode [52]. The velocity of a squirmer decreases with the addition of Brinkman medium, but the hydrodynamic efficiency increases [53]. Viscoelasticity of fluid strengthens the rotational diffusion [54]. Squirmers move faster in a third-order viscoelastic fluid than in a Giesekus fluid [55]. Viscoelasticity of fluid slows down and speeds up the pullers and pushers, respectively, but it has no effect on neutral squirmers [56]. Squirmers consume more energy in Newtonian fluid and has higher hydrodynamic efficiency in viscoelastic fluid. For squirmers swimming in upper-convected Maxwell/Oldroyd-B fluids, the elastic stress becomes singular at a critical Weissenberg number [57]. However, the singularity is removed when the exponential Phan–Thien and Tanner model is utilized [58]. The above research was based on the Stokes flow with zero or minimum Reynolds number, i.e., the inertia effect is neglected. In fact, many flows are beyond the range of the Stokes flow [59]. Therefore, the influence of elastic number, which represents the ratio of elasticity to inertia, on squirmer velocity and hydrodynamic efficiency remains to be studied.

5. Characteristics of Suspension Flow of Axisymmetric Squirmers

Natural and synthetic self-driven particles are not limited to spherical shapes. Axisymmetric shape is a typical shape except for sphere, such as protozoa (e.g., Ciliata), biopolymer (e.g., Myoglobin, movement proteins), Bacillus (e.g., Bacillus subtitles, Escherichia coli) in nature and some synthetic self-driven particles (as shown in Figure 8).
The study of the shape effect on the flow with self-driven particles is helpful to deepen the understanding of the natural selection tendency of some organisms and the design of micro artificial swimming devices. The non-isotropic shape of axisymmetric particles leads to more complex motion modes and abundant self-organization phenomena, e.g., the swimming and feeding changes were found to be more significant in oblate spheroids than prolate spheroids [60]. For spherical squirmers, the self-organization phenomenon is caused by the interference and blocking between squirmers, but the self-organization of axisymmetric squirmers is caused by its array formed by the interactions. For example, Bacillus subtilis forms polar clusters [61]; Bacteroides dendritic moves along the narrow, long and high-density zone [62]; Tubulin-dynamin forms vortex arrays [63]; and Escherichia coli forms long range array [64].
Axisymmetric particles are more prone to self-organization than spherical particles. Moreover, shape-induced phase separation will occur at low Peclet numbers. However, whether the direct contact between particles or the long-range hydrodynamic effect is the main reason for the self-organization of squirmers has been debated. Based on the Langevin dynamics simulation of columnar [65] and filamentous squirmers [66], it was found that the formation of self-organization is dominated by the interactions between squirmers. Moreover, Brownian dynamics’ simulation considering long-range hydrodynamic action [67] also captured the self-organization. However, Pandey et al. [68] believed that particle shape and long-range hydrodynamic force were the factors to form the self-organization.
It has been found that suspension flow of axisymmetric squirmers is more prone to turbulence than that of spherical one [5]. It is generally believed that the turbulence emergence of particle suspension is related to the Peclet number. When the Peclet number is large, i.e., convection is stronger than diffusion, turbulence is more likely to appear. However, it was found that there is no turbulence when the Peclet number in the suspension flow of axisymmetric squirmers is as high as 50 [69]. Therefore, there are other factors at work, e.g., the hydrodynamic interactions, rotating dipoles formed by rotating flagella bundle and a counter-rotating body propulsion of squirmers [70], and induced circular motions [71] on the surface. Reinken et al. [72] pointed out that axisymmetric squirmers had a stronger rotation effect than spherical ones; the coupling of rotation and translation made turbulence appear. If the axisymmetric squirmers are asymmetric along the axis, the order of its orientation would be weakened, which also leads to turbulence [73]. Therefore, the factors leading to turbulence of suspension contain Peclet number, squirmer flexibility, hydrodynamic interactions, effects of strong rotation, anteroposterior asymmetry of squirmers, and self-driving modes. Thus, it is important to explore the correlation between various factors and turbulence.

6. Conclusions and Prospects

In this paper, the interaction between self-propelled/self-rotary particles and passive particles, the aggregation and phase separation of squirmer, the effect of sedimentation and rheological property on the motion of squirmer, and the characteristics of suspension with axisymmetric squirmers are reviewed. The following research can be carried out in the future:
(1) For the interactions between self-propelled/self-rotary particles and passive particles, the following issues should be focused on: the dynamic characteristics and configuration of the passive particles; the critical parameters of configuration transformation under different parameters; and the trajectory and morphology of self-propelled/self- rotary particles near the rough wall with variable roughness element.
(2) For the aggregation, phase separation and sedimentation of squirmers, the following issues should be focused on: the characteristics of aggregation and phase separation of squirmers under different driving modes; the distribution of cluster scale, mean cluster scale for pushers, pullers and neutral squirmers under different parameters; the characteristics of sedimentation under the combined action of different parameters; and the settlement trajectory and morphology under different parameters.
(3) For the influence of rheological property on squirmer movement, the following issues should be focused on: the motion of squirmers in a shear-thickening fluid; the hydrodynamic efficiency under different parameters; the characteristics of squirmer motion in a viscoelastic fluid by taking inertia into account; and the motion modes and velocity variation near the wall with different Weissenberg numbers and elasticity numbers.
(4) For the suspensions with axisymmetric squirmers, the following issues should be focused on: the characteristics of self-organization of the axisymmetric squirmer group; the turbulent characteristics of suspension with axisymmetric squirmers; and the root mean square of flow and squirmer velocity, Reynolds stress, and probability density distribution function of orientation under different parameters.
(5) Scientists have put effort into computational studies on hydrodynamic interactions of self-driven particles. The immersed boundary–lattice Boltzmann method and the direct-forcing fictitious domain method are the main methods in the studies of multiphase flow with self-driven particles. Both methods have been successfully applied to the study of Newtonian fluid. When the methods are applied to viscoelastic fluid, especially the fluid with a large Weissenberg number, there will be a stress boundary layer. Therefore, how to choose the appropriate discrete scheme is worth studying.

Author Contributions

Conceptualization, J.L.; writing, C.L. and J.L.; review, J.L. Both authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (Grant no. 11632016).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

There are no conflict of interest regarding the publication of this paper.

References

  1. Taketoshi, N.; Omori, T.; Ishikawa, T. Elasto-hydrodynamic interaction of two swimming spermatozoa. Phys. Fluids 2020, 32, 101901. [Google Scholar] [CrossRef]
  2. Japaridze, A.; Gogou, C.; Kerssemakers, J.W.J.; Nguyen, H.M. Direct observation of independently moving replisomes in Escherichia coli. Nat. Commun. 2021, 11, 3109. [Google Scholar] [CrossRef]
  3. Pedley, T.J. Spherical squirmers: Models for swimming micro-organisms. IMA J. Appl. Math. 2016, 81, 488–521. [Google Scholar] [CrossRef] [Green Version]
  4. Poddar, A.; Bandopadhyay, A.; Chakraborty, S. Near-wall hydrodynamic slip triggers swimming state transition of micro-organisms. J. Fluid Mech. 2020, 894, A11-1-34. [Google Scholar] [CrossRef]
  5. Wensink, H.H.; Dunkel, J.; Heidenreich, S.; Drescher, K.; Goldstein, R.E.; Lowen, H.; Yeomans, J.M. Meso-scale turbulence in living fluids. Proc. Natl. Acad. Sci. USA 2012, 109, 14308. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Ouyang, Z.Y.; Lin, J.Z.; Ku, X.K. Dynamics of a self-propulsion particle under different driving modes in a channel flow. Chin. Phys. B 2017, 26, 014701. [Google Scholar] [CrossRef]
  7. Ouyang, Z.Y.; Lin, J.Z.; Ku, X.K. Hydrodynamic interactions between a self-rotation rotator and passive particles. Phys. Fluids 2017, 29, 103301. [Google Scholar] [CrossRef]
  8. Kümmel, F.; Shabestari, P.; Lozano, C.; Volpe, G.; Bechinger, C. Formation, compression and surface melting of colloidal clusters by active particles. Soft Matter 2015, 11, 6187–6191. [Google Scholar] [CrossRef] [Green Version]
  9. Bechinger, C.; Di Leonardo, R.; Löwen, H.; Reichhardt, C.; Volpe, G.; Volpe, G. Active Particles in Complex and Crowded Environments. Rev. Mod. Phys. 2016, 88, 045006. [Google Scholar] [CrossRef]
  10. Miki, K.; Clapham, D.E. Rheotaxis Guides Mammalian Sperm. Curr. Biol. 2013, 23, 443–452. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Li, G.; Ardekani, A.M. Hydrodynamic interaction of microswimmers near a wall. Phys. Rev. E 2014, 90, 013010. [Google Scholar] [CrossRef] [Green Version]
  12. Li, Y.; Zhai, H.; Sanchez, S.; Kearns, D.B.; Wu, Y. Noncontact Cohesive Swimming of Bacteria in Two-Dimensional Liquid Films. Phys. Rev. Lett. 2017, 119, 018101. [Google Scholar] [CrossRef] [Green Version]
  13. Guidobaldi, H.A.; Jeyaram, Y.; Condat, C.; Oviedo, M.; Berdakin, I.; Moshchalkov, V.; Giojalas, I.; Silhanek, A.; Marconi, V. Disrupting the wall accumulation of human sperm cells by artificial corrugation. Biomico Fluidics 2015, 9, 024122. [Google Scholar] [CrossRef] [PubMed]
  14. Duan, W.; Wang, W.; Das, S.; Yadav, V.; Mallouk, T.E.; Sen, A. Synthetic nano- and micro-machines in analytical chemistry: Sensing, migration, capture, delivery, and separation. Annu. Rev. Analyt. Chem. 2015, 8, 311–333. [Google Scholar] [CrossRef] [Green Version]
  15. Campuzano, S.; De Ávila, B.E.F.; Yáñez-sedeño, P.; Pingarrón, J.M.; Wang, J. Nano/microvehicles for efficient delivery and (bio) sensing at the cellular level. Chem. Sci. 2017, 8, 6750–6763. [Google Scholar] [CrossRef] [Green Version]
  16. Richard, C.; Simmchen, J.; Eychmüller, A. Photocatalytic Iron Oxide Micro-Swimmers for Environmental Remediation. Z. Phys. Chem. 2018, 232, 747–757. [Google Scholar] [CrossRef]
  17. Ketzetzi, S.; De Graaf, J.; Doherty, R.P.; Kraft, D.J. Slip Length Dependent Propulsion Speed of Catalytic Colloidal Swimmers near Walls. Phys. Rev. Lett. 2020, 124, 048002. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Berke, A.P.; Turner, L.; Berg, H.C.; Lauga, E. Hydrodynamic Attraction of Swimming Microorganisms by Surfaces. Phys. Rev. Lett. 2008, 101, 038102. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Dey, P.; Saha, S.K.; Chakraborty, S. Confluence of channel dimensions and groove width dictates slippery hydrodynamics in grooved hydrophobic confinements. Microfluid. Nanofluidics 2020, 24, 1–15. [Google Scholar] [CrossRef]
  20. Ishimoto, K. Guidance of microswimmers by wall and flow: Thigmotaxis and rheotaxis of unsteady squirmers in two and three dimensions. Phys. Rev. E 2017, 96, 043103. [Google Scholar] [CrossRef]
  21. Nizkaya, T.V.; Asmolov, E.S.; Harting, J.; Vinogradova, O.I. Inertial migration of neutrally buoyant particles in superhydrophobic channels. Phys. Rev. Fluids 2020, 5, 014201. [Google Scholar] [CrossRef] [Green Version]
  22. Marchetti, M.C.; Fily, Y.; Henkes, S.; Patch, A.; Yllanes, D. Minimal model of active colloids highlights the role of mechanical interactions in controlling the emergent behavior of active matter. Curr. Opin. Colloid Interface Sci. 2016, 21, 34–43. [Google Scholar] [CrossRef]
  23. Wysocki, A.; Winkler, R.G.; Gompper, G. Propagating interfaces in mixtures of active and passive Brownian particles. New J. Phys. 2016, 18, 123030. [Google Scholar] [CrossRef] [Green Version]
  24. Stenhammar, J.; Marenduzzo, D.; Allen, R.J.; Cates, M.E. Phase behaviour of active Brownian particles: The role of dimensionality. Soft Matter 2014, 10, 1489–1499. [Google Scholar] [CrossRef] [Green Version]
  25. Cates, M.E.; Tailleur, J. Motility-Induced Phase Separation. Annu. Rev. Condens. Matter Phys. 2015, 6, 219–244. [Google Scholar] [CrossRef] [Green Version]
  26. Abkenar, M.; Marx, K.; Auth, T.; Gompper, G. Collective behavior of penetrable self-propelled rods in two dimensions. Phys. Rev. E 2013, 88, 062314. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Alarcón, F.; Valeriani, C.; Pagonabarraga, I. Morphology of clusters of attractive dry and wet self-propelled spherical particle suspensions. Soft Matter 2017, 13, 814–826. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Delmotte, B.; Keaveny, E.E.; Plouraboué, F.; Climent, E. Large-scale simulation of steady and time-dependent active suspensions with the force-coupling method. J. Comput. Phys. 2015, 302, 524–547. [Google Scholar] [CrossRef] [Green Version]
  29. Matas-Navarro, R.; Golestanian, R.; Liverpool, T.B.; Fielding, S.M. Hydrodynamic suppression of phase separation in active suspensions. Phys. Rev. E 2014, 90, 032304. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Yoshinaga, N.; Liverpool, T.B. Hydrodynamic interactions in dense active suspensions: From polar order to dynamical clusters. Phys. Rev. E 2017, 96, 020603. [Google Scholar] [CrossRef] [Green Version]
  31. Theers, M.; Westphal, E.; Qi, K.; Winkler, R.G.; Gompper, G. Clustering of microswimmers: Interplay of shape and hydrodynamics. Soft Matter 2018, 14, 8590–8603. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Zöttl, A.; Stark, H. Hydrodynamics Determines Collective Motion and Phase Behavior of Active Colloids in Quasi-Two-Dimensional Confinement. Phys. Rev. Lett. 2014, 112, 118101. [Google Scholar] [CrossRef]
  33. Blaschke, J.; Maurer, M.; Menon, K.; Zöttl, A.; Stark, H. Phase separation and coexistence of hydrodynamically interacting microswimmers. Soft Matter 2016, 12, 9821–9831. [Google Scholar] [CrossRef] [Green Version]
  34. Rühle, F.; Blaschke, J.; Kuhr, J.; Stark, H. Gravity-induced dynamics of a squirmer microswimmer in wall proximity. New J. Phys. 2018, 20, 025003. [Google Scholar] [CrossRef]
  35. Fadda, F.; Molina, J.J.; Yamamoto, R. Dynamics of a chiral swimmer sedimenting on a flat plate. Phys. Rev. E 2020, 101, 052608. [Google Scholar] [CrossRef]
  36. Ishikawa, T.; Hota, M. Interaction of two swimming Paramecia. J. Exp. Biol. 2006, 209, 4452–4463. [Google Scholar] [CrossRef] [Green Version]
  37. Yu, Z.S.; Lin, J.Z. Numerical research on the coherent structure in the viscoelastic second-order mixing layers. Appl. Math. Mech. 1998, 19, 717–723. [Google Scholar]
  38. Datt, C.; Natale, G.; Hatzikiriakos, S.G.; Elfring, G.J. An active particle in a complex fluid. J. Fluid Mech. 2017, 823, 675–688. [Google Scholar] [CrossRef] [Green Version]
  39. Yu, Z.; Wang, P.; Lin, J.Z.; Hu, H. Equilibrium positions of the elasto-inertial particle migration in rectangular channel flow of Oldroyd-B viscoelastic fluids. J. Fluid Mech. 2019, 868, 316–340. [Google Scholar] [CrossRef]
  40. D’Avino, G.; Maffettone, P.L. Particle dynamics in viscoelastic liquids. J. Non-Newton. Fluid Mech. 2015, 215, 80–104. [Google Scholar] [CrossRef]
  41. Zöttl, A.; Yeomans, J.M. Enhanced bacterial swimming speeds in macromolecular polymer solutions. Nat. Phys. 2019, 15, 554–563. [Google Scholar] [CrossRef] [Green Version]
  42. Pietrzyk, K.; Nganguia, H.; Datt, C.; Zhu, L.L.; Elfring, G.J.; Pak, O.S. Flow around a squirmer in a shear-thinning fluid. J. Non-Newton Fluid Mech. 2019, 268, 101–110. [Google Scholar] [CrossRef]
  43. Datt, C.; Zhu, L.; Elfring, G.J.; Pak, O.S. Squirming through shear-thinning fluids. J. Fluid Mech. 2015, 784, R1. [Google Scholar] [CrossRef] [Green Version]
  44. Nganguia, H.; Pietrzyk, K.; Pak, O.S. Swimming efficiency in a shear-thinning fluid. Phys. Rev. E 2017, 96, 062606. [Google Scholar] [CrossRef] [Green Version]
  45. Ouyang, Z.Y.; Lin, J.Z.; Ku, X.K. The hydrodynamic behavior of a squirmer swimming in power-law fluid. Phys. Fluids 2018, 30, 083301. [Google Scholar] [CrossRef]
  46. Ouyang, Z.Y.; Lin, J.Z.; Ku, X.K. Hydrodynamic properties of squirmer swimming in power-law fluids near a wall. Rheol. Acta 2018, 57, 655–671. [Google Scholar] [CrossRef]
  47. Ouyang, Z.Y.; Lin, J.Y.; Ku, X.K. Hydrodynamic interaction between a pair of swimmers in the power-law fluid. Int. J. Non-linear Mech. 2019, 108, 72–80. [Google Scholar] [CrossRef]
  48. Keim, N.C.; Garcia, M.; Arratia, P.E. Fluid elasticity can enable propulsion at low Reynolds number. Phys. Fluids 2012, 24, 081703. [Google Scholar] [CrossRef]
  49. Yazdi, S.; Ardekani, A.M.; Borhan, A. Locomotion of microorganisms near a no-slip boundary in a viscoelastic fluid. Phys. Rev. E 2014, 90, 043002. [Google Scholar] [CrossRef] [Green Version]
  50. Yazdi, S.; Ardekani, A.M.; Borhan, A. Swimming Dynamics near a Wall in a Weakly Elastic Fluid. J. Nonlinear Sci. 2015, 25, 1153–1167. [Google Scholar] [CrossRef]
  51. Li, G.; Ardekani, A.M. Near wall motion of undulatory swimmers in non-Newtonian fluids. Eur. J. Comput. Mech. 2017, 26, 44–60. [Google Scholar] [CrossRef]
  52. Corato, M.D.; D’Avino, G. Dynamics of a microorganism in a sheared viscoelastic liquid. Soft Matter 2017, 131, 96–211. [Google Scholar] [CrossRef]
  53. Nganguia, H.; Pak, O.S. Squirming motion in a Brinkman medium. J. Fluid Mech. 2018, 855, 554–573. [Google Scholar] [CrossRef] [Green Version]
  54. Qi, K.; Westphal, E.; Gompper, G.; Winkler, R.G. Enhanced Rotational Motion of Spherical Squirmer in Polymer Solutions. Phys. Rev. Lett. 2020, 124, 068001. [Google Scholar] [CrossRef] [Green Version]
  55. Datt, C.; Elfring, G.J. A note on higher-order perturbative corrections to squirming speed in weakly viscoelastic fluids. J. Non-Newton. Fluid Mech. 2019, 270, 51–55. [Google Scholar] [CrossRef] [Green Version]
  56. De Corato, M.; Greco, F.; Maffettone, P.L. Locomotion of a microorganism in weakly viscoelastic liquids. Phys. Rev. E 2015, 92. [Google Scholar] [CrossRef] [Green Version]
  57. Housiadas, K.D.; Binagia, J.P.; Shaqfeh, E.S.G. Squirmers with swirl at low Weissenberg number. J. Fluid Mech. 2021, 911, A16. [Google Scholar] [CrossRef]
  58. Housiadas, K.D. An active body in a Phan-Thien and Tanner fluid: The effect of the third polar squirming mode. Phys. Fluids 2021, 33, 043110. [Google Scholar] [CrossRef]
  59. Gagnon, D.A.; Keim, N.C.; Arratia, P.E. Undulatory swimming in shear-thinning fluids: Experiments with Caenorhabditis elegans. J. Fluid Mech. 2014, 758, R3. [Google Scholar] [CrossRef] [Green Version]
  60. Eastham, P.S.; Shoele, K. Axisymmetric squirmers in Stokes fluid with nonuniform viscosity. Phys. Rev. Fluids 2020, 5, 063102. [Google Scholar] [CrossRef]
  61. Zhang, Z.; Igoshin, O.A.; Cotter, C.R.; Shimkets, L.J. Agent-Based Modeling Reveals Possible Mechanisms for Observed Aggregation Cell Behaviors. Biophys. J. 2018, 1152, 499–511. [Google Scholar] [CrossRef] [Green Version]
  62. Be’Er, A.; Strain, S.K.; Hernández, R.A.; Ben-Jacob, E.; Florin, E.L. Periodic Reversals in Paenibacillus dendritiformis Swarming. J. Bacteriol. 2013, 195, 2709–2717. [Google Scholar] [CrossRef] [Green Version]
  63. Sumino, Y.; Nagai, K.H.; Shitaka, Y.; Tanaka, D.; Yoshikawa, K.; Chaté, H.; Oiwa, K. Large-scale vortex lattice emerging from collectively moving microtubules. Nature 2012, 483, 448–452. [Google Scholar] [CrossRef]
  64. Nishiguchi, D.; Nagai, K.H.; Chaté, H.; Sano, M. Long-range nematic order and anomalous fluctuations in suspensions of swimming filamentous bacteria. Phys. Rev. E 2017, 95020601. [Google Scholar] [CrossRef] [Green Version]
  65. Bott, M.C.; Brader, J.M.; Wittmann, R.; Winterhalter, F.; Marechal, M.; Sharma, A. Isotropic-nematic transition of self-propelled rods in three dimensions. Phys. Rev. E 2018, 98, 012601. [Google Scholar] [CrossRef] [Green Version]
  66. Prathyusha, K.; Henkes, S.; Sknepnek, R. Dynamically generated patterns in dense suspensions of active filaments. Phys. Rev. E 2018, 97, 022606. [Google Scholar] [CrossRef] [Green Version]
  67. Hennes, M.; Wolff, K.; Stark, H. Self-Induced Polar Order of Active Brownian Particles in a Harmonic Trap. Phys. Rev. Lett. 2014, 112, 238104. [Google Scholar] [CrossRef] [Green Version]
  68. Pandey, A.; Kumar, P.B.S.; Adhikari, R. Flow-induced nonequilibrium self-assembly in suspensions of stiff, apolar, active filaments. Soft Matter 2016, 12, 9068–9076. [Google Scholar] [CrossRef] [Green Version]
  69. Drescher, K.; Dunkel, J.; Cisneros, L.H.; Ganguly, S.; Goldstein, R.E. Fluid dynamics and noise in bacterial cell-cell and cell-surface scattering. Proc. Natl. Acad. Sci. USA 2011, 108, 10940–10945. [Google Scholar] [CrossRef] [Green Version]
  70. Hu, J.; Wysocki, A.; Winkler, R.G.; Gompper, G. Physical Sensing of Surface Properties by Microswimmers—Directing Bacterial Motion via Wall Slip. Sci. Rep. 2015, 5, 9586. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Lemelle, L.; Palierne, J.F.; Chatre, E.; Vaillant, C.; Place, C. Curvature reversal of the circular motion of swimming bacteria probes for slip at solid/liquid interfaces. Soft Matter 2013, 9, 9759–9762. [Google Scholar] [CrossRef]
  72. Reinken, H.; Klapp, S.H.L.; Bär, M.; Heidenreich, S. Derivation of a hydrodynamic theory for mesoscale dynamics in microswimmer suspensions. Phys. Rev. E 2018, 97, 022613. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Chen, Q.S.; Patelli, A.; Chaté, H.; Ma, Y.Q.; Shi, X.Q. Fore-aft asymmetric flocking. Phys. Rev. E 2017, 96, 020601. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Self-driven particles in nature. (a) Spermatozoa and E. coli; (b) Spirillum; (c) Dinoflagellates.
Figure 1. Self-driven particles in nature. (a) Spermatozoa and E. coli; (b) Spirillum; (c) Dinoflagellates.
Applsci 11 07361 g001
Figure 2. Synthesized self-driven particles. (a) Artificial cell; (b) biomarkers; (c) miniature swimmer.
Figure 2. Synthesized self-driven particles. (a) Artificial cell; (b) biomarkers; (c) miniature swimmer.
Applsci 11 07361 g002
Figure 3. (ad): Passive particles (red) aggregated under the function of self-driven particles (blue) [8]. Adapted with permission from Kümmel et al. (2015).
Figure 3. (ad): Passive particles (red) aggregated under the function of self-driven particles (blue) [8]. Adapted with permission from Kümmel et al. (2015).
Applsci 11 07361 g003
Figure 4. Chlamydomonas cells disperse near the wall [9]. Adapted with permission from Bechinger et al. (2016).
Figure 4. Chlamydomonas cells disperse near the wall [9]. Adapted with permission from Bechinger et al. (2016).
Applsci 11 07361 g004
Figure 5. Squirmer oscillating along the wall [9]. Adapted with permission from Bechinger et al. (2016).
Figure 5. Squirmer oscillating along the wall [9]. Adapted with permission from Bechinger et al. (2016).
Applsci 11 07361 g005
Figure 6. Slip length.
Figure 6. Slip length.
Applsci 11 07361 g006
Figure 7. (a) Escherichia coli with honeycomb structure; (b) Brevibacterium with rafted structure.
Figure 7. (a) Escherichia coli with honeycomb structure; (b) Brevibacterium with rafted structure.
Applsci 11 07361 g007
Figure 8. (a) Ciliata; (b) Bacillus; (c) miniature artificial swimming devices; (d) artificial Janus granules.
Figure 8. (a) Ciliata; (b) Bacillus; (c) miniature artificial swimming devices; (d) artificial Janus granules.
Applsci 11 07361 g008
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liu, C.; Lin, J. A Review on the Some Issues of Multiphase Flow with Self-Driven Particles. Appl. Sci. 2021, 11, 7361. https://doi.org/10.3390/app11167361

AMA Style

Liu C, Lin J. A Review on the Some Issues of Multiphase Flow with Self-Driven Particles. Applied Sciences. 2021; 11(16):7361. https://doi.org/10.3390/app11167361

Chicago/Turabian Style

Liu, Chen, and Jianzhong Lin. 2021. "A Review on the Some Issues of Multiphase Flow with Self-Driven Particles" Applied Sciences 11, no. 16: 7361. https://doi.org/10.3390/app11167361

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop