Soft Matter 14, 490 (2018)

Velocity force curves, laning, and jamming for oppositely driven disk systems

C. Reichhardt and C.J.O. Reichhardt*

Theoretical Division, Los Alamos National Laboratory, Los Alamos, New Mexico 87545, USA. E-mail: cjrx@lanl.gov; Fax: +1 505 606 0917; Tel: +1 505 665 1134

Received 2nd November 2017,
Accepted 1st December 2017
DOI: 10.1039/c7sm02162c
rsc.li/soft-matter-journal
Using simulations we examine a two-dimensional disk system in which two disk species are driven in opposite directions. We measure the average velocity of one of the species versus the applied driving force and identify four phases as a function of drive and disk density: a jammed state, a completely phase separated state, a continuously mixing phase, and a laning phase. The transitions between these phase are correlated with jumps in the velocity-force curves that are similar to the behavior observed at dynamical phase transitions in driven particle systems with quenched disorder such as vortices in type-II superconductors. In some cases the transitions between phases are associated with negative differential mobility in which the average absolute velocity of either species decreases with increasing drive. We also consider the situation where the drive is applied to only one species as well as systems in which both species are driven in the same direction with different drive amplitudes. We show that the phases are robust against the addition of thermal fluctuations. Finally, we discuss how the transitions we observe could be related to absorbing phase transitions where a system in a phase separated or laning regime organizes to a state in which contacts between the disks no longer occur and dynamical fluctuations are lost.
1 Introduction
2 Simulation
3 Velocity force curves and dynamic phases
4 Varied species ratios and driving force ratios
5 Summary
References

1  Introduction

A wide variety of systems can be modeled as a collection of interacting particles that, when driven over a quenched substrate, exhibit depinning and dynamical transitions as a function of increasing driving force [1]. Such systems include vortices in type-II superconductors [2, 3, 4], electron crystals [5], driven colloidal systems [6, 7, 8, 9, 10], and sliding friction [11, 12]. At low drives these systems are in a pinned state where the velocity is zero, while above a critical driving force, the particles become depinned and slide. Within the moving states there can be different dynamical modes of motion such as a plastic phase with strong fluctuations in the particle positions and velocities [1, 2, 4, 6, 7]. At higher drives the system can organize to a dynamically ordered state such as a moving crystal [2, 3, 13] or moving smectic [4, 13, 14, 15, 16]. For particles driven over a periodic substrate, additional types of dynamic phases can appear such as soliton motion or one-dimensional (1D) to two-dimensional (2D) transitions, along with ordered and disordered flow phases [1, 8, 9, 10, 11, 12, 17, 18, 19], negative mobility [20], and sorting dynamics [21].
The transitions between these different dynamical states are associated with cusps, jumps, or dips in the velocity force curves, as well as with global changes in the ordering of the particle configurations or the amount of dynamical fluctuations. In all these systems the pinning arises from quenched disorder that is fixed in space; however, there can also be cases where the pinning is not fixed but can move in response to the driven particles. For example, if a number of particles that are not coupled to the external drive can block the motion of particles that are coupled to the external drive, the driven particles can move the blocking particles and over time rearrange them to create a new landscape or pattern [22, 23]. A previously studied system that closely resembles this case is two species of interacting particles driven in opposite directions that exhibit a variety of dynamical behavior, including a transition to a laning state [24, 25, 26, 27, 28, 29, 30, 31] where the particles separate into quasi one-dimensional chains of the same species, as well as regimes in which the particles mix and undergo disordered flow. Such phases have been observed in experiments on colloids moving in opposite directions [32, 33] and dusty plasma systems [34]. This type of system can also exhibit pattern forming states [27, 28, 35, 36, 37] and jammed or clogged states [27, 38]. Certain active matter systems have similar laning behavior [39], such as pedestrian flows that can be mimicked by particles moving in opposite directions [40]. The velocity-force relations for laning systems have never been systematically studied, nor have parallels between laning systems and systems with quenched disorder that exhibit depinning and distinct dynamic phases been drawn.
In this work we examine the velocity-force relations for a two-dimensional system of disks under lane formation conditions. Half the disks are driven in one direction and the other half in the opposite direction, and we measure the net velocity of one disk species along with the amount of six-fold ordering as a function of the driving force and disk density. Previous studies of laning transitions have generally focused on systems of colloids or particles with Yukawa interactions [24, 26]. In our case the system is close to the hard disk limit and the density ϕ is defined as the area covered by the disks. For ϕ > 0.9 the disks form a triangular solid or jammed state [41]. For ϕ ≥ 0.55 we find that the disks can organize into four possible dynamic phases: a jammed phase (I) where all the disks are in contact forming a triangular solid with zero net velocity; a fully phase separated state (II) where the disks organize into two bands with crystalline order moving in opposite directions and where disk-disk collisions do not occur; a strongly fluctuating disordered phase (III) where disk collisions are continuously occurring and the system has liquidlike features; and a laning state (IV) where the disks form a series of lanes, disk-disk collisions are absent, velocity fluctuations drop to zero, and the system has smectic properties. The transitions between these phases correlate with changes in the net velocity of each disk species as well as with changes in the disk ordering and the nature of the dynamic fluctuations. The mobility in the phase separated and laning states is high since the disks can move freely past one another without collisions, while the transition to the disordered phase is accompanied by a drop in the net velocity, leading to a region of negative differential mobility similar to that found in transitions from laminar flow to turbulent flow as a function of increasing drive in certain systems with quenched disorder [17, 18, 19]. For ϕ < 0.55 the system always organizes into a laning state where all disk collisions are lost. We show that when finite thermal fluctuations are introduced, all the phases are robust at low temperatures. As the temperature increases, phase II vanishes, while the III-IV transition remains robust but shifts to higher values of the external drive. We also examine the situation where only one species is driven and show that the same four phases can arise in the high density limit. Here the jammed state consists of a drifting solid phase where the non-driven disks lock to the driven disks, while the phase separated and laning states are composed of assemblies of driven disks moving past stationary regions of non-driven disks. These phases can even occur when both species are driven in the same direction with different couplings to the external drive.
We conjecture that some of the transitions between phases fall into the class of absorbing phase transitions [42] when the system reaches a state where the disk collisions and fluctuations are completely lost, similar to the recently observed irreversible-reversible transitions in periodically sheared disk systems [43] and systems with quenched disorder [44, 45, 46, 47] that exhibit depinning transitions or transitions into a reversible state. We measure the transient time τ required for the system to organize into phase IV and find that as the III-IV transition is approached from above, τ increases as a power law with an exponent of ν = 1.35, in agreement with previous studies of systems with absorbing phase transitions.

2  Simulation

We model a 2D system of size L ×L where L=36 with periodic boundary conditions in the x and y directions containing Nd disks of radius Rd=0.5. The disk-disk interaction is modeled as a finite range repulsive harmonic spring. The overdamped equation of motion for disk i is
η d Ri

dt
= Fidd + FDi .
(1)
The disk-disk interaction force is Fdd = ∑ijNdk(2Rd − |rij|)Θ(2Rd − |rij|) rij, where rij = RiRj, rij = rij/|rij|, and Θ is the Heaviside step function. The spring constant k = 60, and we find negligible changes in the dynamics for larger values of k. Each disk is subjected to an applied driving force FDi = AiFDx, and N1 disks are driven in the positive x direction with Ai = C, where C=1.0 unless otherwise noted. The remaining N2=NdN1 disks are driven in the negative x direction with Ai = −1.0. After applying the drive, we wait for the system to settle into a steady state. This transient waiting time is a strong function of disk density, and under some conditions can be as large as 1 ×108 simulation time steps. After the system reaches a steady state we measure the average disk velocity for each species and normalize it by N1(2) to obtain the average velocity per disk 〈V1(2)〉 = N1(2)−1N1(2)i=1vi·x , where vi is the instantaneous velocity of disk i. Since we use overdamped dynamics with a damping constant of η = 1.0, in the free flow limit the disks move at a velocity of 〈V1〉 = CFD and 〈V2〉 = −FD. The density ϕ is defined as the area coverage of the disks, ϕ = NdπR2d/L2, and in the absence of driving the system forms a uniform crystalline solid at ϕ = 0.9.

3  Velocity force curves and dynamic phases

Fig1.png
Fig.  1: (a) The average velocity per disk 〈V1〉, measuring only the N1 disks driven in the +x direction, vs FD for a system with oppositely driven disks and N1=N2=0.5Nd at ϕ = 0.848. (b) dV1〉/dFD vs FD for the same system. (c) The corresponding fraction of sixfold coordinated disks P6 vs FD. We find four phases: I (jammed), II (phase separated), III (disordered flow), and IV (laning). The transitions between the states appear as jumps or dips in the various measures.
Fig2.png
Fig.  2: The disk configurations from the system in Fig. 1. The blue disks (species 1) are driven in the +x direction and the red disks (species 2) are driven in the −x direction. (a) The jammed phase I at FD = 0.15. (b) The phase separated state II at FD = 0.75. (c) The disordered phase III at FD = 3.0. (d) The laning phase IV at FD = 6.5.
We first consider samples in which N1=N2=0.5Nd. In Fig. 1(a) we plot 〈V1〉 versus FD for the disks driven in the +x direction for a system with ϕ = 0.848, and in Fig. 1(b) we show dV1〉/dFD versus FD. We also measure the fraction P6 of sixfold coordinated disks for all disks, P6=∑iNdδ(zi−6) where the coordination number zi of disk i is obtained using a Voronoi construction, and plot P6 versus FD in Fig. 1(c). The corresponding 〈V2〉 versus FD curve for disks driven in the −x direction looks exactly the same as the curve in Fig. 1(a) but is negative.
We identify four distinct dynamic phases based on transitions in the velocity-force curve. The jammed phase (I) appears for 0 < FD < 0.3 and has no disk motion, 〈V1〉 = 0, and strong sixfold disk ordering, P6 ≈ 0.95. In phase I, the disks form a dense cluster with triangular order, as illustrated in the disk configuration image in Fig. 2(a) for FD = 0.15. Within the jammed phase the local disk density ϕloc is close to ϕloc=0.9, and since this is lower than the total disk density, there is a small region containing no particles (ϕloc=0). We note that in phase I P6 < 1 since the disks on the edge of the jammed cluster do not have six neighbors.
At FD = 0.3, a jump in 〈V1〉 and a peak in dV1〉/dFD indicate the transition from phase I to phase II, which is similar to the peak in dV〉/dFD observed in systems with quenched disorder at the pinned to sliding transition [1]. Phase II, the phase separated state, extends over the range 0.3 < FD < 1.15, and in this phase 〈V1〉 increases linearly with increasing FD and dV1〉/FD = 1.0, indicating that the particles are in a free flow regime. Additionally, P6 ≈ 0.92 and both species exhibit triangular ordering as shown in Fig. 2(b) at FD = 0.75.
At FD = 1.15 we find a transition from phase II to phase III accompanied by a drop in 〈V1〉 which produces a negative spike in dV1〉/dFD. This is an example of negative differential mobility where the disk velocity decreases with increasing drive. The II-III transition is also associated with a drop in P6 when the system enters the disordered flow phase. In systems with quenched disorder, negative differential mobility has also been reported at transitions from ordered to disordered or turbulent flow phases [11, 17, 18, 19]. Phase III appears for 1.15 < FD < 5.45, and is characterized by strongly fluctuating structures of the type shown in Fig. 2(c) for FD = 3.0. The two particle species are strongly mixed and the sixfold ordering is lost. The disks continuously undergo collisions and show strong velocity deviations in the y direction, transverse to the drive. Within phase III, P6 gradually decreases to P6 ≈ 0.55 near FD = 5.45.
The transition from phase III to phase IV appears as an upward jump in both 〈V1〉 and P6 along with a positive peak in dV1〉/dFD. As shown in Fig. 2(d) at FD=6.5, in phase IV the disks form multiple oppositely moving lanes. There is considerable triangular ordering of the disks within each lane which produces the increase in P6 at the III-IV transition. For higher values of FD, the system maintains phase IV flow and dV1〉/dFD = 1.0, indicating that the disks are in a free flow regime.
Fig3.png
Fig.  3: The structure factor S(k) for the four phases in Fig. 2. (a) The jammed phase I at FD=0.15 has triangular ordering. (b) The phase separated state II at FD=0.75 has triangular ordering. (c) The disordered flow phase III at FD=3.0 shows a ring shape indicating liquid ordering. (d) The laning phase IV at FD=6.5 has a smectic character with weak triangular ordering.
We can characterize the structure of the disks in the different phases using the structure factor S(k)=Nd−1|∑iNdexp(−ik ·ri)|2 . In Fig. 3(a) we plot S(k) for the system in Fig. 2(a) in phase I at FD = 0.75, where we find six peaks indicative of triangular ordering. In phase II, Fig. 3(b) shows a sixfold pattern of peaks with a small amount of smearing of the peaks. The disordered flow phase III in Fig. 3(c) has a ring pattern indicative of liquid ordering, while in Fig. 3(d), the laning phase IV has two strong peaks at kx = 0.0 and four weaker side peaks consistent with a moving smectic structure containing some local triangular ordering. The changes in P6 and S(k) from a liquid structure to a moving smectic signature as a function of increasing FD are similar to the transitions observed for particles moving over quenched disorder such as vortices in type-II superconductors [2, 3, 15, 16].
Fig4.png
Fig.  4: The instantaneous velocity V1 per disk vs time for species 1 for the system in Fig. 1 at ϕ = 0.848. (a) Phase I at FD = 0.15, where V1 goes to zero. (b) Phase II at FD = 0.75, where V1 saturates to a fluctuation-free flowing state with V1 = 0.75. (c) Phase III at FD = 1.5, where the system remains in a strongly fluctuating state with 〈V1〉 = 0.56. (d) Phase IV at FD = 5.57, where the system is initially in a fluctuating state and organizes at later times into a fluctuation-free flowing state with V1 = 5.57.
To identify the dynamic fluctuations of the different phases, we examine the time series of the instantaneous velocity V1(t) of species 1 disks in the different phases. In Fig. 4(a) we plot V1 versus time in phase I at FD = 0.15. Initially V1 is in a fluctuating transient state indicating that the disks are moving, but at later times the system organizes into a jammed state with V1 = 0. In phase III at FD=0.75, Fig. 4(b) shows that there are initially strong fluctuations in V1 but that at later times the system settles into a fluctuation-free flowing state with V1 = 0.75. This corresponds to the formation of the phase separated state, where V2=−0.75. The absence of fluctuations in V1 and the fact that V1=FD indicate that the disks are in a completely free flow state and that disk-disk collisions no longer occur. In Fig. 4(c) for phase III at FD = 1.5, V1 strongly fluctuates between V1=0.4 and V1=0.9, and the average velocity 〈V1〉 = 0.56 is almost three times smaller than the free flow value of 〈V1〉 = 1.5. In this phase, the velocity continues to fluctuate out to the longest simulation times we consider, and the strong fluctuations indicate that there are continuous disk-disk collisions that impede the flow of the disks in both directions. Figure 4(d) shows phase IV flow at FD = 5.57, where the system is initially in a fluctuating flow phase similar to phase III, but organizes at later times into a non-fluctuating free flow state where all disk-disk collisions are lost.
Fig5.png
Fig.  5: (a) 〈V1〉 vs FD for a system with N1=N2=0.5Nd at ϕ = 0.727. (b) The corresponding mobility M1=〈V1〉/FD vs FD shows that the I-II and III-IV transitions are shifted to lower values of FD compared to the ϕ = 0.848 system. (c) 〈V1〉 and (d) the mobility M1 vs FD for a system with ϕ = 0.54, where the disks are always in phase IV, the velocity-force curve is linear, and M1=1.
We investigate the evolution of the phases as a function of ϕ and find that for decreasing ϕ the III-IV transition drops to lower values of FD. In Fig. 5(a) we plot 〈V1〉 versus FD for the system in Fig. 1 at ϕ = 0.727, while Fig. 5(b) shows the corresponding mobility M1=〈V1〉/FD versus FD. The I-II transition has dropped to FD=0.1 and the mobility in phase I is M1=0. The II-III transition occurs at roughly the same value of FD as in the ϕ = 0.848 system, but the III-IV transition drops to FD = 4.0. When the system is in free flow in phases II and IV, M1=1, but the mobility is substantially reduced in the disordered flow phase III. In Fig. 5(c,d) we plot 〈V1〉 and the mobility M1 versus FD for a sample with ϕ = 0.54, where the disks always organize into phase IV flow, the velocity-force curve is linear, and M1=1.
Fig6.png
Fig.  6: Dynamic phase diagram as a function of FD vs ϕ. I: jammed phase; II: phase separated state; III: disordered flow phase; IV: laning phase. For ϕ < 0.55 the system is always in phase IV, while phase III grows in extent with increasing ϕ for ϕ > 0.55. Inset: Instantaneous velocity distributions P(V1) at FD = 1.5 in phase III (blue curve) at ϕ = 0.727, where the distribution is broad, and in phase IV (green curve) at ϕ = 0.545, where the velocities are sharply peaked at V1 = 1.5.
By conducting a series of simulations we map the evolution of the different phases as a function of FD vs ϕ, as shown in the dynamic phase diagram in Fig. 6. For ϕ > 0.55, phases II, III, and IV all occur, while for ϕ < 0.55 only phase IV is present. For ϕ > 0.55, the extent of phase III grows with increasing ϕ, and we only observe phase I for ϕ > 0.7. The II-III transition occurs at roughly the same value of FD as ϕ varies. The minimum value of ϕ = 0.55 at which phases II and III first appear may be related to a contact percolation transition. In compressional simulations of 2D monodisperse frictionless disks, Shen et al. [48] found a contact percolation transition at ϕp = 0.549 and argued that this transition is connected to the onset of a non-trivial mechanical response or stress in the system.
We can also characterize the phases by examining the instantaneous velocity distributions P(V1) and P(V2). In phases I, II, and IV, P(V1) and P(V2) are delta function distributed at the driving force FD or −FD, respectively. In contrast, in phase III the velocity distributions are broad but bounded with a peak falling well below FD. In the inset of Fig. 6 we plot P(V1) at FD = 1.5 for phase III at ϕ = 0.727 and phase IV at ϕ = 0.545. In phase III, P(V1) peaks at v = 0.625 and has a full width at half maximum of 0.25. In phase IV, P(V1) is sharply peaked at V1 = 1.5, indicating that all of the species 1 disks are moving at the drive velocity. We note that for ϕ > 0.91, the system should form a jammed crystalline solid. In this work we remain below the jamming limit so that there is always some room for the disks to move and rearrange.
Fig7.png
Fig.  7: 〈V1〉 vs FD for ϕ = 0.727 in samples with finite thermal fluctuations of magnitude FT = 0.0, 2.0, 4.0, 6.0, and 8.0, from top to bottom. As temperature increases, the III-IV transition shifts to higher FD and phase II disappears.
Thermal fluctuations are relevant in many real soft matter systems, so to test the robustness of the phases we observe against such fluctuations, we conduct a series of simulations in which we add a Langevin noise term FiT to the equation of motion, where 〈FiT〉 = 0 and 〈FTi(t)FTj(t′)〉 = 2 kBTδijδ(tt′). In Fig. 7 we plot 〈V1〉 versus FD at ϕ = 0.727 in samples with FT = 0.0, 2.0, 4.0, 6.0, and 8.0. The four phases described above are present for FT=0, while for FT = 2.0, phases I and II vanish and are replaced by phase III, which still has a sharp boundary at the transition to phase IV. For increasing FT, the III-IV transition remains sharp and shifts to higher values of FD, while 〈V1〉 for a given FD generally decreases.
Fig8.png
Fig.  8: (a) 〈V1〉 vs FT for the finite temperature system in Fig. 7 at FD = 5.0 showing a IV-III transition near FT = 5.0. (b) The same at FD = 0.75 showing a II-III transition near FT = 1.5.
In Fig. 8(a) we plot 〈V1〉 versus FT for the system in Fig. 7 at a fixed FD = 5.0. The zero temperature phase IV flow persists up to FT = 5.0, above which we find disordered phase III flow accompanied by a drop in velocity. In Fig. 8(b), the same system at FD = 0.75 is in phase II up to FT = 1.5, above which there is a drop in velocity as the system enters phase III. Phase IV is more robust against increasing temperature since the disks are moving fast enough in the direction of drive that the disk-disk collisions responsible for maintaining the lane structure occur too rapidly to permit individual disks to diffuse transverse to the direction of drive from one lane to another. In contrast, in the phase II flow that occurs at lower drives, disk-disk collisions are more infrequent so an individual disk can more readily diffuse across the border separating the two disk species, destroying the lane structure and producing disordered phase III flow. We find that all four dynamic phases remain accessible for low but finite thermal fluctuations, and that the multiple lanes of phase IV remain robust at higher temperatures. We have performed limited simulations with larger systems. The transient times required for phases II and IV to form increase with system size, but the disk densities at which the transitions between different phases occur remain robust against changes in system size, suggesting that the transitions are strongly connected to the global packing density.
Fig9.png
Fig.  9: (a) 〈V1〉 vs FD for a system with ϕ = 0.727. Solid line: forward sweep of FD. Dashed line: Reverse sweep of FD starting from FD=6 when the system is in phase IV, showing that phase IV flow persists all the way down to FD=0. (b) The transient time τ in phase IV vs FDFC where FC=3.95 is the drive at which the III-IV transition occurs. The dashed line is a power law fit with exponent ν = 1.35.
Several of the phase transitions we observe have similarities to the irreversible-reversible transition observed in periodically sheared disk suspensions, where the disks transition from an irreversible fluctuating state in which disk-disk collisions occur to a reversible state where collisions are lost [43, 49]. In our system, collisions disappear in phases II and IV, which behave like reversible states, while phase III corresponds to an irreversible state. If the reversible state is absorbing, then once the system has entered the reversible state it should remain trapped regardless of any changes in the value of FD. To test this, we initialize a zero temperature ϕ = 0.727 system in phase IV flow at FD = 6.0 and gradually reduce FD to zero in a series of steps. We find that the system remains in phase IV all the way down to FD = 0, as indicated in Fig. 9(a) by the dashed line. If we start the system in phase III at FD = 2.0 and reduce the drive, the system jumps into phase II near FD = 1.0 and remains in phase II down to FD = 0. Under finite temperature such hysteresis will depend strongly on the rate at which we change FD, which will be the subject of a future work.
In the periodically sheared colloidal system of Ref. [43], the system always starts in a fluctuating state and over time organizes into either a steady fluctuating state or a nonfluctuating absorbing state. The transient time τ required to reach either state diverges as a power law at the transition point. Our system always starts in phase III and either remains in phase III or organizes into phase II or IV depending on the value of FD. Focusing on the III-IV transition at FC=3.95 in a system with ϕ = 0.727, we conduct a series of simulations with FD > FC and measure the transient times τ required for phase IV flow to develop from the initial phase III state as FD approaches FC. In Fig. 9(b) we plot τ averaged over different initial state realizations versus FDFC. The dashed line is a power law fit with ν = 1.35 ±0.15. A similar transient time exponent appears in the irreversible-reversible transition and depinning systems, and our value of ν agrees with the exponent expected for an absorbing phase transition [42, 43, 44, 45, 46, 47, 49]. In the colloidal shearing system, transient times could be measured on both sides of the reversible-irreversible transition; however, we are unable to identify a clear relaxation time within phase III for FD < FC. The idea that the disks have reached an absorbing state is only applicable for T = 0 or very low temperatures, since once thermal fluctuations are introduced, it may become possible for the system to escape from a state that was absorbing at zero temperature.

4  Varied species ratios and driving force ratios

Fig10.png
Fig.  10: (a) 〈V1〉 vs FD for a sample with ϕ = 0.848 in which N1=0.2Nd and N2=0.8Nd. (b) The corresponding mobility M1 vs FD. Here M1=1.0 in phases II and IV, but there is a portion of phase III for which the mobility is nearly zero. Phase I is no longer a jammed phase with 〈V1〉 = 0. Instead, a rigid cluster forms that drifts in the −x direction, producing a negative mobility.
We have also considered different ratios of N1 to N2 and find that the same general phases appear. In Fig. 10(a) we plot 〈V1〉 versus FD for a system with ϕ = 0.848, N1=0.2Nd, and N2=0.8Nd, while Fig. 10(b) shows the corresponding mobility M1 versus FD. Here the same four phases arise and the overall shape of the curves is similar to that of the N1=N2=0.5Nd system shown in Fig. 1; however, in phase III the average mobility drops nearly to zero since the additional collisions suppress the flow of the disks in the +x direction. In phases II and IV, M1=1, indicating that the system can still organize into a collisionless state where the disks undergo free flow motion. A notable difference is that for phase I in the N1=0.2Nd sample, the mobility M1 is negative. The jammed phase I for the N1=0.5Nd sample has 〈V1〉 = 0, but in the N1=0.2Nd sample, the jammed state consists of a rigid solid that translates in the −x direction since N2 > N1. In general, for any ratio other than N1/N2=1, the jammed phase I has a net drift in the direction of the majority species. As N1/N2 is decreased further, in phases II and IV we always find M1=1, while for phases I and III the mobility decreases or becomes more negative.
We can also change the ratio of the relative driving force on the two different species. We consider a system with N1=N2=0.5Nd at ϕ = 0.848 and vary the value of C controlling the amplitude of the driving force for species 1, while keeping the drive on species 2 fixed at FDi=−Fdx. Here C is chosen in the range −1.0 ≤ C ≤ +1.0. The exactly oppositely driven disks in Fig. 1 corresponds to the case C=+1.0. For C = 0, half of the disks do not couple to the external drive while the other half are driven in the negative x direction, and for −1.0 ≤ C < 0, both species are driven in the −x direction with different forces. At C = −1.0, all the particles are driven in the −x direction with the same force. In Fig. 11 we show a dynamic phase diagram as a function of FD versus C. All of the phase transitions shift to higher values of FD as C decreases. It is interesting to note that for C=0, where only species 2 is driven, all four dynamic phases still occur.
Fig11.png
Fig.  11: Dynamic phase diagram as a function of FD vs C, where C is the coefficient controlling the amplitude of the driving force for species 1 disks, in a sample with N1=N2=0.5Nd and ϕ = 0.848. The value C = 1.0 corresponds to the case shown in Fig. 1 for exactly oppositely driven particles. Yellow circles: jammed phase I; red circles: phase separated state II; blue circles: disordered flow phase III; and green circles: laning phase IV. At C = 0 where only species 2 couples to the drive, all four phases can still occur.
Fig12.png
Fig.  12: A sample from Fig. 11 with C=0 and FD=1.0. (a) The instantaneous velocity per disk V1 vs time in simulation time steps (upper blue curve) for the non-driven disks and the corresponding V2 (lower pink curve) vs time for the disks driven in the −x direction. Here the system organizes into phase II flow with V1 = 0 and V2 = −FD=−1.0. (b) P6 vs time for all the disks in the same system showing the transition into phase II. (c) V1 and V2 vs time for the same system at FD = 0.25 and C = 0. The disks organize into phase I where both species become locked together and move at V1=V2 = −0.125 = FD/2. (d) The corresponding P6 vs time curve shows that the system organizes into a mostly triangular state at the transition to phase I.
In Fig. 12(a) we plot the instantaneous disk velocities V1 and V2 versus time for C = 0 and FD=1.0. Here, species 1 is not driven. The system organizes into phase II as indicated by the transition to V1=0 and V2=−1.0. The velocities rapidly approach the phase II values at short times, but significant velocity fluctuations persist and appear as jumps in both V1 and V2 in the transient fluctuating state. In Fig. 12(b), the corresponding P6 versus FD for all the disks shows that the large fluctuations in V1,2 correlate with drops in P6, and that when the system fully settles into phase II flow, P6 saturates to P6=0.925, indicating mostly triangular ordering. In Fig. 13(a) we show the disk configurations and trajectories in the phase II flow for the C=0 system in Fig. 12(a). The disks phase separate into an immobile triangular packing surrounded by a moving triangular lattice. At high drives for C=0, a laning phase IV appears where half of the lanes are immobile and the other half are moving.
Fig13.png
Fig.  13: The disk configurations for the system in Fig. 12 with C=0. The blue disks (species 1) are not driven and the red disks (species 2) are driven in the −x direction. (a) The phase separated state II for the system in Fig. 12(a,b) at FD = 1.0. The black lines are the disk trajectories indicating that the system has phase separated into moving and non-moving spatial regions. (b) The jammed state I at FD = 0.25 from Fig. 12(c,d) where the entire system is moving in the negative x-direction.
The jammed phase I at C=0 consists of a moving solid drifting in the negative x-direction, as illustrated in Fig. 12(c) where we plot V1 and V2 versus time at FD = 0.25. Here the velocities of both disk species converge to V1=V2 = −0.125 = −FD/2 in phase I since the non-driven disks reduce the velocity of the driven disks to half of its free-flow value. Figure 12(d) shows the corresponding P6 versus time, where at the transition into phase I, P6=0.93 indicating the formation of a triangular drifting solid. In Fig. 13(b) we show the disk configurations in phase I for the system in Fig. 12(c), where the driven disks drag the non-driven disks.
Phases I and II persist for −1.0 < C < 0.0 in the range of FD that we have investigated, and at C = −1 all the disks move in unison. The fact that these phases can occur over a range of relative drives indicates that such phases may be general to other systems in which the particles are not exactly oppositely driven but have differences in their relative driving. These differences could arise for particles with different drag coefficients, different coupling to a substrate, different amounts of charge, different shapes, and so forth. Such effects could also be realized for active matter systems where one species couples to an external drift force or where only one species is active. There are already examples of mixtures of active and non-active particles that undergo phase separation [50]. Another interesting feature of this system at C = 0 is that certain regimes such as phase II flow exhibit a time dependent resistance. When the drive is initially applied, the sample enters a high resistance state in which many collisions occur, but over time the disks approach a low resistance state as the number of collisions are reduced. This suggests that if finite temperature were included, then when the drive is shut off there would be a finite time during which the memory of the low-resistance organized state is preserved. Such a system would have features similar to those of a memristor [51].

5  Summary

We have numerically investigated a two-dimensional disk system in which two disk species are driven in opposite directions. We characterize the system by measuring the average velocity of one disk species as a function of drive to create a velocity-force curve analogous to what is observed in driven systems with quenched disorder. The density ϕ is the total area coverage of the disks, and for ϕ > 0.55 we identify four dynamical phases: a crystalline jammed phase I, a fully phase separated state II where the particles are not in contact but exhibit six-fold ordering, a strongly fluctuating liquid phase III where continuous collisions are occurring, and a laning phase IV or smectic state where collisions are absent. The transitions between these different phases are associated with jumps or dips in the velocity-force curves and the differential mobility along with global changes in the disk configurations. At the transition into phase III we find negative differential mobility. The transitions also correlate with jumps in the amount of six-fold ordering. For ϕ < 0.55, the system always organizes into the laning phase IV and the velocity-force curves are linear. We vary the relative driving forces on the two species, and consider the case where one species is driven while the other species does not couple to the drive as well as the case where both species are driven in the same direction with different relative forces. We find that the same four driven phases found for the oppositely driven case still occur. By measuring the instantaneous disk velocities, we find that in phases II and IV, the disks organize into a freely flowing state in which disk-disk collisions no longer occur. The phases are robust against the addition of finite temperature, but phase II disappears at much lower temperatures than phase IV. We discuss how the transitions into phases II and IV may be related to the irreversible-reversible transitions recently observed for periodically sheared colloidal or granular systems, which organize to an absorbing state in which collisions between particles are lost. If we start the system at a drive that generates phase II or phase IV flow, which act as reversible or absorbing states, it will remain in that phase as the drive is decreased all the way to zero. We also find that near the III-IV transition, the transient times required to reach a steady state grow as a power law with exponent ν = 1.35, in agreement with the predictions for absorbing phase transitions. Possible physical realizations of our system include noncharged flowing colloids where one species has a larger damping constant than the other, sedimenting colloids with different sedimentation speeds, mixtures of magnetic and nonmagnetic colloids, and certain active matter systems.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

We thank N. Bain for useful discussions. This work was carried out under the auspices of the NNSA of the U.S. DoE at LANL under Contract No. DE-AC52-06NA25396.

References

[1]
C. Reichhardt and C.J.O. Reichhardt, Rep. Prog. Phys., 2017, 80, 026501.
[2]
S. Bhattacharya and M. J. Higgins, Phys. Rev. Lett., 1993, 70, 2617.
[3]
A. E. Koshelev and V. M. Vinokur, Phys. Rev. Lett., 1994, 73, 3580.
[4]
C.J. Olson, C. Reichhardt, and F. Nori, Phys. Rev. Lett., 1998, 81, 3757.
[5]
F.I.B. Williams, P.A. Wright, R.G. Clark, E.Y. Andrei, G. Deville, D.C. Glattli, O. Probst, B. Etienne, C. Dorin, C.T. Foxon, and J.J. Harris, Phys. Rev. Lett., 1991, 66, 3285.
[6]
A. Pertsinidis and X.S. Ling, Phys. Rev. Lett., 2008, 100, 028303.
[7]
P. Tierno, Phys. Rev. Lett., 2012, 109, 198304.
[8]
T. Bohlein, J. Mikhael, and C. Bechinger, Nature Mater., 2012, 11, 126.
[9]
C. Reichhardt and C.J.O. Reichhardt, Phys. Rev. E , 2012, 85, 051401.
[10]
D. McDermott, J. Amelang, C.J.O. Reichhardt, and C. Reichhardt, Phys. Rev. E, 2013, 88, 062301.
[11]
O.M. Braun, T. Dauxois, M.V. Paliy, and M. Peyrard, Phys. Rev. Lett., 1997, 78, 1295.
[12]
A. Vanossi, N. Manini, M. Urbakh, S. Zapperi, and E. Tosatti, Rev. Mod. Phys., 2013, 85, 529.
[13]
P. Le Doussal and T. Giamarchi, Phys. Rev. B, 1998, 57, 11356.
[14]
L. Balents, M.C. Marchetti, and L. Radzihovsky, Phys. Rev. B, 1998, 57, 7705.
[15]
F. Pardo, F. de la Cruz, P.L. Gammel, E. Bucher, and D.J. Bishop, Nature (London) , 1998, 396, 348.
[16]
A. Kolton, D. Domínguez, and N. Grønbech-Jensen, Phys. Rev. Lett., 1999, 83, 3061.
[17]
C. Reichhardt, C. J. Olson, and F. Nori, Phys. Rev. Lett., 1997, 78, 2648.
[18]
J. Gutierrez, A. V. Silhanek, J. Van de Vondel, W. Gillijns, and V. Moshchalkov, Phys. Rev. B, 2009, 80, 140514.
[19]
C. Reichhardt and C.J.O. Reichhardt, Phys. Rev. B, 2008, 78, 224511.
[20]
R. Eichhorn, J. Regtmeier, D. Anselmettib, and P. Reimann, Soft Matter, 2010, 6, 1858.
[21]
F. Martinez-Pedrero et al Phys. Chem. Chem. Phys, 2016, 18, 26353
[22]
C. Reichhardt and C.J.O. Reichhardt, Phys. Rev. E, 2006, 74, 011403.
[23]
R. Candelier and O. Dauchot, Phys. Rev. E, 2010, 81, 011304.
[24]
J. Dzubiella, G.P. Hoffmann, and H. Löwen, Phys. Rev. E, 2002, 65, 021402.
[25]
R.R. Netz, Europhys. Lett., 2003, 63, 616.
[26]
T. Glanz and H. Löwen, J. Phys.: Condens. Matter, 2012, 24, 464114.
[27]
K. Klymko, P.L. Geissler, and S. Whitelam, Phys. Rev. E, 2016, 94, 022608.
[28]
E. M. Foulaadvand and B. Aghaee, Eur. Phys. J. E , 2016, 39, 37.
[29]
C.W. Wächtler, F. Kogler, and S.H.L. Klapp, Phys. Rev. E, 2016, 94, 052603.
[30]
A. Poncet, O. Bénichou, V. Démery, and G. Oshanin, Phys. Rev. Lett., 2017, 118, 118002.
[31]
K. Ikeda and K. Kim J. Phys. Soc. Jpn., 2017 86, 044004
[32]
T. Vissers, A. Wysocki, M. Rex, H. Löwen, C.P. Royall, A. Imhof, and A. van Blaaderen, Soft Matter, 2011, 7, 2352.
[33]
T. Vissers, A. van Blaaderen, and A. Imhof Phys. Rev. Lett., 2011, 106, 228303.
[34]
K. Sütterlin, A. Wysocki, A. Ivlev, C. Räth, H. Thomas, M. Rubin-Zuzic, W. Goedheer, V. Fortov, A. Lipaev, V. Molotkov, O. Petrov, G. Morfill, and H. Löwen, Phys. Rev. Lett., 2009, 102, 085003.
[35]
B. Schmittmann, K. Hwang, and R.K.P. Zia, Europhys. Lett., 1992, 19, 19.
[36]
M. Ikeda, H. Wada, and H. Hayakawa, EPL, 2012, 99, 68005.
[37]
O. A. Vasilyev Olivier Benichou, C. Mejia-Monasterio , E. R. Weeks and G. Oshanin Soft Matter, 2017, 13, 7617
[38]
D. Helbing, I.J. Farkas, and T. Vicsek, Phys. Rev. Lett., 2000, 84, 1240.
[39]
N. Bain and D. Bartolo, Nat. Commun. (2017). 15969 8, 15969.
[40]
D. Helbing, Rev. Mod. Phys., 2001, 73, 1067.
[41]
C. Reichhardt and C.J.O. Reichhardt, Soft Matter, 2014, 10, 2932.
[42]
H. Hinrichsen, Adv. Phys., 2000, 49, 815.
[43]
L. Corté, P.M. Chaikin, J.P. Gollub, and D.J. Pine, Nature Phys., 2008, 4, 420.
[44]
C. Reichhardt and C.J.O. Reichhardt, Phys. Rev. Lett., 2009, 103, 168301.
[45]
S. Okuma, Y. Tsugawa, and A. Motohashi, Phys. Rev. B, 2011, 83, 012503.
[46]
G. Shaw, P. Mandal, S.S. Banerjee, A. Niazi, A.K. Rastogi, A.K. Sood, S. Ramakrishnan, and A.K. Grover, Phys. Rev. B, 2012, 85, 174517.
[47]
C. Reichhardt and C.J.O. Reichhardt, Soft Matter, 2014, 10, 7502.
[48]
T. Shen, C.S. O'Hern, and M.D. Shattuck, Phys. Rev. E, 2012, 85, 011308.
[49]
C.F. Schreck, R.S. Hoy, M.D. Shattuck, and C.S. O'Hern, Phys. Rev. E, 2013, 88, 052205.
[50]
C. Bechinger, R. Di Leonardo, H. Löwen, C. Reichhardt, G. Volpe, and G. Volpe, Rev. Mod. Phys., 2016, 88, 045006.
[51]
L.O. Chua, Appl. Phys. A, 2011, 102, 765.



File translated from TEX by TTHgold, version 4.00.
Back to Home