Europhys. Lett. 75, 489 (2006)

Statics and Dynamics of Two-Dimensional Vortex Liquid Crystals

C. Reichhardt and C.J. Olson Reichhardt

Theoretical Division and Center for Nonlinear Studies, Los Alamos National Laboratory, Los Alamos, New Mexico 87545

received 28 November 2005; accepted in final form 7 June 2006
published online 28 June 2006

PACS. 74.25.Qt - Vortex lattices, flux pinning, flux creep.
Abstract. - With numerical simulations we examine static and dynamic properties of two-dimensional vortices with anisotropic interactions. We find evidence for a smectic-A phase in the absence of pinning. Quenched disorder can induce a smectic type state even at T=0. When an external drive is applied, a variety of novel anisotropic dynamical flow states with distinct voltage signatures occur, including elastic depinning in the hard direction and plastic depinning in the easy direction. We compare our work to recent experiments on two-dimensional colloids with anisotropic interactions. We also discuss the implications of the anisotropic transport for other systems which exhibit depinning phenomena, such as stripes and electron liquid crystals.
Recently, a new state of vortex matter termed a vortex liquid crystal was argued to occur in superconductors with anisotropic vortex-vortex interactions [1]. In such systems, the vortex lattice first melts in the soft direction, giving rise to an intermediate vortex smectic-A state, followed at higher temperatures by a melting into a nematic state. The smectic-A state is distinct from smectic-C states observed for vortices or colloids interacting with one-dimensional (1D) periodically modulated substrates [2]. In the smectic-A case, the anisotropy arises from the elliptical vortex cross section produced by the anisotropic superfluid stiffness which leads to different effective masses in the three crystalline directions [1]. The prediction in Ref. [1] combined an elastic model with the Lindemann criterion for melting, but it was later argued in Ref. [3] that this approach is not sufficient to determine whether the smectic-A state exists, since scaling theory alone predicts that the smectic-A state will not occur. The Lindemann criterion does not take into account the proliferation of dislocations which occurs in the smectic state. It is possible that once dislocations appear, melting will immediately occur in both directions, precluding the smectic-A state. Although a true 2D smectic-A state has been predicted to be unstable for rod like colloids [4], recent experiments on anisotropically interacting 2D colloids have found evidence for a smectic state in which the dislocations are aligned only along the symmetry directions of the lattice [5]. Here, the anisotropic particle interactions may be producing collective motions that allow for only one type of dislocation. It is thus an open question whether the smectic-A state exists.
The physics of a 2D vortex system with anisotropic interactions should be generic to the class of problems involving anisotropically interacting 2D particles, such as colloidal systems with anisotropic magnetic interactions, as well as 2D electron crystal states formed by anisotropic interactions in classical electron crystals [6]. Evidence for such states has been observed in transport measurements which show hard and soft directions for flow [7,8]. The interaction of the proposed anisotropic vortex lattice with quenched disorder is unexplored, and it is not known whether disorder would completely destroy possible smectic type orderings. The transport properties of vortex systems with anisotropic interactions have not been studied previously; these could be used to identify new states in experiment, and may show new dynamical phenomena. Of broader interest is the fact that understanding the behavior of vortex smectic states in the presence of disorder can offer insight into the effect of disorder on the general class of systems of particles with anisotropic repulsive interactions, including dynamics in electron liquid crystal states.
We consider a 2D system of Nv interacting vortices with periodic boundary conditions in the x and y directions. The overdamped equation of motion for a single vortex i is
η dri

dt
= fivv + fTi + fpi + fdi
(1)
The damping constant η = 1. The vortex-vortex interaction force is fivv = ∑NvjiAvK1(rij/λ)rij, where K1 is the modified Bessel function appropriate for stiff, 3D vortex lines [9], which decays exponentially for large distances, λ is the London penetration depth, Av02/(2πμ0λ3), and rij is the distance between vortices i and j. The thermal force fTi arises from random Langevin kicks with the properties 〈fTi〉 = 0 and 〈fiT(t)fjT(t)〉 = 2ηkBT δ(ttij. The quenched disorder fip is modeled as random pinning sites in the form of attractive parabolic traps of radius rp=0.2λ and strength fp. The Lorentz driving force from an external applied current is fd. The system size is measured in units of λ, forces in units of Av, energies in Avλ, and temperature in Avλ/kB. The anisotropic interactions are introduced by multiplying the vortex-vortex interaction force in the x and y directions by a vector (Cx, Cy), where the anisotropy C=Cx/Cy. In this work we concentrate on the case C = 1/√{10} considered in Ref. [1]. We take the x axis to be the soft direction and the y axis as the hard direction. We have also modeled vortices in a thin film superconductor, where the vortex-vortex interaction has the form fvv = Avrij/rij, with Av = Φ200πΛ and where Λ is the thin film screening length [10]. To evaluate the long-range interactions we use a fast summation method [11]. We find the same qualitative features with this potential, and also with a screened Coulomb potential of inverse screening length κ, exp(−rκ)/r, appropriate for charge-stabilized colloids. The initial vortex configurations are obtained through simulated annealing.
Fig1.png
Figure 1: (a,c,e) Black dots: vortices; black lines: vortex trajectories. (b,d,f) Delaunay triangulation, with topological defects (5 and 7-fold coordinated particles) marked as filled circles. (a,b) T = 0.5; (c,d) T = 1.2; (e,f) T = 1.35.
We first consider the case where the pinning and the external driving force are absent. In Fig. 1 we illustrate the melting of a 24λ×24λ system with a vortex density of ρv=1.2/λ2. Figure 1(a) shows the vortex positions (dots) and trajectories (lines) for a fixed period of time with a fixed T = 0.5, and Fig. 1(b) shows a corresponding Delaunay triangulation. At this temperature, the system remains in a crystalline state with no dislocations. The vortices are undergoing larger random displacements in the soft (x) direction than in the hard (y) direction; however, there is no long time diffusion of the particles. Figures 1(c) and 1(d) present the smectic-A state at T=1.2. Here the trajectories have a 1D liquid structure with motion along the soft x direction and no significant translation of the vortices in the y direction. The Delaunay triangulation indicates the presence of dislocations with aligned Burgers vectors, which is characteristic of the smectic-A state and is similar to the recent colloid experiments [5]. In Fig. 2(d) we plot the density of sixfold coordinated particles P6 versus system size L in the crystal and smectic-A states, showing the saturation of P6 for all but the smallest samples. In the smectic-A state, motion in the soft direction occurs in the form of a pulse in which the vortices translate by a single lattice constant a in the +x or −x direction. Similar pulse-like motion has been observed in vortex chain states [12]. It is possible for neighboring vortices in the same row to trade places by moving into the y direction a distance smaller than a. Figures 1(e) and 1(f) illustrate the vortex liquid phase at T = 1.35. The vortex trajectories show clear diffusion in both the x and y directions, with more pronounced motion in the x direction. The dislocations are no longer aligned, indicating the loss of long-range order in both the x and y directions. We note that when the anisotropy ratio C is too small, the two-step melting transition illustrated here is lost.
To further characterize the smectic state, in Fig. 2 we plot the average particle displacements for the x and y directions, dx = 〈∑iNv|xi(0) − xi(t)|〉/Nv and dy = 〈∑iNv|yi(0) − yi(t)|〉/Nv. In the smectic phase at T = 1.21, shown in Fig. 2(a), dx/a increases much more rapidly than dy/a, and does not saturate but increases to a value over 1, indicating that the vortices can diffuse more than a lattice constant in the x direction over time. This is due to the formation of dislocations which allow adjacent rows of vortices to slip past each other while remaining confined in the y direction. The saturation value of dy/a is approximately 1/5, larger than the Lindemann criterion value of 1/10. Excess motion in the y direction occurs during a sliding event when two rows slip past each other and the vortices in each row are temporarily displaced in the direction perpendicular to the slip plane. This transverse motion is too small to permit the formation of dislocations aligned in the hard direction. Similar behavior was observed experimentally in the smectic phase in Ref. [5]. For times longer than illustrated in the figure, dy saturates completely.
In the anisotropic liquid phase, shown in Fig. 2(b) at T = 1.35, dx still increases more rapidly than dy; however, the continuous increase of both quantities indicates that the particles are diffusing throughout the entire system. We have not determined whether the diffusion is normal or anomalous in the smectic region, but in the liquid region it appears normal.
In Fig. 2(c) we plot the peak values of the structure factor S(q) = (1/L2)∑i,jexp(iq·[ri(t) −rj(t)]) for the two different directions. Fig. 2(c) shows that the peak corresponding to the soft direction, S(qs), decreases in magnitude much more rapidly with T than the peak corresponding to the hard direction, S(qh). Near T = 1.0 the value S(qs) drops markedly while S(qh) does not undergo a steep drop until T = 1.32, showing that the system is in a smectic phase for 1.0 < T < 1.32. For T > 1.32 the system is in the anisotropic liquid phase. The smectic-A state illustrated in Fig. 1 also appears for 1/r vortex-vortex interactions appropriate for thin film superconductors. Due to the reduced shear modulus, the smectic-A phase occurs over a lower range of temperatures; however, a similar sequence of phases occurs.
Fig2.png
Figure 2: Average particle displacements in each direction, dx and dy, normalized by the lattice constant a, vs time, measured in molecular dynamics steps. (a) Smectic-A state at T = 1.21. (b) Anisotropic liquid phase at T=1.35. (c) The peaks in the structure factor vs temperature T for the soft direction S(qs) (squares) and hard direction S(qh) (circles). (d) P6 vs system size L at T=0.5 (upper curve) and T=1.21 (lower curve) for the system in Fig. 1. (e) dP6/dT vs T. (f) Tc1 (squares) and Tc2 (circles) vs L.
We have also measured P6 vs T which shows a two step feature with an initial dip at the onset of the smectic phase followed at higher temperature by a larger dip when the system enters the anisotropic liquid phase, as illustrated in Fig. 2(e). For larger systems the two peaks become more pronounced. We extract the temperatures for the onset of the smectic phase, Tc1 and the liquid phase, Tc2, for different system sizes and find that for large systems Tc2Tc1 saturates to a constant nonzero value, as shown in Fig. 2(f). We do not find hysteresis in any of these quantities if we cycle the temperature through these transitions.
We next consider the effect of random disorder by adding Np=2Nv randomly located pinning sites to the same system studied in Fig. 1, and then conducting a series of simulations at varied T and varied pinning strength fp. For high temperatures we obtain a liquid phase, while for low T and small fp we observe what we term a pinned smectic phase similar to that shown in Fig. 1(c,d), where the vortex lattice contains a small number of dislocations aligned in the soft direction. In this pinned state, diffusion along the rows is suppressed by the pinning. For higher fp we observe that dislocations which are not aligned with the soft direction start to appear, and the system enters a disordered phase.
In Fig. 3(a) we indicate the regions in which the smectic and disordered phases appear as a function of temperature and pinning strength. The boundaries in Fig. 3(a) are identified via Delaunay triangulations, which enable both the orientation of the dislocations and the density of sixfold coordinated particles P6 to be measured. In the crystal phase, there are no defects and P6=1. In the smectic phase, P6=0.91 to 0.95, and in the disordered phase P6 < 0.9 and misoriented defects appear. In the inset of Fig. 3(a) we plot P6 vs T for two different disorder strengths. For fp = 0.025 (upper line) the system is in the pinned smectic state at T = 0. As T increases, there is a clear transition to the disordered state, as indicated by the drop in P6 near T = 1.19. The lower line shows P6 for fp=0.2, when the pinning is strong enough to disorder the system even at T = 0. These results suggest that weak random disorder can increase the extent of the regions where the smectic-A phase occurs when there are anisotropic interactions, by suppressing the crystalline phase at low temperatures and raising the melting temperature of the smectic state.
Fig3.png
Figure 3: (a) Regions in which the smectic and disordered phases occur in a system with quenched disorder as a function of temperature T and pinning strength fp. Dashed line roughly indicates the weak pinning region in which a 2D anisotropic Bragg glass forms. Inset: the density of six-fold coordinated particles P6 vs T for (top curve) fp = 0.025 and (bottom curve) fp = 0.2. (b) P6 vs driving force fd for a system with fp = 0.04 at T=0. Upper curve: Py6 for fd=fdy. Lower curve: Px6 for fd=fdx. (c) Average velocities vs fd for the same system. Upper curve: Vy for fd=fdy. Lower curve: Vx for fd=fdx. (d) Vx and Vy vs fd for a system with fp = 0.2. (e) A blowup of (c) in the region near depinning showing the crossing of the velocity force curves.
One question is whether the smectic state with random disorder is stable for very large systems. It has been argued that dislocations are always present for weak disorder in 2D isotropic systems; however, the distance between between the dislocations can become arbitrarily large compared to the range of the translational order, so that for a wide range of temperatures and disorder strengths the system behaves as a dislocation free 2D Bragg glass [13]. In the vortex liquid crystal case there are two length scales associated with the hard and soft directions, and the disorder induced dislocations first form in the soft direction. It may be possible that, on very large length scales, dislocations in the hard direction will also appear and create a true smectic state. For the parameters considered here, dislocations are present except at the lowest pinning strengths, roughly indicated by a dashed line in Fig. 3(a), where a 2D anisotropic Bragg glass forms. The positions of the lines in Fig. 3(a) do not shift with system size.
We next consider dynamical effects in the presence of pinning. In the smectic state, there should be distinct transport signatures for the hard and soft directions. When the disorder is strong enough to destroy the smectic phase, there may still be an anisotropic transport signature. We first consider the pinned smectic state found at fp = 0.04 and T = 0. We perform separate simulations for driving in the soft direction, fd=fdx, and the hard direction, fd=fdy, increasing the applied drive for a total of 5 ×107 MD steps which is slow enough to avoid any transient effects. In Fig. 3(c) we plot Vy=(1/Nv)〈∑iNvvy〉 (upper curve) for driving in the hard direction and Vx=(1/Nv)〈∑iNvvx〉 (lower curve) for driving in the soft direction. Above depinning in Fig. 3(c), Vx < Vy, indicating that pinning has the largest effect on motion in the soft direction. In Fig. 4(b) we plot P6 for the two different driving directions. At fd = 0, P6 is slightly less than one due to the presence of a small number of dislocations in the smectic state. For driving in the hard direction, P6y (upper curve), the system depins elastically without additional proliferation of defects, and the vortices do not exchange neighbors as they move. For driving in the easy direction, P6x (lower curve) drops substantially when the vortices depin plastically, and a portion of the vortices remain pinned while others flow past. A similar proliferation of defects has been associated with the so called peak effect, where the effective pinning force suddenly increases as a function of temperature or applied magnetic field [14]. Further, the elastic and plastic depinning transitions produce different scaling responses in the velocity force curves. At depinning, the velocity scales with the driving force in the form V = (fdfc)β [16]. For the plastic flow regime we find β > 1.0 while in the elastic flow regime we find β < 1.0, in agreement with theoretical expectations. Figure 4(b) shows that at higher drives, the system dynamically reorders [15], as indicated by the increase in P6x, as well as by the merging of Vx and Vy in Fig. 4(c).
At finite temperatures and for fp large enough that we observe only plastic depinning in both directions, we observe that the depinning force in the soft direction, fcx, is lower than the depinning force in the hard direction, fcy, even though Vx < Vy at intermediate drives. This implies that the anisotropic flow exhibits a reversal from Vx > Vy to Vx < Vy at low drives. We explicitly demonstrate this effect for a system with fp=0.25 and T=0.25 in Fig. 4(d,e). Here, fcx=0.015 and fcy=0.11, and the depinning is plastic in both directions. There are fewer dislocations for fd=fdy and the system reorders at fdy = 0.2. For fd=fdx, the system does not reorder until fdx = 0.8. There is a clear crossing of the velocity force curves at fd = 0.13 so that the flow is easier in the soft direction for fd < 0.13 and easier in the hard direction for fd > 0.13. In Fig. 4(e) we show a blowup of this region. The crossing of the velocity force curves can be understood by considering that the depinning in the soft direction is plastic. At low drives, individual vortices can be thermally activated, giving rise to creep. For driving in the hard direction, the depinning is elastic and individual vortex hopping is not possible, so that only collective creep can occur. In the case of the disordered phase, when there is some plastic flow in the y-direction, there is still a large correlated length scale that must move so thermal effects are greatly reduced. Thus, creep in the pinned smectic phase and pinned disordered phase is enhanced in the soft direction compared to the hard direction.
In conclusion, using numerical simulations we find evidence that an intermediate vortex smectic-A state can occur when the vortex-vortex interactions are anisotropic for both short and long range interaction potentials. The smectic-A state contains a small fraction of dislocations which are all aligned in the soft direction. In the presence of disorder, a pinned smectic state can occur where dislocations oriented only in the soft direction appear for the system sizes we have considered. The system depins plastically in the soft direction but elastically in the hard direction. We predict that, for equal intermediate drives, the velocity in the soft direction will be lower than in the hard direction. At finite temperatures the creep is much larger in the soft direction due to the fact that individual vortex hopping can occur, whereas creep is suppressed in the hard direction since the vortex motion is much more correlated. For high temperatures and disorder strengths, the system is disordered. For strong quenched disorder, anisotropic transport should still be observable. Our results may also apply to electron liquid crystals.
We thank E. Carlson for useful discussions. This work was supported by the U.S. Department of Energy under Contract No. W-7405-ENG-36.

References

[1]
Carlson E.W., Castro Neto A.H., and Campbell D.K. Phys. Rev. Lett. 90 (2003) 087001; Phys. Rev. Lett. 92 (2004) 209702.
[2]
Balents L and Nelson D.R. Phys. Rev. B 52 (1995) 12951; Wei Q.-H., Bechinger C., Rudhardt D., and Leiderer P. Phys. Rev. Lett. 81 (1998) 2606; Radzihovsky L., Frey E., and Nelson D.R. Phys. Rev. E 63 (2001) 031503.
[3]
Hu X. and Chen Q.H. Phys. Rev. Lett. 92 (2004) 209701.
[4]
Toner J. and Nelson D.R. Phys. Rev. B 23 (1981) 316; Ostlund S. and Halperin B.I. Phys. Rev. B 23 (1981) 335.
[5]
Eisenmann C., Gasser U., Keim P., and Maret G. Phys. Rev. Lett. 93 (2004) 105702.
[6]
Fradkin E. and Kivelson S.A. Phys. Rev. B 59 (1999) 8065; Radzihovsky L. and Dorsey A.T. Phys. Rev. Lett. 88 (2002) 216802.
[7]
Lilly M.P., Cooper K.B., Eisenstein J.P., Pfeiffer L.N. and West K.W. Phys. Rev. Lett. 83 (1999) 824.
[8]
Cooper K.B., Lilly M.P., Eisenstein J.P., Pfeiffer L.N. and West K.W. Phys. Rev. B 65 (2002) 241313.
[9]
Brandt E.H. J. Low Temp. Phys. 53 (1983) 41, 71.
[10]
Clem J.R. Phys. Rev. B 43 (1991) 7837.
[11]
Grønbech-Jensen N. Comput. Phys. Commun. 119 (1999) 115.
[12]
Matsuda T. et al. Science 294 (2001) 2136; Reichhardt C. and Reichhardt C.J.O. Phys. Rev. B 66 (2002) 172504.
[13]
Le Doussal P. and Giamarchi T. Physica C 331 (2000) 233.
[14]
Higgins M.J. and Bhattacharya S. Physica C 257 (1996) 232.
[15]
Moon K., Scalettar R.T. and Zimányi G.T. Phys. Rev. Lett. 77 (1998) 2778; Olson C.J., Reichhardt C. and Nori F. Phys. Rev. Lett. 81 (1998) 3757.
[16]
Fisher D.S. Phys. Rev. B 31 (1985) 1396.



File translated from TEX by TTHgold, version 4.00.

Back to Home