kibble zurek mechanism in colloidal monolayers - pnas · kibble–zurek mechanism in colloidal...

6
KibbleZurek mechanism in colloidal monolayers Sven Deutschländer, Patrick Dillmann, Georg Maret, and Peter Keim 1 Department of Physics, University of Konstanz, 78464 Konstanz, Germany Edited by Tom C. Lubensky, University of Pennsylvania, Philadelphia, PA, and approved March 27, 2015 (received for review January 13, 2015) The KibbleZurek mechanism describes the evolution of topolog- ical defect structures like domain walls, strings, and monopoles when a system is driven through a second-order phase transition. The model is used on very different scales like the Higgs field in the early universe or quantum fluids in condensed matter systems. A defect structure naturally arises during cooling if separated re- gions are too far apart to communicate (e.g., about their orienta- tion or phase) due to finite signal velocity. This lack of causality results in separated domains with different (degenerated) locally broken symmetry. Within this picture, we investigate the nonequi- librium dynamics in a condensed matter analog, a 2D ensemble of colloidal particles. In equilibrium, it obeys the so-called KosterlitzThoulessHalperinNelsonYoung (KTHNY) melting scenario with continuous (second order-like) phase transitions. The ensemble is exposed to a set of finite cooling rates covering roughly three orders of magnitude. Along this process, we analyze the defect and domain structure quantitatively via video microscopy and de- termine the scaling of the corresponding length scales as a func- tion of the cooling rate. We indeed observe the scaling predicted by the KibbleZurek mechanism for the KTHNY universality class. KibbleZurek mechanism | nonequilibrium dynamics | spontaneous symmetry breaking | KTHNY theory | colloids I n the formalism of gauge theory with spontaneously broken symmetry, Zeldovich et al. and Kibble postulated a cosmo- logical phase transition during the cooling down of the early universe. This transition leads to degenerated states of vacua below a critical temperature, separated or dispersed by defect structures as domain walls, strings, or monopoles (13). In the course of the transition, the vacuum can be described via an N-component, scalar order parameter ϕ (known as the Higgs field) underlying an effective potential V = aϕ 2 + b ϕ 2 η 2 0 2 , [1] where a is temperature dependent, b is a constant, and η 0 is the modulus of hϕi at T = 0. For high temperatures, V has a single minimum at ϕ = 0 (high symmetry) but develops a minimum landscapeof degenerated vacua below a critical temperature T c (e.g., the so-called sombrero shape for N = 2). Cooling down from the high symmetry phase, the system undergoes a phase transition at T c into an ordered (low symmetry) phase with non- zero hϕi . For T < T c it holds hϕi 2 = η 2 0 1 T 2 T 2 c = η 2 ðTÞ. [2] Caused by thermal fluctuations, one can expect that below T c , hϕi takes different nonzero values in regions that are not con- nected by causality. The question now arising concerns the de- termination of the typical length scale ξ d of these regions and their separation. For a finite cooling rate, ξ d is limited by the speed of propagating information, which is given by the finite speed of light defining an ultimate event horizon. Independent of the nature of the limiting causality, Kibble argued that as long as the difference in free energy ΔF (of a certain system volume) between its high symmetry state hϕi = 0 and a possible finite value of hϕi just below T c is less than k B T, the volume can jump between both phases. The temperature at which ΔF = k B T is called the Ginzburg temperature T G , and the length scale ξ d of the initial (proto)domains is supposed to be equal to the corre- lation length at that temperature: ξ d = ξðT G Þ (2). The geometry of the defect network that separates the un- correlated domains is given by the topology of the manifold of degenerated states that can exist in the low symmetry phase. Thus, it depends strongly on the dimensionality of the system D and on the dimension N of the order parameter itself. Regarding the square root of Eq. 2, the expectation value of a one-com- ponent order parameter (N = 1) can only take two different low symmetry values hϕi ηðTÞ (e.g., the magnetization in a 2D or 3D Ising model): the manifold of the possible states is discon- nected. This topological constraint has a crucial effect if one considers a mesh of symmetry broken domains where hϕi is chosen randomly as either +η or η. If two neighboring (but uncorrelated) domains have the same expectation value of hϕi , they can merge. In contrast, domains with an opposite expecta- tion value will be separated by a domain wall in 3D (or a domain line in 2D). At its center, the domain wall attains a value of hϕi = 0, providing a continuous crossover of the expectation value between the domains (Fig. 1A). Consider now N = 2: hϕi can take any value on a circle, e.g., hϕi 2 = hϕ x i 2 + hϕ y i 2 = η 2 ðTÞ (all of the order parameter values that are lying on the minimum circle of the sombrero are degenerated). Because the manifold of possible low symmetry states is now connected, hϕi can vary smoothly along a path (Fig. 1B). In a network of symmetry broken domains in two dimensions, at least three domains (in Fig. 1B separated by dashed lines) meet at a mutual edge. On a closed path around the edge, the expectation value hϕi might be either constant along the path (for a global, uniform ϕ) but can also vary by a multiple of 2π (in analogy to the winding numbers in liquid crystals). In the first case, the closed path can be re- duced to a point with hϕi 0, and no defect is built. If the path is shrunk in the second case, the field eventually has to attain Significance Spontaneous symmetry breaking describes a variety of trans- formations from high- to low-temperature phases and applies to cosmological concepts as well as atomic systems. T. W. B. Kibble suggested defect structures (domain walls, strings, and mono- poles) to appear during the expansion and cooling of the early universe. The lack of such defects within the visible horizon of the universe mainly motivated inflationary Big Bang theories. W. H. Zurek pointed out that the same principles are relevant within the laboratory when a system obeying a second-order phase transition is cooled at finite rates into the low symmetry phase. Using a colloidal system, we visualize the KibbleZurek mechanism on single particle level and clarify its nature in the background of an established microscopic melting formalism. Author contributions: P.K. designed research; S.D. and P.K. performed research; P.D. contributed new reagents/analytic tools; S.D. analyzed data; and S.D., G.M., and P.K. wrote the paper. The authors declare no conflict of interest. This article is a PNAS Direct Submission. Freely available online through the PNAS open access option. See Commentary on page 6780. 1 To whom correspondence should be addressed. Email: [email protected]. www.pnas.org/cgi/doi/10.1073/pnas.1500763112 PNAS | June 2, 2015 | vol. 112 | no. 22 | 69256930 PHYSICS SEE COMMENTARY

Upload: hoangdung

Post on 19-Aug-2018

225 views

Category:

Documents


1 download

TRANSCRIPT

Page 1: Kibble Zurek mechanism in colloidal monolayers - PNAS · Kibble–Zurek mechanism in colloidal monolayers Sven Deutschländer, Patrick Dillmann, Georg Maret, and Peter Keim1 Department

Kibble–Zurek mechanism in colloidal monolayersSven Deutschländer, Patrick Dillmann, Georg Maret, and Peter Keim1

Department of Physics, University of Konstanz, 78464 Konstanz, Germany

Edited by Tom C. Lubensky, University of Pennsylvania, Philadelphia, PA, and approved March 27, 2015 (received for review January 13, 2015)

The Kibble–Zurek mechanism describes the evolution of topolog-ical defect structures like domain walls, strings, and monopoleswhen a system is driven through a second-order phase transition.The model is used on very different scales like the Higgs field inthe early universe or quantum fluids in condensed matter systems.A defect structure naturally arises during cooling if separated re-gions are too far apart to communicate (e.g., about their orienta-tion or phase) due to finite signal velocity. This lack of causalityresults in separated domains with different (degenerated) locallybroken symmetry. Within this picture, we investigate the nonequi-librium dynamics in a condensed matter analog, a 2D ensemble ofcolloidal particles. In equilibrium, it obeys the so-called Kosterlitz–Thouless–Halperin–Nelson–Young (KTHNY) melting scenario withcontinuous (second order-like) phase transitions. The ensemble isexposed to a set of finite cooling rates covering roughly threeorders of magnitude. Along this process, we analyze the defectand domain structure quantitatively via video microscopy and de-termine the scaling of the corresponding length scales as a func-tion of the cooling rate. We indeed observe the scaling predictedby the Kibble–Zurek mechanism for the KTHNY universality class.

Kibble–Zurek mechanism | nonequilibrium dynamics | spontaneoussymmetry breaking | KTHNY theory | colloids

In the formalism of gauge theory with spontaneously brokensymmetry, Zel’dovich et al. and Kibble postulated a cosmo-

logical phase transition during the cooling down of the earlyuniverse. This transition leads to degenerated states of vacuabelow a critical temperature, separated or dispersed by defectstructures as domain walls, strings, or monopoles (1–3). In thecourse of the transition, the vacuum can be described via anN-component, scalar order parameter ϕ (known as the Higgsfield) underlying an effective potential

V = aϕ2 + b�ϕ2 − η20

�2, [1]

where a is temperature dependent, b is a constant, and η0 is themodulus of hϕi at T = 0. For high temperatures, V has a singleminimum at ϕ= 0 (high symmetry) but develops a minimum“landscape” of degenerated vacua below a critical temperatureTc (e.g., the so-called sombrero shape for N = 2). Cooling downfrom the high symmetry phase, the system undergoes a phasetransition at Tc into an ordered (low symmetry) phase with non-zero hϕi. For T <Tc it holds

hϕi2 = η20�1−T2�T2

c

�= η2ðTÞ. [2]

Caused by thermal fluctuations, one can expect that below Tc,hϕi takes different nonzero values in regions that are not con-nected by causality. The question now arising concerns the de-termination of the typical length scale ξd of these regions andtheir separation. For a finite cooling rate, ξd is limited by thespeed of propagating information, which is given by the finitespeed of light defining an ultimate event horizon. Independent ofthe nature of the limiting causality, Kibble argued that as long asthe difference in free energy ΔF (of a certain system volume)between its high symmetry state hϕi= 0 and a possible finitevalue of hϕi just below Tc is less than kBT, the volume can jumpbetween both phases. The temperature at which ΔF = kBT is

called the Ginzburg temperature TG, and the length scale ξd ofthe initial (proto)domains is supposed to be equal to the corre-lation length at that temperature: ξd = ξðTGÞ (2).The geometry of the defect network that separates the un-

correlated domains is given by the topology of the manifold ofdegenerated states that can exist in the low symmetry phase.Thus, it depends strongly on the dimensionality of the system Dand on the dimension N of the order parameter itself. Regardingthe square root of Eq. 2, the expectation value of a one-com-ponent order parameter (N = 1) can only take two different lowsymmetry values hϕi=±ηðTÞ (e.g., the magnetization in a 2D or3D Ising model): the manifold of the possible states is discon-nected. This topological constraint has a crucial effect if oneconsiders a mesh of symmetry broken domains where hϕi ischosen randomly as either +η or −η. If two neighboring (butuncorrelated) domains have the same expectation value of hϕi,they can merge. In contrast, domains with an opposite expecta-tion value will be separated by a domain wall in 3D (or a domainline in 2D). At its center, the domain wall attains a value ofhϕi= 0, providing a continuous crossover of the expectationvalue between the domains (Fig. 1A). Consider now N = 2: hϕican take any value on a circle, e.g., hϕi2 = hϕxi2 + hϕyi2 = η2ðTÞ(all of the order parameter values that are lying on the minimumcircle of the sombrero are degenerated). Because the manifoldof possible low symmetry states is now connected, hϕi can varysmoothly along a path (Fig. 1B). In a network of symmetrybroken domains in two dimensions, at least three domains (inFig. 1B separated by dashed lines) meet at a mutual edge. On aclosed path around the edge, the expectation value hϕimight beeither constant along the path (for a global, uniform ϕ) but canalso vary by a multiple of 2π (in analogy to the winding numbersin liquid crystals). In the first case, the closed path can be re-duced to a point with hϕi≠ 0, and no defect is built. If the path isshrunk in the second case, the field eventually has to attain

Significance

Spontaneous symmetry breaking describes a variety of trans-formations from high- to low-temperature phases and applies tocosmological concepts as well as atomic systems. T. W. B. Kibblesuggested defect structures (domain walls, strings, and mono-poles) to appear during the expansion and cooling of the earlyuniverse. The lack of such defects within the visible horizon ofthe universe mainly motivated inflationary Big Bang theories.W. H. Zurek pointed out that the same principles are relevantwithin the laboratory when a system obeying a second-orderphase transition is cooled at finite rates into the low symmetryphase. Using a colloidal system, we visualize the Kibble–Zurekmechanism on single particle level and clarify its nature in thebackground of an established microscopic melting formalism.

Author contributions: P.K. designed research; S.D. and P.K. performed research; P.D.contributed new reagents/analytic tools; S.D. analyzed data; and S.D., G.M., and P.K.wrote the paper.

The authors declare no conflict of interest.

This article is a PNAS Direct Submission.

Freely available online through the PNAS open access option.

See Commentary on page 6780.1To whom correspondence should be addressed. Email: [email protected].

www.pnas.org/cgi/doi/10.1073/pnas.1500763112 PNAS | June 2, 2015 | vol. 112 | no. 22 | 6925–6930

PHYS

ICS

SEECO

MMEN

TARY

Page 2: Kibble Zurek mechanism in colloidal monolayers - PNAS · Kibble–Zurek mechanism in colloidal monolayers Sven Deutschländer, Patrick Dillmann, Georg Maret, and Peter Keim1 Department

hϕi= 0 within the path, and one remains with a monopole forD= 2 or a string for D= 3 (2, 3). A condensed matter analogwould be a vortex of normal fluid with quantized circulation insuperfluid helium. For N = 3 and D= 3, four domains can meet ata mutual point, and the degenerated solutions of the low tem-perature phase lie on a sphere: hϕi2 = hϕxi2 + hϕyi2 + hϕzi2. If thefield now again varies circularly on a spherical path (all fieldarrows point radially outward), a shrinking of this sphere leads toa monopole in three dimensions (2, 3).Zurek extended Kibble’s predictions and transferred his con-

siderations to quantum condensed matter systems. He suggestedthat 4He should intrinsically develop a defect structure whenquenched from the normal to the superfluid phase (4, 5). Forsuperfluid 4He, the order parameter ψ = jψ jexpðiΘÞ is complexwith two independent components: magnitude jψ j and phase Θ(the superfluid density is given by jψ j2). A nontrivial, static so-lution of the equation of state with a Ginzburg–Landau potentialyields ψ =ψ0ðrÞexpðinφÞ, where r and φ are cylindrical co-ordinates, n∈Z, and ψ0ð0Þ= 0. This solution is called a vortexline, which is topologically equivalent to a string for the caseN = 2 we discussed before. In the vicinity of the critical tem-perature during a quench from the normal fluid to the superfluidstate, ψ will be chosen randomly in uncorrelated regions, leadingto a string network of normal fluid vortices. In condensed mattersystems, the role of the limiting speed of light is taken by thesound velocity (in 4He, the second sound). This upper boundaryleads to a finite speed of the propagation of order parameterfluctuations and sets a “sonic horizon.”Zurek argued that the correlation length is frozen-out close to

the transition point or even far before depending on the coolingrate (4, 5). Consider the divergence of the correlation length ξ fora second-order transition, e.g., ξ= ξ0jej−ν, where e= ðT −TcÞ=Tcis the reduced temperature. If the cooling is infinitely slow, thesystem behaves as in equilibrium: ξ will diverge close to thetransition and the system is a monodomain. For an instantaneousquench, the system has minimal time to adapt to its surrounding:ξ will be frozen-out at the beginning of the quench. For second-order phase transitions, the divergence of correlation lengths isaccompanied by the divergence of the correlation time τ= τ0jej−μ,which is due to the critical slowing down of order parameterfluctuations. If the time t it takes to reach Tc for a given coolingrate is larger than the correlation time, the system stays in equi-librium and the dynamic is adiabatic. Nonetheless, for every finitebut nonzero cooling rate, t eventually becomes smaller than τ, andthe system falls out of equilibrium before Tc is reached. Themoment when both the correlation time and the time it takes toreach Tc coincide defines the freeze-out time.

t= τðtÞ. [3]

The frozen-out correlation length ξ is then set at the temperaturee of the corresponding freeze-out time: ξ= ξðeÞ= ξðtÞ. For a lin-ear temperature quench

e= ðT −TcÞ=Tc = t�τq, [4]

with the quench time scale τq, one observes t= ðτ0τμqÞ1=ð1+μÞ and

ξ= ξðtÞ= ξ0�τq�τ0�ν=ð1+μÞ. [5]

For the Ginzburg–Landau model (ν= 1=2, μ= 1), one finds thescaling ξ∼ τ1=4q , whereas a renormalization group correction(ν= 2=3) leads to ξ∼ τ1=3q (4, 5).A frequently used approximation is that when the adiabatic

regime ends at t before the transition, the correlation lengthcannot follow the critical behavior until τ again exceeds the timet when Tc is passed. Given a symmetric divergence of τ aroundTc, this is the time t after the transition. The period in between isknown as the impulse regime, in which the correlation length isassumed not to evolve further. A recent analytical investigation,however, suggests that in this period, the system falls into a re-gime of critical coarse graining (6). There, the typical lengthscale of correlated domains continues to grow because localfluctuations are still allowed, and the system is out of equilib-rium. On the other side, numerical studies in which dissipativecontributions and cooling rates were alternatively varied beforeand after the transition indicate that the final length scale of thedefect and domain network is entirely determined after thetransition (7). Several efforts have been made to provide ex-perimental verification of the Kibble–Zurek mechanism in avariety of systems, e.g., in liquid crystals (8) (the transition isweakly first order but the defect network can easily observedwith cross polarization microscopy), superfluid 3He (9), super-conducting systems (10), convective, intrinsically out of equilib-rium systems (11), multiferroics (12), quantum systems (13), ioncrystals (14, 15), and Bose–Einstein condensates (16) (the lattertwo systems contain the effect of inhomogeneities due to, forexample, temperature gradients). A detailed review concerningthe significance and limitations of these experiments can befound in ref. 17.In this experimental study, we test the validity and applicability

of the Kibble–Zurek mechanism in a 2D colloidal model systemwhose equilibrium thermodynamics follow the microscopicallymotivated Kosterlitz–Thouless–Halperin–Nelson–Young (KTHNY)theory. This theory predicts a continuous, two-step melting be-havior whose dynamics, however, are quantitatively differentfrom phenomenological second-order phase transitions describedby the Ginzburg–Landau model. We applied cooling rates overroughly three orders of magnitude, for which we changed thecontrol parameter with high resolution and homogeneouslythroughout the sample without temperature gradients. Singleparticle resolution provides a quantitative determination of de-fect and domain structures during the entire quench procedure,and the precise knowledge of the equilibrium dynamics allowsdetermination of the scaling behavior of corresponding lengthscales at the freeze-out times. In the following, we validate thatthe Kibble–Zurek mechanism can be successfully applied to theKTHNY universality class.

Defects and Symmetry Breaking in 2D CrystallizationThe closed packed crystalline structure in two dimensions is ahexagonal crystal with sixfold symmetry. The thermodynamics ofsuch a crystal can analytically be described via the KTHNYtheory, a microscopic, two-step melting scenario (including twocontinuous transitions) that is based on elasticity theory and a

<φ> = 0

<φ> = -η

<φx>2+ <φy>2 = η2

path

<φ>

= 0

<φ> = +η

‘path’

A B

Fig. 1. Emergence of defects in the Higgs field that is illustrated with redvectors (shown in 2D for simplicity). (A) For N= 1 and D= 3, domain walls canappear (strings for D= 2). (B) For N= 2 and D= 3, nontrivial topologies arestrings (monopoles for D= 2). The defects are regions where the order pa-rameter ϕ retains the high symmetry phase (hϕi= 0) to moderate betweendifferent degenerated orientations of the symmetry broken field.

6926 | www.pnas.org/cgi/doi/10.1073/pnas.1500763112 Deutschländer et al.

Page 3: Kibble Zurek mechanism in colloidal monolayers - PNAS · Kibble–Zurek mechanism in colloidal monolayers Sven Deutschländer, Patrick Dillmann, Georg Maret, and Peter Keim1 Department

renormalization group analysis of topological defects (18–20). Inthe KTHNY formalism, the orientationally long range-orderedcrystalline phase melts at a temperature Tm via the dissociationof pairs of dislocations into a hexatic fluid, which is unknown in3D systems. This fluid is characterized by quasi-long-range ori-entational but short-range translational order. In a triangularlattice, dislocations are point defects and consist of two neigh-boring particles with five and seven nearest neighbors, re-spectively, surrounded by sixfold coordinated particles. At ahigher temperature Ti, dislocations start to unbind further intoisolated disclinations (a disclination is a particle with five orseven nearest neighbors surrounded by sixfold coordinated par-ticles), and the system enters an isotropic fluid with short-rangeorientational and translational order. A suitable orientationalorder parameter is the local bond order field ψ6ð~rj, tÞ=n−1j

Pke

i6θjkðtÞ =��ψ6ð~rj, tÞ

��eiΘjðtÞ, which is a complex number withmagnitude

��ψ6ð~rj, tÞ�� and phase ΘjðtÞ defined at the discrete par-

ticle positions~rj. ΘjðtÞ is the average bond orientation for a spe-cific particle. The k-sum runs over all nj nearest neighbors ofparticle j, and θjk is the angle of the kth bond with respect to acertain reference axis. If particle j is perfectly sixfold coordinated[e.g., all θjkðtÞ equal an ascending multiple of π=3], the local bondorder parameter attains

��ψ6ð~rj, tÞ��= 1. A five- or sevenfold co-

ordinated particle yields��ψ6ð~rj, tÞ

��J 0. The three differentphases can be distinguished via the spatial correlation g6ðrÞ=hψp

6ð~0Þψ6ð rj!Þi or temporal correlation g6ðtÞ= hψp6ð0Þψ6ðtÞi of the

local bond order parameter. For large r and t, respectively, eachcorrelation attains a finite value in the (mono)crystalline phase, de-cays algebraically in the hexatic fluid, and exponentially∼ expð−r=ξ6Þand ∼ expð−t=τ6Þ in the isotropic fluid (20, 21). Unlike second-order phase transitions where correlations typically diverge alge-braically, the orientational correlation length ξ6 and time τ6 di-verge in the KTHNY formalism exponentially at Ti

ξ6 ∼ exp�ajej−1=2� and τ6 ∼ exp

�bjej−1=2�, [6]

where e= ðT −TiÞ=Ti, and a and b are constants (20, 22). Thispeculiarity is the reason why KTHNY melting is named contin-uous instead of second order. In equilibrium, the KTHNY sce-nario has been verified successfully for our colloidal system invarious experimental studies (23–25).To transfer this structural 2D phase behavior into the frame-

work of the Kibble–Zurek mechanism, we start in the hightemperature phase (isotropic fluid) and describe the symmetrybreaking with the spatial distribution of the bond order parameter.Because in 2D the local symmetry is sixfold in the crystal and thefluid, the isotropic phase is a mixture of sixfold and equally num-bered five- and sevenfold particles (other coordination numbersare extremely rare and can be neglected). During cooling, isolateddisclinations combine to dislocations that, for infinite slow coolingrates, can annihilate into sixfold particles with a uniform directorfield. This uniformity is given by a global phase, characterizing the

orientation of the crystal axis. Spontaneous symmetry breakingimplies that all possible global crystal orientations are degenerated,and the Kibble–Zurek mechanism predicts that in the presence ofcritical fluctuations the system cannot gain a global phase at finitecooling rates: Locally, symmetry broken domains will emerge,which will have different orientations in causally separated regions.The final state is a polycrystalline network with frozen-in defects.As in the case of superfluid 4He, ψ6ð~rj, tÞ is complex with twoindependent components (N = 2). Consequently, we expect toobserve monopoles in two dimensions. The phase of ψ6ð~rj, tÞ isinvariant under a change in the particular bond angles ofΔθjkðtÞ=±nπ=3 (n∈N), which is caused by the sixfold orientationof the triangular lattice. Similar to the Higgs field or the superfluid,one cannot consider a closed (discrete) path in ψ6ð~r, tÞ on whichθjkðtÞ changes by an amount of ±π=3, leaving the orientational fieldinvariant. Reducing this path to a point, ψ6ð~r, tÞ must tend towardzero at the center to maintain continuity. Because the orientationalfield is defined at discrete positions, the defect is a single particlemarked as a monopole of the high symmetry phase. In fact, thiscoincides with the definition of disclinations in the KTHNY for-malism (20): The particle at the center is an isolated five- or sev-enfold coordinated site. Fig. 2 illustrates this for a bond on a closedpath. Going counterclockwise around the defect, the bond anglechanges by an amount of +π=3 for a fivefold (Fig. 2A) and by −π=3for a sevenfold site (Fig. 2B). [In principle, also larger changes inθjkðtÞ are possible, e.g., for n= 2, a four- or eightfold oriented site,but these are extremely rare.] In KTHNY theory the monopoles(disclinations) combine to dipoles (dislocations) that can only an-nihilate completely if their orientation is exactly antiparallel. Atfinite cooling rates, they arrange in chains, separating symmetrybroken domains of different orientation: chains of dislocations canbe regarded as strings or 2D domain walls.

Colloidal Monolayer and Cooling ProcedureOur colloidal model system consists of polystyrene beads withdiameter σ = 4.5 μm, dispersed in water and sterically stabilizedwith the soap SDS. The beads are doped with iron oxide nano-particles that result in a superparamagnetic behavior and a massdensity of 1.7 kg/dm3. The colloidal suspension is sealed within amillimeter-sized glass cell where sedimentation leads to the for-mation of a monolayer of beads on the bottom glass plate. Thewhole layer consists of >105 particles and in a 1,158 μm × 865 μmsubwindow, ≈ 5,700 particles are tracked with a spatial resolutionof submicrometers and a time resolution in the order of seconds.The system is kept at room temperature and exempt from densitygradients due to a months-long precise control of the horizontalinclination down to microradian. The potential energy can be

+π/3

-π/3

1

2

3

2 5

4 1

3

14

13 12

19

18 17

4

16

A B

Fig. 2. Sketch of a fivefold oriented (A) and sevenfold oriented (B) dis-clination. The red arrows illustrate the change in bond angle (blue) whencircling on an anticlockwise path around the defect.

30 35 40 45 50 55 60 650

1000

2000

3000

4000

5000

6000 Γ [1/s]0.0000420.000370.00110.00230.00600.0110.0200.033

t

Γ ( 1/Τ)0

1000

2000

3000

4000

5000

6000τ6fit

6

Fig. 3. Orientational correlation time τ6 (experimental data and fit accordingto Eq. 6) and the time t left until the transition temperature is reached fordifferent cooling rates (colored straight lines, Eq. 9) as a function of inversetemperature Γ (small Γ correspond to large temperatures and vice versa). Theintersections define the freeze-out interaction values Γ=ΓðtÞ.

Deutschländer et al. PNAS | June 2, 2015 | vol. 112 | no. 22 | 6927

PHYS

ICS

SEECO

MMEN

TARY

Page 4: Kibble Zurek mechanism in colloidal monolayers - PNAS · Kibble–Zurek mechanism in colloidal monolayers Sven Deutschländer, Patrick Dillmann, Georg Maret, and Peter Keim1 Department

tuned by an external magnetic field H applied perpendicular tothe monolayer, which induces a repulsive dipole–dipole inter-action between the particles. The ratio between potential energyEmag and thermal energy kBT

Γ=Emag

kBT=μ0ðπnÞ3=2ð χHÞ2

4πkBT, [7]

acts as inverse temperature (or dimensional pressure for fixedvolume and particle number). n= 1=a20 is the 2D particle den-sity with a mean particle distance a0 ≈ 13 μm, and χ = 1.9×10−11  Am2=T is the magnetic susceptibility of the beads. Γ is thethermodynamic control parameter: A small magnetic field corre-sponds to a large temperature and vice versa. Measured valuesof the equilibrium melting temperatures are Γm ≈ 70.3 for thecrystal/hexatic transition and Γi ≈ 67.3 for the hexatic/isotropictransition (25). The cooling procedure is as follows: we equili-brate the system deep in the isotropic liquid at Γ0 ≈ 25 and applylinear cooling rates _Γ=ΔΓ=Δt deep into the crystalline phase up toΓend ≈ 100; thenceforward, we let the system equilibrate. We per-form different rates, ranging over almost three decades from_Γ= 0.000042 1=s up to _Γ= 0.0326 1=s. The slowest cooling ratecorresponds to a quench time of ≈ 19 days and the fastest one to≈ 40min. We would like to emphasize that with the given controlparameter there is no heat transport from the surface as in 3Dbulk material. The lack of gradients rules out a temperaturegradient-assisted annealing of defects that might be present ininhomogeneous systems.

Structure and Dynamics of Defects and DomainsThe key element of the Kibble–Zurek mechanism is a frozen-outcorrelation length ξ as the system falls out of equilibrium at thefreeze-out time t. For slow cooling rates, the system can followadiabatically closer to the transition (large ξ) than for fast rateswhere the systems reaches the freeze-out time earlier (small ξ).To find t, we determine the orientational correlation time τ6according the KTHNY theory by fitting g6ðtÞ∼ expð−t=τ6Þ in the

isotropic fluid for independent equilibrium measurements. Themeasured values for the correlation time τ6 as well as a fit with

τ6 = τ0 exp�bτj1=Γ− 1=Γcj−1=2

�, [8]

is shown in Fig. 3 (right ordinate). The time left to the isotropic–hexatic transition is given by

t= ðΓi −ΓÞ� _Γ, [9]

and is also plotted in Fig. 3 (left ordinate) for various cooling ratesincluding the slowest and the fastest one. The points of intersection

t= τ0 exphbτ��1�Γ�t, _Γ�− 1=Γc

��−1=2i, [10]

define the freeze-out temperatures Γ=ΓðtÞ.The length scale of the defect network can be measured by the

overall defect concentration ρ (counting all particles being notsixfold, normalized by the total number of particles) in theψ6ð~r, tÞ field. Fig. 4 (Upper) shows the evolution of ρ for the samecooling rates _Γ as in Fig. 3, as well as for the equilibrium(melting) behavior (25). One recognizes that the course of ρdeviates from the equilibrium case in advance of the isotropic/hexatic transition at Γi ≈ 67.3. This deviation appears at differenttimes for distinct cooling rates and marks the end of the adia-batic regime. Within the noise, deviations from the equilibriumbehavior start at the temperature Γ given by the freeze-out timet (big colored dots). Beyond the adiabatic regime, the defectdensity decreases, which is an indication of critical coarse grain-ing as predicted in ref. 6. At Γi (and also Γm), the slope of thecurves increases with decreasing cooling rate, indicating a furtherevolution, but the system cannot perform critical fluctuations.

0.0

0.1

0.2

0.3

0.4 Γm equilibriumΓi

Γ [1/s]0.0330.0200.0110.00600.00230.00110.000370.000042

ρ

30 40 50 60 70 80 902

4

6

8

10

12

<A>/a2 0

Fig. 4. Defect number density ρ and average domain size hAi (in units of a20)as a function of Γ (∝ 1=T) during cooling from small Γ (=hot on the left side)to large Γ (=cold on the right side). The curves cover the complete range ofcooling rates from _Γ= 0.000042 to _Γ= 0.033 and are averaged within aninterval ΔΓ= 0.4. Big dots mark the freeze-out temperatures Γ=ΓðtÞ (coloredcorrespondingly to _Γ). Open symbols show the equilibrium melting behavior(lines are a guide to the eye).

A B

C D

Fig. 5. Snapshot sections of the colloidal ensemble (992× 960 μm2, ≈ 4,000particles) illustrating the defect (A and C) and domain configurations (B andD) at the freeze out temperature Γ for the fastest (A and B: _Γ=0.0326 1=s,Γ≈ 30.3) and slowest cooling rate (C and D: _Γ= 0.000042 1=s, Γ≈ 66.8). Thedefects are marked as follows. Particles with the five nearest neighbors arered, seven nearest neighbors are green, and other defects are blue. Sixfoldcoordinated particles are gray. Different symmetry broken domains are col-ored individually, and high symmetry particles are displayed by smaller circles.

6928 | www.pnas.org/cgi/doi/10.1073/pnas.1500763112 Deutschländer et al.

Page 5: Kibble Zurek mechanism in colloidal monolayers - PNAS · Kibble–Zurek mechanism in colloidal monolayers Sven Deutschländer, Patrick Dillmann, Georg Maret, and Peter Keim1 Department

The domain structure, on the other hand, can be characterizedquantitatively by analyzing symmetry broken domains with asimilar phase of ψ6ð~rj, tÞ. According to ref. 26, we define a particleto be part of a symmetry broken domain if the following threeconditions are fulfilled for the particle itself and at least onenearest neighbor: (i) the magnitude

��ψ6ð~rj, tÞ�� of the local bond

order parameter must exceed 0.6 for both neighboring particles;(ii) the bond length deviation of neighboring particles is less than10% of the average particle distance a0; and (iii) the variation inthe average bond orientation ΔΘijðtÞ=

��Im½ψ6ð~riÞ�− Im½ψ6ð~rjÞ��� of

neighboring particles i and j must be less than 14° (less than 14°=6in real space). Simply connected domains of particles that fulfillall three criteria are merged to a local symmetry broken domain.If a particle does not satisfy these conditions in respect to aneighboring particle, it is assigned to the high symmetry phase[almost all defects are identified as such due to their small valueof

��ψ6ð~rj, tÞ��]. Fig. 4 (Lower) shows the evolution of the ensemble

average domain size hAi as a function of Γ. We observe a behavioranalog to ρ: domain formation significantly deviates from theequilibrium case before Γi, namely around the freeze-out tem-perature Γ of the corresponding cooling rates _Γ. To compare bothnetworks in the following, we define the dimensionless lengthsξdef = ρ−1=2 and ξdom = ðhAi=a20Þ1=2, which display the characteris-tic length scales in units of a0.Colloidal ensembles offer the unique possibility to monitor the

system and its domain and defect structure on single particlelevel. Fig. 5 illustrates both (Left column for defects and rightcolumn for domains) at the freeze-out temperature Γ for thefastest (Fig. 5 A and B) and the slowest (Fig. 5 C and D) coolingrate. For _Γ= 0.0326 1=s (Fig. 5 A and B) where t is alreadyreached at Γ= 30.3, the defect density is large, as is the numberof high symmetry particles. However, there is a significantnumber of sixfold coordinated particles and a few orientationallyordered domains (to accord for finite size effects, we will excludedomains that hit the border of the field of view when evaluatingξdom at Γ). At this point, the length scales are ξdef = 1.56± 0.01

and ξdom = 1.56± 0.03. For the slowest cooling rate _Γ=0.000042 1=s (Fig. 5 C and D) where Γ= 66.7, the mean dis-tance between defects, as well as the typical domain size, is signif-icantly larger compared with the fastest cooling rate. We observeξdef = 2.36± 0.07 and ξdom = 2.30± 0.09.To allow relaxation of the defect and domain structure after the

freeze-out time (6), we keep the temperature constant afterΓend ≈ 100 is reached. Fig. 6 shows the defect and domain config-urations after an equilibration time of ≈ 5 h for the fastest coolingrate (Fig. 6 A and B) where the quench time was ≈ 40 min andafter an equilibration time of ≈ 3 days for the slowest rate (Fig. 6 Cand D) where the quench time was ≈ 19 days. The different lengthscale of the defect and symmetry broken domain network in re-spect to the cooling rate is clearly visible: although we observe alarge number of domains for fast cooling, slow cooling results inmerely two large domains separated by a single grain boundary.The final evolution will be given by classical coarse graining. Theground state is known to be a monodomain but its observation liesbeyond experimental accessible times for our system.

Scaling BehaviorThe main prediction of the Kibble–Zurek mechanism is a powerlaw dependence of the frozen-out correlation length ξ as afunction of τq (Eq. 5), which results from the algebraic di-vergence of the correlation, presuming Eqs. 3 and 4. In KTHNYmelting, ξ6 and τ6 diverge exponentially, and one has to solve Eq.10 to find the implicit dependency tð _ΓÞ. We did this numericallyfor discrete values in the range of 4× 10−5 ≤ _Γ≤ 4× 10−2 anddetermined the frozen-out orientational correlation length ξ6 fora scaling τ6=τB = c  ξz6 with the dynamical exponent z (22). Here,τB ≈ 171.6 s is the Brownian time that is the time a single particleneeds to diffuse its own diameter. Using Eq. 8, one finds withΓðt, _ΓÞ from Eq. 9 the expression

ξ6�_Γ�=�τ0cτB

1=zexp

"bτz

�����Γc −Γi + _Γt�_Γ�

ΓcΓi −Γc _Γt�_Γ������−1=2#

. [11]

This function is plotted in Fig. 7 for z= 4.5 and c= 0.83 (redcurves) on a double logarithmic scale together with ξdef and

A B

C D

Fig. 6. Snapshot sections of the colloidal ensemble illustrating the defect(A and C) and domain configurations (B and D) after quasi-equilibration of thesystem for the fastest (A and B: _Γ= 0.0326  1=s, Γend ≈ 105) and slowest coolingrate (C and D: _Γ= 0.000042  1=s, Γend ≈ 98). The system size and the labeling ofdefects and domains are the same as in Fig. 5.

10-4 10-3 10-2

1.6

2.0

2.4

Γ

ξ dom

~ Γ -0.061(1)

1.6

2.0

2.4 •

~ Γ -0.061(1)

ξ def

Fig. 7. Length scale of the defect ξdef and domain network ξdom is plotted asa function of the cooling rate _Γ (open symbols). Red lines are numericalsolutions of the transcendental equation following the freeze-out conditionfor the KTHNY-like divergences (see text for definition). For comparison,dashed blue lines are power law fits predicted by the standard Kibble–Zurekmechanism that show the same algebraic exponent κ ≈0.06 for ξdef and ξdom.

Deutschländer et al. PNAS | June 2, 2015 | vol. 112 | no. 22 | 6929

PHYS

ICS

SEECO

MMEN

TARY

Page 6: Kibble Zurek mechanism in colloidal monolayers - PNAS · Kibble–Zurek mechanism in colloidal monolayers Sven Deutschländer, Patrick Dillmann, Georg Maret, and Peter Keim1 Department

ξdom at the freeze-out temperature Γ. We find very good agree-ment. Nonetheless, we fit for comparison the data via an alge-braic scaling (blue dotted lines) of the form f ð _ΓÞ= a _Γ−κ

, forwhich we observe κdef = 0.061± 0.001 and κdom = 0.061± 0.002.The data are compatible with the algebraic decay only for in-termediate cooling rates. The deviations from the standard Kib-ble–Zurek mechanism for systems with second-order transitionsare in line with the temperature-quenched 2D XY model (27)also having nonalgebraic divergences of the correlation length inequilibrium and thus being in a similar universality class. Thesmall algebraic exponent κ can be explained by the relativelylarge value of the dynamical exponent z, which regulates theslope of ξ6ð _ΓÞ (in ref. 22, a value z= 2.5 was proposed for thehard-disk system). The large value of z is due to quite longcorrelation times in this colloidal system (Fig. 3), which are causedby its overdamped dynamics. Note that the sonic horizon is set bythe sound velocity of the colloidal monolayer (and not the solvent)being approximately millimeters per second, which is six orders ofmagnitude slower compared with atomic systems.

ConclusionsWe presented a colloidal model system, where structure formationin spontaneously symmetry broken systems can be investigated with

single particle resolution. The theoretical framework is given by theKibble–Zurek mechanism, which describes domain formation ondifferent scales like the Higgs field shortly after the Big Bang or thevortex network in 4He quenched into the superfluid state. Alongvarious cooling rates, we analyzed the development of defects andsymmetry broken domains when the systems fall out of equilibriumand fluctuations of the order parameter cannot follow adiabaticallydue to critical slowing down. Although 2D melting in the colloidalmonolayer is described by KTHNY theory where the divergence ofthe relevant correlation lengths in equilibrium is exponential(rather than algebraic as typically found in 3D systems), the centralidea of the Kibble–Zurek mechanism still holds, and the scaling ofthe observed domain network is correctly described. Implicitly, thisshows that existence of grain boundaries cannot solely be used ascriterion for first-order phase transitions and nucleation or tofalsify second/continuous-order transitions because they natu-rally arise for nonzero cooling rates. Those will always be presenton finite time scales in experiments and computer simulationsafter preparation of the system.

ACKNOWLEDGMENTS. P.K. thanks Sébastien Balibar for fruitful discussion. P.K.received financial support from German Research Foundation Grant KE 1168/8-1.

1. Zel’dovich YaB, Kobzarev IYu, Okun’ LB (1975) Cosmological consequences of aspontaneous breakdown of a discrete symmetry. Sov Phys JETP 40(1):1–5.

2. Kibble TWB (1976) Topology of cosmic domains and strings. J Phys Math Gen 9(8):1387–1398.

3. Kibble TWB (1980) Some implications of a cosmological phase transition. Phys Rep67(1):183–199.

4. Zurek WH (1985) Cosmological experiments in superfluid helium? Nature 317:505–508.

5. Zurek WH (1993) Cosmic strings in laboratory superfluids and the topological rem-nants of other phase transitions. Acta Phys Pol B 24(7):1301–1311.

6. Biroli G, Cugliandolo LF, Sicilia A (2010) Kibble-Zurek mechanism and infinitely slowannealing through critical points. Phys Rev E Stat Nonlin Soft Matter Phys 81(5 Pt 1):050101.

7. Antunes ND, Gandra P, Rivers RJ (2006) Is domain formation decided before or afterthe transition? Phys Rev D Part Fields Gravit Cosmol 73(12):125003.

8. Chuang I, Durrer R, Turok N, Yurke B (1991) Cosmology in the laboratory: Defectdynamics in liquid crystals. Science 251(4999):1336–1342.

9. Bäuerle C, Bunkov YuM, Fischer SN, Godfrin H, Pickett GR (1996) Laboratory simula-tion of cosmic string formation in the early universe using superfluid 3He. Nature 382:332–334.

10. Carmi R, Polturak E, Koren G (2000) Observation of spontaneous flux generation in amulti-josephson-junction loop. Phys Rev Lett 84(21):4966–4969.

11. Miranda MA, Burguete J, Mancini H, Gonzles-Vinas W (2013) Frozen dynamics andsynchronization through a secondary symmetry-breaking bifurcation. Phys Rev E StatNonlin Soft Matter Phys 87(3):032902.

12. Chae SC, et al. (2012) Direct observation of the proliferation of ferroelectric loopdomains and vortex-antivortex pairs. Phys Rev Lett 108(16):167603.

13. Xu X-Y, et al. (2014) Quantum simulation of Landau-Zener model dynamics sup-porting the Kibble-Zurek mechanism. Phys Rev Lett 112(3):035701.

14. Ulm S, et al. (2013) Observation of the Kibble-Zurek scaling law for defect formationin ion crystals. Nat Commun 4:2290.

15. Pyka K, et al. (2013) Topological defect formation and spontaneous symmetrybreaking in ion Coulomb crystals. Nat Commun 4:2291.

16. Lamporesi G, Donadello S, Serafini S, Dalfovo F, Ferrari G (2013) Spontaneous creationof Kibble-Zurek solitons in a Bose-Einstein condensate. Nat Phys 9:656.

17. del Campo A, Zurek WH (2014) Universality of phase transition dynamics: Topologicaldefects from symmetry breaking. Int J Mod Phys A 29(8):1430018.

18. Kosterlitz JM, Thouless DJ (1973) Ordering, metastability and phase transitions intwo-dimensional systems. J Phys C: Solid State Phys 6:1181–1203.

19. Young AP (1979) Melting and the vector Coulomb gas in two dimensions. Phys Rev B19(4):1855–1866.

20. Nelson DR, Halperin BI (1979) Dislocation-mediated melting in two dimensions. PhysRev B 19(5):2457–2484.

21. Nelson DR (1983) Phase Transition and Critical Phenomena, eds Domb C, Lebowitz JL(Academic Press, London), pp 2–99.

22. Watanabe H, Yukawa S, Ozeki Y, Ito N (2004) Critical exponents of isotropic-hexaticphase transition in the hard-disk system. Phys Rev E Stat Nonlin Soft Matter Phys 69(4Pt 2):045103.

23. Zahn K, Lenke R, Maret G (1999) Two-stage melting of paramagnetic colloidal crystalsin two dimensions. Phys Rev Lett 82(13):2721–2724.

24. Keim P, Maret G, von Grünberg HH (2007) Frank’s constant in the hexatic phase. PhysRev E Stat Nonlin Soft Matter Phys 75(3 Pt 1):031402.

25. Deutschländer S, Puertas AM, Maret G, Keim P (2014) Specific heat in two-dimensional melting. Phys Rev Lett 113(12):127801.

26. Dillmann P, Maret G, Keim P (2013) Two-dimensional colloidal systems in time-dependent magnetic fields. Eur Phys J Spec Top 222:2941–2959.

27. Jelic A, Cugliandolo LF (2011) Quench dynamics of the 2d XY model. J Stat Mech2011(2):1–26.

6930 | www.pnas.org/cgi/doi/10.1073/pnas.1500763112 Deutschländer et al.