Вы находитесь на странице: 1из 13

Shape and symmetry determine two-dimensional melting transitions of hard regular

polygons
Joshua A. Anderson,1 James Antonaglia,2 Jaime A. Millan,3 Michael Engel,1, 4 and Sharon C. Glotzer1, 2, 3, 5,
1
Department of Chemical Engineering, University of Michigan, Ann Arbor, MI 48109, USA
2
Department of Physics, University of Michigan, Ann Arbor, MI 48109, USA
3
Department of Materials Science and Engineering,
University of Michigan, Ann Arbor, MI 48109, USA
4
Institute for Multiscale Simulation, Friedrich-Alexander-Universitat Erlangen-N
urnberg, 91058 Erlangen, Germany
5
Biointerfaces Institute, University of Michigan, Ann Arbor, MI 48109, USA
The melting transition of two-dimensional (2D) systems is a fundamental problem in condensed
matter and statistical physics that has advanced significantly through the application of computa-
arXiv:1606.00687v2 [cond-mat.soft] 23 Dec 2016

tional resources and algorithms. 2D systems present the opportunity for novel phases and phase
transition scenarios not observed in 3D systems, but these phases depend sensitively on the system
and thus predicting how any given 2D system will behave remains a challenge. Here we report a
comprehensive simulation study of the phase behavior near the melting transition of all hard reg-
ular polygons with 3 n 14 vertices using massively parallel Monte Carlo simulations of up to
one million particles. By investigating this family of shapes, we show that the melting transition
depends upon both particle shape and symmetry considerations, which together can predict which
of three different melting scenarios will occur for a given n. We show that systems of polygons with
as few as seven edges behave like hard disks; they melt continuously from a solid to a hexatic fluid
and then undergo a first-order transition from the hexatic phase to the fluid phase. We show that
this behavior, which holds for all 7 n 14, arises from weak entropic forces among the particles.
Strong directional entropic forces align polygons with fewer than seven edges and impose local order
in the fluid. These forces can enhance or suppress the discontinuous character of the transition
depending on whether the local order in the fluid is compatible with the local order in the solid. As
a result, systems of triangles, squares, and hexagons exhibit a KTHNY-type continuous transition
between fluid and hexatic, tetratic, and hexatic phases, respectively, and a continuous transition
from the appropriate x-atic to the solid. In particular, we find that systems of hexagons display
KTHNY melting. In contrast, due to symmetry incompatibility between the ordered fluid and solid,
systems of pentagons and plane-filling 4-fold pentilles display a one-step first-order melting of the
solid to the fluid with no intermediate phase.

The phase behavior of two-dimensional (2D) solids is the transition from solid to hexatic, and then pairs of
a fundamental, long-standing problem in statistical me- disclinations unbind to drive the transition from hexatic
chanics. Whereas three-dimensional (3D) solids charac- to fluid. Two, the system can exhibit a first-order fluid-
teristically exhibit first order (or discontinuous) melting to-solid (or solid-to-fluid) transition, with no intervening
transitions, 2D solids can melt by either continuous or phase. Both this and the first scenario were realized in a
first order melting transitions and may exhibit an inter- system of charged polystyrene microspheres, depending
mediate, so-called x-atic ordered phase that is some- on the particle diameter, which was postulated to have
where between a fluid and a solid. Previous studies [118] an effect on defect core energies [7]. Three, the system
that examine two-dimensional melting find three distinct can exhibit a first-order fluid-to-x-atic and a subsequent
scenarios [19, 20]. One, the system can exhibit a contin- continuous x-atic-to-solid transition. This combination
uous fluid-to-x-atic-to-solid transition. The x-atic phase of transitions is intermediate to the one-step fluid-to-
has quasi-long-range (power-law decay) correlations in solid first order transition and the two-step continuous
the bond order but only short-range (exponential de- KTHNY behavior. It was first experimentally observed
cay) correlations in positional order. The hexatic phase, in neutral micron-scale colloidal spheres [5], and has been
with six-fold bond order, is the most well-known exam- observed recently in simulations of hard disks in two di-
ple. The existence of continuous fluid-to-solid transi- mensions [12, 24] and under quasi-2D confinement of hard
tions was predicted by the Kosterlitz-Thouless-Halperin- spheres where out-of-plane fluctuations are limited [14].
Nelson-Young (KTHNY) theory [2123] and has been All three melting scenarios have been observed in ex-
confirmed in experiments with electrons [1] and spher- perimental studies of different systems, with a variety
ical colloids [2, 4, 8, 9]. The KTHNY theory of two-step of long and short range interactions. Recent simulation
melting is based upon the behavior of topological defects work [18] finds two of the three scenarios: point particles
in the form of dislocations and disclinations. The the- with hard core repulsion interactions follow the third sce-
ory envisions that pairs of dislocations unbind to drive nario and softer potentials lead to continuous melting. In
this paper, we report the occurrence of all three distinct
melting scenarios in a single family of hard, regular poly-
sglotzer@umich.edu gons. Hard polygons have a rotational degree of freedom,
2

which creates the possibility for more complex entropic


forces and more diverse solid phases than observed for KTHNY 4 6
3
hard disks. By varying the number of polygon edges, we
show that the melting transition scenario for a system Space-filling
of any given polygon is determined by the anisotropy of First order 5 5*
emergent entropic interactions along with the symmetry Disk-like
of the particles relative to that of the solid phase. In
particular, we show that systems of hexagons are a per- 7 8 9 10 11 12 13 14
fect realization of the KTHNY theory, exhibiting melt-
ing from the solid to the hexatic phase with an increase
in the dislocation density, then from the hexatic to the FIG. 1. The polygons studied in this work are the regular
fluid with an increase in the disclination density. We find n-gons with 3 n 14 and the 4-fold pentille labelled 5 .
that systems of triangles and squares also show a contin-
uous KTHNY-type melting transition, while systems of
pentagons and 4-fold pentilles have a first order melting A. Computational strategy
transition that occurs in a single step. Finally, we show
that systems of regular polygons with n 7 behave like We executed Monte Carlo (MC) simulations as large
disks with a first order fluid-to-hexatic and continuous as 10242 particles to obtain high precision equations of
KTHNY-type hexatic-to-solid transition. state and sample long range correlation functions to con-
We focus our study on hard, convex, regular poly- clusively identify hexatic and solid phases, which recent
gons because we aim to discover general unifying prin- hard disk simulations [12, 24] found necessary. The MC
ciples of 2D melting by filling in gaps in existing litera- simulations were performed in the isochoric (constant
ture: regular triangles [13, 16, 25], squares [11, 17, 26], area) ensemble using the HPMC [29] module of HOOMD-
pentagons [3, 10], and heptagons [10] have been previ- blue [3032]. HPMC (for hard particle Monte Carlo) is
ously studied by both experiment and simulation. These a parallel simulation implementation on many CPUs or
studies were instrumental in identifying and character- GPUs using MPI domain decomposition.
izing possible intermediate phases (triatic, tetratic, and
Each simulation begins with the particles placed in a
hexatic). We present new results for regular hexagons,
square, periodic simulation box. We perform simulations
octagons, etc. up to 14-gons and clarify the results of
for a long enough time to reach and statistically sam-
previous simulations of hard polygons using very large
ple thermodynamic equilibrium, see Supplementary In-
simulations to conclusively determine the orders of the
formation section II for a complete simulation protocol
various melting transitions, where previous studies were
description, and figure S14 for an example. Typical sim-
too small to be conclusive. Section II includes detailed
ulation runs initially form many domains in the system,
comparisons between our results and previous simulation
which coalesce together over a long equilibration period
and experimental results.
of several hundred million trial moves per particle.
We demonstrate that changing only the number of
Occasionally, two infinite, twinned domains form in the
edges on convex regular polygons is sufficient to gener-
system. Such configurations are metastable, they are at a
ate a rich array of different melting behavior, including
higher pressure than a corresponding single domain and
all three known 2D melting scenarios. We leave for fu-
the larger domain grows very slowly into the smaller one
ture studies investigation of, e.g. rounding of polygons,
as the simulation progresses. We remove stuck simula-
where experiment [13] and simulation [27] have revealed
tions and rerun them with different random number seeds
additional phases.
until we obtain a cleanly equilibrated single domain sam-
ple, except in a few cases where multiple attempts to do
so failed (e.g. = 0.714 in figure S10).
I. METHODS
This work utilized significant computational resources
on XSEDE [33] Stampede (222,000 SUs), OLCF Eos,
We investigate large systems of N identical polygons OLCF Titan (115 million Titan-core-hours), and the Uni-
with n edges that interact solely through excluded vol- versity of Michigan Flux cluster (100,000 GPU hours).
ume interactions in a box of area Abox . Particle a has
position ~ra and orientation angle a . The circumcircle
diameter of the polygons is denoted as . The majority
of our work focuses on regular polygons (n-gons) with the B. Equation of state
area of a single particle A = 2 n/2 sin(2/n). We also
include in our study, the 4-fold pentille [28], which is the We compute the equation of state P () using vol-
Voronoi cell of the Cairo pentagon pattern and thus tiles ume perturbation techniques [34, 35] to measure pressure
space (see figure S16 in the Supplementary Information in isochoric simulations. We report pressure in reduced
for the tiling configuration). Figure 1 shows all thirteen units P = P 2 /kB T . In addition to the system density
polygons and summarizes their melting behavior. = N A/Abox , we determine the averaged local density
3

field on a grid,
a
PN ~q0
A a=1 H(rc |~ra ~ri |)
(~ri ) = , (1) ~rb
rc2

where we choose the cut-off rc = 20 and H is the Heav-


iside step function. y
With the fluid density f and the density of the solid ab
(or hexatic) phase s the latent heat L = P Abox of a
first-order phase transition is a
  ~ra x
L A 1 1
= Pc 2 , (2)
N kB T f s
Im
where Pc is the coexistence pressure estimated by
b a
Maxwell construction. Bond order: a
k,p e ihk,p i
a
Positional order: a e ih i
Re
ihsa i
C. Order parameters Body order: sa e

We use three order parameters that were previously


FIG. 2. a) Pictorial definitions of ~ra , ~rb , a , ab , and ~
q0 b)
effective at identifying the hexatic phase in the hard disk
We map order parameters to a color wheel for visualization.
system [12, 24]. Each order parameter is a complex num- Each order parameter is rotated by its average in a given
ber on the unit circle. We visualize order parameter fields frame so that the average order parameter is colored green.
directly in the x-y-plane of the system by mapping the The color wheel is the color part of the cubehelix [37] color
complex values of an order parameter to a color wheel map at constant apparent luminance (v = 0.5, = 1, s = 4.0,
(see Figure 2). Short-range order shows up as colors r = 1, h = 1).
rapidly cycling through the color wheel, quasi-long-range
order appears as patches, long-range order appears as a
single solid color across the box, and two-phase regions identifies the orientation of the p nearest neighbors
show two separate behaviors in a single simulation box. around particle a. Here ab is the angle the separation
Independent simulation runs result in different system vector ~rb ~ra makes with respect to the positive x-axis,
orientations in the box. We rotate the order parameters and NNp (a) is the set of p nearest neighbors of a, see
so that the average in a given frame is colored green so Figure 2 for a graphical definition. We omit p in the sub-
that images may be compared by eye. script when it is equal to k and write ka = k,k a
. Some
The positional order parameter authors suggest using a morphometric approach [36] to
compute the bond orientation order parameter, which
a = ei~ra ~q0 (3) requires computing a Voronoi diagram of the system of
particles. We do not adopt this scheme because we find
identifies how well the position ~ra of particle a fits on the use of p fixed neighbors sufficient as it generates order
a perfect lattice with reciprocal lattice vector ~q0 , as de- parameter fields fully consistent with the defects present
picted in Figure 2. When all particles have the same in the system.
phases in , they are in a perfect solid. Defects cause The body orientation order parameter
to rotate. We choose ~q0 as the brightest peak in the struc-
ture factor [12] computed with the following procedure: sa = eisa (5)
(i) Initialize a density grid with roughly 8 8 pixels per
particle. (ii) At the center of each particle, place a Gaus- identifies the orientation of a particle accounting for s-
1
sian on this grid with standard deviation 10 . (iii) Take fold symmetry. a is the angle that rotates particle a
the fast Fourier transform (FFT) of the density grid to from a reference frame into a global coordinate system
get the discretized S(~q). (iv) Smooth with a Guassian (see Figure 2). It allows us to analyze the presence of
filter, standard deviation 2 pixels S(~q) Ssmooth (~q). rotator phases in which sa decays to zero rapidly as a
(v) Choose ~q0 from the location of the brightest pixel function of the separation distance.
in Ssmooth (~q).
The bond orientation order parameter for k-fold rota-
tional symmetry D. Correlation functions

1 X
a
k,p = eikab (4) Correlation functions measure the behavior of the or-
p der parameters as the separation rba = |~rb ~ra | between a
bNNp (a)
4

pair of particles increases. The correlation functions for box of length L is divided into squares of side length Lb .
bond orientation order and body orientation order are Within each box, we calculate
implemented in our analysis code as * NL +

1 Xb a
P PN v(Lb ) = v , (10)
bS(m,N ) v a (v b ) (r rba )
a=1 N a=1
Cv (r) = P PN , (6)
bS(m,N ) a=1 (r rba )
where v = k,p or , NLb is the number of particles in
a sub-block and h i denotes averaging over sub-blocks in
where v = k,p or n and v is the complex conjugate.
the same snapshot. Standard errors are calculated by
The sampling S(m, N ) randomly selects m particles out
averaging the values of v(Lb ) over independent frames.
of N without replacement.
One can compute a correlation function from the po-
sitional order parameter , but it is extremely sensitive F. Topological defect analysis via cell-edge
to the choice of ~q0 . We observe that peaks misidentified counting
by only a single pixel result in an apparent lack of quasi-
long-range positional order. Instead, we follow Ref. [12]
We generate statistics on topological defects using a
and compute positional correlation functions from
Voronoi tessellation of the set of particle centers to count
hg ((r, 0), )i 1 edges of Voronoi cells. Each particle a is assigned the
Cg (r) = , (7) number of adjacent Voronoi cells na and the disclina-
max(hg ((r, 0), )i 1) tion charge qa = na 6. In the presence of well-
separated topological defects, disclinations can be found
which oscillates, so we show only identified peaks. The
by identifying particles with non-zero disclination charge,
signal then decays to about 103 at large r. The function
and dislocations correspond to bounds pairs of a five-
g (~r, ) is the discretized 2D pair correlation function ob-
coordinated particle (q = 1) and a seven-coordinated
tained by correlated averaging over individual measure-
particle (q = +1). However, while disclination charges
ments with index i,
with |q| > 1 are very rare in our hard particle systems,
P PN particles with non-zero disclination charge are often not
bS(m,N ) a=1 d (R()~rba ~ri ) clearly separated or bound into pairs but instead agglom-
g (~ri , ) = , (8)
m/A erate into larger clusters. This makes it ambiguous to
identify the locations of individual disclinations and dis-
where R() is the rotation matrix that rotates a vector locations.
by the angle , r is the bin size, and d is the coarse- To overcome this ambiguity, we cluster defects if they
grained delta function are in adjacent Voronoi P cells and calculate the total
 disclination charge q = qa of the cluster [38]. For
(r)2 if 0 rx , ry < r, disclination-neutral clusters (q = 0), the total Burgers
d (~r) = (9)
0 otherwise. vector ~b of the cluster can be determined from the discli-
nation charges and their positions,
Since the system rotates from frame to frame, it must be X
aligned to the bond orientation order parameter before ~b = z qa~ra , (11)
averaging. To do this, we compute g (~ri , ) over many a
frames with large m. We align each frame using the bond where z is the unit vector along the out-of-plane axis
orientation order parameter averaged over all particles and the cross product is performed in 3D space. Because
in the frame, = arg(hk,p i). Averaging the separately Burgers vectors are topologically restricted to be lattice
aligned g (~ri , ) and then computing Cg (r) significantly vectors, the vector we obtain from this computation is
reduces noise [24]. then snapped to its closest lattice point in a hexagonal
lattice whose lattice spacing is that of the ideal lattice at
the density of the simulation frame. We count the num-
E. Sub-block scaling analysis ber of particles in three types of defect clusters, overall
neutral (q = 0, ~b = 0), Burgers-charged (q = 0, ~b 6= 0),
The sub-block scaling of is a sensitive measure of and disclination-charged (q 6= 0) as a function of density.
the density of the hexatic-to-solid transition [6, 15]. The
first density that sits on or under a line of slope 1/3
is at the hexatic-to-solid transition. Similarly, sub-block G. Phase determination
scaling in determines the density at which the hexatic
melts into the fluid, with a line of slope 1/4. We identify first-order transitions using the following
We perform a sub-block scaling analysis on the po- criteria: (i) a two-phase region is evident in isochoric sim-
sitional order parameter and the bond orientational ulations at large N , (ii) the two phases have different den-
a
order parameter k,p . For this analysis the simulation sities, and (iii) the equation of state has a Mayer-Wood
5

a b
= 0.708 = 0.694 = 0.682

Histogram
= 0.702 = 0.686 = 0.678

0.670 0.678 0.686 0.694 0.702 0.710



4
ln (v [Lb ]/v [L])

c Slope -1/3 Slope -1/4


3
2 6
1
0
-4 -3 -2 -1 0 -4 -3 -2 -1 0
= 0.686 = 0.702 = 0.708
ln (Lb /L) ln (Lb /L) hexatic hexatic solid
d = 0.696 = 0.692 = 0.680 e
Histogram

= 0.694 = 0.688 = 0.676

Fluid Solid 6

0.670 0.678 0.686 0.694



4
ln (v [Lb ]/v [L])

Slope -1/3 Slope -1/4


f 3
2 6
1
0
-4 -3 -2 -1 0 -4 -3 -2 -1 0
= 0.680 = 0.688 = 0.694
ln (Lb /L) ln (Lb /L) fluid/solid fluid/solid solid
g = 0.704 = 0.692 = 0.686 h
Histogram

= 0.700 = 0.698 = 0.684

Fluid Hexatic 6

0.670 0.678 0.686 0.694 0.702



4
ln (v [Lb ]/v [L])

Slope -1/3 Slope -1/4


i 3
6
2
1
0
-4 -3 -2 -1 0 -4 -3 -2 -1 0 = 0.692 = 0.700 = 0.704
ln (Lb /L) ln (Lb /L) fluid/hexatic hexatic solid

FIG. 3. Example phase transitions from the three melting scenarios. We show data for (a-c) hexagons (N = 5122 ) following the
KTHNY scenario, (d-f) pentagons (N = 10242 ) exhibiting a first-order fluid-to-solid transition, and (g-i) octagons (N = 10242 )
following an intermediate scenario exhibiting a first-order fluid-to-hexatic transition. For each shape, the top left panels show
local density histograms (a,d,g). The right panels show bond and positional order parameters at selected densities (b,e,h). The
bottom left panels show sub-block scaling analysis (c,f,i).
6

loop [12, 24] which decreases in height as N increases. In for the triangle, square, and hexagon data, respectively.
contrast, continuous transitions have a monotonic equa- Triangles and hexagons have hexatic order with a signal
tion of state and only a single phase at a single density in 6 , and squares have tetratic order with a signal in 4 .
is present in each frame across the entire transition. The Figure 3ac demonstrates the continuous transitions for
signature of two phase regions includes a bimodal local hexagons, with a single density at each state point and
density histogram along with two different correlation gradually developing order in the order parameter fields.
length behaviors seen in the and order parameter
fields coincident with the low and high local densities. In
all cases where we find a Mayer-Wood loop, we observe 1. Decay of correlation functions
two-phase regions, and in all cases where we find a con-
tinuous equation of state, we observe only single phases The first state point showing quasi-long-range bond
across the transition. orientation order in the hexagon system is = 0.686. At
Correlation lengths vary significantly in the simula- this density, positional order is extremely short-ranged,
tions. In the fluid phase, positional order decorrelates persisting for only a few particle diameters. As density
instantly with correlation lengths of only 1. Bond orien- increases, the decay length of the positional order be-
tational order k,p persists a bit further with correlation comes longer, up to hundreds of particle diameters at
lengths of tens of but still exhibits clear short-range or- = 0.702, and begins to diverge. At = 0.708, posi-
der. Previous authors have used the dynamic Lindemann tional order switches to quasi-long-range and the system
criterion to locate melting transition points [8, 39]. We is in the solid phase as determined by sub-block scaling.
focus on the correlation functions here because of their KTHNY theory predicts a slope of 1/4 in 6 scaling
utility in discriminating hexatic and solid phases, which at the fluid-to-hexatic transition, a perfect match for the
the dynamic Lindemann criterion is unable to do. In the = 0.686 line in Figure 3c. The theory also predicts a
hexatic phase, oscillates visibly through the full color slope of 1/3 for scaling, which similarly matches the
wheel along a given direction forming stripes (Fig. 3b), scaling for = 0.708. At the same density, the directly
a behavior indicative of short-range order. Unbound dis- computed positional correlation Cg (r) length is begin-
locations exist at the end of each stripe. For continu- ning to diverge so it is difficult to determine the exact
ous phase transitions, the oscillation period gets larger density of the hexatic-to-solid transition from correlation
as density increases. Solid phases lead to patchy motifs functions alone.
in (quasi-long-range order). Two-phase regions show a
combination of two of these motifs in a single system.
Our observations confirm in all cases that we have run 2. Topological defects and local order
simulations sufficiently large to allow the fluctuations in
the order parameter fields of the fluids and x-atics to
The KTHNY theory envisions a picture of tightly
occur several times across the box. Only in the solid
bound defects transitioning to free dislocations at the
phase regions are the order parameter fields essentially
solid-hexatic transition, and to free disclinations at the
constant and vary by less than one period.
hexatic-fluid transition. Interestingly, both Bernard et
al. [12] and this work show that defects are not free,
but rather form large and complex clusters, often highly
II. RESULTS AND DISCUSSION
anisotropic, indicating medium-ranged effective inter-
defect interactions. In the extreme case that defects clus-
We performed identical analyses for each of the thir- ter into one-dimensional strings, such a scenario would
teen polygons investigated in this work. Within the main lead to grain-boundary induced melting [40, 41].
text we include representative plots and snapshots to il- Figure 4c shows the density dependence of the num-
lustrate examples and explanations of the three melting ber of particles comprising clusters of defects that are
scenarios. The Supplementary Information contains de- overall neutral, those with a Burgers charge, and those
tailed plots for all of the shapes in figures S1S13, includ- with disclination charge. A large number of free dislo-
ing the equation of state, local density histogram, sub- cation clusters with non-zero Burgers vectors simultane-
block scaling analysis, correlation functions, snapshots ous with few free disclinations is evidence for the hexatic
colored by all of the order parameters and structure fac- phase. We observe that the count of particles in Burgers-
tors. charged clusters begins increasing just below = 0.710
at the solid-to-hexatic transition. The count of particles
in disclination-charged clusters remains low, and starts
A. KTHNY behavior ramping up slowly in the middle of the hexatic phase
at = 0.696. Throughout the hexatic phase, we find
Our data show that systems of triangles, squares, and many more particles in Burgers-charged clusters than in
hexagons have continuous fluid-to-x-atic and continuous disclination-charged clusters. This is consistent with the
x-atic-to-solid transitions. We refer to the monotonic two-step KTHNY melting scenario and the continuous
equations of state in Figure S1, Figure S2, and Figure S5 phase transition we observe for hexagons.
7

a 2500 b 2500 c 2500


Neutral
2000 Dislocation 2000 2000
Disclination
Particle count

Particle count

Particle count
1500 1500 1500
Fluid Two-phase Solid Fluid Two-phase Solid Fluid Hexatic Solid
1000 1000 1000

500 500 500

0 0 0
0.672 0.678 0.684 0.690 0.696 0.702 0.708 0.718 0.721 0.724 0.727 0.730 0.733 0.674 0.682 0.690 0.698 0.706 0.714

d 2500 e 2500 f 2000
1800
2000 2000 1600
1400
Particle count

Particle count

Particle count
1500 1500 1200
Hexatic

Hexatic

Hexatic
Fluid
Fluid Two-phase Solid Fluid Two-phase Solid 1000 Two-phase Solid
1000 1000 800
600
500 500 400
200
0 0 0
0.672 0.678 0.684 0.690 0.696 0.702 0.708 0.672 0.678 0.684 0.690 0.696 0.702 0.708 0.684 0.690 0.696 0.702 0.708 0.714

g 2500 h 2500 i 2000
1800
2000 2000 1600
1400
Particle count

Particle count

Particle count
1500 1500 1200
Hexatic

Hexatic

Hexatic
Fluid Two-phase Solid Fluid Two-phase 1000 Fluid Two-phase
1000 1000 800
600
500 500 400
200
0 0 0
0.684 0.690 0.696 0.702 0.708 0.714 0.684 0.690 0.696 0.702 0.708 0.714 0.690 0.696 0.702 0.708 0.714

j 2500 k 2500

2000 2000
Particle count

Particle count

1500 1500
Hexatic

Hexatic
Fluid

Two-phase Solid Fluid Two-phase Solid


1000 1000

500 500

0 0
0.686 0.692 0.698 0.704 0.710 0.716 0.690 0.696 0.702 0.708 0.714 0.720

FIG. 4. Defect counts as a function of density obtained with cell-edge counting. The particle count is the number of particles
that belong to any defect cluster classified into neutral, Burgers-charged and disclination-charged clusters. The figures show
results of the defect analysis plots for polygons n = 5, 6, 7, 8, 12, averaged over 48 runs with 1949 frames per run at the system
size N = 1282 . The defect count algorithm is expensive, so we do not run it on the largest size simulations we performed. We
also show the phase diagram overlaid on the counts. Error bars at one standard deviation are smaller than the symbol size.

Figure 5a shows the structure of the defects in a sys- 4A/P 2 for each Voronoi cell, where A is the area and
tem of hexagons at = 0.690, in the hexatic phase. The P is the perimeter of the cell.
 Regular polygons have
density of free dislocations (red and green pairs) is much the value IQ = / n tan n . More elongated Voronoi
higher than that of free disclinations (there is a single cells have low IQ and indicate locations of large crystal
lone green particle in the image), but the structure is deformity, which makes IQ an approximate indicator of
more complicated because the defects combine into large defects. Figure 5b and the right panel of Supplementary
clusters. The left panel of Supplementary Movie 1 shows Movie 1 color particles by hIQiIQ. Blue indicates large
how defects migrate as the Monte Carlo simulation pro- deviation from hIQi and yellow indicates little deviation.
gresses via local moves for a simulation of hexagons at If IQ > hIQi, the particle is darkened, as these particles
= 0.690. Dislocations are unstable and highly mobile, are closest to having regular crystal environments.
quickly hopping between sites and popping in and out of The coloring scheme based on the isoperimetric quo-
existence. We find that the count of nearest neighbors is tient IQ is a continuous quantity and, as seen in the
very sensitive to small changes in particle coordinates. movie, is less sensitive to small particle displacements. In
We also compute the isoperimetric quotient IQ = contrast, the number of edges of a Voronoi cell changes
8

a
Phase
Polygon Cross-correlation Fluid x-atic Solid
Triangle |h6,3 6 i| 0.117 0.221 0.236
Square |h4 4 i| 0.449 0.790 0.810

Pentagon |h6 10 i| 0.001 - 0.001

Hexagon |h6 6 i| 0.244 0.294 0.299

Heptagon |h6 14 i| 0.001 0.001 0.001
Octagon |h6 8 i| 0.001 0.001 0.001

Dodecagon |h6 12 i| 0.001 0.039 0.039

TABLE I. Cross-correlation between bond and body orien-


tation in the fluid, x-atic, and solid phases. We select the
highest density pure fluid phase, the highest density pure x-
atic phase, and the lowest density pure solid phase. Errors of
two standard deviations of the mean are approximately 0.001
for all values.
b

that of the bond order parameter k,p . We quantify


this by computing the cross-correlation between the two
quantities |hk,p s i| in the fluid, hexatic, and solid phases
(Table I). We expect high bond-body cross-correlation at
high density when particle rotation becomes locked in, as
found in previous studies [10]. However, a strong signal
is present even in the fluid for triangles, squares, and
hexagons.
The cross-correlation results demonstrate alignment of
the edges of neighboring particles. This can be inter-
preted as an effect of directional entropic forces [42]. We
quantify the alignment with the potential of mean force
and torque (PMFT) [43], computed from runs in the
highest density pure fluid phase for each polygon. For
each particle, we determine the relative position of its
neighbors and construct histograms using many frames
to obtain reliable averages. The free energy F of a given
FIG. 5. Representative snapshot of the hexagon system
N = 1282 at density = 0.690. (a) Particles are colored by configuration is related to the negative logarithm of the
the number of edges in the Voronoi cell. Red, blue, and green histogram. We average an area in the PMFT 25 from
cells have seven, six, and five sides respectively. (b) Particles the center to set F = 0 as a baseline.
are colored by hIQi IQ. Blue indicates large deviation from Figure 6 shows the results of this calculation. All low-
hIQi and yellow indicates little deviation. symmetry regular polygons (n < 7) have distinctly sep-
arated wells for edge-to-edge contacts at F = 1.5kB T .
This indicates preferential edge-to-edge alignment in the
discontinuously across Monte Carlo steps. This suggests solid by strong directional entropic forces. But only when
hIQiIQ might be a more robust indicator of defect con- the symmetry of the particle shape is compatible with
centrations on short time scales, while cell-edge counting the symmetry of the bond order, as found in hexagons,
is simple and reliable for long time averages over uncor- triangles, and squares, does such edge-to-edge alignment
related frames. We also note that in systems of trian- promote local solid motifs in the fluid. Such a pre-ordered
gles and squares, defects are even more delocalized and fluid allows the correlation length to increase smoothly
Voronoi cell-edge counting is unable to identify defects, from short-range to quasi-long-range and the transition
while hIQi IQ can still identify areas where defects are to the solid to be continuous.
present. Triangles and squares have a distinct overall behavior
from hexagons due to the delocalized defects. They have
a smeared out phase transition between = 0.71 and
0.8 (Figure 7). Their equations of state are smooth
3. Particle alignment and slowly increasing with barely detectable kinks (Fig-
ures S1 and S2). In comparison, the hexagon equation of
At all densities we simulate for triangles, squares, and state has sharp kinks and almost levels off through the
hexagons, the body order parameter s closely matches transition (Figure S5). Interestingly the melting transi-
9

1.5 1.5 1.5 1.5

0.5 0.5 0.5 0.5

0.5 0.5 0.5 0.5

1.5 1.5 1.5 1.5

1.5 0.5 0.5 1.5 1.5 0.5 0.5 1.5 1.5 0.5 0.5 1.5 1.5 0.5 0.5 1.5

Triangle Square Pentagon 4-fold pentile


1.5 1.5 1.5 1.5

0.5 0.5 0.5 0.5

0.5 0.5 0.5 0.5

1.5 1.5 1.5 1.5

1.5 0.5 0.5 1.5 1.5 0.5 0.5 1.5 1.5 0.5 0.5 1.5 1.5 0.5 0.5 1.5

Hexagon Heptagon Octagon Nonagon


1.5 1.5 1.5 1.5

0.5 0.5 0.5 0.5

0.5 0.5 0.5 0.5

1.5 1.5 1.5 1.5

1.5 0.5 0.5 1.5 1.5 0.5 0.5 1.5 1.5 0.5 0.5 1.5 1.5 0.5 0.5 1.5

Decagon Hendecagon Dodecagon Tridecagon


F /kB T = 1.5 F /kB T = 1.0 F /kB T = 0 F /kB T = 4

FIG. 6. Potential of mean force and torque plots for polygons n=313 computed at the highest density fluid for each polygon.
We choose the zero energy reference as the average over a large region 25 from the center (not shown). The data are shown
as a contour plot with selected free energy or entropy well contours. Computed from N = 2562 simulations.

tion of hexagons becomes more evidently continuous as B. First-order fluid-to-solid


the system size increases and could, in fact, be mistaken
for a first order transition in small systems of only a few Our data clearly show that systems of regular pen-
thousand particles. tagons have a first-order transition directly from the fluid
to the solid. This is seen in Figure 3d-f at = 0.688.
There is a stripe of low density (local density = 0.676)
next to a stripe of solid ( = 0.694) with quasi-long-
Our data agree qualitatively with previous studies. range positional order. The two phase region starts at
Simulations of equilateral triangles using an approximate = 0.680 and the pure solid phase starts at = 0.694.
event-driven molecular dynamics method showed a con- The two phase region is coincident with a Mayer-Wood
tinuous fluid-to-liquid crystal-like phase transition [25]. loop in the equation of state (Figure S3). The symme-
Monte Carlo simulations revisiting that system reported try of the pentagon is not compatible with hexagonal
a fluid-to-hexatic transition at = 0.70 and a hexatic-to- ordering in the solid, so even though it has strong en-
solid transition at = 0.87 [16]. Ref. [16] finds a chiral tropic edge-edge bonds, the body-bond cross-correlation
phase at 0.89, which we do not observe as we focus on (Table I) is zero and the phase transition becomes first-
the melting behavior at lower density < 0.80. Although order.
no experiments have yet been reported on hard, mathe- The one-step melting process of pentagons also shows
matically regular triangles, rounded triangular colloidal up in the defect counts Figure 4a,b. We observe sharp
platelets exhibit a fluid-to-hexatic transition, but the or- kinks just below the pure solid phase at = 0.692 for
der of the transition was not determined [13]. Ref. [13] counts in clusters with Burgers and disclination charges.
also finds a phase with local chiral symmetry breaking One-step melting is supported by this coincident increase
that we do not observe for hard regular triangles at in the number of free dislocations and free disclinations.
< 0.85. Monte Carlo simulations of squares show a Our results are consistent with previous studies of pen-
continuous fluid-to-tetratic transition at = 0.7 [26]. tagons. Monte Carlo simulations on small systems of
Experiments on vibrated granular squares (LEGOs) find pentagons showed a transition from a fluid to a hexagonal
tetratic orientational order in the range = 0.70 to rotator crystal at = 0.68 [10]. The same fluid to rota-
0.74 [17]. We are not aware of any studies on systems tor crystal transition was also reported with rounded col-
of hard hexagons to compare with. loidal pentagons at = 0.66 [44]. The data in that work
10

24
... 22 0.32
0.28

NkB T
14 20

L
0.24
13 18 0.2


12

Pc

s
16

1
0.03
11


0.02
n

Two-phase 14

f
10 Fluid

1
Solid 0.01
9 12

2
A
0.0
5 7 9 11 13 ...
8 10
7 8
6 Hexatic 5 7 9 11 13 ...
5 n

0.676 0.682 0.688 0.694 0.700 0.706 0.712 0.718 FIG. 8. Latent heat (green diamonds) of the first-order phase
5 transition and the two component factors plotted separately.
Pc (red circles) is the coexistence pressure and the area dif-
4
n

ference per particle A is shown below (blue squares). Filled


3 symbols are regular polygons, and the open symbols are the
0.70 0.72 0.74 0.76 0.78 0.80 0.82 0.84 4-fold pentille. Error bars are estimated at 0.001 error in the
transition densities. Data for disks (n ) is from [12].

FIG. 7. Phase diagram of hard polygon melting behavior.


Disk results (n ) are from [12, 24]. The label 5 refers C. Disk-like behavior
to the 4-fold pentille. The n = 3 solid is a honeycomb lattice
with alternating triangle orientations, the n = 4 the solid is a Figure 7 summarizes the phases found for all regular
square lattice, and the n 5 solids are all hexagonal. polygons studied in this work and includes the hard disk
phase diagram (n ) for comparison. Disks have a
first-order fluid-to-hexatic transition, a narrow region of
stability for the hexatic phase, and a continuous hexatic-
suggests a possible hexatic phase, but is inconclusive due to-solid transition. A similar smooth behavior with n is
to small system sizes within the field of view of the cam- found in the latent heat of the first-order transitions (Fig-
era. Our simulations indicate there should be no hex- ure 8). The coexistence pressure decreases monotonically
atic phase with zero rounding. Pentagon phase behav- as n increases, approaching that of disks. The 4-fold pen-
ior has also been studied in systems of 5-fold symmetric tille is an outlier compared to the regular polygons with
molecules [45] and with vibrating shaking tables [3, 46]. almost a factor of two higher coexistence pressure.
None of these previous studies, however, conclusively We executed simulations up to n = 14 in this work and
demonstrated the first order, one step nature of the melt- find that all regular polygons with n 7 have first-order
ing transition in pentagons. fluid-to-hexatic and continuous hexatic-to-solid phase
transitions. Figure 3g-i illustrates this for a system of
Of all the regular shapes we studied, only triangles, octagons. In the two-phase region we see a mixture of
squares, and hexagons fill space. These are also the only fluid and hexatic phases. This is the same behavior as
three KTHNY-type shapes. To test if another space fill- seen for disks, the only difference is that the transition
ing polygon behaves similarly, we conducted simulations shifts to lower packing fraction as n decreases. This shift
of the 4-fold pentille [28], an irregular pentagon with two is expected from the increasing anisotropy and relative
different edge lengths that tiles space with 4-fold symme- strength of directional entropic forces with decreasing n.
try. We find that the 4-fold pentille behaves like the reg- At n = 7 the start of the phase transition is at = 0.680,
ular pentagon, with a first-order fluid-to-solid transition increases to = 0.692 by n = 12, 14 then again increases
and no intermediate hexatic, though the transition oc- to = 0.702 for disks. We observe a small jump in
curs at a higher pressure (Figure S4). At high fluid den- critical density from n = 14 to disks, despite the very
sities, directional entropic forces (Figure 6) are blurred close coexistence pressures. We surmise that all regular
by the edge lengths and the resulting ten-fold particle polygons with n 7 have a first-order fluid-to-hexatic
body order is not compatible with either the tiling or the transition followed by a continuous hexatic-to-solid tran-
hexagonal solid motifs, so the transition is first-order. sition, with the transition range shifting higher in density
This suggests the space-filling property itself is not the with larger n.
factor triggering KTHNY melting but instead the simi- Of the studied polygons with n 7, only the do-
larity between the local order in the dense fluid and the decagon with n = 12 has a particle symmetry compati-
local order in the solid. ble with the bond order in the solid. Although for this
11

polygon strong directional entropic forces have the po- ible with the solid. Pentagons and plane-filling 4-fold
tential to drive a continuous transition from the fluid to pentilles have similar strong anistropic aligning forces in
the hexatic and solid phase, the simulations show a clear the fluid, but their symmetry is incompatible with the
first-order transition for dodecagons. There is no body- hexagonal order of their solid phase. As a result they
bond cross-correlation in the dodecagon fluid (Table I), display a clear one-step first-order melting of the solid to
but it does appear weakly in the hexatic phase. The the fluid with no intermediate phase. Regular polygons
relatively short edge lengths in the dodecagon lower the with sufficiently many (n 7) edges have no preferential
strength of edge-edge alignment to the point where it is alignment. They show a close resemblance to hard disk
not strong enough to encourage local hexagonal motifs behavior and exhibit a first order fluid-to-hexatic phase
in the fluid, and as a result the transition from bond ori- transition and a continuous hexatic-to-solid transition,
entation disorder to order becomes sharp. All regular which is the intermediate scenario between KTHNY and
polygons with n 7 have smeared out F = 1kB T con- one-step first-order melting.
tact wells (Figure 6) that encircle the polygon, indicating Our results show that some 2D particles exhibit a
weaker edge-edge alignment than triangles, squares, and KTHNY melting scenario when local coupling is strong
hexagons, allowing a much more isotropic behavior. This and a hard-disk melting scenario when local coupling is
is why all regular polygons n 7 share the same fluid- weak. Our findings agree with a recent study [18] of disks
solid transition properties as disks. interacting via soft rm potentials; for m > 6, disks are
Regular polygons with n 7 exhibit an intermediate hard enough to exhibit hard-disk melting, whereas for
hexatic phase with a very narrow region of stability and a m 6, they exhibit KTHNY melting. In our system,
first-order transition to the fluid. As a result, the defect we correlate softness achieved with n 6 with strong
counts (Figure 4d-k) for these polygons is not as clear. anisotropy in the entropic force field. For polygons with
The Burgers-charged and disclination-charged counts for n 6, the anisometry is sufficient to allow other particles
these polygons have sharper kinks than the hexagons, to approach at distances well inside the corresponding
but they do not parallel each other as closely as in the circumcircle of the polygon. This overlap creates the
pentagons. Despite this, the hexatic-to-solid transition equivalent of a soft effective interaction. Rounding poly-
follows KTHNY predictions. In particular, the sub-block gon edges limits how close neighboring particles may ap-
scaling analysis for predicts the hexatic-to-solid tran- proach, and e.g. systems of sufficiently rounded hexagons
sition density with high accuracy, as shown in Figure 3i should demonstrate disk-like behavior we leave such a
where the scaling line for = 0.704 falls almost exactly study for future work. More generally, the melting tran-
on the dotted line predicted by theory. sition depends strongly on the symmetry of local inter-
actions, not just the strength, when one has anisotropic
interactions. In the present study, this anisotropic cou-
III. CONCLUSION pling is provided by emergent entropic forces, but we
predict that anisotropic coupling of any origin is suffi-
cient to produce the variety of melting scenarios we have
In 1988, Strandburg wrote [19]: For a decade now
observed in polygons.
the nature of the two-dimensional melting transition has
remained controversial. At that time the three fluid-
to-solid transition scenarios discussed here were consid-
ered, but in many cases the available evidence was in- IV. ACKNOWLEDGMENTS
conclusive as to which system shows which scenario and
whether all scenarios indeed occur. 2D simulations have This work was supported by the National Science
come a long way since 1988. With the advancement Foundation, Division of Materials Research Award
of high-performance computing we can now obtain cor- #DMR 1409620 (to J.A.A. and S.C.G.), US Army
relation functions with high enough precision and rely Research Office Grant W911NF-10-1-0518 (to S.C.G.
on additional analysis techniques like order parameter and M.E.), National Science Foundation Graduate Re-
fields, sub-block scaling, and cell-edge counting. To- search Fellowship Grant DGE 1256260 (to J.A.), and
gether, these tools reliably identify x-atic phases and re- the DOD/ASD(R&E) under Award No. N00244-09-1-
solve the nature of the 2D melting transition. Today, we 0062. Any opinions, findings, and conclusions or rec-
can confidently state that the three scenarios discussed ommendations expressed in this publication are those
already 30 years ago indeed occur and, in fact, can be of the authors and do not necessarily reflect the views
observed in a single system, namely the hard polygon of the DOD/ASD(R&E). M.E. acknowledges funding by
family studied in the present work. Deutsche Forschungsgemeinschaft through the Cluster of
As we have demonstrated, the polygons shape sym- Excellence Engineering of Advanced Materials.
metry with respect to the lattice of the solid phase to- All data analysis in this work was performed us-
gether determine the melting scenario. Systems of trian- ing Freud, an in-house Python-driven high-performance
gles, squares, and hexagons follow the KTHNY scenario toolkit developed by the Glotzer Group. Matthew
well. They promote strong directional entropic forces Spellings implemented the complex correlation function
that preorder the fluid into symmetries that are compat- routine and the cubeellipse color map. Eric S. Harper
12

implemented the PMFT analysis code and parallelized and services provided by Advanced Research Computing
all modules. We thank Wenbo Shen for suggesting that Technology Services at the University of Michigan, Ann
we examine correlations in the body orientation order pa- Arbor.
rameter, and Greg van Anders for a critical read of the
manuscript.
V. AUTHOR CONTRIBUTIONS
This work used resources of the Oak Ridge Lead-
ership Computing Facility at the Oak Ridge National J. A. Anderson ran the large scale simulations and ana-
Laboratory, which is supported by the Office of Sci- lyzed the data. J. Antonaglia analyzed the defects. J. A.
ence of the U.S. Department of Energy under Contract Millan executed small scale simulations and determined
No. DE-AC05-00OR22725, on the Extreme Science and regions of interest. J. A. Anderson and M. Engel con-
Engineering Discovery Environment (XSEDE), which is ceived the project. S. C. Glotzer supervised the research.
supported by National Science Foundation grant num- All authors discussed methods and results and partici-
ber ACI-1053575, and also on computational resources pated in writing the manuscript.

[1] C. J. Guo, D. B. Mast, R. Mehrotra, Y. Z. Ruan, Transition, Physical Review Letters 107, 155704 (2011).
M. A. Stan, and A. J. Dahm, Evidence in Support [13] Kun Zhao, Robijn Bruinsma, and Thomas G. Mason,
of Dislocation-Mediated Melting of a Two-Dimensional Local chiral symmetry breaking in triatic liquid crys-
Electron Lattice, Physical Review Letters 51, 1461 tals, Nature Communications 3, 801 (2012).
1464 (1983). [14] Weikai Qi, Anjan P. Gantapara, and Marjolein Dijkstra,
[2] C. A. Murray and D. H. Van Winkle, Experimen- Two-stage melting induced by dislocations and grain
tal observation of two-stage melting in a classical two- boundaries in monolayers of hard spheres, Soft matter
dimensional screened Coulomb system, Physical Review 10, 54495457 (2014).
Letters 58, 12001203 (1987). [15] Weikai Qi and Marjolein Dijkstra, Destabilisation of the
[3] Subir Sachdev and David R. Nelson, Statistical mechan- hexatic phase in systems of hard disks by quenched disor-
ics of pentagonal and icosahedral order in dense liquids, der due to pinning on a lattice. Soft matter 11, 28526
Physical Review B 32, 14801502 (1985). (2015).
[4] R. E. Kusner, J. A. Mann, J. Kerins, and A. J. Dahm, [16] Anjan P. Gantapara, Weikai Qi, and Marjolein Dijkstra,
Two-Stage Melting of a Two-Dimensional Collodial Lat- A Novel Chiral Phase of Achiral Hard Triangles and an
tice with Dipole Interactions, Physical Review Letters Entropy-Driven Demixing of Enantiomers, Soft Matter
73, 31133116 (1994). 11, 19 (2015).
[5] Andrew Marcus and Stuart Rice, Observations of First- [17] Lee Walsh and Narayanan Menon, Ordering and
Order Liquid-to-Hexatic and Hexatic-to-Solid Phase Dynamics of Vibrated Hard Squares, (2015),
Transitions in a Confined Colloid Suspension, Physical arXiv:1510.00656.
Review Letters 77, 25772580 (1996). [18] Sebastian C. Kapfer and Werner Krauth, Two-
[6] Ken Bagchi, Hans C. Andersen, and William Swope, Dimensional Melting: From Liquid-Hexatic Coexistence
Computer Simulation Study of the Melting Transition to Continuous Transitions, Physical Review Letters
in Two Dimensions, Physical Review Letters 76, 255 114, 035702 (2015).
258 (1996). [19] Katherine J. Strandburg, Two-dimensional melting,
[7] a J Armstrong, R C Mockler, and W J OSullivan, Reviews of Modern Physics 60, 161207 (1988).
Isothermal-expansion melting of two-dimensional col- [20] David R. Nelson, Defects and Geometry in Condensed
loidal monolayers on the surface of water, Journal of Matter Physics (Cambridge University Press, 2002).
Physics: Condensed Matter 1, 17071730 (1999). [21] J. M. Kosterlitz and D. J. Thouless, Ordering, metasta-
[8] K. Zahn, R. Lenke, and G. Maret, Two-Stage Melt- bility and phase transitions in two-dimensional systems,
ing of Paramagnetic Colloidal Crystals in Two Dimen- Journal of Physics C: Solid State Physics 6, 1181 (1973).
sions, Physical Review Letters 82, 27212724 (1999), [22] B. I. Halperin and David R. Nelson, Theory of Two-
arXiv:9905238 [cond-mat]. Dimensional Melting, Physical Review Letters 41, 121
[9] C. Eisenmann, U. Gasser, P. Keim, and G. Maret, 124 (1978).
Anisotropic Defect-Mediated Melting of Two- [23] A. P. Young, Melting and the vector Coulomb gas in two
Dimensional Colloidal Crystals, Physical Review dimensions, Physical Review B 19, 18551866 (1979).
Letters 93, 105702 (2004). [24] Michael Engel, Joshua A. Anderson, Sharon C. Glotzer,
[10] Tanja Schilling, Sander Pronk, Bela Mulder, and Daan Masaharu Isobe, Etienne P. Bernard, and Werner
Frenkel, Monte Carlo study of hard pentagons, Physi- Krauth, Hard-disk equation of state: First-order liquid-
cal Review E 71, 036138 (2005). hexatic transition in two dimensions with three simula-
[11] Kun Zhao, Robijn Bruinsma, and Thomas G Ma- tion methods, Physical Review E 87, 042134 (2013).
son, Entropic crystal crystal transitions of Brownian [25] Mark Benedict and John Maguire, Molecular dynam-
squares, Molecules , 14 (2010). ics simulation of nanomaterials using an artificial neural
[12] Etienne P. Bernard and Werner Krauth, Two-Step net, Physical Review B 70, 174112 (2004).
Melting in Two Dimensions: First-Order Liquid-Hexatic
13

[26] K.W. Wojciechowski and D. Frenkel, Tetratic Phase in entational order parameters for the analysis of disordered
the Planar Hard Square System? Computational Meth- particulate matter, The Journal of Chemical Physics
ods in Science and Technology 10, 235255 (2004). 138, 044501 (2013).
[27] Carlos Avenda no and Fernando A. Escobedo, Phase be- [37] Dave A. Green, A colour scheme for the display of astro-
havior of rounded hard-squares, Soft Matter 8, 4675 nomical intensity images, Bulletin of the Astronomical
(2012). Society of India 39, 289-295 (2011), arXiv:1108.5083.
[28] John H Conway and Heidi Burgiel, The Symmetries of [38] M.F. Ashby, F. Spaepen, and S. Williams, The struc-
Things (A K Peters/CRC Press, 2008). ture of grain boundaries described as a packing of poly-
[29] Joshua A. Anderson, M. Eric Irrgang, and Sharon C. hedra, Acta Metallurgica 26, 16471663 (1978).
Glotzer, Scalable Metropolis Monte Carlo for simulation [39] X. H Zheng and J. C Earnshaw, On the Lindemann
of hard shapes, submitted to computer physics commu- criterion in 2D, Europhysics Letters (EPL) 41, 635640
nications (2015), arXiv:1509.04692. (1998).
[30] Joshua A. Anderson, Christian D. Lorenz, and Alex [40] Daniel S. Fisher, B. I. Halperin, and R. Morf, Defects
Travesset, General purpose molecular dynamics simu- in the two-dimensional electron solid and implications for
lations fully implemented on graphics processing units, melting, Physical Review B 20, 46924712 (1979).
Journal of Computational Physics 227, 53425359 [41] S. T. Chui, Grain-boundary theory of melting in two
(2008). dimensions, Physical Review B 28, 178194 (1983).
[31] Jens Glaser, Trung Dac Nguyen, Joshua A. Anderson, [42] Pablo F. Damasceno, Michael Engel, and Sharon C.
Pak Lui, Filippo Spiga, Jaime A. Millan, David C. Morse, Glotzer, Crystalline assemblies and densest packings of
and Sharon C. Glotzer, Strong scaling of general- a family of truncated tetrahedra and the role of direc-
purpose molecular dynamics simulations on GPUs, tional entropic forces, ACS Nano 6, 60914 (2012).
Computer Physics Communications 192, 97107 (2015). [43] Greg van Anders, Daphne Klotsa, N. Khalid Ahmed,
[32] HOOMD-blue, http://glotzerlab.engin.umich.edu/hoomd- Michael Engel, and Sharon C. Glotzer, Understanding
blue (2016). shape entropy through local dense packing, Proceedings
[33] John Towns, Timothy Cockerill, Maytal Dahan, Ian Fos- of the National Academy of Sciences 111, E4812E4821
ter, Kelly Gaither, Andrew Grimshaw, Victor Hazle- (2014).
wood, Scott Lathrop, Dave Lifka, Gregory D. Peterson, [44] Kun Zhao and Thomas G. Mason, Frustrated rotator
Ralph Roskies, J. Ray Scott, and Nancy Wilkens-Diehr, crystals and glasses of brownian pentagons, Physical Re-
XSEDE: Accelerating Scientific Discovery, Computing view Letters 103, 1316 (2009).
in Science & Engineering 16, 6274 (2014). [45] Tobias Bauert, Leo Merz, Davide Bandera, Manfred
[34] R. Eppenga and D. Frenkel, Monte Carlo study of Parschau, Jay S. Siegel, and Karl-Heinz Ernst, Building
the isotropic and nematic phases of infinitely thin hard 2D Crystals from 5-Fold-Symmetric Molecules, Jour-
platelets, Molecular Physics 52, 13031334 (1984). nal of the American Chemical Society 131, 34603461
[35] Paul E. Brumby, Andrew J. Haslam, Enrique de Miguel, (2009).
and George Jackson, Subtleties in the calculation of the [46] Y Limon Duparcmeur, A Gervois, and J P Troadec,
pressure and pressure tensor of anisotropic particles from Crystallization of pentagon packings, Journal of
volume-perturbation methods and the apparent asym- Physics: Condensed Matter 7, 34213430 (1995).
metry of the compressive and expansive contributions,
Molecular Physics 109, 169189 (2011).
[36] Walter Mickel, Sebastian C. Kapfer, Gerd E. Schroder-
Turk, and Klaus Mecke, Shortcomings of the bond ori-

Вам также может понравиться