Entropy of fully packed hard rigid rods on -dimensional hyper-cubic lattices
Abstract
We determine the asymptotic behavior of the entropy of full coverings of a square lattice by rods of size and , in the limit of large . We show that full coverage is possible only if at least one of and is a multiple of , and that all allowed configurations can be reached from a standard configuration of all rods being parallel, using only basic flip moves that replace a square of parallel horizontal rods by vertical rods, and vice versa. In the limit of large , we show that the entropy per site tends to , with . We conjecture, based on a perturbative series expansion, that this large- behavior of entropy per site is super-universal and continues to hold on all -dimensional hyper-cubic lattices, with .
I Introduction
Systems of particles with only hard core interactions between them have been studied as prototypical models for phase transitions in equilibrium statistical mechanics as well as for understanding aspects of non-equilibrium statistical mechanics. In equilibrium statistical mechanics, hard sphere systems serve as minimal models of solid to fluid transition in molecular solids [1, 2, 3], and in colloidal crystals [4]. Dimer models are equivalent to the Ising model, and anisotropic hard particles can effectively model different phases and phase transitions in liquid crystals [5, 6, 7, 8, 9, 10]. In non-equilibrium statistical mechanics, hard core models like symmetric or asymmetric exclusion processes provide basic models for driven systems and jamming in granular systems [11, 12, 13].
Lattice models of hard-core particles have been of particular interest, as they are analytically more tractable. The phases of assemblies of particles of many different shapes have been studied. Examples include squares [14, 15, 16, 17, 18, 19], triangles [20], hexagons [21], long rods [22, 23, 24], rectangles [25, 26, 27], Y-shaped molecules [28, 29, 30], tetraminoes [31], lattice gases with exclusion upto th nearest neighbors [32, 33, 34, 35, 36, 37], cubes [38], plates [39], etc. An analytical exact solution has been possible only for the case of hard hexagons so far [21]. Phase transitions have also been studied in mixtures of different shapes, for example squares and dimers [17, 40], rods of different lengths [41, 42], polydispersed spheres [43], etc. For the mixture of squares and dimers, it was shown that the critical exponents of the order-disorder transition depends continuously on the relative concentration of the components. Despite a long history, many basic questions about these systems remain open; for example, for a given shape of particles, what are the possible ordered phases, and in which sequence will they appear on increasing the density?
System of hard rods/cylinders have attracted a lot of interest, starting with the pioneering work of Onsager, who showed that a system of thin, long cylinders in three dimensional continuum undergo a phase transition from a disordered phase to an orientationally ordered nematic phase [45]. The study of lattice models of linear hard rods (-mers) started with the work of Flory [22] and Zwanzig [44]. On a -dimensional hyper-cubic lattice, rods can only orient in one of the directions. It was realized in Ref. [23], based on Monte Carlo simulations and high density expansions, that nematic order is present at intermediate densities for large enough , and that the lattice model at high densities must undergo a second disordering transition at a critical density for large , when the nematic order is lost. Usual Monte Carlo techniques with local moves are rather inefficient in sampling states at high-density due to high rates of rejection of moves due to jamming, but recently-introduced strip-update Monte Carlo technique has made it possible to reach densities within a few percent of maximum packing density [46, 47]. Using these techniques, it is found that on the square lattice, for , there is no phase transition, but for , as density is increased, there are three phases: the low-density disordered phase, intermediate-density nematic phase, and the high-density phase in which there is no long ranged positional or orientational order [47]. The existence of the transition may be rigorously proved [48]. The first phase transition belongs to the Ising [49, 50, 51] or three-state Potts universality classes [49, 50, 52] depending on whether the rods are on a square or triangular/honeycomb lattice. The nature of the second transition is not so clear. There is some indication of the high density phase having power law correlations [47] with the second transition not being in the Ising universality class [47, 53], while the exact solution of soft repulsive rods on a tree-like lattice [54] suggests otherwise. More recently, the transitions in two dimensions have been studied using measures such as the classical ‘entanglement’ entropy, mutability, Shannon entropy and data compression [55, 53, 56].
In three dimensions, there is no phase transition for . For , the system undergoes phase transitions from disordered to nematic to a layered disordered phase as density is increased. In the layered disordered phase, the system breaks up into very weakly interacting two dimensional planes within which the rods are disordered. For , there is no nematic phase, and a single phase transition from a disordered to a layered disordered phase [57, 58].
In this paper, we focus on the fully packed limit of linear rods on the square lattice and give heuristic arguments to extend the results to higher dimensions. In particular, we focus on the entropy per site . We note that, in addition to understanding phase transitions in lattice systems of rods with finite density of vacancies, the study of the fully packed phase is relevant for other physical systems. For instance, it would help understand the tetratic order in self assembly of squares and rectangles in the continuum [59]. It may also help in our understanding of phase transitions in other strongly correlated systems. For example, the binding unbinding transition of quarks as a function of the density of hadrons in nuclear matter, studied in QCD, is similar to the binding-unbinding transition of species of holes during the transition from the disordered high density phase to the nematic phase of -mers.
For the case of dimers ( - the only case that is exactly solvable [60, 61, 62, 63] - the entropy per site for square lattice is , where is the Catalan’s constant [60]. On square lattice, the orientation-orientation correlation of two dimers separated by a distance decays as a power law for large [64], while on a triangular lattices, these correlations are short-ranged [65]. A review of the method of solution of dimer problems on planar lattices may be found in Ref. [66]. In three dimensions, there is a class of lattices (not cubic lattice) for which an exact solution can be found and the correlations are strictly finite ranged [67], while for the cubic lattice, the orientational correlations decay as a power law [6]. The entropy of fully packed trimer () tilings on square lattice have also been studied [68]. By numerically diagonalising the transfer matrices for strips, the entropy per site was found to be . Much less is known for higher values of . It is known that the tilings admit a vector height field representation [69]. For larger values of , Gagunashvili and Priezzhev obtained an upper bound for the entropy on the square lattice: , where , with being the Catalan’s constant [70]. It is clear that the full-packing constraint induces strong correlations in the orientations of rods, and one would generally expect orientation-orientation correlations to decrease with distance as a power-law.
The full packing constraint severely limits the allowed configurations. One way to generate a large number of such configurations, satisfying all these constraints is to break the system into parallel 2-dimensional layers, and fully pack each layer with rods. Since rods on different layers do not interact, configurations on different layers can be independently generated, giving a large entropy. Indeed, there is evidence from Monte Carlo simulations (the simulations are done not at fully packing, but for densities close to full-packing) that the high density phase of long rods in three dimensions shows two-dimensional layering [57], and our perturbation expansion suggests that, in the fully packed limit, configurations in even higher dimensions would be dominated by layered two-dimensional configurations.
In this paper, we determine the asymptotic behavior of the entropy of the fully packed configurations in the limit of large rod lengths : first in two dimensions, and then generalized to higher dimensions. The number of coverings depends strongly on the boundary conditions imposed. We will consider configurations of a finite rectangular portion of square lattice fully covered by rectangles of size or . Equivalently, we can consider this a lattice model, with all sites covered using straight rigid rods of length . We will call this open boundary conditions. We prove that full coverage in the open boundary case is possible only if at least one of and is a multiple of . All the allowed configurations for this case can be reached from the standard configuration of all horizontal rods, using only basic flip moves that flip a square of parallel horizontal rods by vertical rods, and vice versa. Using rigorous upper and lower bound estimates, we show that , to leading order in , equals with .
Based on a perturbation series expansion, we conjecture that in higher dimensions, the entropy for the fully packed phase, for large , would be dominated by configurations where the rods arrange themselves in stacked two dimensional layers. Thus, we conjecture that the large- behavior of entropy per site is ‘super-universal’, and continues to hold on -dimensional hypercubical lattices for all and
(1) |
independent of .
The remainder of the paper is organized as follows. In Sec. II, we define the problem precisely. We derive some basic properties of the fully packed phase by showing that an rectangle can be completely covered by -mers, only if at least one of or is multiple of and that all full packing configurations on an open rectangle can be obtained from the standard configuration of all horizontal rods by a combination of basic flip moves. In Sec. III, we obtain lower bounds for entropy by solving exactly for the entropy of rods on semi-infinite strips and . These results are generalised to arbitrary strips by considering truncated generating functions. In Sec. IV, we combine the lower bounds for entropy with existing upper bounds to obtain Eq. (1). In Sec. V, we use heuristic arguments based on perturbation theory to support the conjecture that that this result should also hold for all -dimensional hypercubical lattices with . Section VI contain some concluding remarks.
II Preliminaries
We consider tilings of a rectangle, with , positive integers, by and rectangles (-mers). Each -mer can only be in one of two orientations: horizontal or vertical. An example is shown in Fig. 1 for the case . Equivalently, we can consider this a lattice model, with all sites covered using straight rigid rods of length . Let be the number of such tilings.

II.1 Divisibility of , by
We first show that is non-zero, if and only if at least one of and is divisible by . The ‘if’ part is trivial. For the other part, clearly has to be a multiple of , for full coverage. We now argue that in this case, at least one of and has to be a multiple of .
Assign one of the colors, called here to each of the squares of the lattice, with square given color . The coloring of the squares for the case is shown in Fig. 2. Then each -mer covers exactly one square of each color. Let , , with . Divide the rectangle into three smaller rectangles of sizes and , as shown in Fig. 3. Then, clearly the rectangles of size , and can be covered by -mers, implying that the number of squares of different colors in these two rectangles are equal. However, the small rectangle of size has squares of same color along the diagonal. To cover them would require at least rods, with total area . Equating this to the total area , we obtain . This contradicts the assumption that . Hence, the rectangle can not be fully covered by -mers, unless either or is divisible by .


For simplicity of presentation, in the following, we shall assume that both and are multiples of .
II.2 Ergodicity of the flip moves
In this subsection, we show that all configurations of rods can be reached from any configurations by just using the flip move (defined below).
We define the standard tiling configuration of rectangle by -mers as one using only horizontal -mers. A basic flip move is defined as replacing a square filled with vertical -mers by one with horizontal -mers, and vice versa, as illustrated in Fig. 4.

A combination of two flip moves defines a ‘slide’ move, where a vertical -mer next to a flippable square exchanges position (see Fig. 5), and the vertical -mer will be said to slide across the flippable square.

We now argue that that any full tiling of rectangle by -mers may be reached from the standard configuration by using only the basic flip and slide moves.
Proof: Look at the lowest row. If it consists of only horizontal -mers, then we ignore this row, and the problem reduces to one with a smaller . Else, it would have horizontal -mers, and vertical -mers. In Fig. 6, we have shown an example of a -mer tiling of a rectangle, where , . We move to the left any block of horizontal flippable rods we find between these vertical -mers, using the slide move, and make the vertical rods closer to each other. If now there is any block of consecutive vertical -mers, we can flip these to horizontal, and reduce the problem to one with fewer number of vertical -mers.

If there is no such horizontal flippable block of rods, we look at the bottom row. Let us say that it has segments of horizontal rods, interspersed with vertical rods. [In Fig. 6, there are segments, with , , , ]. Clearly, these are bordered by vertical -mers at the ends, unless the segment itself is at the end of the rectangle. Then we look at the sub-rectangles of sizes made up of these segments and bounded by vertical boundaries. In the example shown in Fig. 6, these rectangles are shown with orange boundaries.
We now argue that there will be a flippable block within each of these small rectangles. This is clear if the width of the rectangle is exactly . Then the sites just above can only be covered by a horizontal rod, or vertical rods. In the latter case, it forms a vertical flippable rectangle. If not, then eventually, we will have horizontal rods just above each other, and form a horizontal flippable rectangle.
If the width is greater than , and the row just above is not made of all horizontal rods, then it will be made up of a number of horizontal segments, separated by vertical rods. And we can repeat the argument with this smaller set. This process can not continue for ever, as the total width is finite, and the width decreases at each step.
Thus, we will be able to find a flippable box at each stage, and eventually, the number of vertical rods becomes zero, and the standard tiling of all horizontal -mers is reached. Since all moves are reversible, and any valid configuration of full-packing can be changed to standard configuration, we can go from any full packing configuration on the rectangle to any other using only flip moves.
III Lower bound for entropy for large
We first show that at full packing, there is a finite entropy per site. We divide the lattice into squares. There are such squares, and each can be tiled in two ways, independent of the others. Then the total number of such tilings is (see Fig. 7). Of course, more complicated tilings are possible, as shown in Fig. 1, and the above only provides a lower bound. We define entropy per site
(2) |
Then, .

III.1 Entropy of strips
We can easily obtain a better lower bound on . Break the lattice in strips of width each. Let be denoted by . Since by breaking into strips of width , we disallow configurations where rods cross the boundary, leading to undercounting, we obtain the inequality
(3) |
obeys a simple recursion relation. Consider the packing of a rectangle. The first row can be covered by a horizontal -mer (reducing by one) or the first square can be covered parallel vertical -mers (reducing by ). Thus, ’s satisfy the recursion relation
(4) |
This implies that increases as where is the largest root of the equation
(5) |
For large , to leading order . A little bit of algebra shows that In the limit of large , the subleading terms take the form
(6) |
where is the solution of the equation
(7) |
The function is called the Lambert function [71]. To leading order, for large (to see this, take logarithm on both sides of Eq. (7) and compare the terms of leading order). The sub-leading term can be similarly obtained to give for large
(8) |
with corrections that only grow slower than . Thus we obtain
(9) |
The entropy per site for the strip is . Thus,
(10) |
Since is a lower bound for , we obtain the leading behavior:
(11) |
III.2 Entropy of strips
In this subsection, we describe the exact calculation of the entropy of tilings of the semi-infinite stripe with -mers, where the -coordinate is , and the -coordinate lies in the range . We define the generating function as the sum over all covering of rectangles of size , summed over all positive integer values of , where the weight of a covering with tiles is . Then, we have
(12) |
We also define a partial covering of the strip with rods to the bottom of some reference line , so that no site with -coordinate less than is left uncovered (see Fig. 8), and all rods must cover at least one site with -coordinate less than . Clearly, all rods that do not lie completely to the bottom of the must be vertical. A partial covering is a rectangular covering iff no site with -coordinate larger than is covered.

A partial covering may be characterized by its top boundary , for to , where specifies how many sites to the top of the reference line are covered in the column with coordinate . We will choose to be as large as possible, so that at least one of the ’s has to be zero, and , for all . For example, the boundary of the configuration shown in Fig. 8 is specified by . In a more compact notation, we will write this as .
Not all height configurations are allowed. A bit of thought shows that for a partial covering of the stripe, the only allowed height configurations are , , , or , with and taking values from to .
We define the generating functions as the generating function of all possible ways of completing a partial tilings with a given height profile , where the completed covering is rectangular, and the weight of tiling in which we add extra rods is . Therefore, for example,
(13) | |||||
(14) |
Consider a particular height configuration . We can write recursion equations for the corresponding generating function , by considering all possible ways of filling the column of sites immediately to the top of the reference line by -mers, such that the top edge of the full tiling is horizontal, and no sites are left uncovered.
For example, it is easily seen that (see Fig. 9)
The different terms in this equation correspond to the cases where the next row is left empty, or filled by two horizontal rods, or by vertical rods, or by first vertical rods, then a horizontal rod, then vertical rods.

Writing such generating functions for all possible height configurations, we obtain a set of inhomogeneous linear equations in approximately variables. This may be written as a transfer matrix of dimension . However, using the symmetries of the problem, this number can be considerably reduced.
We note that the recursion equations for the generating function is
(16) |
This equation has no -dependence. Hence, we may expect that is independent of . It can be checked that this ansatz is consistent with the remaining recursion equations. Similarly, we find that is also independent of . With this simplification, the number of independent variables reduces to approximately .
The remaining recursion equations are easily written down, we obtain for all ,
(17) |
and
(18) |
Substituting for in Eq. (16) from Eq. (17), we obtain
(19) |
which is immediately solved to give
(20) |
Substituting for in Eq. (III.2) from Eq. (20) and simplifying, we obtain
(21) |
To close the equations, we have to determine in terms of . The values of for one value of are related by Eq. (18) to arguments and to . This seems complicated, but it is easily checked that the ansatz
(22) |
satisfies Eq. (18), so long as
(23) |
Eliminating from Eq (23), we obtain
(24) |
This is a quadratic equation in , and determines for any given value of . Explicitly, we obtain
(25) |
Using Eq. (23), we can express and in terms of a single variable :
(26) | |||
(27) |
The actual values of and can be determined from the boundary condition at :
(28) |
We obtain
(29) |
Substituting for in Eqs. (26) and (27), we obtain and in terms of .
(30) | |||
(31) |
Finally, substituting the value of and in Eq. (22), we obtain
(32) |
Equation (32) may be simplified by substituting for from Eq. (24):
(33) | |||
Note the explicit symmetry of the expression under the exchange of .
Substituting the expressions for , in Eq. (21), we obtain an explicit expression for of the form
(34) |
where the denominator equals
(35) |
The entropy is given by , where is the singularity of that is closest to the origin. We will show below that asymptotic behavior of entropy for strips of width is the same as that of strips of width . The explicit values of the entropies for strips and for upto are given in Table 1, and compared with the asymptotic result in Eq. (1).
We now determine the leading singularity of in the limit . To do so, consider the denominator . It has a square root singularity at when the discriminant in Eq. (25) equals zero. By factorising the discriminant and writing in terms of the modulus of , we obtain that satisfies the equation
(36) |
identical to that satisfied by for the strip [see Eq. (5) with ]. For large , it has the solution
(37) |
We now show that has a zero at . Following a bit of algebra, it can be shown that
(38) |
Clearly, , as . At the same time, it is clear that since . Therefore, there must be a zero in , leading to a higher entropy for strips , as evident in Table 1. Also as for large , we expect that for large .
We now show that the leading behavior of and are identical for large . We look for solutions
(39) |
where for any , and for . In this limit, , and after some algebra, may be simplified to give
(40) |
Equating to zero, we obtain
(41) |
From direct substitution, it is straightforward to check that Eq. (41) is satisfied to leading order by . Equation (39) then gives
(42) |
Since the entropy is given by , we obtain
(43) |
The leading behaviour of coincides with that of [see Eq. (10)]. The subleading term is different. We thus obtain the same lower bound as given in Eq. (11).
III.3 Entropy of strips
For general , though one can write down the recursion relations obeyed by different generating functions, it is not possible to find a closed form solution for them. Here, we provide an alternate analysis by determining a lower bound for the asymptotic behaviour of entropy of strips, which will happen to coincide with the exact results for the strips and
We define the generating function
(44) |
with , by convention. Then, by direct enumeration,
(45) | |||||
(46) |
is sum of weights of all configurations of rods on a semi-infinite strip of width , with the weight of a configuration of rods being . We have the further constraint that all rods must lie fully in a rectangular region, with both bottom and top edge horizontal, and no uncovered regions within the rectangle. As a slightly more complicated example, it is easily seen that is the generating function for strips, and hence,
(47) |
In determining admissible tilings, a useful concept is that of concatenation. Given two tilings of rectangles of sizes and , we define the vertical concatenation of these tilings as the tiling of size , obtained by just putting the rectangles on top of each other in the order of concatenation. An illustration of vertical concatenation of three tilings is shown in Fig. 10. A horizontal concatenation is defined similarly.

A tiling is said to be vertically indecomposable, if it cannot be expressed as a vertical concatenation of two admissible tilings. For a vertically decomposable tiling, there is a horizontal line that divides the rectangle into two smaller tilings, such that no rod crosses the horizontal line.
We now define as the sum of weights of all vertically indecomposable tilings of rectangles of width . Clearly, is a series in powers of , with all coefficients as non-negative integers. Then, we have
(48) |
Let the radius of convergence of the power series of be . Then,
(49) |
Since is a series of positive coefficients, we may truncate the series at any order, and obtain an upper bound estimate of , and hence of , the limit of , for large . This, in turn, will provide a lower bound for the entropy.
We will take to be a multiple of , as these give the best bounds. The simplest case is . In this case, is a finite polynomial, and we have
(50) |
We see that this is consistent with Eq. (47).
Now, let us consider the more complicated case . In this case, is not a finite polynomial. But it has an interesting structure: the lowest order term is , corresponding to a configuration of two horizontal k-mers side by side. But, then terms of order are all zero, as the corresponding tilings are decomposable. The first non-zero term is of order , which corresponds to configurations consisting of a plaquette of aligned horizontal rods and vertical rods tiling the remaining area, and another of vertical -mers. With a small amount of brute force enumeration, it is easily seen that the plaquette can be placed in ways to give
(51) |
If we truncate the equation at order , we obtain an upper bound estimate for . It turns out that for large , the terms that have been dropped make only a negligible contribution to at . We will verify this claim later. First, we solve the truncated equation for :
(52) |
Writing , we see that , if is large, to leading order in , satisfies the equation
(53) |
which has the solution [ being the Lambert function], which for large has the leading behavior
(54) |
Since , we obtain
(55) |
This is a bit larger than the estimate using strips of width [see Eq. (10)], but for large , the leading behavior remains the same with the difference showing only in the sub-leading correction of order . We also note that the bound for obtained by truncation coincides with the exact analysis [see Eq. (43)].
Now, the term of order in is only , and its contribution to sum is smaller than that of the term of order by a factor . At higher orders, the term of order has a coefficient of order . Using the fact that is of order , the net contribution of this term decreases as . This also does not change the leading order -dependence of .
A similar argument works for other values of . We will only sketch the arguments here. The series expansion for in powers of is of the form
(56) |
Here is an -dependent coefficient. The leading contribution to this term comes from configurations depicted in Fig. 11, consisting of of plaquettes of aligned horizontal -mers, interspersed with vertical rods. The number of such configurations is . Thus, keeping only the first two non-trivial terms in the expansion for , we write
(57) |
Solving this equation, we see that its smallest positive root has the leading -dependence given by
(58) |
Since the entropy , we obtain that
(59) |
We conclude that
(60) |

IV Upper bound for entropy for large and the main result
Gagunashvili and Priezzhev obtained an upper bound for [70]. They considered a subset of sites of the square lattice whose coordinates are multiple of , and assumed that we are given the configuration of -mers that cover these sites. Then, they proved that there is at most one way to cover the remaining sites with -mers. Then, the number of coverings allowed is bounded from above by the number of ways the subset of sites can be covered by -mers. But each of these can be covered in at most ways. Since there are at most such sites, they obtain
(61) |
This implies that , or equivalently
(62) |
In fact, Gagunashvili and Priezzhev proved a stronger upper bound which for large is , where , with being the Catalan’s constant. Numerically, . However, the weaker bound is adequate for our purpose here
We now combine the lower bound obtained for entropy in Eqs. (11) and (60), and the upper bound obtained for entropy in Eq. (62). Since, these two bounds are the same, we conclude that the entropy for fully packed rods on a square lattice has the asymptotic behavior
(63) |
as given in Eq. (1), thus proving our main result.
We now look at how the bounds converge to the asymptotic result. The entropies on the strips and provide lower bounds for the entropy on infinite lattices. Ref. [70] gives an upper bound for large as , where . Since the leading form is the same for both the upper bounds as well as lower bounds, we divide it out by considering , which converges to for large . The strips provide lower bounds while provides an upper bound for this quantity. These bounds are shown in Fig. 12.

In addition, it is also possible to put a bound on the subleading corrections to the entropy. Let be the coefficient of the subleading term in the asymptotic expansion of the entropy such that
(64) |
From Eq. (59), by taking the limit , it follows that
(65) |
V Extension to higher dimensions
We now present a heuristic argument that extends the above result to higher dimensions . For a system of -mers on a -dimensional hypercubical lattice, with , we argue below that configurations of rods that are fully layered provide a good starting point to calculate the entropy of the system. In fact, this approach becomes better for larger and one can develop a series expansion in the number of rods that are between the layered planes. Then, in a typical state, there are only a few such rods, and the full state will show spontaneous symmetry breaking, with most of the rods in the configuration being one of the orientations. In this way, one obtains the full packing constraint satisfied within a 2-dimensional layer, and different layers can be occupied independently, leading to a large entropy. We note that the existence of layering in the high-density phase has been seen in simulations at densities close to full packing for rods of length larger than or equal to five in [57].
We first consider the case . The argument is easily extended to higher .
We will use an adaption of the series expansion technique developed in the context of hard square lattice gases [15, 14, 16] and hard rectangle lattice gases [72]. For the simple cubic lattice, we consider different activities for rods oriented along the -, -, and - directions. Let the corresponding partition function for an lattice be denoted by .
We will consider a perturbation expansion of this in powers of . We start with the case , and . The the grand partition function of a cuboid can be written as product of -dimensional partition functions
(66) |
where is the partition function of a full packing of layer by -mers. We write as a perturbation expansion in about . If this series expansion is well behaved, this implies that for large enough , the full packed configuration will show spontaneous symmetry breaking, and for large , the fraction of non-planar rods in a random full planar configuration will tend to zero. The first term in must be proportional to so that the entropy is extensive. We, thus, write the expansion as
(67) |
Taking logarithm and dividing by , we obtain the expansion for the entropy:
(68) |
and we assume that this systematic expansion is well behaved, and converges for small .
The first nontrivial term in Eq. (67) is proportional to , and the coefficient is determined in terms of the number of configurations of -type rods (to be also called vertical rods), with the rest of the rods being of the - and - type. These vertical rods will have to be in the same vertical slab of height . Let the - and -coordinates of the lowest point of these vertical rods , to . Let the number of possible coverings of rods in one plane, given unoccupied sites , be . Then the number of coverings of the cuboid is proportional to , and the relative weight of this term will be . We note that may be considered as proportional to the probability distribution of the bound state of holes. We then sum over different possible . We expect to be less than 1, and to decrease as a power law of the distances between , but with a power large enough so that
(69) |
where the primed sum is over , with fixed. Then the sum over and gives a factor proportional to .
This is a complicated problem, for which we do not know the exact closed form expression. However, we note that each term decreases exponentially with for large since we expect . We note that would be expected to be largest, when the holes are near each other. In fact, the closest they can be be is in a continuous single line of points, which may be created by removing a single rod from the -covering. In this case, the contribution of the term is .
If we sum to all orders, the dominant contribution will be expected to be of the same form. We thus conclude that for large ,
(70) |
Then , as a function of , and expand in powers of , order by order, each term in the perturbation series will give an exponentially small contribution in the large- limit.
Taking derivative of with respect to , we obtain the fractional number of rods in the -direction at full packing in this ensemble, and we see that this fraction tends to zero as increases, and the departure from perfect layering decreases for larger .
The argument is immediately extended to higher , and leads us to the conjecture in Eq. (1).
VI Concluding remarks
In this paper, we studied the tiling of a finite rectangular part of the plane by rectangles of size and . We showed that in order to get a full coverage, one of the sides of the rectangle to be covered should a multiple of . We also studied the structure of the tilings of rectangles, and showed that all tilings can be obtained from each other by a sequence of basic flip move that exchanges a small small square made of parallel vertical rectangles into horizontal ones, and vice versa. We also showed provided non-rigorous perturbation theory based arguments for the conjecture that , the entropy per site for -mers on a -dimensional hypercubical lattice covered by straight rods of length , for all satisfies Eq. (1). We emphasize that while the perturbation theory argument seems quite plausible, there is no proof that such a perturbation expansion is convergent. If the series expansion does not converge, or converges to a wrong value, the argument given here would break down.
The fact that this limit is independent of dimension deserves some comment. In general, we would expect the coefficients of logarithms encountered in the study of critical phenomena to be ‘universal’, because by definition, they do not change under a change of length scale in a renormalization transformation. But, such coefficients are in general not dimension independent. In fact, here, has a multiplying factor , which indicates that the relevant quantity is the number of allowed configurations per unit square of length (which is the natural length scale in the problem). This number is proportional to , and the entropy is proportional to . The fact that this is independent of dimension is only reflecting the fact that for large , the problem essentially reduces a two-dimensional problem, because of spontaneous symmetry breaking, and most of the configurations at full packing are the ones where the system breaks up into disjoint two-dimensional layers. Consequently, for large , the leading behavior of entropy in higher dimensions is same as the two-dimensional case.
VII Acknowledgments
DD’s work was partially supported by the grant DST-SR-S2/JCB-24/2005 of the Government of India.
References
- Alder and Wainwright [1957] B. J. Alder and T. E. Wainwright, Phase transition for a hard sphere system, J. Chem. Phys. 27, 1208 (1957).
- Alder and Wainwright [1962] B. J. Alder and T. E. Wainwright, Phase transition in elastic disks, Phys. Rev. 127, 359 (1962).
- Noya et al. [2008] E. G. Noya, C. Vega, and E. de Miguel, Determination of the melting point of hard spheres from direct coexistence simulation methods, J. Chem. Phys. 128, 154507 (2008).
- Pusey and van Megen [1986] P. N. Pusey and W. van Megen, Phase behaviour of concentrated suspensions of nearly hard colloidal spheres, Nature 320, 340 (1986).
- Fisher [1966] M. E. Fisher, On the dimer solution of planar ising models, J. Math. Phys. 7, 1776 (1966).
- Huse et al. [2003] D. A. Huse, W. Krauth, R. Moessner, and S. L. Sondhi, Coulomb and liquid dimer models in three dimensions, Phys. Rev. Lett. 91, 167004 (2003).
- Möessner and Sondhi [2003] R. Möessner and S. L. Sondhi, Ising and dimer models in two and three dimensions, Phys. Rev. B 68, 054405 (2003).
- de Gennes and Prost [1995] P. G. de Gennes and J. Prost, The physics of liquid crystals, Vol. 83 (Oxford university press, 1995).
- Frenkel [1988] D. Frenkel, Structure of hard-core models for liquid crystals, J. Phys. Chem. 92, 3280 (1988).
- Care and Cleaver [2005] C. Care and D. Cleaver, Computer simulation of liquid crystals, Rep. Prog. Phys. 68, 2665 (2005).
- Donev et al. [2004] A. Donev, S. Torquato, F. H. Stillinger, and R. Connelly, Jamming in hard sphere and disk packings, J. Appl. Phys. 95, 989 (2004).
- Liggett [2012] T. M. Liggett, Interacting particle systems, Vol. 276 (Springer Science & Business Media, 2012).
- Mallick [2015] K. Mallick, The exclusion process: A paradigm for non-equilibrium behaviour, Physica A 418, 17 (2015).
- Bellemans and Nigam [1966] A. Bellemans and R. K. Nigam, Phase transitions in the hard-square lattice gas, Phys. Rev. Lett. 16, 1038 (1966).
- Bellemans and Nigam [1967] A. Bellemans and R. K. Nigam, Phase transitions in two‐dimensional lattice gases of hard‐square molecules, J. Chem. Phys. 46, 2922 (1967).
- Ramola and Dhar [2012] K. Ramola and D. Dhar, High-activity perturbation expansion for the hard square lattice gas, Phys. Rev. E 86, 031135 (2012).
- Ramola et al. [2015] K. Ramola, K. Damle, and D. Dhar, Columnar order and ashkin-teller criticality in mixtures of hard squares and dimers, Phys. Rev. Lett. 114, 190601 (2015).
- Nath et al. [2016] T. Nath, D. Dhar, and R. Rajesh, Stability of columnar order in assemblies of hard rectangles or squares, Eur. Phys. Lett. 114, 10003 (2016).
- Mandal et al. [2017] D. Mandal, T. Nath, and R. Rajesh, Estimating the critical parameters of the hard square lattice gas model, J. Stat. Mech. 2017, 043201 (2017).
- Verberkmoes and Nienhuis [1999] A. Verberkmoes and B. Nienhuis, Triangular trimers on the triangular lattice: An exact solution, Phys. Rev. Lett. 83, 3986 (1999).
- Baxter [1980] R. J. Baxter, Hard hexagons: exact solution, J. Phys. A 13, L61 (1980).
- Flory [1956] P. J. Flory, Phase equilibria in solutions of rod-like particles, Proc. Roy. Soc. A 234, 73 (1956).
- Ghosh and Dhar [2007] A. Ghosh and D. Dhar, On the orientational ordering of long rods on a lattice, Eur. Phys. Lett. 78, 20003 (2007).
- Dhar et al. [2011] D. Dhar, R. Rajesh, and J. F. Stilck, Hard rigid rods on a bethe-like lattice, Phys. Rev. E 84, 011140 (2011).
- Kundu and Rajesh [2014] J. Kundu and R. Rajesh, Phase transitions in a system of hard rectangles on the square lattice, Phys. Rev. E 89, 052124 (2014).
- Kundu and Rajesh [2015a] J. Kundu and R. Rajesh, Phase transitions in systems of hard rectangles with non-integer aspect ratio, Eur. Phys. J. B 88, 133 (2015a).
- Kundu and Rajesh [2015b] J. Kundu and R. Rajesh, Asymptotic behavior of the isotropic-nematic and nematic-columnar phase boundaries for the system of hard rectangles on a square lattice, Phys. Rev. E 91, 012105 (2015b).
- Szabelski et al. [2013] P. Szabelski, W. Rżysko, T. Pańczyk, E. Ghijsens, K. Tahara, Y. Tobe, and S. De Feyter, Self-assembly of molecular tripods in two dimensions: structure and thermodynamics from computer simulations, RSC Adv. 3, 25159 (2013).
- Ruth et al. [2015] D. Ruth, R. Toral, D. Holz, J. Rickman, and J. Gunton, Impact of surface interactions on the phase behavior of y-shaped molecules, Thin Solid Films 597, 188 (2015).
- Mandal et al. [2018] D. Mandal, T. Nath, and R. Rajesh, Phase transitions in a system of hard y-shaped particles on the triangular lattice, Phys. Rev. E 97, 032131 (2018).
- Barnes et al. [2009] B. C. Barnes, D. W. Siderius, and L. D. Gelb, Structure, thermodynamics, and solubility in tetromino fluids, Langmuir 25, 6702 (2009).
- Panagiotopoulos [2005] A. Z. Panagiotopoulos, Thermodynamic properties of lattice hard-sphere models, J. Chem. Phys. 123, 104504 (2005).
- Fernandes et al. [2007] H. C. M. Fernandes, J. J. Arenzon, and Y. Levin, Monte carlo simulations of two-dimensional hard core lattice gases, J. Chem. Phys. 126, 114508 (2007).
- Nath and Rajesh [2014] T. Nath and R. Rajesh, Multiple phase transitions in extended hard-core lattice gas models in two dimensions, Phys. Rev. E 90, 012120 (2014).
- Nath and Rajesh [2016] T. Nath and R. Rajesh, The high density phase of the k -nn hard core lattice gas model, J. Stat. Mech. 2016, 073203 (2016).
- Akimenko et al. [2019] S. S. Akimenko, V. A. Gorbunov, A. V. Myshlyavtsev, and P. V. Stishenko, Tensor renormalization group study of hard-disk models on a triangular lattice, Physical Review E 100, 022108 (2019).
- Thewes and Fernandes [2020] F. C. Thewes and H. C. M. Fernandes, Phase transitions in hard-core lattice gases on the honeycomb lattice, Phys. Rev. E 101, 062138 (2020).
- Vigneshwar et al. [2019] N. Vigneshwar, D. Mandal, K. Damle, D. Dhar, and R. Rajesh, Phase diagram of a system of hard cubes on the cubic lattice, Phys. Rev. E 99, 052129 (2019).
- Disertori et al. [2020] M. Disertori, A. Giuliani, and I. Jauslin, Plate-nematic phase in three dimensions, Commun. Math. Phys. 373, 327 (2020).
- Mandal and Rajesh [2017] D. Mandal and R. Rajesh, Columnar-disorder phase boundary in a mixture of hard squares and dimers, Phys. Rev. E 96, 012140 (2017).
- Kundu et al. [2016] J. Kundu, J. F. Stilck, and R. Rajesh, Phase diagram of a bidispersed hard-rod lattice gas in two dimensions, Eur. Phys. Lett. 112, 66002 (2016).
- Stilck and Rajesh [2015] J. F. Stilck and R. Rajesh, Polydispersed rods on the square lattice, Phys. Rev. E 91, 012106 (2015).
- Rodrigues and Oliveira [2019] N. T. Rodrigues and T. J. Oliveira, Three stable phases and thermodynamic anomaly in a binary mixture of hard particles, J. Chem. Phys. 151, 024504 (2019).
- Zwanzig [1963] R. Zwanzig, First‐order phase transition in a gas of long thin rods, J. Chem. Phys. 39, 1714 (1963), http://dx.doi.org/10.1063/1.1734518 .
- Onsager [1949] L. Onsager, The effects of shape on the interaction of colloidal particles, Ann. N. Y. Acad. Sci. 51, 627 (1949).
- Kundu et al. [2012] J. Kundu, R. Rajesh, D. Dhar, and J. F. Stilck, A monte carlo algorithm for studying phase transition in systems of hard rigid rods, AIP Conf. Proc. 1447, 113 (2012).
- Kundu et al. [2013] J. Kundu, R. Rajesh, D. Dhar, and J. F. Stilck, Nematic-disordered phase transition in systems of long rigid rods on two-dimensional lattices, Phys. Rev. E 87, 032103 (2013).
- Disertori and Giuliani [2013] M. Disertori and A. Giuliani, The nematic phase of a system of long hard rods, Commun. Math. Phys. 323, 143 (2013).
- Matoz-Fernandez et al. [2008a] D. A. Matoz-Fernandez, D. H. Linares, and A. J. Ramirez-Pastor, Determination of the critical exponents for the isotropic-nematic phase transition in a system of long rods on two-dimensional lattices: Universality of the transition, Eur. Phys. Lett. 82, 50007 (2008a).
- Matoz-Fernandez et al. [2008b] D. A. Matoz-Fernandez, D. H. Linares, and A. J. Ramirez-Pastor, Critical behavior of long straight rigid rods on two-dimensional lattices: Theory and monte carlo simulations, J. Chem. Phys. 128, 214902 (2008b).
- Fischer and Vink [2009] T. Fischer and R. L. C. Vink, Restricted orientation ”liquid crystal” in two dimensions: Isotropic-nematic transition or liquid-gas one(?), Eur. Phys. Lett. 85, 56003 (2009).
- Matoz-Fernandez et al. [2008c] D. Matoz-Fernandez, D. Linares, and A. Ramirez-Pastor, Critical behavior of long linear k-mers on honeycomb lattices, Physica A 387, 6513 (2008c).
- Chatelain and Gendiar [2020] C. Chatelain and A. Gendiar, Absence of logarithmic divergence of the entanglement entropies at the phase transitions of a 2d classical hard rod model, Eur. Phys. J. B 93, 134 (2020).
- Kundu and Rajesh [2013] J. Kundu and R. Rajesh, Reentrant disordered phase in a system of repulsive rods on a bethe-like lattice, Phys. Rev. E 88, 012134 (2013).
- Vogel et al. [2017] E. E. Vogel, G. Saravia, and A. J. Ramirez-Pastor, Phase transitions in a system of long rods on two-dimensional lattices by means of information theory, Phys. Rev. E 96, 062133 (2017).
- Vogel et al. [2020] E. E. Vogel, G. Saravia, A. J. Ramirez-Pastor, and M. Pasinetti, Alternative characterization of the nematic transition in deposition of rods on two-dimensional lattices, Phys. Rev. E 101, 022104 (2020).
- Vigneshwar et al. [2017] N. Vigneshwar, D. Dhar, and R. Rajesh, Different phases of a system of hard rods on three dimensional cubic lattice, J. Stat. Mech. 2017, 113304 (2017).
- Gschwind et al. [2017] A. Gschwind, M. Klopotek, Y. Ai, and M. Oettel, Isotropic-nematic transition for hard rods on a three-dimensional cubic lattice, Phys. Rev. E 96, 012104 (2017).
- Wojciechowski and Frenkel [2004] K. Wojciechowski and D. Frenkel, Tetratic phase in the planar hard square system, Comput. Meth. Sci. Technol. 10, 235 (2004).
- Kasteleyn [1961] P. Kasteleyn, The statistics of dimers on a lattice, Physica 27, 1209 (1961).
- Kasteleyn [1963] P. W. Kasteleyn, Dimer statistics and phase transitions, J. Math. Phys. 4, 287 (1963).
- Temperley and Fisher [1961] H. N. V. Temperley and M. E. Fisher, Dimer problem in statistical mechanics-an exact result, Phil. Mag. 6, 1061 (1961).
- Fisher [1961] M. E. Fisher, Statistical mechanics of dimers on a plane lattice, Phys. Rev. 124, 1664 (1961).
- Fisher and Stephenson [1963] M. E. Fisher and J. Stephenson, Statistical mechanics of dimers on a plane lattice. ii. dimer correlations and monomers, Phys. Rev. 132, 1411 (1963).
- Fendley et al. [2002] P. Fendley, R. Moessner, and S. L. Sondhi, Classical dimers on the triangular lattice, Phys. Rev. B 66, 214513 (2002).
- Nagle et al. [1989] J. F. Nagle, C. S. Yokoi, and S. M. Bhattacharjee, Dimer models on anisotropic lattices, Phase transitions and critical phenomena 13, 235 (1989).
- Dhar and Chandra [2008] D. Dhar and S. Chandra, Exact entropy of dimer coverings for a class of lattices in three or more dimensions, Phys. Rev. Lett. 100, 120602 (2008).
- Ghosh et al. [2007] A. Ghosh, D. Dhar, and J. L. Jacobsen, Random trimer tilings, Phys. Rev. E 75, 011115 (2007).
- Kenyon [2000] R. Kenyon, Conformal invariance of domino tiling, Ann. Prob. , 759 (2000).
- Gagunashvili and Priezzhev [1979] N. D. Gagunashvili and V. B. Priezzhev, Close packing of rectilinear polymers on a square lattice, Theor Math Phys 39, 507–510 (1979).
- Corless et al. [1996] R. M. Corless, G. H. Gonnet, D. E. Hare, D. J. Jeffrey, and D. E. Knuth, On the lambert-w function, Advances in Computational mathematics 5, 329 (1996).
- Nath et al. [2015] T. Nath, J. Kundu, and R. Rajesh, High-activity expansion for the columnar phase of the hard rectangle gas, J. Stat. Phys. 160, 1173 (2015).