Skip to main content
Proceedings of the National Academy of Sciences of the United States of America logoLink to Proceedings of the National Academy of Sciences of the United States of America
. 2012 Jun 25;109(31):12369-12374. doi: 10.1073/pnas.1119941109

Surface phonons, elastic response, and conformal invariance in twisted kagome lattices

Kai Sun a, Anton Souslov b,1, Xiaoming Mao b, T C Lubensky b,2
PMCID: PMC3412002  PMID: 22733727

Abstract

Model lattices consisting of balls connected by central-force springs provide much of our understanding of mechanical response and phonon structure of real materials. Their stability depends critically on their coordination number z. d-dimensional lattices with z = 2d are at the threshold of mechanical stability and are isostatic. Lattices with z < 2d exhibit zero-frequency “floppy” modes that provide avenues for lattice collapse. The physics of systems as diverse as architectural structures, network glasses, randomly packed spheres, and biopolymer networks is strongly influenced by a nearby isostatic lattice. We explore elasticity and phonons of a special class of two-dimensional isostatic lattices constructed by distorting the kagome lattice. We show that the phonon structure of these lattices, characterized by vanishing bulk moduli and thus negative Poisson ratios (equivalently, auxetic elasticity), depends sensitively on boundary conditions and on the nature of the kagome distortions. We construct lattices that under free boundary conditions exhibit surface floppy modes only or a combination of both surface and bulk floppy modes; and we show that bulk floppy modes present under free boundary conditions are also present under periodic boundary conditions but that surface modes are not. In the long-wavelength limit, the elastic theory of all these lattices is a conformally invariant field theory with holographic properties (characteristics of the bulk are encoded on the sample boundary), and the surface waves are Rayleigh waves. We discuss our results in relation to recent work on jammed systems. Our results highlight the importance of network architecture in determining floppy-mode structure.

Keywords: auxetic response, self stress, conformal field theory, Cosserat elasticity


Networks of balls and springs or frames of nodes connected by compressible struts provide realistic models for physical systems from bridges to condensed solids. Their elastic properties depend on their coordination number z—the average number of nodes each node is connected to. If z is small enough, the networks have deformation modes of zero energy—they are floppy (16). As z is increased, a critical value, zc, is reached at which springs provide just enough constraints that the system has no zero-energy floppy modes (7) (or mechanisms, ref. 2, in the engineering literature), and the system is isostatic. For z > zc, networks with appropriate geometry (see below) are rigid in the sense that they have no zero modes other than those associated with trivial rigid translations and rotations. If a network with z > zc is homogeneously distributed in space, it can be viewed as an elastic solid whose long-wavelength mechanical properties are described by a continuum elastic energy with nonvanishing elastic moduli (711). The phenomenon of rigidity percolation (12, 13) whereby a sample spanning rigid cluster develops upon the addition of springs is one version of this floppy-to-rigid transition. The coordination numbers of whole classes of systems, including engineering structures (14, 15) (bridges and buildings), randomly packed spheres near jamming (10, 11, 16, 17), network glasses (7, 8), cristobalites (18), zeolites (19, 20), and biopolymer networks (2124) are close enough to zc that their elasticity and mode structure is strongly influenced by those of the isostatic lattice.

Though the isostatic point always separates rigid from floppy behavior, the properties of isostatic lattices are not universal; rather they depend on lattice architecture. Here we explore the unusual properties of a particular class of periodic isostatic lattices derived from the two-dimensional kagome lattice by rigidly rotating triangles through an angle α without changing bond lengths as shown in Fig. 1. The bulk modulus B of these lattices is rigorously zero for all α ≠ 0. As a result, their Poisson ratio acquires its limit value of -1; when stretched in one direction, they expand by an equal amount in the orthogonal direction: They are maximally auxetic (2528). These modes represent collapse pathways (29, 30) of the kagome lattice. Modes of isostatic systems are generally very sensitive to boundary conditions (9, 31, 32), but the degree of sensitivity depends on the details of lattice structure. For reasons we will discuss more fully below, modes of the square lattice, which is isostatic, are in fact insensitive to changes from free boundary conditions (FBCs) to periodic boundary conditions (PBCs), whereas those of the undistorted kagome lattice are only mildly so. The modes of both, however, change significantly when rigid boundary conditions (RBCs) are applied. We show here that, in all families of the twisted kagome lattice, modes depend sensitively on whether FBCs, PCBs, or RBCs are applied: Finite lattices with free boundaries have floppy surface modes that are not present in their periodic or rigid spectrum or in that of finite undistorted kagome lattices. In the long-wavelength limit, the surface floppy modes, which are present in any 2d material with B = 0, reduce to surface Rayleigh waves (33) described by a conformally invariant energy whose analytic eigenfunctions are fully determined by boundary conditions. At shorter wavelengths, the surface waves become sensitive to lattice structure and remain confined to within a distance of the surface that diverges as the undistorted kagome lattice is approached. In the simplest twisted kagome lattice, all floppy modes are surface modes, but in more complicated lattices, including ones with uniaxial symmetry that we construct, there are both surface and bulk floppy modes.

Fig. 1.

Fig. 1.

(A) Section of a kagome lattice with Nx = Ny = 4 and Nc = NxNy three-site unit cells. Nearest-neighbor bonds, occupied by harmonic springs, are of length a. The rotated row (second row from the top) represents a floppy mode. Next-nearest-neighbor bonds are shown as dotted lines in the lower left hexagon. The vectors e1, e2, and e3 indicate symmetry directions of the lattice. The numbers in the triangles indicate those that twist together under PBCs in zero modes along the three symmetry direction. Note that there are only four of these modes. (B) Section of a square lattice depicting a floppy mode in which all sites along a line are displaced uniformly. (C) Twisted kagome lattice, with lattice constant aL = 2a cos α, derived from the undistorted lattice by rigidly rotating triangles through an angle α. A unit cell, bounded by dashed lines, is shown in violet. Arrows depict site displacements for the zone-center (i.e., zero wavenumber) ϕ mode which has zero (nonzero) frequency under free (periodic) boundary conditions. Sites 1, 2, and 3 undergo no collective rotation about their center of mass, whereas sites 1, 2, and 3 do. (D) Superposed snapshots of the twisted lattice showing decreasing areas with increasing α.

Arguments due to Maxwell (1) provide a criterion for network stability: Networks in d dimensions consisting of N nodes, each connected with central-force springs to an average of z neighbors, have Inline graphic zero-energy modes when z < 2d (in the absence of redundant bonds—see below). Of these, a number, Ntr, which depends on boundary conditions, are trivial rigid translations and rotations, and the remainder are floppy modes of internal structural rearrangement. Under FBCs and PBCs, Ntr equals d(d + 1)/2 and d, respectively. With increasing z, mechanical stability is reached at the isostatic point at which N0 = Ntr. The Maxwell argument is a global one; it does not provide information about the nature of the floppy modes and does not distinguish between bulk or surface modes.

Kagome Zero Modes and Elasticity

The kagome lattice of central-force springs shown in Fig. 1A is one of many locally isostatic lattices, including the familiar square lattice in two dimensions (Fig. 1B) and the cubic and pyrochlore lattices in three dimensions, with exactly z = 2d nearest-neighbor (NN) bonds connected to each site not at a boundary. Under PBCs, there are no boundaries, and every site has exactly 2d neighbors. Finite, N-site sections of these lattices have surface sites with fewer than 2d neighbors and of order Inline graphic zero modes. The free kagome lattice with Nx and Ny unit cells along its sides (Fig. 1A) has N = 3NxNy sites, NB = 6NxNy - 2(Nx + Ny) + 1 bonds, and N0 = 2(Nx + Ny) - 1 zero modes, all but three of which are floppy modes. These modes, depicted in Fig. 1A, consist of coordinated counterrotations of pairs of triangles along the symmetry axes e1, e2, and e3 of the lattice. There are Nx modes associated with lines parallel to e1, Ny associated with lines parallel to e3, and Nx + Ny - 1 modes associated with lines parallel to e2.

In spite of the large number of floppy modes in the kagome lattice, its longitudinal and shear Lamé coefficients, λ and μ, and its bulk modulus B = λ + μ are nonzero and proportional to the NN spring constant k: Inline graphic and Inline graphic. The zero modes of this lattice can be used to generate an infinite number of distorted lattices with unstretched springs and thus zero energy (30). We consider only periodic lattices, the simplest of which are the twisted kagome lattices obtained by rotating triangles of the kagome unit cell through an angle α as shown in Fig. 1 C and D (30, 34). These lattices have C3v rather than C6v symmetry and, like the undistorted kagome lattice, three sites per unit cell. As Fig. 1D shows, the lattice constant of these lattices is aL = 2a cos α, and their area Aα decreases as cos2α as α increases. The maximum value that α can achieve without bond crossings is π/3, so that the maximum relative area change is Aπ/3/A0 = 1/4. Because all springs maintain their rest length, there is no energy cost for changing α and, as a result, B is zero for every α ≠ 0, whereas the shear modulus Inline graphic remains nonzero and unchanged. Thus, the Poisson ratio σ = (B - μ)/(B + μ) attains its smallest possible value of -1. For any α ≠ 0, the addition of next-nearest-neighbor (NNN) springs, with spring constant k (or of bending forces between springs) stabilizes zero-frequency modes and increases B and σ. Nevertheless, for sufficiently small k, σ remains negative. Fig. 2 shows the region in the k - α plane with negative σ.

Fig. 2.

Fig. 2.

Phase diagram in the α - k plane showing region with negative Poisson ratio σ.

Kagome Phonon Spectrum

We now turn to the linearized phonon spectrum of the kagome and twisted kagome lattices subjected to PBCs. These conditions require displacements at opposite ends of the sample to be identical and thus prohibit distortions of the shape and size of the unit cell and rotations but not uniform translations, leaving two rather than three trivial zero modes. The spectrum (35) of the three lowest frequency modes along symmetry directions of the undistorted kagome lattice with and without NNN springs is shown in Fig. 3A. When k = 0, there is a floppy mode for each wavenumber q ≠ 0 running along the entire length of the three symmetry-equivalent straight lines running from M to Γ to M in the Brillouin zone (see Fig. 3, Inset). When Nx = Ny, there are exactly Nx - 1 wavenumbers with q ≠ 0 along each of these lines for a total of 3(Nx - 1) floppy modes. In addition, there are three zero modes at q = 0 corresponding to two rigid translations and one floppy mode that changes unit cell area at second but not first order in displacements, yielding a total of 3Nx zero modes rather than the 4Nx - 1 modes expected from the Maxwell count under FBCs. This discrepancy is our first indication of the importance of boundary conditions. The addition of NNN springs endows the floppy modes at k = 0 with a characteristic frequency Inline graphic and causes them to hybridize with the acoustic phonon modes (Fig. 3A) (35). The result is an isotropic phonon spectrum up to wavenumber Inline graphic and gaps at Γ and M of order ω. Remarkably, at nonzero α and k = 0, the mode structure is almost identical to that at α = 0 and k > 0 with characteristic frequency Inline graphic and length lα ∼ 1/ωα. In other words, twisting the kagome lattice through an angle α has essentially the same effect on the spectrum as adding NNN springs with spring constant | sin α|2k. Thus, under PBCs, the twisted kagome lattice has no zero modes other than the trivial ones: It is “collectively” jammed in the language of refs. 32 and 36, but because it is not rigid with respect to changing the unit cell size, it is not strictly jammed.

Fig. 3.

Fig. 3.

(A) Phonon spectrum for the undistorted kagome lattice. Dashed lines depict frequencies at k = 0 and full lines at k > 0. The inset shows the Brillouin zone with symmetry points Γ, M, and K. Note the line of zero modes along ΓM when k = 0, all of which develop nonzero frequencies for wavenumber q > 0 when k > 0 reaching Inline graphic on a plateau beginning at Inline graphic defining a length scale l = 1/q. (B) Phonon spectrum for α > 0 and k = 0. The plateau along ΓM defines Inline graphic and its onset at qα ∼ ωα defines a length lα ∼ 1/| sin α|.

Mode Counting and States of Self-Stress

To understand the origin of the differences in the zero-mode count for different boundary conditions, we turn to an elegant formulation (2) of the Maxwell rule that takes into account the existence of redundant bonds (i.e., bonds whose removal does not increase the number of floppy modes; ref. 13) and states in which springs can be under states of self-stress (36). Consider a ring network in two dimensions shown in Fig. 4 with N = 4 nodes and Nb = 4 springs with three springs of length a and one spring of length b. The Maxwell count yields N0 = 4 = 3 + 1 zero modes: two rigid translations, one rigid rotation, and one internal floppy mode—all of which are “finite-amplitude” modes with zero energy even for finite-amplitude displacements. When b = 3a, the Maxwell rule breaks down. In the zero-energy configuration, the long spring and the three short ones are colinear, and a prestressed state in which the b spring is under compression and the three a springs are under tension (or vice versa) but the total force on each node remains zero becomes possible. This is called a state of self-stress. The system still has three finite-amplitude zero modes corresponding to arbitrary rigid translations and rotations, but the finite-amplitude floppy mode has disappeared. In the absence of prestress, it is replaced by two “infinitesimal” floppy modes of displacements of the two internal nodes perpendicular of the now linear network. In the presence of prestress, these two modes have a frequency proportional to the square root of the tension in the springs. Thus, the system now has one state of self-stress and one extra zero mode in the absence of prestress, implying N0 = 2N - NB + S, where S is the number of states of self-stress.

Fig. 4.

Fig. 4.

(A) Ring network with b > 3a showing internal floppy mode. (B) Ring-network with b = 3a showing one of the two infinitesimal modes.

This simple count is more generally valid, as can be shown with the aid of the equilibrium and compatibility matrices (2), denoted, respectively, as H and C ≡ HT. H relates the vector t of NB spring tensions to the vector f of dN forces at nodes via H·t = f, and C relates the vector d of dN node displacements to the vector e of NB spring stretches via C·d = e. The dynamical matrix determining the phonon spectrum is D = kH·HT. Vectors t0 in the null space of H, (H·t0 = 0), describe states of self-stress, whereas vectors d0 in the null space of C represent displacements with no stretch e—i.e., modes of zero energy. Thus the null-space dimensions of H and C are, respectively, S and N0. The rank-nullity theorem of linear algebra (37) states that the rank r of a matrix plus the dimension of its null space equals its column number. Because the rank of a matrix and its transpose are equal, the H and C matrices, respectively, yield the relations r + S = NB and r + N0 = dN, implying N0 = dN - NB + S. Under PBCs, locally isostatic lattices have z = 2d exactly, and the Maxwell rule yields N0 = 0: There should be no zero modes at all. But we have just seen that both the square and undistorted kagome lattices under PBCs have of order Inline graphic zero modes as calculated from the dynamical matrix, which, because it is derived from a harmonic theory, does not distinguish between infinitesimal and finite-amplitude zero modes. Thus, in order for there to be zero modes, there must be states of self-stress, in fact, one state of self-stress for each zero mode.

In the square lattice under FBCs, N = NxNy and NB = 2NxNy - Nx - Ny, there are no states of self-stress, and N0 = Nx + Ny zero modes depicted in Fig. 1B. Under PBCs, the dimension of the null space of H is S = Nx + Ny, and there are also N0 = S = Nx + Ny zero modes that are identical to those under FBCs. We have already seen that there are N0 = 2(Nx + Ny) - 1 zero modes in the free undistorted kagome lattice. Direct evaluations (29) (see SI Text) of the dimension of the null spaces of H and C for the undistorted kagome lattice with PBCs yields S = N0 = 3Nx when Nx = Ny. The zero modes under PBCs are identical to those under FBCs except that the 2Nx - 1 modes associated with lines parallel to e2 under FBCs get reduced to Nx modes because of the identification of apposite sides of the lattice required by the PBCs, as shown in Fig. 1A. Thus the modes of both the square and kagome lattices do not depend strongly on whether FBCs or PBCs are applied. Under RBCs, however, the floppy modes of both disappear. The situation for the twisted kagome lattice is different. There are still 2(Nx + Ny) - 1 zero modes under FBCs, but there are only two states of self-stress under PBCs and thus only N0 = S = 2 zero modes, as a direct evaluation of the null spaces of H and C verifies (SI Text), in agreement with the results obtained via direct evaluation of the eigenvalues of the dynamical matrix (35, 38). All of the floppy modes under FBCs have disappeared.

Effective Theory and Edge Modes

An effective long-wavelength energy Eeff for the low-energy acoustic phonons and nearly floppy distortions provides insight into the nature of the modes of the twisted kagome lattice. The variables in this theory are the vector displacement field u(x) of nodes at undistorted positions x and the scalar field ϕ(x) describing nearly floppy distortions within a unit cell. The detailed form of Eeff depends on which three lattice sites are assigned to a unit cell. Fig. 1C depicts the lattice distortion ϕ for the nearly floppy mode at Γ (with energy proportional to | sin α|2) along with a particular representation of a unit cell, consisting of a central asymmetric hexagon and two equilateral triangles, with eight sites on its boundary. If sites 1, 2, and 3 are assigned to the unit cell, then the distortion ϕ involves no rotations of these sites relative to their center of mass, and the harmonic limit of Eeff depends only on the symmetrized and linearized strain uij = (∂iuj + ∂jui)/2 and on ϕ:

graphic file with name pnas.1119941109eq45.jpg [1]

where Inline graphic is the symmetric-traceless stain tensor, Inline graphic, Inline graphic, Inline graphic, Inline graphic, and Inline graphic. The last term in which Γijk is a third-rank tensor, whose only nonvanishing components are Γxxx = -Γxyy = -Γyyx = -Γxyx = 1, is invariant under operations of the group C6v but not under arbitrary rotations. The Kξϕuii term is the only one that reflects the C3v (rather than C6v) symmetry of the lattice. There are several comments to make about this energy. The gauge-like coupling in which the isotropic strain uii appears only in the combination (ϕ + ξuii)2 guarantees that the bulk modulus vanishes: ϕ will simply relax to -ξuii to reduce to zero the energy of any state with nonvanishing uii. The coefficient K can be calculated directly from the observation that, under ϕ alone, the length of every spring changes by Inline graphic, and this length change is reversed by a homogenous volume change uii = δAα/Aα = -2δa/a. In the α → 0 limit, K → 0, and the energy reduces to that of an isotropic solid with bulk modulus Inline graphic if the V and W terms, which are higher order in gradients, are ignored. The W term gives rise to a term, singular in gradients of u, when ϕ is integrated out that is responsible for the deviations of the finite-wavenumber elastic energy from isotropy. At small α, the length scale lα appears in several places in this energy: in the length ξ and in the ratios Inline graphic, Inline graphic, and Inline graphic. At length scales much larger than lα, the V and W terms can be ignored, and ϕ relaxes to -ξuii, leaving only the shear elastic energy of an elastic solid proportional Inline graphic. At length scales shorter than lα, ϕ deviates from -ξuii and contributes significantly to the form of the energy spectrum. If 1, 2, and 3 in Fig. 1D are assigned to the unit cell, then ϕ involves rotations relative to the lattice axes, and the energy develops a Cosserat-like form (39, 40) that is a function of ϕ - a(∇ × u)z/2 rather than ϕ.

The modes of our elastic energy in the long-wavelength limit (qlα ≪ 1) are simply those of an elastic medium with B = 0. In this limit, there are transverse and longitudinal bulk sound modes with equal sound velocities Inline graphic and Inline graphic, where m is the particle mass at each node and ρ is the mass density. In addition there are Rayleigh surface waves (33) in which there is a single decay length (rather than the two at B > 0), and displacements are proportional to e-qyy cos[qxx] with qy = qx for a semiinfinite sample in the upper half-plane so that the penetration depth into the interior is 1/qx. These waves have zero frequency in two dimensions when B = 0, and they do not appear in the spectrum with PBCs. Thus this simple continuum limit provides us with an explanation for the difference between the spectrum of the free and periodic twisted kagome lattices. Under FBCs, there are zero-frequency surface modes not present under PBCs.

Further insight into how boundary conditions affect spectrum follows from the observation that the continuum elastic theory with B = 0 depends only on Inline graphic. The metric tensor gij(x) of the distorted lattice is related to the strain uij(x) via the simple relation gij(x) = δij + 2uij(x); and Inline graphic, which is zero for gij = δij, is invariant, and thus remains equal to zero, under conformal transformations that take the metric tensor from its reference form δij to h(xij for any continuous function h(x). The zero modes of the theory thus correspond simply to conformal transformations, which in two dimensions are best represented by the complex position and displacement variables z = x + iy and w(z) = ux(z) + iuy(z). All conformal transformations are described by an analytic displacement field w(z). Because, by Cauchy’s theorem, analytic functions in the interior of a domain are determined entirely by their values on the domain’s boundary (the “holographic” property; ref. 41), the zero modes of a given sample are simply those analytic functions that satisfy its boundary conditions. For example, a disc with fixed edges (u = 0) has no zero modes because the only analytic function satisfying this FBC is the trivial one w(z) = 0; but a disc with free edges (stress and thus strain equal to zero) has one zero mode for each of the analytic functions w(z) = anzn for integer n≥0. The boundary conditions Inline graphic and u(x,y) = u(x + L,y) on a semiinfinite cylinder with axis along x are satisfied by the function w(z) = eiqxz = eiqxxe-qxy when qx = 2nπ/L, where n is an integer. This solution is identical to that for classical Rayleigh waves on the same cylinder. Like the Rayleigh theory, the conformal theory puts no restriction on the value of n (or equivalently qx). Both theories break down, however, at Inline graphic, beyond which the full lattice theory, which yields a complex value of Inline graphic, is needed.

Fig. 5A shows an example of a surface wave. At the bottom of this figure, uy(x) is an almost perfect sinusoid. As y decreases toward the surface, the amplitude grows, and in this picture reaches the nonlinear regime by the time the surface at y = 0 is reached. Fig. 5B plots Inline graphic as a function of qx obtained both by direct numerical evaluation and by an analytic transfer matrix procedure (42) for different values of α (SI Text). The Rayleigh limit Inline graphic is reached for all α as qx → 0. Interestingly, the Rayleigh limit remains a good approximation up to values of qx that increase with increasing α. The inset to Fig. 5 plots Inline graphic as a function of η = qxlα and shows that in the limit α → 0 (lα/a → ∞), Inline graphic obeys an α-independent scaling law of the form Inline graphic. The full complex qy obeys a similar equation. This type of behavior is familiar in critical phenomena where scaling occurs when correlation lengths become much larger than microscopic lengths. The function f(η) approaches η as η → 0 and asymptotes to 4/3 for η → ∞. Thus for qxlα ≪ 1, Inline graphic and for qxlα≫1, Inline graphic. As α increases, lα/a is no longer much larger than one, and deviations from the scaling law result. The situation for surfaces along different directions (e.g., along x = 0 rather than y = 0) is more complicated and is beyond the scope of this paper.

Fig. 5.

Fig. 5.

(A) Lattice distortions for a surface wave on a cylinder, showing exponential decay of the surface displacements into the bulk. This figure was constructed by specifying a small sinusoidal modulation on the bottom boundary and propagating lattice-site positions upward to the free boundary at the top under the constraint of constant lengths and periodic boundary conditions around the cylinder. Distortions near and at the top boundary, which have become nonlinear, are not described by our linearized treatment. (B) Inline graphic as a function of qxaL for lattice Rayleigh surface waves for α = π/20, π/10, 3π/20, π/5, π/4, in order from bottom to top. Smooth curves are the analytic results from a transfer matrix calculation, and dots are from direct numerical calculations. The dashed line is the continuum Rayleigh limit Inline graphic. Curves at smaller α break away from this curve at smaller values of qy than do those at large α. At α = π/4, Inline graphic diverges at qyaL = π. The inset plots Inline graphic as a function of qxlα for different α. The lower curve in the inset (black) is the α-independent scaling function of qylα reached in the α → 0 limit. The other curves from top to bottom are for α = π/25, π/12, π/9, and π/6 (chosen to best present results rather than to match the curves in the main figure). Curves for α < π/15 are essentially indistinguishable from the scaling limit. The curve at α = π/6 stops because qy < π/aL.

Other Lattices and Relation to Jamming

The C3v twisted kagome lattice is the simplest of many lattices that can be formed from the kagome and other periodic isostatic lattices. Fig. 6 A and B shows two other examples of isostatic lattices constructed from the kagome lattice. Most intriguing is the lattice with pgg symmetry. Its geometry has uniaxial symmetry, yet its long-wavelength elastic energy is identical to that of the C3v twisted kagome lattice (i.e., it is isotropic with a vanishing bulk modulus), and its mode structure near q = 0 is isotropic as shown in Fig. 6C. Thus, this system loses long-wavelength zero-frequency bulk modes of the undistorted kagome lattice to surface modes. However, at large wavenumber, lattice anisotropy becomes apparent, and (infinitesimal) floppy bulk modes appear. Thus in this and related systems, a fraction of the zero modes under FBCs are bulk modes that are visible under PBCs, and a fraction are surface modes that are not.

Fig. 6.

Fig. 6.

(A) Kagome-based lattice with pgg space group symmetry and uniaxial C2v point group symmetry. (B) Lattice with p6 space symmetry but global C6 point-group symmetry. (C) Density plot of the spectrum of the lowest frequency branch of the pgg uniaxial kagome lattice. The spectrum is absolutely isotropic near the origin point Γ, but it has a zero modes on two symmetry related continuous curves at large values of wavenumber.

Randomly packed spheres above the jamming transition with average coordination number z = 2d + Δz exhibit a characteristic frequency ω ∼ Δz and length l ∼ (Δz)-1 and a transition from a Debye-like (∼ ωd-1) to a flat density of states at ω ≈ ω (43, 44). The square and kagome lattices with randomly added NNN springs have the same properties (45, 46). A general “cutting” argument (9, 31) provides a procedure for perturbing away from the isostatic limit and an explanation for these properties. However, it only applies provided a finite fraction of the of order Ld-1 floppy modes of a sample with sides of length L cut from an isostatic lattice with PBCs are extended—i.e., have wave functions that extend across the sample rather than remaining localized either in the interior or at the surface the sample. Clearly the twisted kagome lattice, whose floppy modes are all surface modes, violates this criterion; and indeed, the density of states of the lattice with Δz = 0 shows Debye behavior crossing over to a flat plateau at ω ≈ ωα. Adding NNN bonds gives rise to a length lc, which is approximately the parallel combination, Inline graphic, of the lengths lα and l arising, respectively, from twisting the lattice and adding NNN springs, and cross-over to the plateau at Inline graphic. The pgg lattice in Fig. 6A, however, has both extended and surface floppy modes, so its cross-over to a flat plateau occurs at ω ≈ ω rather than at ωα or ωc.

Connections to Other Systems

Our study highlights the rich and remarkable variety of physical properties that isostatic systems can exhibit. Under FBCs, floppy modes can adopt a variety of forms, from all being extended to all being localized near surfaces to a mixture of the two. Under PBCs, the presence of floppy modes depends on whether the lattice can or cannot support states of self-stress. When a lattice exhibits a large number of zero-energy edge modes, its mechanical/dynamical properties become extremely sensitive to boundary conditions, much as do the electronic properties of the topological states of matter studied in quantum systems (4750). The zero-energy edge modes observed in our isostatic lattices are collective modes whose amplitudes decay exponentially from the edge with a finite decay length, in direct contrast to the very localized and trivial floppy modes arising from dangling bonds. We focused primarily on high-symmetry lattices derived from the kagome lattice, but the properties they exhibit (namely, a deficit of floppy modes in the bulk and the existence of floppy surface modes) are shared by any two-dimensional system with a vanishing bulk modulus (or the equivalent in anisotropic systems). Three-dimensional analogs of the twisted kagome lattice can be constructed by rotating tetrahedra in pyrochlore and zeolite lattices (19, 20) and in cristobalites (18). These lattices are anisotropic. With NN forces only, they exhibit a vanishing modulus for compression applied in particular planes rather than isotropically, but we expect them to exhibit many of the properties the two-dimensional lattices exhibit. Finally, we note that Maxwell’s ideas can be applied to spin systems such as the Heisenberg antiferromagnet on the kagome lattice (51, 52), and the possibility of unusual edge states in them is intriguing.

Supplementary Material

Supporting Information

Acknowledgments.

We are grateful for informative conversations with Randall Kamien, Andea Liu, and S.D. Guest. This work was supported in part by the National Science Foundation under Grants DMR-0804900, DMR-1104707 (to T.C.L. and X.M.), Materials Research Science and Engineering Center DMR-0520020 (to T.C.L. and A.S.), and JQI-NSF-PFC (to K.S.).

Footnotes

The authors declare no conflict of interest.

This article is a PNAS Direct Submission.

See Commentary on page 12266.

This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.1073/pnas.1119941109/-/DCSupplemental.

References

  • 1.Maxwell JC. On the calculation of the equilibrium and stiffness of frames. Philos Mag. 1865;27:294. [Google Scholar]
  • 2.Calladine CR. Buckminster fuller “tensegrity” structures and Clerk Maxwell rules for the construction of stiff frames. Int J Solids Struct. 1978;14:161–172. [Google Scholar]
  • 3.Asimow L, Roth B. Rigidity of graphs. Trans Am Math Soc. 1978;245:279–289. [Google Scholar]
  • 4.Asimow L, Roth B. Rigidity of graphs 2. J Math Anal Appl. 1979;68:171–190. [Google Scholar]
  • 5.Connelly R. Rigidity and energy. Invent Math. 1982;66:11–33. [Google Scholar]
  • 6.Connelly R. Generic global rigidity. Discrete Comput Geom. 2005;33:549–563. [Google Scholar]
  • 7.Thorpe MF. Continuous deformations in random networks. J Non Cryst Solids. 1983;57:355–370. [Google Scholar]
  • 8.Phillips JC. Topology of covalent non-crystalline solids 2. Medium-range order in chalcogenide alloys and α - Si((Ge) J Non Cryst Solids. 1981;43:37–77. [Google Scholar]
  • 9.Wyart M. On the rigidity of amorphous solids. Annales de Physique. 2005;30:1–96. [Google Scholar]
  • 10.Liu AJ, Nagel SR. Granular and jammed materials. Soft Matt. 2010;6:2869–2870. [Google Scholar]
  • 11.Liu AJ, Nagel SR. Langer J, editor. The jamming transition and the marginally jammed solid. (Series Annual Reviews of Condensed Matter Physics).Web of Science. 2010;1:347–369. [Google Scholar]
  • 12.Feng S, Sen PN. Percolation on elastic networks: New exponents and threshold. Phys Rev Lett. 1984;52:216–219. [Google Scholar]
  • 13.Jacobs DJ, Thorpe MF. Generic rigidity percolation—the pebble game. Phys Rev Lett. 1995;75:4051–4054. doi: 10.1103/PhysRevLett.75.4051. [DOI] [PubMed] [Google Scholar]
  • 14.Heyman J. The Science of Structural Engineering. Stamford, CT: Cengage Learning; 2005. [Google Scholar]
  • 15.Kassimali A. Structural Analysis. Stamford, CT: Cengage Learning; 2005. [Google Scholar]
  • 16.Liu AJ, Nagel SR. Nonlinear dynamics—jamming is not just cool any more. Nature. 1998;396:21–22. [Google Scholar]
  • 17.Torquato S, Stillinger FH. Jammed hard-particle packings: From Kepler to Bernal and beyond. Rev Mod Phys. 2010;82:2633–2672. [Google Scholar]
  • 18.Hammonds KD, Dove MT, Giddy AP, Heine V, Winkler B. Rigid-unit phonon modes and structural phase transitions in framework silicates. Am Mineral. 1996;81:1057–1079. [Google Scholar]
  • 19.Auerbach SM, Carrado KA, Dutta PK, editors. Zeolite Science and Technology. New York: Taylor and Francis; 2005. [Google Scholar]
  • 20.Sartbaeva A, Wells SA, Treacy MMJ, Thorpe MF. The flexibility window in zeolites. Nat Mater. 2006;5:962–965. doi: 10.1038/nmat1784. [DOI] [PubMed] [Google Scholar]
  • 21.Wilhelm J, Frey E. Elasticity of stiff polymer networks. Phys Rev Lett. 2003;91:108103. doi: 10.1103/PhysRevLett.91.108103. [DOI] [PubMed] [Google Scholar]
  • 22.Heussinger C, Frey E. Floppy modes and nonaffine deformations in random fiber networks. Phys Rev Lett. 2006;97 doi: 10.1103/PhysRevLett.97.105501. 105501. [DOI] [PubMed] [Google Scholar]
  • 23.Huisman L, Lubensky T. Internal stresses, normal modes and non-affinity in three-dimensional biopolymer networks. Phys Rev Lett. 2011;106:088301. doi: 10.1103/PhysRevLett.106.088301. [DOI] [PubMed] [Google Scholar]
  • 24.Broedersz C, Mao X, Lubensky TC, MacKintosh F. Criticality and isostaticity in fiber networks. Nat Phys. 2011;7:983–988. 105501. [Google Scholar]
  • 25.Lakes R. Foam structures with a negative poisson’s ratio. Science. 1987;235:1038–1040. doi: 10.1126/science.235.4792.1038. [DOI] [PubMed] [Google Scholar]
  • 26.Evans KE, Nkansah MA, Hutchinson IJ, Rogers SC. Molecular network design. Nature. 1991;353:124–124. [Google Scholar]
  • 27.Chen CP, Lakes RS. Holographic study of conventional and negative poisson ratio metallic foams—elasticity, yield and micro-deformation. J Mater Sci. 1991;26:5397–5402. [Google Scholar]
  • 28.Greaves G, Greer A, Lakes R, Rouxe T. Poisson’s ratio and modern materials. Nat Mater. 2011;10:823–827. doi: 10.1038/nmat3134. [DOI] [PubMed] [Google Scholar]
  • 29.Hutchinson RG, Fleck NA. The structural performance of the periodic truss. J Mech Phys Solids. 2006;54:756–782. [Google Scholar]
  • 30.Kapko V, Treacy MMJ, Thorpe MF, Guest SD. On the collapse of locally isostatic networks. Proc R Soc A. 2009;465:3517–3530. [Google Scholar]
  • 31.Wyart M, Nagel SR, Witten TA. Geometric origin of excess low-frequency vibrational modes in weakly connected amorphous solids. Europhys Lett. 2005;72:486–492. [Google Scholar]
  • 32.Torquato S, Stillinger FH. Multiplicity of generation, selection, and classification procedures for jammed hard-particle packings. J Phys Chem B. 2001;105:11849–11853. [Google Scholar]
  • 33.Landau L, Lifshitz E. Theory of Elasticity. 3rd Ed. New York: Pergamon; 1986. [Google Scholar]
  • 34.Grima JN, Alderson A, Evans KE. Auxetic behaviour from, rotating rigid units. Phys Status Solidi B. 2005;242:561–575. [Google Scholar]
  • 35.Souslov A, Liu AJ, Lubensky TC. Elasticity and response in nearly isostatic periodic lattices. Phys Rev Lett. 2009;103:205503. doi: 10.1103/PhysRevLett.103.205503. [DOI] [PubMed] [Google Scholar]
  • 36.Donev A, Torquato S, Stillinger FH, Connelly R. Jamming in hard sphere and disk packings. J Appl Phys. 2004;95:989–999. [Google Scholar]
  • 37.Birkhoff G, MacLane S. A Survey of Modern Algebra. New York: Taylor and Francis; 1998. [Google Scholar]
  • 38.Guest SD, Hutchinson JW. On the determinacy of repetitive structures. J Mech Phys Solids. 2003;51:383–391. [Google Scholar]
  • 39.Cosserat E, Cosserat FC. Théorie des Corps Déformables. Paris: Hermann et Fils; 1909. [Google Scholar]
  • 40.Kunin I. Elstic Media and Microstructure II. Berlin: Springer; 1983. [Google Scholar]
  • 41.Susskind L. The world as a hologram. J Math Phys. 1995;36:6377–6396. [Google Scholar]
  • 42.Lee DH, Joannopoulos JD. Simple scheme for surface-band calculations. I. Phys Rev B Condens Matter Mater Phys. 1981;23:4988–4996. [Google Scholar]
  • 43.O’Hern CS, Silbert LE, Liu AJ, Nagel SR. Jamming at zero temperature and zero applied stress: The epitome of disorder. Phys Rev E Stat Nonlinear Soft Matter Phys. 2003;68:011306. doi: 10.1103/PhysRevE.68.011306. [DOI] [PubMed] [Google Scholar]
  • 44.Silbert LE, Liu AJ, Nagel SR. Vibrations and diverging length scales near the unjamming transition. Phys Rev Lett. 2005;95:098301. doi: 10.1103/PhysRevLett.95.098301. [DOI] [PubMed] [Google Scholar]
  • 45.Mao XM, Xu N, Lubensky TC. Soft modes and elasticity of nearly isostatic lattices: Randomness and dissipation. Phys Rev Lett. 2010;104:085504. doi: 10.1103/PhysRevLett.104.085504. [DOI] [PubMed] [Google Scholar]
  • 46.Mao XM, Lubensky TC. Coherent potential approximation of random nearly isostatic kagome lattice. Phys Rev E Stat Nonlinear Soft Matter Phys. 2011;83:011111. doi: 10.1103/PhysRevE.83.011111. [DOI] [PubMed] [Google Scholar]
  • 47.Kane C, Fisher MP. In: Perspectives in Quantum Hall Effects. Das Sarma S, Pinczuk A, editors. New York: Wiley; 1997. [Google Scholar]
  • 48.Jain JK. Composite Fermions. New York: Cambridge Univ Press; 2007. [Google Scholar]
  • 49.Hasan MZ, Kane CL. Colloquium: Topological insulators. Rev Mod Phys. 2010;82:3045–3067. [Google Scholar]
  • 50.Qi XL, Zhang SC. Topological insulators and superconductors. Rev Mod Phys. 2011;83:1057–1110. [Google Scholar]
  • 51.Moessner R, Chalker JT. Low-temperature properties of classical geometrically frustrated antiferromagnets. Phys Rev B Condens Matter Mater Phys. 1998;58:12049–12062. [Google Scholar]
  • 52.Lawler MJ. Emergent gauge dynamics in highly frustrated magnet. 2011. arXiv:1104.0721.

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supporting Information

Articles from Proceedings of the National Academy of Sciences of the United States of America are provided here courtesy of National Academy of Sciences

RESOURCES