arXiv:2107.07216v1 [physics.optics] 15 Jul 2021

13
Gap Sensitivity Reveals Universal Behaviors in Optimized Photonic Crystal and Disordered Networks Michael A. Klatt, 1, 2, * Paul J. Steinhardt, 1, and Salvatore Torquato 1, 3, 1 Department of Physics, Princeton University, Princeton, New Jersey 08544, USA 2 Institut f¨ ur Theoretische Physik, University of Erlangen-N¨ urnberg, Staudtstr. 7, 91058 Erlangen, Germany 3 Department of Chemistry, Princeton Institute for the Science and Technology of Materials, and Program in Applied and Computational Mathematics, Princeton University, Princeton, New Jersey 08544, USA (Dated: July 16, 2021) Through an extensive series of high-precision numerical computations of the optimal complete pho- tonic band gap (PBG) as a function of dielectric contrast α for a variety of crystal and disordered heterostructures, we reveal striking universal behaviors of the gap sensitivity S (α) dΔ(α)/dα, the first derivative of the optimal gap-to-midgap ratio Δ(α). In particular, for all our crystal networks, S (α) takes a universal form that is well approximated by the analytic formula for a one-dimensional quarter-wave stack, SQWS(α). Even more surprisingly, the values of S (α) for our disordered net- works converge to SQWS(α) for sufficiently large α. A deeper understanding of the simplicity of this universal behavior may provide fundamental insights about PBG formation and guidance in the design of novel photonic heterostructures. Introduction. A complete photonic band gap (PBG) prohibits the propagation of light in all directions and for all polarizations for a substantial continuous range of frequencies [1, 2]. PBGs can occur in heterostructures composed of two or more substances with different dielec- tric constants such as silicon and air; see Fig. 1. In early studies, PBGs were exclusively found in crystalline struc- tures, such as the diamond crystal network [3]. Later it was discovered that quasicrystals without long-range pe- riodic translational order [4, 5] and isotropic “disordered” solids can also exhibit complete PBGs [6–16]. In this paper, we present evidence for a subtle, unan- ticipated universal behavior of the maximum complete PBG as a function of dielectric contrast among network heterostructures spanning a wide variety of symmetries and topologies. For this purpose, we follow a multistage procedure that begins with identifying different “candi- date classes” of networks distinguished by their nearest- neighbor table properties (e.g., coordination ), transla- tional order and rotational symmetry which are known from past work to include examples with large complete (a) (b) FIG. 1. Photonic networks: (a) the crystal diamond network, where cylindrical rods connect nearest neighbors of a diamond lattice, and (b) the disordered nearly hyperuniform network (NHN) model; see also video S1. PBGs [1–3, 6, 7, 17]. For each candidate class, we aim to identify the member with the largest complete PBG, which are, based on experience to date, among those that combine the highest degree of hyperuniformity [18–21] with sufficiently narrow distributions of bond lengths and angles. Then, we optimize the microscopic properties of the selected networks. For this, we assume the networks are composed of rods and spheres with dielectric constant ε 2 embedded in a bulk that has a smaller dielectric constant ε 1 . We also assume the rods have circular cross-section with radius R joined at sphere-shaped vertices with the same radius R. For each fixed dielectric contrast ratio α := ε 2 1 , we vary R to find the maximal value of the photonic band gap-to-midgap ratio Δ(α); that is, Δ(α) := max R0 Δω ω m ( R, α ) . (1) We call the optimized radius R opt (α) and the correspond- ing optimized volume fraction φ opt (α). Since our purpose is to study cases with large PBGs, we only consider candidate classes known to have some networks satisfying this condition. In some cases, this condition is not straightforward to check. For exam- ple, disordered networks typically have many localized defect modes that break up what would otherwise be a large PBG into many smaller PBGs. As a last step of the optimization, we check whether such defects can be removed via bond switching; if so, then the candidate class is included in the study and the modified network is treated according to the same optimization procedure as described above. Critical to our analysis are high-precision calculations of the optimal gap-to-midgap ratios, Δ(α), that are un- precedented in their scope, both for the wide variety of heterostructures considered and the wide range of dielec- arXiv:2107.07216v1 [physics.optics] 15 Jul 2021

Transcript of arXiv:2107.07216v1 [physics.optics] 15 Jul 2021

Gap Sensitivity Reveals Universal Behaviors in Optimized PhotonicCrystal and Disordered Networks

Michael A. Klatt,1, 2, ∗ Paul J. Steinhardt,1, † and Salvatore Torquato1, 3, ‡

1Department of Physics, Princeton University, Princeton, New Jersey 08544, USA2Institut fur Theoretische Physik, University of Erlangen-Nurnberg, Staudtstr. 7, 91058 Erlangen, Germany

3Department of Chemistry, Princeton Institute for the Science and Technology of Materials,and Program in Applied and Computational Mathematics,Princeton University, Princeton, New Jersey 08544, USA

(Dated: July 16, 2021)

Through an extensive series of high-precision numerical computations of the optimal complete pho-tonic band gap (PBG) as a function of dielectric contrast α for a variety of crystal and disorderedheterostructures, we reveal striking universal behaviors of the gap sensitivity S(α) ≡ d∆(α)/dα, thefirst derivative of the optimal gap-to-midgap ratio ∆(α). In particular, for all our crystal networks,S(α) takes a universal form that is well approximated by the analytic formula for a one-dimensionalquarter-wave stack, SQWS(α). Even more surprisingly, the values of S(α) for our disordered net-works converge to SQWS(α) for sufficiently large α. A deeper understanding of the simplicity of thisuniversal behavior may provide fundamental insights about PBG formation and guidance in thedesign of novel photonic heterostructures.

Introduction. A complete photonic band gap (PBG)prohibits the propagation of light in all directions andfor all polarizations for a substantial continuous rangeof frequencies [1, 2]. PBGs can occur in heterostructurescomposed of two or more substances with different dielec-tric constants such as silicon and air; see Fig. 1. In earlystudies, PBGs were exclusively found in crystalline struc-tures, such as the diamond crystal network [3]. Later itwas discovered that quasicrystals without long-range pe-riodic translational order [4, 5] and isotropic “disordered”solids can also exhibit complete PBGs [6–16].

In this paper, we present evidence for a subtle, unan-ticipated universal behavior of the maximum completePBG as a function of dielectric contrast among networkheterostructures spanning a wide variety of symmetriesand topologies. For this purpose, we follow a multistageprocedure that begins with identifying different “candi-date classes” of networks distinguished by their nearest-neighbor table properties (e.g., coordination), transla-tional order and rotational symmetry which are knownfrom past work to include examples with large complete

(a) (b)

FIG. 1. Photonic networks: (a) the crystal diamond network,where cylindrical rods connect nearest neighbors of a diamondlattice, and (b) the disordered nearly hyperuniform network(NHN) model; see also video S1.

PBGs [1–3, 6, 7, 17]. For each candidate class, we aimto identify the member with the largest complete PBG,which are, based on experience to date, among those thatcombine the highest degree of hyperuniformity [18–21]with sufficiently narrow distributions of bond lengths andangles.

Then, we optimize the microscopic properties of theselected networks. For this, we assume the networks arecomposed of rods and spheres with dielectric constant ε2embedded in a bulk that has a smaller dielectric constantε1. We also assume the rods have circular cross-sectionwith radius R joined at sphere-shaped vertices with thesame radius R. For each fixed dielectric contrast ratioα := ε2/ε1, we vary R to find the maximal value of thephotonic band gap-to-midgap ratio ∆(α); that is,

∆(α) := maxR≥0

{∆ω

ωm

(R,α

)}. (1)

We call the optimized radius Ropt(α) and the correspond-ing optimized volume fraction φopt(α).

Since our purpose is to study cases with large PBGs,we only consider candidate classes known to have somenetworks satisfying this condition. In some cases, thiscondition is not straightforward to check. For exam-ple, disordered networks typically have many localizeddefect modes that break up what would otherwise be alarge PBG into many smaller PBGs. As a last step ofthe optimization, we check whether such defects can beremoved via bond switching; if so, then the candidateclass is included in the study and the modified networkis treated according to the same optimization procedureas described above.

Critical to our analysis are high-precision calculationsof the optimal gap-to-midgap ratios, ∆(α), that are un-precedented in their scope, both for the wide variety ofheterostructures considered and the wide range of dielec-

arX

iv:2

107.

0721

6v1

[ph

ysic

s.op

tics]

15

Jul 2

021

2

tric contrasts. Performing this computation for disor-dered networks with 1000 vertices is only first possiblenow using state-of-the-art computational techniques andcomputer clusters. To get reliable results for ∆(α) andhence S(α), it is essential to (i) accurately determinethe optimized volume fraction φopt(α); (ii) compute thestopgap along a number of different directions of propa-gation; (iii) use a sufficiently high spatial resolution; and(iv) use a precise plane wave expansion method to solveMaxwell’s equation [22, 23].

A cursory view of plots of ∆(α) as a function of αalready suggests certain common trends across the dif-ferent types of networks (despite the distinctly differentfunctional values). What proves to be the key to reveal-ing the universal behaviors is the gap-sensitivity, S(α),defined as

S(α) :=d∆(α)

dα. (2)

As we show below, for all our optimized three-dimensional crystal networks, S(α) is well approximatedfor all α by

SQWS(α) :=2

π

1

α3/4(1 +√α), (3)

where SQWS(α) is the precise analytic result for the gap-sensitivity of a one-dimensional quarter-wave stack com-posed of alternating layers with dielectric constants ε1and ε2 and each of quarter-wavelength thickness (whichresults in an optimized volume fraction) where the prop-agation of electromagnetic waves is perpendicular to thealternating dielectric layers. For a derivation of the cor-responding formula for ∆(α); see, e.g., Refs. [24, 25].

Perhaps even more striking is the fact that, for all ouroptimized disordered networks, the values of S(α) alsoconverge to SQWS(α) at large α. In other words, at suffi-ciently large α, all our optimized crystal and disorderedheterostructures — despite the differences in symmetry,topology, and long-range order — exhibit the same uni-versal behavior: S(α) ≈ SQWS(α).

Notably, the story is different for small α = O(1) nearthe critical value αc (defined as the minimal contrast atwhich a complete PBG first appears). Over this range,S(α) displays a common pattern for all our optimizeddisordered networks that is clearly distinct from the com-mon pattern for crystal networks. Namely, whereas S(α)for the crystal networks decreases monotonically evenfor small α, S(α) for disordered networks first increases,reaches a maximum, and then decreases, ultimately con-verging to SQWS(α) as α continues to increase.

Candidate classes of networks. Among the abundantvariety of known photonic crystals and disordered het-erostructures, we select, for our computationally inten-sive optimization, candidate classes of networks thatrepresent a broad spectrum of symmetries and coor-dination numbers. The diamond crystal network ex-hibits the largest known optimal gap-to-midgap ratio

∆(α) [2, 3, 17]. The hexagonal-diamond network hasthe same topology and perfectly tetrahedral vertices buta different symmetry. It has substantial PBGs, al-though its optimal gap-to-midgap ratios ∆(α) are dis-tinctly smaller than those of the diamond network [26]. Abetter performing photonic crystal is the rod-connectednetwork based on the Laves graph, which has the sametopology as the single gyroid [2, 15, 27, 28]. The diamondand Laves graphs are the only two crystal lattices in threedimensions with the “strong isotropic property” [29], i.e.,a symmetry under permutations of neighboring edges.The key differences are that the diamond network is four-fold coordinated and nonchiral with a face-centered cu-bic symmetry and the Laves graph is trivalent and chiralwith a body-centered cubic symmetry. We also study asimple cubic (SC) heterostructure, a sixfold-coordinatednetwork consisting of rods connecting nearest neighborsin a SC lattice [30, 31].

Our isotropic disordered networks are not based onany underlying lattice. They exhibit a correlated disor-der with a varying degree of both local and global order.Based on experience to date, the member in a candidateclass with the largest PBG has bond length and angle dis-tributions with a standard deviation less than 15% of themean. For larger variations, the PBG becomes smallerand the absolute difference between S(α) and SQWS(α)becomes larger.

One class of disordered models that we consider arecontinuous random networks (CRNs), i.e., idealized mod-els for amorphous tetrahedrally coordinated solids (likeamorphous silicon) [19, 32]. The first complete PBGof three-dimensional disordered networks was found forCRNs [7]. Here, we study the PBGs of the nearly hyper-uniform network (NHN) model [19]. Starting from a clas-sical CRN, the model is carefully annealed to suppresslarge-scale density fluctuations. The model is thus driventowards a vanishing of density fluctuations in the infinite-wavelength limit, known as hyperuniformity [18, 33]. Intwo dimensions, hyperuniformity has been found advan-tageous for opening up large complete PBGs [6].

Another disordered network that we consider is basedon an alternative structural model of amorphous siliconthat was simulated by a slow quench of a liquid usingmolecular dynamics (MD) [20]. When we connected eachatom with its four nearest neighbors, we obtained a dis-ordered photonic network, but it had defect modes thatappeared within a large PBG. We, therefore, switchedbonds to remove the defect modes, nearly doubling thegap size.

Finally, we introduce quantizer-based networks(QBNs). We begin from amorphous, nearly hyperuni-form inherent structures of the quantizer energy startingfrom some random initial point configuration [21]. Formore details on the quantizer energy functional; see sup-plemental material (SM) Sec. 1 and Refs. [21, 34–41].Next, we construct the corresponding Delaunay tessella-

3

0%

15%

30%

45%

4 10 16 22

NHN

MD quench

QBN

Diamond

Laves

Hex. diamond

SC

1D quarter-wave stack

0%

0.7%

1.4%

4.5 5

NHN Laves

MD quench

∆(α

)

α := ε2/ε1

FIG. 2. The gap plot shows the optimal gap-to-midgap ratio∆(α) as a function of the dielectric contrast α. The plot com-pares three-dimensional photonic crystals (open symbols) tothree-dimensional disordered networks (solid symbols). Thedashed-dotted line shows ∆QWS(α) for the one-dimensionalquarter-wave stack. The inset zooms in on small values of αwhere the PBGs first open up for the NHN, MD quench, andLaves networks.

tion. A QBN then connects the centroids of neighboringcells, in analogy to the tiling procedure for hyperuniformdisordered solids (HUDS) in two dimensions described inRef. [6]. In three dimensions, the tiling procedure wasalready applied to hard-sphere packings [9, 12, 42] andvertex models [16]. To avoid defect modes for α > 13,we removed each vertex where two triangles met by bondswitching.

Our three disordered networks fulfill all of our selectioncriteria. In particular, the ratio of standard deviations tomean values for the bond lengths and angles are 3.6% and7.4% for the NHN, 1.9% and 8.5% for the MD quench,and 7.4% and 10.7% for the QBN.

Protocol. All samples are constructed with periodicboundary conditions to enable band structure calcula-tions. The PBGs are accurately determined using theplane-wave expansion method to solve the frequency-domain eigenproblem implemented in the MIT PhotonicBands (MPB) software package [22]. The plane-waveexpansion is the best available method for achieving thehigh precision needed for the optimization and analysispresented here. It is not clear whether the same accu-racy can be achieved using time-domain eigensolvers, likethe finite-difference time-domain (FDTD) method, whichmay miss eigenfrequencies or locate spurious ones [1].We define a complete PBG by a range of frequencies forwhich there are no states no matter the direction of thewave vector. In practice, we compute the stopgaps fora finite number of directions along the edges of the irre-ducible Brillouin zone.

For each sample and each dielectric contrast α, we op-timize the rod radius R and hence the volume fractionφ of the high dielectric phase. We first determine the

0

0.02

0.04

0.06

4 10 16 22

S(α

):=

d∆

(α)/

α

NHN

QBN

MD quench

Diamond

Hex. diamond

Laves

SC

1D quarter-wave stack

FIG. 3. The gap-sensitivity plot shows the slope S(α) :=d∆(α)/dα for each of the curves in Fig. 2. For our three-dimensional crystal networks, S(α) is well approximated bySQWS(α) for all α. For our disordered networks, S(α) hasa maximum at low α = O(1) and only converges (approxi-mately) to SQWS(α) at larger α� 1.

gap-to-midgap ratios ∆ω/ωm for several radii close tothe putative optimum PBG. Next, we fit a parabola toestimate the optimized radius Ropt(α), at which we thenrepeat the band structure calculations. Thus, we deter-mine the optimal gap-to-midgap ratio ∆(α); see Eq. (1).We estimate S(α), see Eq. (2), using the symmetric dif-ference quotient.

The combined absolute error depends on the typeof network and dielectric contrast and is significantlysmaller at low than at high α. Therefore, we limit our in-vestigation to values of α ≤ 26. The error in determining∆(α) is less than 0.7% for α ≤ 13 and less than about1.4% for α > 13. As shown in Fig. S4, the results are notsensitive to system size.

Gap plots and gap-sensitivity plots. The key results ofthis paper derive from Figs. 2 and 3. Figure 2 shows plotsof the optimal gap-to-midgap ratio ∆(α) as a function ofα (referred to henceforth as the gap plot) for all of ourcrystal and disordered networks with different symme-tries and topological characteristics. Figure 3 (the gap-sensitivity plot) shows the slope S(α) := d∆(α)/dα forthe same networks.

Results. Photonic band structures of optimized crys-tal and disordered networks considered in this work areshown in Fig. S1 and video S2. The gap plot in Fig. 2shows that the diamond network exhibits the largestPBG for all dielectric contrasts, closely followed by theLaves network. At α > 10, disordered networks havesmaller PBGs than the Laves and hexagonal-diamondnetworks. However, ∆(α) of the disordered networks in-creases more slowly as a function of α than those of thecrystal networks, leading to a notable crossing of curvesin the gap plot. In fact, we find a smaller critical contrast,αc = 4.1, for the NHN model and MD quench than forthe hexagonal-diamond network (αc = 5.8) or even the

4

Laves network (αc = 4.5). Only the diamond networkhas a smaller critical contrast (αc = 3.5). Moreover, thegap plot shows that the NHN model has a larger optimalgap-to-midgap ratio [∆(13) = 19.7%] than the currentlyreported record for CRNs (18%) at α = 13 [7].

Turning now to the gap-sensitivity plot in Fig. 3, wenote that for all our crystal networks and for all α, S(α) iswell approximated without any fit parameter by SQWS(α)from Eq. (3), the analytic result for the quarter-wavestack. Remarkably, even for our isotropic disordered net-works, S(α) converges at large α to SQWS(α) to a goodapproximation. More precisely, we use “convergence” inthis paper to mean S(α) = (1± 0.2)SQWS(α) for α & 20,corresponding to an absolute error of 2 × 10−3, whichwe estimate to be the systematic error associated withthe optimization protocol used to identify the memberof a candidate class with the topology, bond length andangle distributions, and volume fraction that maximizesthe PBG. Convergence to within this 20% uncertaintyis impressive given the diverse candidate classes thathave been considered: classes with highly isotropic andanisotropic PBGs; with uniform and irregular topolo-gies; with translational order and disorder; and with one-dimensional layered structure and three-dimensional net-work structure.

We emphasize that the multistage optimization pro-tocol is essential to achieving this level of convergence.Without any optimization, the difference between S(α)and SQWS(α) would increase by an order of magnitudeor more. Even though S(α) is approximately the samefor our three, four, and sixfold coordinated networks, weonly find the universal behavior for networks with rela-tively uniform bond lengths and angles, as required inthe first stage of our procedure. To examine the effect ofvariations in bond lengths and angles, we randomly per-turbed the vertices of a diamond network using indepen-dent and isotropic Gaussian displacements. The strengthof perturbations is controlled by the ratio a of the stan-dard deviation of the displacements to the bond lengthin the unperturbed network. For a ≤ 10%, the perturbeddiamond network fulfills our selection criterion for bondlength and angles, as mentioned above, and S(α) approx-imately converges to SQWS(α) at large α. However, withincreasing strength of perturbations a, we find that thedifference between S(α) and SQWS(α) increasingly growsuntil the PBG closes altogether when a ≈ 0.5. We findthe same behavior for a perturbed CRN [32]. The qual-itative behavior of S(α) remains the same, but S(α) de-viates from SQWS(α) approximately by a constant factorthat grows as a increases.

The second stage of our procedure entails finding theoptimized volume fraction. This step is crucial becauseφopt(α) varies between 8.0% and 38.4% for our networksand values of α so that, without optimizing φ, we obtainboth quantitatively and qualitatively different results. Ifwe only optimized φ to within a certain percentage of the

true optimal value, we observe similar effects as for largevariations in the bond lengths and angles: the differencebetween S(α) and SQWS(α) increases with an increasingpercentage difference between φ and φopt(α). Alterna-tively, if we fix φ independent of α, we observe a moredramatic effect: the gap-to-midgap ratio quickly flattensfor large α at a value well below the optimal ∆(α) of thatcandidate class; that is, S(α) falls rapidly to zero at largeα.

Finally, the third stage is necessary to take account oflocalized defect modes that break up an otherwise largePBG. These defects not only decrease ∆(α) but also leadto large deviations between S(α) and SQWS(α). We ob-served such localized defects in the initial QBN modelwherever two triangles met at one vertex. When we re-moved these triangle defects by bond switching, ∆(α)increased significantly and S(α) approached SQWS(α) atlarge α.

Although we do not have a theoretical explanation ofthe universal behavior of S(α) at large α and the sys-tematic differences between crystal versus disordered net-works at small α, we find an interesting correlation withthe behavior of the square of the magnitude of the electricfield eigenmodes E(r, α) just below and above the PBG.More precisely, we find that ‖E(r, α)‖2 changes more atsmall α for a given change in dielectric contrast α thanit does at large α and that this effect is much more pro-nounced for the disordered than for the crystal networks.To quantify this effect, we introduce, as a heuristic mea-sure, the average change of ‖E(r, α)‖2 with the dielectriccontrast: DE(α) :=

∫ ∣∣∂‖E(r, α)‖2/∂α∣∣ dr. Figure S2

shows that DE(α) is, apart from statistical and numer-ical fluctuations, a decreasing function. At low α, thechanges are distinctly stronger for the disordered thanthe crystal networks. In contrast, at large α, where weobserve the universal behavior of S(α), DE(α) convergeswithin the computational uncertainty for the crystal anddisordered networks.

This observation suggests the following argument: Atlarge α, the electric field is strongly confined by the high-dielectric material, and the configuration changes onlyslowly with the dielectric contrast. At small α, the elec-tric field is generically less confined to the high-dielectricmaterial allowing more degrees of freedom to be accessedwhen optimizing the PBG. The high degree of transla-tional order in crystal systems constrains the field config-uration possibilities, so the corresponding curves of S(α)still follow Eq. (3). For disordered systems, though, thereare not the same symmetry constraints, allowing morefield configurations that enable the optimized PBG todecrease more slowly as α decreases.

We also observe that the derivative of the optimalvolume fraction, dφopt(α)/dα, approximately agrees forall our three-dimensional crystal networks at all α andis, in fact, well approximated by the analytic result forthe one-dimensional quarter-wave stack: dφQWS(α)/dα =

5

−1/(2α3/2+4α+2√α). Similarly, for our disordered net-

works, dφopt(α)/dα approximately converges to the sameformula at large α; see Fig. S3. This behavior correlateswith our finding that the optimization of the volume frac-tion is an essential part of our multistage procedure toreveal the universal behavior of S(α).

Conclusions and Outlook. For all our candidateclasses, including ordered and disordered varieties withdifferent symmetries and topologies, S(α) of our op-timized three-dimensional networks approximately con-verges at large α to SQWS(α) of the one-dimensionalquarter-wave stack. A physical explanation of this uni-versal behavior has to apply to both crystal and disor-dered networks even though the PBGs in these two caseshave distinctly different physical properties. The crys-tal networks have anisotropic PBGs bounded above atone k point and below by a different k point along theband gap. The slope S(α) of the single stop gaps doesnot agree as well with the one-dimensional quarter-wavestack formula as that of the three-dimensional completePBG. In contrast, the PBGs of the disordered networksare isotropic, i.e., they have the same stop gap in everydirection (apart from negligible statistical fluctuations)so that the upper and lower boundaries of the PBG areboth set by any single k point. Moreover, the eigenmodesabove and below the PBGs of the disordered networks arelocalized — again in contrast to the crystal networks. Fi-nally, a physical explanation has to explicate why it onlyapplies if the networks are optimized according to thethree stages of our procedure.

We believe the similar behaviors of S(α), DE(α), andφopt(α) provide important clues for understanding theuniversal curves of S(α) reported here, but we have notyet found a solid theoretical explanation that ties thesedifferent observations together to explain the universalbehaviors. The challenge is to identify how, from a com-bination of highly nonlinear physics and a diverse rangeof network geometries and topologies explored here, thesebehaviors emerge.

As a practical application, the discovery of this univer-sal behavior at large α makes it possible to estimate fora given type of optimized heterostructure the photonicband gap-to-midgap ratio ∆(α) for all α once one hasdetermined it for a single α without any further extensivecomputations. On the more theoretical side, the fact thatthe behavior applies to optimized structures in three di-mensions, whether crystalline or disordered, as well as toperiodic layered media, indicates an unanticipated sim-plicity, despite the apparent nonlinear mathematics andcomplex physics involved in the PBG computation andthe optimization procedure. A better understanding ofthis behavior will ultimately help to identify the relationbetween the structural features of optimized heterostruc-tures and the formation of large PBGs, which may guidethe design of improved photonic heterostructures.

We thank G. T. Barkema and N. Mousseau for pro-

viding samples of CRNs and the Princeton Institutefor Computational Science and Engineering (PICSciE)for the computational resources. This work was par-tially supported by the Princeton University InnovationFund for New Ideas in the Natural Sciences. S. T. alsogratefully acknowledges the support of the Air ForceOffice of Scientific Research Program on Mechanicsof Multifunctional Materials and Microsystems underGrant No. FA9550-18-1-0514. M. A. K. also acknowledgesfunding by the Volkswagenstiftung via the Experiment-Projekt Mecke.

[email protected][email protected][email protected]

[1] J. D. Joannopoulos, S. G. Johnson, J. N. Winn, and R. D.Meade, Photonic Crystals: Molding the Flow of Light,2nd ed. (Princeton University Press, Princeton, 2008).

[2] M. Maldovan and E. L. Thomas, Nat. Mater. 3, 593(2004).

[3] C. T. Chan, K. M. Ho, and C. M. Soukoulis, Europhys.Lett. 16, 563 (1991).

[4] W. Man, M. Megens, P. J. Steinhardt, and P. M. Chaikin,Nature 436, 993 (2005).

[5] M. C. Rechtsman, H.-C. Jeong, P. M. Chaikin,S. Torquato, and P. J. Steinhardt, Phys. Rev. Lett. 101,073902 (2008).

[6] M. Florescu, S. Torquato, and P. J. Steinhardt, Proc.Natl. Acad. Sci. U.S.A. 106, 20658 (2009).

[7] K. Edagawa, S. Kanoko, and M. Notomi, Phys. Rev. Lett.100, 013901 (2008).

[8] S. Imagawa, K. Edagawa, K. Morita, T. Niino, Y. Ka-gawa, and M. Notomi, Phys. Rev. B 82, 115116 (2010).

[9] S. F. Liew, J.-K. Yang, H. Noh, C. F. Schreck, E. R.Dufresne, C. S. O’Hern, and H. Cao, Phys. Rev. A 84,063818 (2011).

[10] H. Yin, B. Dong, X. Liu, T. Zhan, L. Shi, J. Zi, andE. Yablonovitch, Proc. Natl. Acad. Sci. U.S.A. 109,10798 (2012).

[11] K. Edagawa, Sci. Technol. Adv. Mater. 15, 034805(2014).

[12] N. Muller, J. Haberko, C. Marichy, and F. Scheffold, Adv.Opt. Mater. 2, 115 (2014).

[13] S. Tsitrin, E. P. Williamson, T. Amoah, G. Nahal, H. L.Chan, M. Florescu, and W. Man, Sci. Rep. 5, 13301(2015).

[14] S. Imagawa and K. Edagawa, Appl. Phys. A 123, 41(2017).

[15] S. R. Sellers, W. Man, S. Sahba, and M. Florescu, Nat.Commun. 8, 14439 (2017).

[16] X. Li, A. Das, and D. Bi, Proc. Natl. Acad. Sci. U.S.A.115, 6650 (2018).

[17] H. Men, K. Y. K. Lee, R. M. Freund, J. Peraire, andS. G. Johnson, Opt. Express 22, 22632 (2014).

[18] S. Torquato, Phys. Rep. 745, 1 (2018).[19] M. Hejna, P. J. Steinhardt, and S. Torquato, Phys. Rev.

B 87, 245204 (2013).[20] V. L. Deringer, N. Bernstein, A. P. Bartok, M. J. Cliffe,

R. N. Kerber, L. E. Marbella, C. P. Grey, S. R. Elliott,

6

and G. Csanyi, J. Phys. Chem. Lett. 9, 2879 (2018).[21] M. A. Klatt, J. Lovric, D. Chen, S. C. Kapfer, F. M.

Schaller, P. W. A. Schonhofer, B. S. Gardiner, A.-S.Smith, G. E. Schroder-Turk, and S. Torquato, Nat. Com-mun. 10, 811 (2019).

[22] S. Johnson and J. Joannopoulos, Opt. Express 8, 173(2001).

[23] The overall CPU time utilized for this study was about70 years on Intel Skylake CPUs.

[24] P. Yeh, Optical Waves in Layered Media (Wiley, Hobo-ken, NJ, 1988).

[25] H. A. Macleod, Thin-Film Optical Filters, 4th ed. (CRCPress, Boca Raton; London; New York, 2010).

[26] M. A. Klatt, P. J. Steinhardt, and S. Torquato, Proc.Natl. Acad. Sci. U.S.A. 116, 23480 (2019).

[27] S. T. Hyde, M. O’Keeffe, and D. M. Proserpio, Angew.Chem. Int. Ed. 47, 7996 (2008).

[28] B. D. Wilts, B. Apeleo Zubiri, M. A. Klatt, B. Butz,M. G. Fischer, S. T. Kelly, E. Spiecker, U. Steiner, andG. E. Schroder-Turk, Sci. Adv. 3, e1603119 (2017).

[29] T. Sunada, Not. Am. Math. Soc. 55, 8 (2008).[30] H. S. Sozuer and J. W. Haus, J. Opt. Soc. Am. B 10, 296

(1993).

[31] M. Maldovan and E. L. Thomas, J. Opt. Soc. Am. B 22,466 (2005).

[32] G. T. Barkema and N. Mousseau, Phys. Rev. B 62, 4985(2000).

[33] S. Torquato and F. H. Stillinger, Phys. Rev. E 68, 041113(2003).

[34] S. Torquato, Phys. Rev. E 82, 056109 (2010).[35] S. Lloyd, IEEE Trans. Inf. Theory 28, 129 (1982).[36] Y. Liu, W. Wang, B. Levy, F. Sun, D.-M. Yan, L. Lu,

and C. Yang, ACM Trans. Graph. 28, 101:1 (2009).[37] Q. Du, M. Gunzburger, and L. Ju, Numer. Math. 3, 119

(2010).[38] J. Zhang, M. Emelianenko, and Q. Du, Int. J. Numer.

Anal. Model. 9, 950 (2012).[39] C. Ruscher, J. Baschnagel, and J. Farago, Europhys.

Lett. 112, 66003 (2015).[40] C. Ruscher, S. Ciarella, C. Luo, L. M. C. Janssen,

J. Farago, and J. Baschnagel, J. Phys. Condens. Matter33, 064001 (2021).

[41] T. M. Hain, M. A. Klatt, and G. E. Schroder-Turk, J.Chem. Phys. 153, 234505 (2020).

[42] J. Haberko and F. Scheffold, Opt. Express 21, 1057(2013).

Supplemental Material:Gap Sensitivity Reveals Universal Behaviors in Optimized Photonic

Crystal and Disordered Networks

Michael A. Klatt,1, 2 Paul J. Steinhardt,3 and Salvatore Torquato3, 4

1Department of Physics, Princeton University, Princeton, New Jersey 08544, USA2Institut fur Theoretische Physik, University of Erlangen-Nurnberg, Staudtstr. 7, 91058 Erlangen, Germany

3Department of Physics, Princeton University, Princeton, NJ 08544, USA4Department of Chemistry, Princeton Institute for the Science and Technology of Materials,

and Program in Applied and Computational Mathematics,Princeton University, Princeton, New Jersey 08544, USA

(Dated: July 16, 2021)

S1. Simulation details

Here we provide additional parameters for our network models and the computation of their (optimal) photonicband gaps (PBGs).

Network models

For the nearly hyperuniform network (NHN), we used the most annealed sample with 1000 vertices from Ref. [1]. Forthe network based on the molecular dynamics (MD) quench, we used the sample GAP-MD_quench_1e11_lastMDstep

published together with Ref. [2]. The definition of our quantizer-based networks (QBNs) is based on the amorphousinherent structures of the quantizer energy from Ref. [3]; for references on the quantizer energy, see Refs. [3–10]. Toconstruct our QBN, we start from a binomial point process (that is, a snapshot of the ideal gas in the canonicalensemble) and minimize the quantizer energy using the so-called Lloyd algorithm [11]: In each step, the Voronoicenter of each cell is replaced by the center of mass of the cell. The algorithm converges to an amorphous inherentstructure [3]. Here, we apply 10,000 steps of the Lloyd algorithm. The network is then constructed using the Delaunaytessellation as explained in the main text. Our definition of the QBNs allows the simulation of extended networkswith a million vertices. For computational reasons, we here simulated a sample with 444 vertices.

The unit of length was chosen for each sample such that the number of vertices equals the volume of the sample.This choice corresponds to a unit number density ρ = 1 for the vertices. The number of vertices per sample is 1000for the NHN, 444 for the QBN, 512 for the MD quench, and 216 for the perturbed diamond networks. To check forsystem size effects, we also analyzed a continuous random network (CRN) by Barkema and Mousseau [12] with 216vertices and a perturbed diamond network (a = 0.2) with 1000 vertices; see Fig. S4.

A single sample of a perturbed diamond or disordered network can have several complete PBGs between differentbands. Typically, one of these PBGs is much larger than the others; see video S2 and Fig. S1. For each sample andeach value of α, we optimize the PBG for which we find the largest value of ∆(α) (with few exceptions as explainedbelow). The number of bands below this largest PBG is 1000 for the NHN, 509 for the MD quench, 444 for the QBNfor α > 9.3 (for α ≤ 9.3, we used the same band number although the PBG at band number 442 is slightly larger),1000 for the perturbed diamond network with 1000 vertices, 216 for the perturbed diamond networks with 216 verticesand with a = 0.1, 0.2, and 0.3, and 215 for the perturbed diamond network with a = 0.4. For the last sample, wefound several small complete PBGs of roughly the same size so that the number of bands below the largest gap candiffer for different values of R and α. The same applies to our perturbed CRN with a = 0.3.

Parameters

The resolution R, which is defined in the MIT Photonic Bands (MPB) software package as the number of voxelsper unit length, was set to the following values during the optimization of the radii:

• R = 16 for the disordered networks and perturbed diamond networks,

• R = 20 for the diamond, hexagonal-diamond, and Laves networks, and

arX

iv:2

107.

0721

6v1

[ph

ysic

s.op

tics]

15

Jul 2

021

2

• R = 80 for the simple cubic (SC) network.

For the final values of Figs. 2 and 3 in the main text, as well as for Fig. S4, we used the following resolutions:

• R = 16 for the perturbed diamond network with 1000 vertices,

• R = 20 for the disordered networks and perturbed diamond networks,

• R = 80 for the diamond, hexagonal-diamond, and Laves networks, and

• R = 160 for the simple cubic (SC) network.

The number of k points for which we computed the eigenfrequencies are

• 10 for the NHN, MD quench, and the QBN,

• 22 for the perturbed diamond networks and the perturbed CRNs.

For the disordered networks, the k points always include the high symmetry points (of the simple cubic simulationbox). For the crystal networks, we computed the eigenfrequencies at 8 points between pairs of high symmetry points.The tolerance of the MPB eigensolver was 10−5.

The calculations at strong contrasts are more prone to voxelization errors because of the smaller values of theoptimal radii. The estimation of the optimal radii is also more difficult at strong contrasts because the optimal gapsize as a function of the radius R becomes a skewed function (even close to the optimum). Finally, single defectsin disordered or perturbed networks can strongly influence the optimal gap size. We, therefore, used the followingimproved parameters for some of the calculations at strong contrasts (to check their influence on the results, whichwas not found to be strong). To determine the optimal radii for the crystal networks, we used a resolution of 80 forthe diamond network with α ≥ 20 and the laves network with α ≥ 18. Moreover, we used a resolution of 20 for theCRN with α = 8.3, 11.3, and 13.0.

Finally, in our study of system size effects in Fig. S4, the computation for the perturbed diamond network with1000 vertices was computationally particularly expensive because of the large number of k-points that have to beconsidered (because the perturbed diamond network has an anisotropic band structure—in contrast to the NHN).Therefore, we did not optimize the radii separately but used the optimal radii of the perturbed diamond networkwith 216 vertices. We checked for α = 4.8 and 13 that these radii are too a good approximation also optimal forthe network with 1000 vertices. We also checked that the resolution R = 16 was sufficient for our calculations bycomparing the results for two stop gaps with the corresponding results for a resolution R = 20. The differences inthe gap-to-midgap ratios were about 0.1%, that is, within the accuracy of our study.

Volume fractions

We estimate the volume fractions φ of our networks using the dielectric filling fraction implemented in MPB.The MPB software smooths the discontinuous dielectric function at the resolution of the grid (to avoid convergenceproblems caused by the discretization) [13]. The dielectric filling fraction is then defined as:

ϕ :=〈ε〉 −min ε

max ε−min ε, (S1)

where 〈ε〉, min ε, and max ε are the average, minimal, and maximal dielectric constant over all voxels. For a two-phasemedium that consists of a high and a low dielectric material, the filling fraction ϕ is equivalent to the volume fractionφ of the high dielectric phase.

[1] M. Hejna, P. J. Steinhardt, and S. Torquato, Phys. Rev. B 87, 245204 (2013).[2] V. L. Deringer, N. Bernstein, A. P. Bartok, M. J. Cliffe, R. N. Kerber, L. E. Marbella, C. P. Grey, S. R. Elliott, and

G. Csanyi, J. Phys. Chem. Lett. 9, 2879 (2018).[3] M. A. Klatt, J. Lovric, D. Chen, S. C. Kapfer, F. M. Schaller, P. W. A. Schonhofer, B. S. Gardiner, A.-S. Smith, G. E.

Schroder-Turk, and S. Torquato, Nat. Commun. 10, 811 (2019).[4] Y. Liu, W. Wang, B. Levy, F. Sun, D.-M. Yan, L. Lu, and C. Yang, ACM Trans. Graph. 28, 101:1 (2009).

3

[5] Q. Du, M. Gunzburger, and L. Ju, Numer. Math. 3, 119 (2010).[6] S. Torquato, Phys. Rev. E 82, 056109 (2010).[7] J. Zhang, M. Emelianenko, and Q. Du, Int. J. Numer. Anal. Model. 9, 950 (2012).[8] C. Ruscher, J. Baschnagel, and J. Farago, Europhys. Lett. 112, 66003 (2015).[9] C. Ruscher, S. Ciarella, C. Luo, L. M. C. Janssen, J. Farago, and J. Baschnagel, J. Phys. Condens. Matter 33, 064001

(2021).[10] T. M. Hain, M. A. Klatt, and G. E. Schroder-Turk, J. Chem. Phys. 153, 234505 (2020).[11] S. Lloyd, IEEE Trans. Inf. Theory 28, 129 (1982).[12] G. T. Barkema and N. Mousseau, Phys. Rev. B 62, 4985 (2000).[13] S. Johnson and J. Joannopoulos, Opt. Express 8, 173 (2001).

4

S2. Photonic band structures

0

0.3

0.6

Γ X WK Γ L UW L K U X

Fre

quen

cyω

a/2

πc

Diamond network with α = 4.6

0

0.3

0.6

Γ X WK Γ L UW L K U X

Fre

quen

cyω

a/2

πc

Diamond network with α = 13.0

0

0.15

0.3

0.45

0.6

0.75

Γ M K Γ A L H A M L K H

Fre

quen

cyω

a/2

πc

Hexagonal diamond network with α = 4.6

0

0.15

0.3

0.45

0.6

0.75

Γ M K Γ A L H A M L K H

Fre

quen

cyω

a/2

πc

Hexagonal diamond network with α = 13.0

0

0.15

0.3

Γ X M R Γ M X R

Fre

quen

cyω

a/2

πc

Perturbed dia. network (a = 0.2) with α = 4.6

0

0.15

0.3

Γ X M R Γ M X R

Fre

quen

cyω

a/2

πc

Perturbed dia. network (a = 0.2) with α = 13.0

FIG. S1. Photonic band structures for some of the (anisotropic) crystal and perturbed crystal networks at different dielectriccontrasts α; see also video S2.

5

0

0.15

0.3

Γ X M R Γ M X R

Fre

quen

cyω

a/2

πc

NHN with α = 4.6

0

0.15

0.3

Γ X M R Γ M X R

Fre

quen

cyω

a/2

πc

NHN with α = 13.0

0

0.15

0.3

Γ X M R Γ M X R

Fre

quen

cyω

a/2

πc

MD quench with α = 4.6

0

0.15

0.3

Γ X M R Γ M X R

Fre

quen

cyω

a/2

πc

MD quench with α = 13.0

0

0.15

0.3

Γ X M R Γ M X R

Fre

quen

cyω

a/2

πc

QBN with α = 9.6

0

0.15

0.3

Γ X M R Γ M X R

Fre

quen

cyω

a/2

πc

QBN with α = 19.0

FIG. S1. (Continued) Photonic band structures of (isotropic) disordered networks at different dielectric contrasts α; see alsovideo S2.

6

S3. Derivative of the optimal volume fraction

0

1

2

4 10 16 22

Disordered

networks

Crystals

DE

(α)

α

Diamond

Laves

Hex. diamond

SC

NHN

MD quench

QBN

FIG. S2. The average change of the electric field intensity with the dielectric contrast α for the bands below (open symbols)and above (solid symbols) the PBG. At low α, the configuration of the electric field changes more strongly for the disorderednetworks than for the photonic crystals.

S4. Derivative of the optimal volume fraction

-0.04

-0.02

0

4 10 16 22

op

t(α

)dα

α

1D Quarter-wave stack

NHN

MD quench

QBN

Diamond

Laves

Hex. diamond

SC

FIG. S3. The derivative of the optimal volume fraction, dφopt(α)/dα for our crystal and disordered networks approximately

agrees at large α with the analytic result for the 1D quarter-wave stack, dφQWS(α)/dα = −1/(2α3/2 + 4α + 2√α), shown by

the dashed-dotted line.

7

S5. System size effects

0%

10%

20%

30%

4 5 6 7 8 9 10 11 12 13

0%

0.7%

1.4%

4.2 4.5 4.8∆(α

)

α := ε2/ε1

Pert. Dia. a = 0.2, 216 nodes

Pert. Dia. a = 0.2, 1000 nodes

CRN, 216 nodes

NHN, 1000 nodes

0

0.02

0.04

0.06

4 5 6 7 8 9 10 11 12 13

S(α

):=

d∆

(α)/

α := ε2/ε1

Pert. Dia. a = 0.2, 216 nodes

Pert. Dia. a = 0.2, 1000 nodes

CRN, 216 nodes

NHN, 1000 nodes

FIG. S4. The gap plot (top) and gap-sensitivity plot (bottom) for the NHN (1000 vertices) and the perturbed diamond networkwith a = 0.2 (216 vertices) are compared to similar networks at different system sizes: a CRN with 216 vertices and a perturbeddiamond network with 1000 vertices, respectively. The curves at different system sizes agree within the systematic errors andstatistical fluctuations that can be expected between different samples; that is, no strong system size effect is observed.