ArticlePDF Available

Abstract and Figures

The interaction of a material with light is intimately related to its wavelength-scale structure. Simple connections between structure and optical response empower us with essential intuition to engineer complex optical functionalities. Here we develop local self-uniformity (LSU) as a measure of a random network’s internal structural similarity, ranking networks on a continuous scale from crystalline, through glassy intermediate states, to chaotic configurations. We demonstrate that complete photonic bandgap structures possess substantial LSU and validate LSU’s importance in gap formation through design of amorphous gyroid structures. Amorphous gyroid samples are fabricated via three-dimensional ceramic printing and the bandgaps experimentally verified. We explore also the wing-scale structuring in the butterfly Pseudolycaena marsyas and show that it possesses substantial amorphous gyroid character, demonstrating the subtle order achieved by evolutionary optimization and the possibility of an amorphous gyroid’s self-assembly.
Content may be subject to copyright.
ARTICLE
Received 13 Jul 2016 |Accepted 30 Dec 2016 |Published 17 Feb 2017
Local self-uniformity in photonic networks
Steven R. Sellers1, Weining Man2, Shervin Sahba2& Marian Florescu1
The interaction of a material with light is intimately related to its wavelength-scale structure.
Simple connections between structure and optical response empower us with essential
intuition to engineer complex optical functionalities. Here we develop local self-uniformity
(LSU) as a measure of a random network’s internal structural similarity, ranking networks on
a continuous scale from crystalline, through glassy intermediate states, to chaotic config-
urations. We demonstrate that complete photonic bandgap structures possess substantial
LSU and validate LSU’s importance in gap formation through design of amorphous gyroid
structures. Amorphous gyroid samples are fabricated via three-dimensional ceramic printing
and the bandgaps experimentally verified. We explore also the wing-scale structuring in the
butterfly Pseudolycaena marsyas and show that it possesses substantial amorphous gyroid
character, demonstrating the subtle order achieved by evolutionary optimization and the
possibility of an amorphous gyroid’s self-assembly.
DOI: 10.1038/ncomms14439 OPEN
1Advanced Technology Institute and Department of Physics, University of Surrey, Guildford GU2 7XH, UK. 2Department of Physics and Astronomy, San
Francisco State University, 1600 Holloway Avenue, San Francisco, California 94132, USA. Correspondence and requests for materials should be addressed to
S.R.S. (email: steven.sellers@surrey.ac.uk) or to M.F. (email: m.florescu@surrey.ac.uk).
NATURE COMMUNICATIONS | 8:14439 | DOI: 10.1038/ncomms14439 | www.nature.com/naturecommunications 1
Acomplete photonic bandgap (PBG) is frequency window
within which a material, by virtue of its structure,
supports no propagating electromagnetic modes. Typi-
cally, structures which possess complete PBGs are periodic arrays
of dielectric material; such arrays are called photonic crystals
(PhCs). PhCs have the potential to play a key role in the
development of next-generation photonic integrated circuits1–3.
However, although the complexity of PhC-based technologies
continues to grow, questions regarding the fundamental
mechanisms of PBG formation remain unresolved4–6.
The formation of PBGs is conventionally interpreted as a result
of coherent scattering by a PhC’s lattice planes7,8. In this picture,
a plane wave may be scattered onto its counter-propagating
equivalent when the wavevector of the initial state lies on the edge
of the PhC’s Brillouin zone (BZ). When this condition is met, a
pair of orthogonal standing wave modes, each possessing a
distinct electromagnetic field profile, is formed5,9,10. Energetic
interaction between the electric field and the underlying dielectric
distribution then breaks the degeneracy of the standing wave
states. For the specific propagation direction under consideration,
the resulting forbidden spectral range defines a photonic stop gap.
To engineer a complete PBG, photonic stop gaps must open
along all propagation directions. Further, these stops gaps must
be spectrally aligned. Both these considerations can be addressed
by designing PhCs to possess maximally spherical BZs11. The
search for the first complete PBG thus focussed on face-centred
cubic crystals—the most isotropic of the three-dimensional (3D)
Bravais lattices—and discovered a large complete gap in a
diamond-like network of dielectric material12. In spite of the
many PhC designs that have since been discovered, those based
on the diamond network remain the champion, possessing the
largest complete PBGs13,14.
There is, however, much evidence to suggest that PBG
formation is governed by more than just coherent scattering
processes. PhCs derived from the body-centred cubic single-
network gyroid (SNG) structure (triamond) and low-symmetry
diamond embeddings all possess near-champion PBGs in spite of
their less spherical BZs13. Although based on a face-centred cubic
lattice, the inverse opal network exhibits a complete PBG only
one quarter the size of the champion diamond gap1. Most
tellingly, a glassy 3D network—dubbed photonic amorphous
diamond (PAD)—exhibits a sizeable complete PBG4. This gap
exists despite PAD’s diffuse primary diffraction maximum which
spreads the structure’s coherent scattering power of a range of
wavevectors5.
Similar evidence is found in two-dimensional structures that
possess PBGs for light with a transverse electric (TE) polarization.
Amongst these structures, the champion PhC design is a honey-
comb network of dielectric material15. Many glassy networks that
can be broadly styled as ‘hyperuniform disordered honeycombs’
have been found to possess sizeable TE PBGs6,16–20. As with
PAD, these gaps exist despite the diminished coherent scattering
power of each structure.
Here, we address the mechanisms governing PBG formation by
re-formulating the ideal structural properties of a PBG-forming
network. To achieve this, we introduce the concept of local self-
uniformity (LSU). LSU measures the geometrical and topological
similarities of the local environments in a connected network of
uniform valency. We note that existing sizeable PBG networks
possess significant LSU. We demonstrate the connection between
LSU and PBG forming ability by designing novel amorphous
gyroid (amorphous triamond) connected networks. Specifically,
amorphous gyroids can possess sizeable PBGs and an amorphous
gyroid’s LSU is strongly correlated with its PBG width.
This correlation is explained by recognizing the contribution of
spatially localized resonant scattering processes to PBG formation
in connected networks. Locally self-uniform ceramic 3D-printed
amorphous gyroids are characterized through microwave trans-
mission experiments. We explore also the possibility that
amorphous gyroid exists within the wing scales of butterflies. In
particular, we reveal that the microstructure in the scales of
Pseudolycaena marsyas possesses substantial amorphous gyroid
character and demonstrate that the butterfly’s reflectance
spectrum can be effectively reproduced by amorphous gyroid
microstructures.
Results
Local self-uniformity. The exact structure of glasses has long
been debated21. Recent research has demonstrated the complex
interplay of ordered and disordered phases in the vitreous
state22–24. The disparate variety of bulk metallic glasses in
particular has challenged researchers to develop predictive
theories of an alloy’s glass-forming ability25. The existence of
sizeable PBGs in glassy networks exposes an analogous deficiency
in current understanding of a structure’s PBG forming ability.
Unlike silicate and metallic glasses, the structures of glassy PBG
materials are designed. Nevertheless, the structural characteristics
that render these glasses amenable to PBG formation remain
mysterious. With the aim of clarifying these PBG forming
characteristics, here we develop LSU a general measure of
structural order in connected networks.
A typical PhC consists of a connected distribution of dielectric
material surrounded by air. As an example, Fig. 1a,b show the
champion PhC diamond; it is a connected network of dielectric
cylinders (green) arranged as in a diamond crystal. A PBG’s size
is usually measured as a dimensionless width given by Do/o
0
the frequency width divided by the central frequency. Figure 1c
displays an amorphous version of the diamond network. The
complete PBG in diamond has been shown to have a width of
30% for dielectric material of refractive index 3.6 (ref. 13). We
focus our discussion of PBG properties hereafter in this high-
refractive index regime.
To describe a general connected network, we decompose it
into a set of vertices and edges. A vertex is a point at which two
or more distinct lobes of material intersect. An edge is a vector
between two vertices that specifies the central axis of a lobe of
material. As an illustration, Fig. 1a shows the fundamental unit
of a diamond network in its Wigner–Seitz (WS) cell. A vertex,
which sits at the centre of the cell, has four edges connecting it
to its nearest neighbour vertices. The four nearest neighbours sit
at the corners of a tetrahedron and the four edges define a
tetrahedral unit. Similarly, Fig. 1d shows the fundamental unit
of a SNG structure in its WS cell. The central vertex has three
nearest neighbour vertices to which it is connected by three
edges. These edges are of equal length and are separated by
inter-edge angles of 120°;wecallthiscongurationatrihedral
unit.
The WS cell is the basic building block from which an extended
periodic network can be assembled. Stacking the WS cells of
diamond and SNG produces extended diamond (Fig. 1b) and
gyroid (Fig. 1e) networks, respectively. An extended periodic
network of this type is a highly ordered case of a continuous
random network (CRN). A general CRN is a collection of
vertices, connected by edges, which are arbitrarily positioned in
space26. CRNs do not typically possess translational periodicities
as the diamond and gyroid crystals do. Instead, Fig. 1c illustrates
a classic CRN—an amorphous diamond network, studied
extensively as a model of amorphous silicon27,28. Every vertex
in amorphous diamond is tetravalent and has four nearest
neighbours that describe the points of a deformed tetrahedron.
The resulting network is a connected assembly of deformed
ARTICLE NATURE COMMUNICATIONS | DOI: 10.1038/ncomms14439
2NATURE COMMUNICATIONS | 8:14439 | DOI: 10.1038/ncomms14439 | www.nature.com/naturecommunications
tetrahedral units. Similarly, an amorphous gyroid is a CRN comp-
rising deformed trihedral units as depicted in Fig. 1f.
We now introduce the concept of a tree within a CRN. An
n-tree on a vertex aof a CRN is the set of edges and vertices that
can be reached from aby traversing no more than nedges. The
tree’s ‘root edges’ are understood to mean the edges belonging to
vertex a. To illustrate this, the diamond WS cell (Fig. 1a) contains
a 1-tree. The SNG WS cell (Fig. 1d) contains a single 2-tree.
A tree thus describes a small network unit within a CRN.
We define LSU as a property of a CRN that describes the extent
to which its component n-trees have shapes similar to that of
nearby n-trees within the network. For the moment, we consider
comparing some example 1-trees. Such trees consist of a central
vertex with a number of edges g. At first, we consider two
trihedral 1-trees which we label T1
aand T1
b.
To measure the extent to which these two trees have similar
shape we follow the process shown in Fig. 2a. First, we label the
root edges of both trees as (1
a
,2
a
,3
a
) and (1
b
,2
b
,3
b
). Second, we
specify a permutation of the labels of tree b, say [213]. Third, we
attempt to align the two trees such that the root edges overlap in
the prescribed permutation. This alignment is achieved by
performing a congruent transformation on tree T1
b. That is, we
physically pick up the tree and perform any combination of a
translation, a rotation and a reflection until the edges of both trees
overlap as desired29. Fourth, we score the extent to which T1
aand
T1
bnow overlap by taking the scalar product of the edge vectors in
pairs as determined by the permutation. The result of this
comparison is normalized such that perfectly overlapping trees
receive a maximum score of 1.
Finally we repeat steps 1 through 4 and score the overlap of T1
a
and T1
bfor all g! permutations of their root-edge alignment. We
define the spatial similarity statistic f
ab
of these two trees as the
average result of the g! comparisons (see the Methods section).
This process can be applied without modification to determine
the spatial similarity statistic of two identical tetrahedral 1-trees,
as shown in Fig. 2b. As before, we label the root edges of the two
trees and specify some arbitrary permutation of the root edges of
tree b; Fig. 2b shows the permutation [2413]. Once permuted, tree
bis congruently transformed by a translation, a reflection and a
rotation to maximally overlap with tree a. The quality of the
overlap is then measured. The average overlap quality for all 4!
root-edge alignment permutations defines the spatial similarity
statistic f
ab
.
In Fig. 2a,b, we note that that the two trihedra and tetrahedra
could be made to overlap exactly for their respective root edge
permutations. In fact, pairs of either trihedra or tetrahedra will
overlap exactly for all root edge permutations and have a spatial
similarity statistic of unity. This property of trihedral and
tetrahedral units is a corollary of work on strong isotropy29.
There exist only three strongly isotropic networks in three
dimensions or less. In three dimensions, these networks are
diamond (Fig. 1b) and single gyroid (Fig. 1e). In two dimensions,
the honeycomb network is the only strongly isotropic network.
We consider now a third set of example 1-trees—two identical
simple cubic network units (Fig. 2c), each with six edges about a
central vertex—and calculate their spatial similarity statistic. We
find that perfect overlap can be achieved for specific root edge
permutations, but in most cases there is no congruent
transformation to align the two trees correctly; we call such a
permutation inaccessible (this is depicted in Fig. 2c). As a result,
the simple cubic crystal does not possess a strong isotropy
property. In the remainder of this paper, we focus on trees
comprising trivalent and tetravalent vertices only.
To build a robust measure of network structural order, we
require a formalism that can compare the trees of arbitrarily
disordered networks. We thus build upon the measurement of
tree spatial similarity statistics in the following ways. We stipulate
that the CRN from which the trees are drawn comprises only
vertices with a fixed number of edges. We then generalize the
measurement of tree spatial similarities to n-trees of any size. In
this case, the correct edge overlap pairings are not known beyond
the root edges. To solve this, we have developed an algorithm
ab c
de f
Figure 1 | Strongly isotropic networks and their amorphous derivatives in three dimensions. The WS cell of a diamond PhC (a) contains a 1-tree of
dielectric material (green); the edges of the tree define a tetrahedral unit. Stacking of the WS cells generates the champion diamond PhC (b). Amorphous
diamond (c) consists of deformed tetrahedral units connected in a CRN. Similarly, the WS cell of a single-network gyroid PhC contains a 2-tree (d),
comprising dielectric material (blue), in which each vertex is a trihedron. When stacked, the WS cells generate a single-network gyroid PhC with a near-
champion PBG (e). Amorphous gyroid (f) comprises deformed trihedral units connected in a CRN.
NATURE COMMUNICATIONS | DOI: 10.1038/ncomms14439 ARTICLE
NATURE COMMUNICATIONS | 8:14439 | DOI: 10.1038/ncomms14439 | www.nature.com/naturecommunications 3
which determines the natural edge overlaps; this is detailed in the
Methods section. We also allow the similarities of arbitrarily
disordered trees to be measured. Here we acknowledge that the
root edges of two trees will not necessarily overlap perfectly.
Instead, for each permutation we find a transformation that yields
an approximate overlap. We then quantify the quality of this
overlap according to a metric of spatial similarity (see Methods
section).
We now define a CRN’s LSU distribution F
nl
to depth nand
locality las the set of n-tree spatial similarities f
ab
for all pairs of
trees whose root vertices are within ledges of one another. The
LSU distribution is a set of spatial similarities that describes the
extent to which the CRN’s geometry is uniform on the length-
scale l. For example, F
22
describes the spatial similarities for all
pairs of trees of depth 2 separated by two edges or less.
We present a set of example LSU distributions in Fig. 3.
Figure 3a shows F
22
for an amorphous diamond network. The
trees in amorphous diamond are non-identical and present a
broad distribution of spatial similarity statistics. However, the
spatial similarities remain relatively large and show that the local
geometries in the CRN are similar. Fig. 3b compares three LSU
distributions, F
12
,F
22
and F
32
, for amorphous diamond.
Amorphous networks have strong positional correlations at local
length-scales that fade with distance; as a result, trees become
Tree a
Tree b
1. Label 1a
2a
3a
4a
5a
6a
1b
2b
3b
4b
6b
5b
Tree a
1a
Tree b
2a
3a
1. Label
4a
1b
2b
3b
4b
Tree a
1a
Tree b
1b
2b
2a
3a
3b
1. Label
2. Permute 1a
2a
3a
6a
4a
5a
1b
2b
3b4b
5b
6b
1a
2a
3a
4a
2. Permute
1b
2b
3b
4b
2. Permute
2b
1b
3b
1a
2a
3a
3. Overlap
3. Overlap
3. Overlap
Perfect
Perfect
Imperfect
a
b
c
Figure 2 | Comparison of 1-trees illustrated with trihedra, tetrahedra and octahedra. Two identical trihedral trees (a) are labelled by their edges. The
edges of tree bare then permuted. Tree bcan then be rotated around the edge 3
b
axis and made to perfectly overlap tree a. Similarly, two identical
tetrahedral trees are labelled (b), and then the edges of tree bpermuted. Reflection of tree B in the plane of edges 2
b
and 4
b
, followed by rotation around the
new edge 3
b
axis, brings the two trees into alignment. Two identical octahedral trees (c) can also be compared. We apply a permutation to tree B’s edges.
When overlapped, the two trees are now mismatched; their yellow and purple edges are not aligned and cannot be made so by any congruent
transformation without creating a new mismatch.
ARTICLE NATURE COMMUNICATIONS | DOI: 10.1038/ncomms14439
4NATURE COMMUNICATIONS | 8:14439 | DOI: 10.1038/ncomms14439 | www.nature.com/naturecommunications
more spatially dissimilar as their depths increase. We note that
strongly isotropic networks are the only CRNs whose F
nl
comprise a single peak at unity for all tree depths and localities;
they are uniquely identified by their LSU distributions. This
property is a consequence of the definition of strong isotropy29.
The LSU distributions are thus naturally interpreted as a
continuous measure of the extent to which a network is
strongly isotropic.
LSU distributions can also be used as a diagnostic of the phases
present in a CRN. Figure 3c presents F
22
for an amorphous
gyroid sample containing a gyroid crystallite surrounded by an
amorphous network. The distribution comprises a background
spectrum from which a clear peak emerges Bf¼0.98. The peak
signifies the existence of network regions comprising a strained
crystalline gyroid phase. The LSU distribution is thus diagnostic
of separate phases within a system and is useful for probing the
local order of complex systems undergoing phase transitions22.
Champion photonic bandgap architectures. 3D architectures
which exhibit large complete PBGs, together with two-dimen-
sional architectures with large TE PBGs, possess a number of
shared structural characteristics. First, successful architectures are
connected networks of dielectric material5,18,30. Second, the
champion structures—the honeycomb and diamond networks in
two and three dimensions, respectively—are both strongly
isotropic. SNG, which is the only other strongly isotropic
network, exhibits a near-champion PBG of width 28% for a
refractive index contrast of 3.6:1 (see Supplementary Fig. 1).
Third, amorphous derivatives of both honeycomb and diamond
networks exhibit PBGs which are sizeable but smaller than the
gaps of their parent crystalline networks. Crucially, both networks
possess significant but imperfect LSU (see Fig. 3a).
Overall, evidence suggests that the extent of a network’s LSU
influences its PBG forming ability. Hence, we expect that a
hypothetical amorphous SNG, analogous to PAD4and disordered
honeycomb6, should possess both a high degree of LSU and a
sizeable complete PBG. To the best of our knowledge, amorphous
gyroids have not been observed in any context. We thus under-
take the design of amorphous gyroids as a means of testing the
relationship between LSU and PBG forming ability. To achieve
this, we apply the Wooten-Winer-Weaire (WWW) algorithm to
anneal amorphous gyroids from random seed networks27,28.In
order to yield faithfully gyroidal local geometries, networks are
annealed using a modified potential energy function that is
distinct to the regular Keating energy27 (see also Supplementary
Methods).
We generated a set of amorphous gyroid models of varying
total size as measured by their number of component vertices N.
In particular, we generated an ensemble of 57, 216-vertex models
across a spectrum of disorder. PBGs were probed by numerical
solution of the Maxwell equations via a plane wave expansion
method31. A refractive index contrast of 3.6:1 was used for all
calculations. Bandgap widths were measured as a percentage of
their central frequency.
We demonstrate that amorphous gyroid networks can possess
sizeable PBGs. The average PBG measured for a set of high-
quality 1,000-vertex networks was 16%. The largest single gap
observed was for a well-annealed 216-vertex network and had a
size of 21%. These gap widths compare favourably with the 18%
gap in PAD at the same refractive index contrast4.
We investigated our ensemble of 216-vertex networks closely,
calculating the LSU F
22
distribution for each network. In Fig. 4a
we plot the mean value
F22 of each distribution against PBG
width for all 57 networks. We see that gap width is strongly
correlated with network LSU. We expect PBGs in amorphous
gyroid to open within the gap region of a SNG PhC of equivalent
index contrast. We thus defined the SNG gap as a critical
frequency region, and counted the number of photonic bands
that each of our 216-vertex networks support within this window.
The number of bands within the critical frequency region can be
interpreted as an integrated density of states (DOS). We plot the
integrated DOS against
F22 in Fig. 4b. We see that a network’s
inability to support electromagnetic modes is strongly correlated
with its LSU (see also Supplementary Figs 2 and 3 and Supple-
mentary Notes 1 and 2). Together these results demonstrate the
power of LSU as a predictor of PBG forming ability.
We now discuss two characteristic types of amorphous gyroid
in detail; we call these type-1 and type-2 networks. We
characterize the networks through histograms of their geome-
trical properties. In particular, we measure edge lengths (d), inter-
edge angles (y), dihedral angles (f) and skew angles (w). Note
that the skew angle measures the coplanarity of a trihedral unit,
with w¼p/2 representing a flat trihedron (see Supplementary
Methods).
Figure 4c,d show typical frequency distributions of d,y,f
and wfor type-1 networks. They are characterized by strongly
peaked edge-length and inter-edge angle distributions, but have
non-uniform dihedral and skew angles. Type-1 networks can
thus be considered high-quality amorphous networks of
trihedral 1-tree units. Figure 4f,g show the same distributions
for a typical type-2 network. Its d,y,fand wdistributions are
all peaked around the ideal values for gyroidal vertices.
Compared with type-1 networks, the local geometries of type-
2 structures are much closer to an ideal strongly isotropic
configuration. They possess gyroidal structural order on the
length-scale of a 2-tree unit—the fundamental building block of
SNG (Fig. 1d). Type-1 networks have
F22 values around 0.72.
Type-2 networks have significantly higher LSUs with
F22s
0.18
Φ12
Φ22
Φ32 0.12
0.10
0.06
0.02
0.06
0.12
0.08
0.04
Relative frequency
0.10
0.06
0.02
Type-1
Type-2
0.7 0.8 0.9 1 0.7 0.8 0.9 1
Spatial similarity
0
ab
cd
Figure 3 | Example LSU distributions. (a) The distribution of spatial
similarity statistics for tree comparisons of depth 2 and locality 2 (F
22
) for
an amorphous diamond network. (b) LSU distributions F
n2
of amorphous
diamond for different tree depths n.(c)F
22
for a 216-vertex sample
comprising a single gyroid crystallite suspended in amorphous gyroid,
demonstrating LSU’s ability to differentiate between phases. (d)F
22
distributions for two characteristic types of amorphous gyroid network—
type-1 (1-tree local order) and type-2 (2-tree local order).
NATURE COMMUNICATIONS | DOI: 10.1038/ncomms14439 ARTICLE
NATURE COMMUNICATIONS | 8:14439 | DOI: 10.1038/ncomms14439 | www.nature.com/naturecommunications 5
around 0.89. Typical F
22
distributions for these networks are
showninFig.3d.
Figure 4e,h show planar slices through the structure factors of
type-1 and type-2 networks, respectively. The structure factor is a
quantity that characterizes a structure’s coherent scattering power
as a function of wavevector. We see that the peak scattering power
of both network types is distributed in a circular ring on account of
the average isotropy of the structures; wavevectors that lie on this
ring are strongly scattered. We averaged the full structure factor
across all propagation directions and measured the peak scattering
power of type-2 networks to be approximately 1.6 times greater
than type-1. In spite of this small increase, type-2 networks take the
PBG width from practically zero to a maximum of 21%. We
attribute these radically different PBG widths to the formation of
locally self-uniform trees within type-2 networks.
We now demonstrate the natural connection between LSU
andPBGformationinconnectednetworks.Inadditionto
the coherent scattering (Bragg) mechanism, there exists
aMiescatteringmechanism
5,32 of PBG formation. The Mie
mechanism is known to be the dominant formation process for
transverse magnetic (TM) polarization gaps in dielectric
cylinder arrays. Specifically, sizeable PBGs exist in periodic10,
quasicrystalline18 and random6cylinder arrangements and are
observed to be of the same origin in each case32.
The gap originates from resonant scattering by the Mie modes
of a single cylinder. For all TM-gap cylinder arrays, it is clear
from the electric field profiles at the edges of the fundamental
PBG that Mie scattering mediates light propagation6,10,18. Just
below the gap, modes are characterized by localization of field
nodes in the vicinity of the cylinder surfaces. Just above the gap,
field nodes consistently bisect the dielectric cylinders. These two
node profiles derive from the interaction of a plane wave with an
isolated cylinder, and are associated with the first and second Mie
resonances, respectively32. Just above the first Mie resonance,
incident and scattered fields are in antiphase at the cylinder
surface; this creates a localized standing wavefront which inhibits
propagation and leads to PBG formation32.
Existing evidence suggests that PBG formation in two-
dimensional and 3D dielectric networks is governed by a similar
mechanism of resonant scattering. Specifically, careful examina-
tion of the gap-edge eigenmodes of all honeycomb-derived
networks presents a consistent picture of the nature of these
scattering resonances. In both crystalline and hyperuniform
disordered honeycombs6,18, modes just below the fundamental
PBG are characterized magnetic field nodes localized within the
dielectric network (conversely, field anti-nodes focus within the
air cells). Just above the PBG, magnetic field nodes pass between
air cells, cutting the inter-vertex dielectric walls almost normally.
These gap-edge node characteristics are consistent across
honeycomb-derived trivalent networks and, by analogy to the
TM case, evidence the significance of spatially localized resonant
scattering processes.
Type-2
Type-1
f. (a.u.)
60
50
40
30
20
10
0
25
20
15
10
5
0
200
150
100
50
0
f. (a.u.)
200
150
100
50
0
d/d0
0.8 1.21
200
150
100
f. (a.u.)
50
0
8
4
0
–4
–8 –1.5
–0.5
0
–1
–1.5
–0.5
0
–1
–8 –4 0 4 8–8 –4 0 4 8
8
4
0
kya
kya
–4
–8
f. (a.u.)
200
150
100
50
0
66666
0
π5π4π3π2ππ
d/d0
0.8 1.21
666
Angle Angle
66
0π5π4π3π2ππ
kxa
0.7 0.75 0.8 0.85 0.9
N () (a.u.)
kxa
I
I0
Φ22
Φ22
0.7 0.75 0.8 0.85 0.9
Δ
0
acf
dg
he
b
Figure 4 | LSU and PBG forming ability. The LSU of amorphous gyroid networks, as measured by the mean of their F
22
distributions, is strongly correlated
with PBG width (a). Similarly, the integrated DOS N(o) decreases smoothly with increasing F
22
(b). Fit lines (cubic polynomials) are for visualization
purposes only. Approximate LSU regions for type-1 and type-2 networks are indicated (a, orange). We present also the edge length frequency (f)
distributions (c,f), inter-edge (y), dihedral (f) and skew (w) angle frequency distributions (d,g) and structure factor slices (e,h) for typical type-1 and type-2
networks, respectively. Structure factor intensities Iare plotted on logarithmic colour scales and data in both panels is normalized to the maximum intensity
I
0
of h.
ARTICLE NATURE COMMUNICATIONS | DOI: 10.1038/ncomms14439
6NATURE COMMUNICATIONS | 8:14439 | DOI: 10.1038/ncomms14439 | www.nature.com/naturecommunications
We therefore view CRNs as a connected ensemble of distinct
scattering units (n-trees) which, in isolation, exhibit a number of
resonant electromagnetic modes. For frequencies just above a
resonance, scattered and incident fields are in antiphase and
interfere destructively, localizing a field node in the vicinity of the
scatterer and suppressing propagation through it.
A potential champion PBG structure comprises geometrically
identical scattering units. All scattering centres thus possess
degenerate electromagnetic resonances, and the spectral ranges
for which each scatterer inhibits transmission are maximally
aligned. Fixed valence networks comprising non-identical
scattering units exhibit smaller PBGs than their crystalline
precursors. Structural deformation of the scattering centres
breaks the degeneracy of their scattering resonances; the resulting
PBG is thus narrowed by imperfect overlap of the spectral ranges
for which each scattering centre suppresses transmission.
The set of networks comprising geometrically identical
scattering units exhibits a clear hierarchy of PBG size33 (see
Supplementary Table 1). Specifically, networks built from vertices
with a low coordination number possess the largest PBGs.
Accordingly, the diamond and SNG architectures are
champion and near-champion, respectively; these networks are
strongly isotropic due to the simplicity of their vertices and
comprise scattering units which are perfectly superimposable
under permutation. This combinatorial symmetry has a
strong influence on the PBG width. We argue that symmetry
under permutation minimizes the number of distinct scattering
resonances that a scattering centre supports. As a result, the
frequency gaps between scattering resonances are maximized,
together with the width of the spectral region above resonance for
which transmission is suppressed.
Fundamentally, both Bragg and resonant scattering mechan-
isms contribute to PBG formation. The largest PBGs are obtained
by optimization of a structure’s dielectric fill fraction to overlap
the spectral range associated with the two mechanisms. We note,
however, that strong diffraction rings in the structure factor of
amorphous materials do not directly lead to PBGs, but reflect the
presence of local order that, depending on its LSU, may favour
gap formation. This observation clarifies the relationship between
LSU and work on PBG formation in hyperuniform structures6.
Architectures derived from disordered hyperuniform point
patterns possess significant local structural correlations and
local geometrical order; these characteristics have proven
essential in establishing sizeable PBGs6. However, hyperuniform
point patterns must be tessellated in an ad-hoc way to produce
viable PBG-forming networks6,34. This tessellation protocol
naturally creates nearly-optimal network topologies, but these
networks are successful only because they possess locally self-
uniform structural order. In contrast to hyperuniformity, LSU
measures both geometrical and topological order simultaneously
and is thus an effective measure of PBG forming ability.
Hyperuniformity and LSU remain compatible; we note the
emergence of a hyperuniform-like exclusion domain around
k¼0 in the structure factor of networks with significant LSU
(Fig. 4h). The association of LSU with PBG formation parallels
the proof that amorphous materials with well-defined atomic
connectivity can possess an electronic bandgap35.
Microwave experiments with amorphous gyroid.Toverifyour
theoretical calculations, we fabricated millimetre-scale amor-
phous gyroid samples and experimentally characterized their
PBGs. Samples were produced at the Fraunhofer Institute for
Ceramic Technologies and Systems using a 3D ceramic printing
technique. The samples were made from alumina (Al
2
O
3
),
whose permittivity was experimentally determined to be
Er¼(9.5±0.3) at frequencies in the microwave Kband
(18–26.5 GHz). Two types of sample were made: cuboidal
samples of SNG (Fig. 5c) and amorphous gyroid (Fig. 5a,d), and
a cylindrical sample of amorphous gyroid (Fig. 5b). The internal
network, comprising cylinders with diameter D¼2.03 mm, was
well formed (Fig. 5c,d). The gyroid primitive cell parameter was
Norm. trans. (dB)
Raw trans. (dB)
–10
0
27 17 f (GHz)
θ
–25
–35
–45
–55
–20
–30
Norm. trans. (dB)
10
0
–10
–20
–30
–40
–50
15 20 25 30
Freq. (GHz)
SNG [111]
Cal. PBG
ae
Norm. trans. (dB)
10
0
–10
–20
–30
–40
–5015 20 25 30
Freq. (GHz)
A-Gyroid
Cal. PBG
f
gh
c
bd
Figure 5 | Microwave experiments with single network and amorphous gyroid structures. Alumina prototypes of amorphous gyroid, both a single piece
cuboid (a) and a compound cylinder (b). View down the (111) axis of our single-network gyroid sample (c), and along an arbitrary axis of amorphous gyroid
(d), showing the network quality. Comparison of transmission between single-network gyroid (SNG [111]) and amorphous gyroid (A-Gyroid) cuboidal
samples (e,f, respectively), with gap edges predicted by plane wave expansion (vertical dashed lines); measured data was smoothed by Fourier filtering for
clarity. Measured (g) and simulated (h) polar false-colour maps of transmission for the amorphous gyroid cylinder, with gap edges calculated by band
structure overlaid as black and white rings, respectively. Scale bars in (c,d) 5 mm.
NATURE COMMUNICATIONS | DOI: 10.1038/ncomms14439 ARTICLE
NATURE COMMUNICATIONS | 8:14439 | DOI: 10.1038/ncomms14439 | www.nature.com/naturecommunications 7
designed to be 3.08 mm and experimentally measured to be
a¼(3.13±0.05) mm.
Measurements were made with microwave radiation using an
established experimental set-up9,19. Each sample was placed
between two facing microwave horn antennae connected to an
HP-8510C vector network analyser (VNA). Figure 5e,f present
the transmission as a function of frequency along the (111) crystal
axis of SNG and an arbitrary amorphous gyroid axis, respectively.
Both samples show wide transmission gaps, with up to 35 dB of
attenuation. Propagation through the periodic sample (Fig. 5e,f)
is predominantly ballistic9; photons are weakly scattered both
above and below the PBG, and the gap edges are well defined. As
a result, the measured transmission gap for the SNG sample
agrees very well with the simulated PBG, highlighted by vertical
dashed lines in the figure.
The measured transmission gap for the amorphous gyroid
sample (Fig. 5f) is also centred at the frequency range calculated
by simulation (shown as dashed lines), although it appears wider;
we attribute this to the limited dynamic range of our experiment
(see Methods section) and strong diffusive scattering9. Scattering
processes are significant for the amorphous network. They scatter
radiation both away from the detector and into different
polarization states. As a result, strongly scattered radiation has
a low coupling efficiency into the receiver horn antenna;
this prevents the transmission from recovering its peak value
for frequencies above the PBG, thus widening the perceived
transmission gap.
We employed the amorphous gyroid cylindrical sample
(Fig. 5b) to investigate the isotropy of the PBG. The sample
was rotated around its cylindrical axis and a transmission
spectrum was recorded every 2°. The resulting set of transmission
spectra is presented in Fig. 5g as a polar, false-color map in which
the radial coordinate represents frequency and the angular
coordinate records the cylinder’s rotation angle36,37. Here, the
rotational isotropy of the PBG is clear; the blue-and-green
ring represents an isotropic transmission gap. The expected
PBG edges, as predicted by band structure calculation, are
overlaid as black solid curves. We corroborate the PBG isotropy
by performing finite-difference time-domain (FDTD) electro-
magnetic simulations. Specifically, we place a number of dipole
sources inside our amorphous gyroid cylinder, and record the
power flux some distance from the cylinder using a circular array
of detectors (see Supplementary Methods). This result is
presented as a second polar false-colour map in Fig. 5h; the
expected PBG edges, according to the band structure, are overlaid
as white solid curves. Although the transmission contrast of the
experimental data is noise-limited (see Supplementary Note 3),
the frequencies of the experimentally measured PBGs accord well
with the results of both band structure and FDTD simulations.
Self-uniformity by evolution. Many plants and animals have
evolved wavelength-scale microstructures as means of producing
colour38–41. Study of these architectures can inspire the develo-
pment of industrial scale fabrication techniques for next-
generation PhC-based technologies42–44.
In this regard, self-assembly is a particularly attractive
fabrication method. The three strongly isotropic networks have
all been observed to self-assemble45–48. Interestingly, amorphous
honeycomb and diamond have also been observed in the natural
world49,50; it is thus clear that a self-assembly pathway capable of
producing complex short-range order exists. Here, we explore
evidence that an amorphous gyroid could be similarly self-
assembled. First, we demonstrate that topological defects exist in
the gyroidal microstructures of green hairstreak butterflies47,51.
We then present a disordered network structure in the scales of
the Cambridge Blue butterfly and model its reflectance spectrum
with an amorphous gyroid structure52.
Many butterfly species are known to derive colouration from
gyroidal networks of chitin within their scales, most famously
Parides sesostris (the emerald-patched cattleheart) and numerous
species of the genus Callophrys (the green hairstreaks). Micro-
scopy has demonstrated that the scales contain numerous
crystallites of a well-ordered network of chitin in air51,53. Small
angle X-ray scattering (SAXS) has assigned the symmetry group
of SNG to the structures47.
However, tomographic reconstruction of the chitin/air inter-
face in Callophrys rubi shows that the network is not everywhere
a perfect topological gyroid51. Evidence suggests that the chitin
network passes continuously between adjacent gyroid grains, and
that this grain-matching is facilitated by topological network
defects51. Further, different crystallites within the same scale
exhibit different chitin fill fractions. Accounting for these obser-
vations, we now show that SNG crystallites incorporating small
amounts of topological and positional disorder are better models
of the scale structure in the green hairstreaks than a perfect gyroid
crystal.
To model the SAXS results, we approximate the scale as an
ensemble of separate crystallites which contribute to the
diffraction spectrum independently. The positional correlations
between adjacent crystal domains are destroyed by the matching
defects at the grain boundaries; the SAXS spectrum can thus be
considered an incoherent superposition of the diffraction patterns
of distinct crystallites. To fit the experimental results, we generate
a number of distinct crystallites across a range of chitin
fill fractions and sum their diffraction patterns incoherently,
optimizing their weights in the summation to minimize the
Reitveld weighted profile Rfactor, R
wp
, of the fit54. We calculate
two types of fits, described in detail in the Methods section. In the
first case, we model the scales as perfect gyroids, employing the
level-set approximation of the gyroid minimal surface (dotted
orange curve in Fig. 6a). In the second case, we model the
tomographic data51 using partially disordered gyroid (PDG)
networks (dashed blue line in Fig. 6a). PDG is generated by
introducing a small number of topological defects into a perfectly
ordered gyroid net. Specifically, one defect is introduced for every
100 network vertices together with a small amount of vertex
positional disorder.
Note that, at publication, no SAXS data for C. rubi was
available. Instead we compare our models to SAXS results for
Callophrys gryneus (solid green line in Fig. 6a); we expect the
microstructure in C. gryneus to comprise gyroid grains connected
by topological matching defects, just as in C. rubi.R
wp
values for
our level-set and PDG fits are 980 and 645, respectively,
suggesting that PDG is a superior model of the green hair-
streak scale structure. The pure level-set model produces overly
prominent peaks, particularly between the (110) and (211),
and (400) and (420) reflections. The inclusion of topological and
positional disorder dampens the network correlations and
reduces this contrast across the whole spectrum. The quality of
the disordered PDG fit is most evident for high (hkl) values; here,
the overall profile of the pattern, in particular the double peak of
the (321) and (400) planes, is well-captured.
Taken together, the tomographic data51 and the results of our
scale modelling (Fig. 6a) suggest that topological imperfections
form, to a limited degree, at the grain boundaries in the gyroidal
microstructures of the green hairstreaks. The existence of these
topological defects renders the hairstreaks (sub-family Theclinae)
a promising family of butterflies within which to search for an
amorphous gyroid.
We now turn our attention to the Cambridge Blue butterfly
(Ps. marsyas, Theclinae, Fig. 6b). Its wing scales contain a unique
ARTICLE NATURE COMMUNICATIONS | DOI: 10.1038/ncomms14439
8NATURE COMMUNICATIONS | 8:14439 | DOI: 10.1038/ncomms14439 | www.nature.com/naturecommunications
aperiodic photonic structuring from which they derive a blue
brilliance. Focussed ion beam (FIB) sections reveal an intricately
connected trivalent network, with fully 3D interconnectivity and
an apparent LSU (Fig. 6c). Computational projections of
amorphous gyroid with a volume fill fraction of 25% strongly
resemble the butterfly network (Fig. 6d,e). Transmission electron
microscopy sections (Fig. 6f) reveal a mixture of interwoven
layering and fully disordered network regions52. In cross-section,
the lower half of the scale (bottom half of Fig. 6f) appears as a
dark band—this increased contrast can be associated with
pigmentation55.
We now model scales of Ps. marsyas as amorphous gyroid
CRNs and investigate the compatibility of these models with
experimental reflectance measurements. We employ a number of
amorphous gyroid CRNs with
F22s around 0.88. We calculate
their far field reflectance spectra by FDTD solution of the
Maxwell equations.
We estimated the scale thickness from FIB sections to be
1300±200 nm. The lower half of the scale was taken to be
absorbing in accordance with the pigment distribution observed
by transmission electron microscopy (Fig. 6f). We suggest a
plausible model for the complex refractive index of the
Ps. marsyas pigment, derived from the extinction coefficient of
the pigment in Papilio nireus (see the Methods section).
Amorphous gyroid edge lengths for our Ps. marsyas sample were
estimated from electron micrographs to be 117±6 nm; this
corresponds to an effective SNG primitive cell parameter of
a¼166±8 nm. The reflectance of a large wing area of
Ps. marsyas was measured by a previous study52; scaling of the
amorphous gyroid (110)-type SAXS peak to the wavelength of
maximum reflectance suggests a¼169 nm. The theoretical
reflectance presented (Fig. 6g) is an average over six amorphous
gyroid models, all scaled by an avalue of 166 nm.
Given the assumptions made in modelling the pigmentation,
the general agreement between theoretical and experimental
reflectance spectra suggests that an amorphous gyroid with a
UV-absorbing pigment is a plausible model of the structure in
Ps. marsyas scales. The divergence at UV wavelengths is
attributable to uncertainty in the exact complex refractive index
of the butterfly scale across a 300 to 700 nm range; future studies
should measure this directly. The small divergence at red
wavelengths is attributable to reflections from unmodelled
melanized ground scales.
Two other neotropical hairstreaks, Arcas imperialis (Theclinae)
and Evenus coronata (Theclinae)—close relatives of the
Cambridge Blue—have previously been surveyed through micro-
scopy56. Their structures appear to possess strong multilayer and
gyroidal characteristics, respectively. It is therefore possible that
amorphous gyroid has evolved in the Cambridge Blue by some
small change in scale cell development conditions, leading to an
evolutionary divergence. Indeed, such a divergence has been
postulated in the case of the gyroid-containing Pa sesostris,
whose-scale structuring has diverged from the perforated
multilayers of many closely related Parides species57.
Our work indicates that the scale structuring in the Cambridge
Blue is related to an amorphous gyroid. Electromagnetic
modelling shows that amorphous gyroid models are consistent
with observed reflectance data and the existence of topological
matching defects between gyroid grains in green hairstreak scales
suggests that the production of amorphous gyroid is devel-
opmentally possible.
Any further search for a natural amorphous gyroid should
not be limited to butterflies. A thorough survey of avian
feather barbs has revealed an abundance of colour-producing
channel-type architectures. Several species—Diglossa cyanae58
(Thraupidae), Passerina cyanea58 (Cardinalidae) and Alcedo
10–3
10–2
10–1
100
2
6
8
14
16
h2 + k2 + l2
PDG
Level-set
C. gryneus SAXS
20
22
24
26
Reflectance
400 500 600
200 300 400 500 600 700 800
(nm)
(nm)
kra
S(kr)
Smax
246810121416
1.7
1.5
0.1
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0
k
n
P. marsyas exp.
A-Gyroid model
n Model (inset)
k Model (inset)
a
bc
de
f
g
Figure 6 | LSU in butterfly scales. Comparison with SAXS data of level-set and partially disordered gyroid models of wing scales in the butterfly C. gryneus
(a). Specimen of Ps. marsyas (b) and FIB section of one of its blue scales (c), imaged with 52°tilt in the FIB. Detail (d), taken from (c), compared with a
projection of an amorphous gyroid model (e). Transmission electron micrograph of a Ps. marsyas scale cross-section (f). Comparison of experimental
reflectance data from a large wing area to an amorphous gyroid scale model (g). Scale bars, 1 cm (b); 1 mm(c); 500 nm (d,e); and 1.5 mm(f). The TEM
section (f) is reproduced with permission from Ve
´rtesy et al.52. TEM, transmission electron microscopy.
NATURE COMMUNICATIONS | DOI: 10.1038/ncomms14439 ARTICLE
NATURE COMMUNICATIONS | 8:14439 | DOI: 10.1038/ncomms14439 | www.nature.com/naturecommunications 9
atthis59 (Alcedinidae)—are themselves a brilliant blue and possess
microstructures with striking similarity to amorphous gyroid.
Experimentally, it may be possible to produce amorphous gyroid
or a similar architecture via block co-polymer self-assembly60.It
is possible to generate mixed lamellar and gyroidal states that may
resemble the Ps. marsyas structure61,62. An alternative pathway
may be to quench the phase transition between the gyroid and
metastable perforated multilayer phases62.
Discussion
We have introduced LSU as means of measuring the extent
to which local environments in a CRN of fixed valency
are spatially similar. When applied to strictly trivalent
and tetravalent networks the states of maximum LSU
are the strongly isotropic single gyroid and diamond
networks, respectively. As a network’s LSU decreases from its
maximum it becomes glassy and eventually chaotic. LSU can
thus classify all CRNs of fixed valency by the extent of their
internal order.
We designed LSU as a means of characterizing the optical
properties of CRNs. In particular, we have shown that networks
with maximal LSU possess champion PBGs in both 2D and 3D.
Further, other known networks endowed with a complete PBG
in 3D or a TE polarization PBG in 2D are characterized by high
levels of LSU. PBGs result from scattering by the electro-
magnetic resonances of a network’s local scattering units. When
these units are spatially similar, their resonances are maximally
degenerate and complete PBG formation is favoured. LSU is
thus a predictive measure of a network’s PBG forming ability.
While here we focussed on trivalent and tetravalent networks,
LSU can be generalized to include networks of arbitrary or
mixed valency.
We have introduced designs of novel amorphous gyroid
CRNs. Here we used ceramic 3D printing to fabricate
amorphous gyroid samples in high-refractive index alumina
and demonstrated their sizeable isotropic PBGs via microwave
transmission experiments. The relevance of amorphous gyroid,
and architectures which can be derived from it, is broad. In
particular, its development, or that of a closely related CRN,
appears to have occurred in the scales of the butterfly Ps.
marsyas. That it is possible to self-assemble such a structure is a
prerequisite for its existence in natural systems. It may be
possible to produce amorphous gyroid networks via block co-
polymer experiments or equivalent self-assembly methods. This
poses an interesting experimental challenge, the solution of
which will facilitate the fabrication of advanced optical
metamaterials for industrial applications43,60.
Fundamentally, we have demonstrated that the tree compar-
ison method is a powerful framework for controlling the LSU of
the scattering centres in a CRN. Sculpting a network’s LSU
distributions translates directly to advanced control over its
optical properties. The optical properties controlled need not be
limited to PBG forming ability; they could include structural
colour, the scattering mean free path63 and random lasing64.
Moreover, similar design principles may be employed to control
other wave phenomena in electronic, phononic, elastic and
acoustic materials.
Methods
Definition of the LSU distributions.Consider a CRN Cwith a set of defined
vertex positions and inter-vertex connectivities, in which each vertex in Chas
exactly gedges. We define an n-tree Ta
non vertex aof Cas the set of vertices within
nedges of a. The root edges of Ta
nare all the edges of Ta
1. Computationally, all
information regarding Ta
ncan be obtained by performing a breadth-first graph
search to depth nfrom vertex a.
We consider now two n-trees of equal depth Ta
nand Tb
n. We label the root edges
of the trees {1
a
,2
a
yg
a
} and {1
b
,2
b
... g
b
}, respectively. The spatial similarity
statistic f
ab
of these two trees is defined as
fab¼1
g!X
g!
i¼1
fT
a
n;Tb
n;si

;ð1Þ
where fis a similarity measure which grades the overlap of Ta
nand Tb
nwhen they
are maximally aligned in a root edge permutation s
i
. We sum the similarities for all
g! overlap permutations of the trees’ root edges and then take the average.
The value of the spatial similarity statistic will depend both on the form that the
measure ftakes and the method that is used to maximally align the two trees for
each permutation. We describe our choice of fand the alignment procedure we
follow in the next section.
The LSU distributions of the CRN are particular sets of spatial similarity
statistics. We define a new tree Ta
lwith depth lon vertex a. The set of vertices in Ta
l
is the local neighbourhood of ato depth l. The LSU distribution F
nl
can then be
written as
Fnl¼fab
fg
:b2Ta
l;a2C:ð2Þ
F
nl
is thus the set of spatial similarity statistics for all trees of depth nwhose root
vertices are within ledges of one another. F
nl
can be plotted as histograms as in
Fig. 3. In Fig. 3 the histogram frequencies are normalized through division by the
total number of spatial similarity statistics in F
nl
.
Calculation of the overlap between two trees.Here we discuss in detail the exact
form of our similarity measure fand the process by which two trees are maximally
overlapped. We note that the choice of fis an important user-controllable degree of
freedom. In general, fmust be maximal when two trees can be overlapped perfectly;
this defines uniquely the maximally self-uniform configuration. However, the exact
manner in which overlap is calculated can be defined to best suit the application
and to possess meaningful properties as the network departs from maximal self-
uniformity. Here, we adopt an intuitive framework in which tree edges are over-
lapped in pairs with their most natural partners.
Consider now an n-tree Ta
nin the CRN C. We label its Ta
n
vertices with index j,
subject to the constraint that indices {1, 2 ... g} represent the tree’s root edges. The
branches of the tree are defined by its edge vectors. For Ta
nwe denote these as ra
j,
which defines the vector to vertex jfrom its parent vertex as defined by a breadth-
first graph search starting on a.
Consider now two n-trees Ta
nand Tb
n. Before they can be compared, they must
be maximally overlapped. This process consists of determining a congruent
transformation which maximizes the overlap of their root edges in some chosen
permutation s
i
. First, Tb
nis translated such that vertices aand boverlap. Second, Tb
n
is rotated until rb
si1ðÞand ra
1are parallel; this is always possible. Third, Tb
nis rotated
about the ra
1axis until rb
si2ðÞand ra
2are maximally parallel, defined to be the
configuration which minimizes their scalar product. For disordered trees, it is not
usually possible to make these two vectors perfectly parallel. For trivalent networks,
we deem these three steps sufficient to maximally align the two trees.
For tetravalent networks it is necessary to introduce a fourth alignment step. In
this case, Tb
nis reflected in the plane defined by ra
1and ra
2so as to bring ra
3into
maximal alignment with rb
si3
ðÞand ra
4with rb
si4
ðÞ
. This step is performed only if the
alignment between Ta
nand Tb
nis improved, as measured by an increase in the value
of ra
3rb
si3ðÞ
þra
4rb
si4ðÞ
. At this point, the two trees are considered to be maximally
aligned.
Once maximally aligned, we define our similarity measure ffor grading the
overlap of two trees as
fT
a
n;Tb
n;si

¼1
Ta
n
1X
j2Ta
n
ra
jrb
kðjÞ
1
4ra
j
þrb
kðjÞ

2:ð3Þ
Overlap is calculated between edge pairs by taking their scalar product and
normalizing it with the square of their mean norm; overlap of a single pairing is
thus distributed on [ 1, 1]. The Ta
n
1 edge pair comparisons are summed and
averaged such that fyields a maximum value of unity for perfectly overlapping
trees.
Edge pairs are overlapped in the combination that is most natural. The whole
overlap calculation is performed recursively in a depth-first sense, greatly
simplifying the process of determining natural pairings. At a particular point in the
algorithm’s execution, it has reached some vertex pair jand k(j) by comparison of
the natural edge pair ra
jwith rb
kjðÞ
. Overlap of all (g1)! possible sets of edge pair
comparisons is calculated. The set that maximizes the sum of edge pair scalar
products is chosen as the set of natural pairings; this result is accepted, and the
algorithm proceeds to calculate overlap for the edges around each of the (g1)
child vertices. This decision process is captured in the notation rb
kjðÞ
, to reflect that
the choice of edge vector in tree bis a function of the edge vector ra
jto which it is
being compared.
Finally, the spatial similarity statistic f
ab
is determined by repeating this process
for all root edge permutations according to equation (1).
Microwave transmission measurement.Gyroid and amorphous gyroid models
were printed and finished in the workshops of the Fraunhofer Institute for Ceramic
Technologies and Systems, Dresden. The gyroid primitive cell parameter was
ARTICLE NATURE COMMUNICATIONS | DOI: 10.1038/ncomms14439
10 NATURE COMMUNICATIONS | 8:14439 | DOI: 10.1038/ncomms14439 | www.nature.com/naturecommunications
designed to be to be 3.08 mm. As part of the printing process, the samples were
sintered at high temperature, shrinking in side-length by B20% and introducing
some uncertainty into their true sizes. Actual scaling values were thus measured
and fill fractions were determined using a water displacement method, yielding an
effective gyroid primitive cell parameter of a¼3.13±0.05 mm, and alumina
volume-fill-fractions between 27 and 29%. Amorphous gyroid samples were
fabricated preferentially. We also produced one SNG model. This model was
designed for measurement of transmission along the [111] axis; this axis was
chosen because of its symmetry (see Supplementary Note 4).
Characterization of the samples was performed with microwave radiation using
an HP-8510C VNA. A single polarization mode was coupled through a pair of
rectangular horn antennae and a pair of custom-made Teflon microwave lenses.
For the measurements shown in Fig. 5e,f, the beam was aligned along the short
edge of the cuboidal samples (Fig. 5a,c) and perpendicular to a flat sample surface,
and the frequency varied from 15 to 35 GHz. Note that frequencies beyond the
standard microwave K band (18–26.5 GHz) have very low efficiency coupling
through the horns and waveguides. For the measurements shown in Fig. 5g, the
cylindrical sample was aligned with the incident beam perpendicular to the
cylinder axis, around which we rotated the sample and recorded the transmission
every 2°. This set of measurements was performed across a frequency range of
17–27 GHz.
The total dynamic range of the HP-8510C VNA is from 0 to 65 dB
(measured dark noise). A large area of microwave-absorbing material, into which a
window was cut to hold the sample, was used to prevent reflection and scattering of
radiation into the environment. The normalized transmission is then defined as the
ratio between detected intensities with and without the sample in place. The
addition of microwave-absorbing material lowers the overall coupling-efficiency
through the pair of horn-antennae. As a result, the actual accessible dynamic range
through this measurement is only B35 dB, which limits the measured gap-depth.
This limited dynamic range is apparent in the amorphous gyroid transmission of
Fig. 5f; the transmission bottoms out and becomes noisy for frequencies between
20 and 26 GHz. For the same reason, transmission results for the amorphous
gyroid cylinder (Fig. 5g) appear noisy in comparison to the theoretical results
(Fig. 5h). An increased dynamic range could be accessed by amplifying the source
power.
SAXS pattern modelling.The total scattered intensity measured in a
SAXS experiment can be written as a Fourier transform of the electron density
function r(r):
IqðÞ¼
r2
e
VZV
rðrÞ
r½eiqrdr
2
;ð4Þ
for qthe scattering vector in reciprocal space, r
e
the radius of the electron and Vthe
total sample volume65. We Fourier transform the difference between r(r) and its
volume average
rto remove the forward scattering peak and access any diffraction
signatures that may be present at small scattering vectors. The r(r) functions for
our 3D network structures were generated by voxelization and were distributed on
[0, 1], representing the air and chitin distributions of the butterfly scale networks,
respectively. Fourier transforms were calculated using a fast Fourier transform
algorithm. Powder-like structure factors were calculated by azimuthal averaging of
the total scattered intensity. Structure factors were normalized to the intensity of
the (110) peak for comparison with experimental data.
To model the observed scattering from C.gryneus, we investigated both partially
disordered gyroid (PDG) networks and level-set surfaces as approximations to the
chitin network. In all cases, cubic crystallites of dimension 10a(1,000 vertex points)
were used. When f¼0, the function
fx;y;zðÞ¼sinðxÞcosðyÞþsinðyÞcosðzÞþsinðzÞcosðxÞð5Þ
defines the level-set approximation to the gyroid surface; it divides space into two
inter-penetrating labyrinths each with a 50% volume fill. For fa0, it defines a
surface with the symmetry of SNG, and alters the relative filling fraction of the two
labyrinths. A set of level-set basis states was generated by sampling these different
fill fractions via a critical value t, assigning unity to r(r) for f4tand zero otherwise.
PDG networks were generated by introducing geometrical and topological
disorder into SNGs. All PDG networks were topologically disordered by the
introduction of 10 Stone–Wales defects. This was performed using our
implementation of the WWW algorithm. Geometrical disorder was added by
perturbation of the vertex points by a random vector whose magnitude was
normally distributed with mean zero and s.d. sup to 40% of the gyroid edge
length. To increase the physicality of this disorder, perturbations were propagated
through the network with their magnitude decaying with depth; vertex positions
thus retained a local correlation. The final set of disordered SNG basis states
sampled various degrees of vertex positional disorder, as defined by their
perturbation standard deviations and rates of perturbation decay, each across a set
of fill fractions.
Final diffraction spectra consist of a weighted sum over the diffraction patterns
of the appropriate basis states. A fit to the experimental data was performed by
optimization of these weights so as to minimize the Rietveld weighted profile factor
of the fit.
Pseudolycaena marsyas scale imaging.A dried male specimen of Ps. marsyas was
acquired online from The Bugmaniac Insect Shop. The specimen was spread and
mounted for sample photography. Single scales were removed, mounted and gold
sputtered to produce a 4 nm coating. Scales were then imaged via electron
microscopy, and sectioned using a FIB.
SNG net scaling parameters were estimated from micrographs under the
assumption that network regions with a locally honeycomb-like appearance
correspond to the nine-segment helices visible along the SNG (111) axis. The
appearance of such regions is similar to views along the (111) axis of gyroid
crystallites in green hairstreak butterflies, supporting this supposition. The radius
rof such a helix satisfies the approximate relation 2pr¼7.732l, for lthe SNG edge
length.
Pseudolycaena marsyas reflectance modelling.Models of Ps. marsyas scales
were generated using type-2 amorphous gyroid networks with
F22s around 0.88.
The effective single-network gyroid primitive unit cell dimension was estimated
from micrographs of Ps. marsyas to be a¼166±8 nm, corresponding to an edge
length of l¼117±6 nm. Approximating the scale’s reflectance as a result of
coherent scattering from (111)-plane-type reflections suggests a¼169 nm.
Six models were made, each of thickness 1,350 nm and scaled with an avalue of
166 nm. To model the apparent variation in cross-member thickness within the
scale’s network, amorphous gyroid point patterns were decorated with cylinders
having normally distributed radii r, with a mean fractional radius r/l¼0.44
and a 15% s.d. The resulting network has a mean volume fill fraction of B25%.
The Maxwell equations were solved using the Lumerical’s implementation of
the FDTD method (Lumerical Solutions, Inc). The previously published
experimental data was gathered through normal illumination of a large wing area
and use of an integrating sphere in reflection. To model this, scales were
illuminated from above at normal incidence with a total field scattered field source.
The scattered fields were recorded by a monitor box, and were projected into the
far field in the upper half-space, back along the axis of incidence. The total power
scattered into the upper half-space was determined and normalized to the injected
power. Reflectance data was calculated across the wavelength range 360–670 nm
using a uniform 11.5 nm mesh.
Cylinders in the upper half of each scale were modelled as a dispersive material
with the refractive index of chitin as measured by polarizing interference
microscopy applied to the butterfly Graphium sarpedon66. In accordance with an
existing study55, the dark banding of the lower half of the Ps. marsyas scale was
interpreted as pigmentation. The absorbance of this pigmentation was not
measured experimentally. Rather, we suggest that the pigment is ultraviolet
absorbing, as has been shown to be the case for Papilio nireus55, the gyroid-
containing butterfly Pa. sesostris53 and many other structurally-coloured
butterflies67.
To make progress in modelling the reflectance of Ps. marsyas scales, we propose
a plausible complex refractive index for the pigmented lower layer. We derive the
extinction coefficient k(o) of the ultraviolet-absorbing pigment in Papilio nireus
using the Beer–Lambert law55. The real component of the refractive index was
taken to be equal to that of the upper layer of the scale. Lumerical’s in-built
refractive index fitting algorithm was used to fit this combination of nand k; extra
weight was given to the imaginary component to accurately model the absorption.
The resulting complex refractive index used for the lower half of the scales is inset
into Fig. 6g.
The reflectance spectrum of Fig. 6g should be interpreted in light of the
assumptions we have made in modelling the complex refractive index of the
pigment. Given these assumptions, the agreement between the theoretical and
experimental reflectance data shows that an amorphous gyroid network is a
plausible model for the structure in Ps. marsyas scales. The divergence between
theoretical and experimental reflectance at ultraviolet wavelengths is a result of
uncertainty in the true refractive index of the Ps. marsyas pigment in this
wavelength range. Further modelling of Ps. marsyas scales should be informed by a
direct measurement of their complex refractive index55.
Data availability.The data underlying the findings of this study are available
without restriction. Details of the data and how to request access are available from
the University of Surrey publications repository: http://dx.doi.org/10.15126/
surreydata.00813094.
References
1. Blanco, A. et al. Large-scale synthesis of a silicon photonic crystal with a
complete three-dimensional bandgap near 1.5 micrometres. Nature 405,
437–440 (2000).
2. Kuramochi, E. et al. Large-scale integration of wavelength-addressable all-
optical memories on a photonic crystal chip. Nat. Photonics 8, 474–481 (2014).
3. Nozaki, K. et al. Ultralow-power all-optical RAM based on nanocavities. Nat.
Photonics 6, 248–252 (2012).
4. Edagawa, K., Kanoko, S. & Notomi, M. Photonic amorphous diamond
structure with a 3D photonic band gap. Phys. Rev. Lett. 100, 013901 (2008).
5. Edagawa, K. Photonic crystals, amorphous materials, and quasicrystals. Sci.
Technol. Adv. Mater. 15, 034805 (2014).
NATURE COMMUNICATIONS | DOI: 10.1038/ncomms14439 ARTICLE
NATURE COMMUNICATIONS | 8:14439 | DOI: 10.1038/ncomms14439 | www.nature.com/naturecommunications 11
6. Florescu, M., Torquato, S. & Steinhardt, P. J. Designer disordered materials
with large, complete photonic band gaps. Proc. Natl Acad. Sci. 106,
20658–20663 (2009).
7. John, S. Strong localization of photons in certain disordered dielectric
superlattices. Phys. Rev. Lett. 58, 2486–2489 (1987).
8. Yablonovitch, E. Inhibited spontaneous emission in solid-state physics and
electronics. Phys. Rev. Lett. 58, 2059–2062 (1987).
9. Imagawa, S. et al. Photonic band-gap formation, light diffusion, and
localization in photonic amorphous diamond structures. Phys. Rev. B 82,
115116 (2010).
10. Joannopoulos, J. D. et al. Photonic Crystals: Molding The Flow of Light
(Princeton University Press, 2011).
11. Yablonovitch, E. Photonic band-gap structures. J. Opt. Soc. Am. B 10, 283–295
(1993).
12. Ho, K. M., Chan, C. T. & Soukoulis, C. M. Existence of a photonic gap in
periodic dielectric structures. Phys. Rev. Lett. 65, 3152–3155 (1990).
13. Maldovan, M. & Thomas, E. L. Diamond-structured photonic crystals. Nat.
Mater. 3, 593–600 (2004).
14. Men, H., Lee, K. Y., Freund, R. M., Peraire, J. & Johnson, S. G. Robust topology
optimization of three-dimensional photonic-crystal band-gap structures. Opt.
Express 224, 214–224 (2014).
15. Fu, H. K., Chen, Y. F., Chern, R. L. & Chang, C. C. Connected hexagonal
photonic crystals with largest full band gap. Opt. Express 13, 7854–7860 (2005).
16. Amoah, T. & Florescu, M. High- Q optical cavities in hyperuniform disordered
materials. Phys. Rev. B Condens. Matter Mater. Phys. 91, 2–6 (2015).
17. Florescu, M., Steinhardt, P. & Torquato, S. Optical cavities and waveguides in
hyperuniform disordered photonic solids. Phys. Rev. B 87, 165116 (2013).
18. Florescu, M., Torquato, S. & Steinhardt, P. J. Complete band gaps in two-
dimensional photonic quasicrystals. Phys. Rev. B 80, 155112 (2009).
19. Tsitrin, S. et al. Unfolding the band structure of non-crystalline photonic band
gap materials. Sci. Rep. 5, 13301 (2015).
20. Zhang, X., Zhang, Z.-Q. & Chan, C. T. Absolute photonic band gaps in 12-fold
symmetric photonic quasicrystals. Phys. Rev. B 63, 081105 (2001).
21. Wright, A. C. & Thorpe, M. F. Eighty years of random networks. Phys. Status
Solidi B 250, 931–936 (2013).
22. Eder, F. R., Kotakoski, J., Kaiser, U. & Meyer., J. C. A journey from order to
disorder—atom by atom transformation from graphene to a 2D carbon glass.
Sci. Rep. 4, 4060 (2014).
23. Treacy, M. M. J. & Borisenko, K. B. The local structure of amorphous silicon.
Science 335, 950–953 (2012).
24. Zhang, S. et al. Polymorphism in glassy silicon: inherited from liquid-liquid
phase transition in supercooled liquid. Sci. Rep. 5, 8590 (2015).
25. Laws, K. J., Miracle, D. B. & Ferry, M. A predictive structural model for bulk
metallic glasses. Nat. Commun. 6, 8123 (2015).
26. Zachariasen, W. H. The atomic arrangement in glass. J. Am. Chem. Soc. 196,
3841–3851 (1932).
27. Barkema, G. T. & Mousseau, N. High-quality continuous random networks.
Phys. Rev. B 62, 4985–4990 (2000).
28. Wooten, F., Winer, K. & Weaire, D. Computer generation of structural models
of amorphous Si and Ge. Phys. Rev. Lett. 54, 1392–1395 (1985).
29. Sunada, T. Crystals that nature might miss creating. Notices AMS 55, 208–215
(2008).
30. Florescu, M., Torquato, S. & Steinhardt, P. J. Effects of random link removal on
the photonic band gaps of honeycomb networks. Appl. Phys. Lett. 97, 10–13
(2010).
31. Johnson, S. & Joannopoulos, J. Block-iterative frequency-domain
methods for Maxwell’s equations in a planewave basis. Opt. Express 8, 173–190
(2001).
32. Rockstuhl, C., Peschel, U. & Lederer, F. Correlation between single-cylinder
properties and bandgap formation in photonic structures. Opt. Lett. 31,
1741–1743 (2006).
33. Men, H., Lee, K. Y. K., Freund, R. M., Peraire, J. & Johnson, S. G. Robust
topology optimization of three-dimensional photonic-crystal band-gap
structures. Opt. Express 22, 22632–22648 (2014).
34. Muller, N., Haberko, J., Marichy, C. & Scheffold, F. Silicon hyperuniform
disordered photonic materials with a pronounced gap in the shortwave
infrared. Adv. Opt. Mater. 2, 115–119 (2014).
35. Weaire, D. Existence of a gap in the electronic density of states of a
tetrahedrally bonded solid of arbitrary structure. Phys. Rev. Lett. 26, 1541–1543
(1971).
36. Man, W., Florescu, M. & Matsuyama., K. Photonic band gap in isotropic
hyperuniform disordered solids with low dielectric contrast. Opt. Express 21,
19972–19981 (2013).
37. Man, W., Megens, M., Steinhardt, P. J. & Chaikin, P. M. Experimental
measurement of the photonic properties of icosahedral quasicrystals. Nature
436, 993–996 (2005).
38. Kinoshita, S., Yoshioka, S. & Miyazaki, J. Physics of structural colors. Rep. Prog.
Phys. 71, 076401 (2008).
39. Seago, A. E., Brady, P., Vigneron, J.-P. & Schultz, T. D. Gold bugs and beyond:
a review of iridescence and structural colour mechanisms in beetles
(Coleoptera). J. R. Soc. Interface 6, S165–S184 (2009).
40. Vignolini, S. et al. Pointillist structural color in Pollia fruit. Proc. Natl Acad. Sci.
USA 109, 15712–15715 (2012).
41. Vukusic, P. & Sambles, J. R. Photonic structures in biology. Nature 424,
852–855 (2003).
42. Mille, C., Tyrode, E. C. & Corkery, R. W. 3D titania photonic crystals replicated
from gyroid structures in butterfly wing scales: approaching full band gaps at
visible wavelengths. RSC Adv. 3, 3109–3117 (2013).
43. Parker, A. R. & Townley, H. E. Making photonic structures via cell culture:
Morpho butterfly scales. Bioinspir. Biomim. Nanobiomater. 4, 1–5 (2014).
44. Zhang, S. & Chen, Y. Nanofabrication and coloration study of artificial Morpho
butterfly wings with aligned lamellae layers. Sci. Rep. 5, 16637 (2015).
45. Almsherqi, Z. A., Kohlwein, S. D. & Deng, Y. Cubic membranes: a legend beyond
the Flatland* of cell membrane organization. J. Cell Biol. 173, 839–844 (2006).
46. Parker, A. R. & Mcphedran, R. C. Aphrodite’s iridescence. Nature 409, 36–37
(2001).
47. Saranathan, V. et al. Structure, function, and self-assembly of single network
gyroid (I4
1
32 ) photonic crystals in butterfly wing scales. Proc. Natl Acad. Sci.
USA 107, 11676–11681 (2010).
48. Wilts, B. D., Michielsen, K., De Raedt, H. & Stavenga, D. G. Hemispherical
Brillouin zone imaging of a diamond-type biological photonic crystal. J. R. Soc.
Interface 9, 1609–1614 (2012).
49. Prum, R. O., Torres, R., Kovach, C., Williamson, S. & Goodman, S. M.
Coherent light scattering by nanostructured collagen arrays in the caruncles of
the malagasy asities (Eurylaimidae: aves). J. Exp. Biol. 202, 3507–3522 (1999).
50. Yin, H. et al. Amorphous diamond-structured photonic crystal in the feather
barbs of the scarlet macaw. Proc. Natl Acad. Sci. USA 109, 12260–12260 (2012).
51. Schro
¨der-Turk, G. E. et al. The chiral structure of porous chitin within the
wing-scales of Callophrys rubi.J. Struct. Biol. 174, 290–295 (2011).
52. Ve
´rtesy, Z. et al. Wing scale microstructures and nanostructures in butterflies-
natural photonic crystals. J. Microsc. 224, 108–110 (2006).
53. Wilts, B. D., Michielsen, K., De Raedt, H. & Stavenga, D. G. Iridescence and
spectral filtering of the gyroid-type photonic crystals in Parides sesostris wing
scales. Interface Focus 2, 681–687 (2012).
54. McCusker, L. B., Von Dreele, R. B., Cox, D. E., Loue
¨r, D. & Scardi, P. Rietveld
refinement guidelines. J. Appl. Crystallogr. 32, 36–50 (1999).
55. Trzeciak, T., Wilts, B. D., Stavenga, D. G. & Vukusic, P. Variable multilayer
reflection together wtith long-pass filtering pigment determines the wing
coloration of papilionid butterflies of the nireus group. Opt. Express 20,
8877–8890 (2012).
56. Ingram, A. L. & Parker, A. R. A review of the diversity and evolution of
photonic structures in butterflies, incorporating the work of John Huxley (The
Natural History Museum, London from 1961 to 1990). Philos. Trans. R. Soc. B
363, 2465–2480 (2008).
57. Wilts, B. D., IJbema, N. & Stavenga, D. G. Pigmentary and photonic coloration
mechanisms reveal taxonomic relationships of the Cattlehearts (Lepidoptera:
Papilionidae Parides). BMC Evol. Biol. 14, 160 (2014).
58. Saranathan, V. et al. Structure and optical function of amorphous photonic
nanostructures from avian feather barbs: a comparative small angle X-ray
scattering (SAXS) analysis of 230 bird species. J. R. Soc. Interface 9, 2563–2580
(2012).
59. Stavenga, D. G., Tinbergen, J., Leertouwer, H. L. & Wilts, B. D. Kingfisher
feathers—colouration by pigments, spongy nanostructures and thin films.
J. Exp. Biol. 214, 3960–3967 (2011).
60. Vignolini, S. et al. A 3D optical metamaterial made by self-assembly. Adv.
Mater. 20, 1–5 (2012).
61. Ahn, J.-H. & Zin, W.-C. Mechanism of morphological transition from lamellar/
perforated layer to gyroid phases. Macromol. Res. 11, 152–156 (2003).
62. Meuler, A. J., Hillmyer, M. A. & Bates, F. S. Ordered network mesostructures in
block polymer materials. Macromolecules 42, 7221–7250 (2009).
63. Burresi, M. et al. Bright-white beetle scales optimise multiple scattering of light.
Sci. Rep. 4, 6075 (2014).
64. Wiersma, D. S. Disordered photonics. Nat. Photonics 7, 188–196 (2013).
65. Imperor-Clerc, M. Three-dimensional periodic complex structures in soft
matter: investigation using scattering methods. Interface Focus 2, 589–601
(2012).
66. Leertouwer, H. L., Wilts, B. D. & Stavenga, D. G. Refractive index and
dispersion of butterfly chitin and bird keratin measured by polarizing
interference microscopy. Opt. Express 19, 24061–24066 (2011).
67. Stavenga, D. G., Leertouwer, H. L. & Wilts, B. D. Coloration principles of
nymphaline butterflies—thin films, melanin, ommochromes and wing scale
stacking. J. Exp. Biol. 217, 2171–2180 (2014).
Acknowledgements
We thank Uwe Scheithauer of Fraunhofer IKTS for technical advice and support related
to the 3D ceramic printing. We thank Vlad Stolojan and David Cox for assistance
ARTICLE NATURE COMMUNICATIONS | DOI: 10.1038/ncomms14439
12 NATURE COMMUNICATIONS | 8:14439 | DOI: 10.1038/ncomms14439 | www.nature.com/naturecommunications
regarding the microscopy of butterfly scales, and Zofia Ve
´rtesy and Laszlo Biro
´for
providing data on Pseudolycaena marsyas. We also thank Miroslav Hejna for advice
regarding implementation of the WWW protocol. This work was partially supported by
the University of Surrey’s IAA award to M.F., the EPSRC (United Kingdom) DTG Grant
No. EP/J500562/1, EPSRC (United Kingdom) Strategic Equipment Grant No. EP/
L02263X/1 (EP/M008576/1) and EPSRC (United Kingdom) Grant EP/M027791/1.
The work was also partially supported by the National Science Foundation’s award
DMR-1308084 (USA) and the American Chemistry Society award ACS PRF 52644-UR6
to W.M.
Author contributions
S.R.S. initiated the project, designed the local self-uniformity measure and the algorithm
for its measurement, performed simulations and wrote the paper. W.M. designed and
performed the microwave experiments, analysed the data and wrote the paper. S.S.
performed the microwave experiments and analysed data. M.F. initiated the programme,
oversaw and directed the project, designed the local self-uniformity measure and wrote
the paper.
Additional information
Supplementary Information accompanies this paper at http://www.nature.com/
naturecommunications
Competing financial interests: The authors declare that British patent application no.
1601838.4 has been filed in the name of the University of Surrey. The contributing
authors Steven Sellers and Marian Florescu are recorded as the co-inventors. Weining
Man and Shervin Sahba declare no competing financial interests.
Reprints and permission information is available online at http://npg.nature.com/
reprintsandpermissions/
How to cite this article: Sellers, S. R. et al. Local self-uniformity in photonic networks.
Nat. Commun. 8, 14439 doi: 10.1038/ncomms14439 (2017).
Publisher’s note: Springer Nature remains neutral with regard to jurisdictional claims in
published maps and institutional affiliations.
This work is licensed under a Creative Commons Attribution 4.0
International License. The images or other third party material in this
article are included in the article’s Creative Commons license, unless indicated otherwise
in the credit line; if the material is not included under the Creative Commons license,
users will need to obtain permission from the license holder to reproduce the material.
To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/
rThe Author(s) 2017
NATURE COMMUNICATIONS | DOI: 10.1038/ncomms14439 ARTICLE
NATURE COMMUNICATIONS | 8:14439 | DOI: 10.1038/ncomms14439 | www.nature.com/naturecommunications 13

Supplementary resource (1)

... 40 Additionally, connected low-coordination-number networks of uniform valency possessing substantial local self-uniformity are considered conducive to the formation of large CPBs. 20,41 To date, the face-centered cubic single diamond (SD, dia net, space group Fd3m, No. 227) structure with tetravalent vertices is considered the "holy grail" of all photonic structures due to the most isotropic and highly geometrical identical framework, exhibiting the widest photonic bandgap and the lowest dielectric contrast requirement. 5,8,20 The body-centered cubic, chiral single gyroid (SG, srs net, space group I4132, No. 214) structure with trivalent nodes, which also exhibits perfect local self-uniformity, also forms a large CPB comparable to but smaller than the diamond. ...
... 5,8,20 The body-centered cubic, chiral single gyroid (SG, srs net, space group I4132, No. 214) structure with trivalent nodes, which also exhibits perfect local self-uniformity, also forms a large CPB comparable to but smaller than the diamond. 10,41 It is worth noting that these two structures have also been found to be the source of structural colors in various biomineralized skeletons and scaffolds. [42][43][44][45][46] In addition, the single primitive (SP, pcu net, space group Pm3m, No. 221) network also produces a relatively smaller CPB in simulations, 10,30 which can be attributed to the fact that it consists of hexavalent nodes with less local self-uniformity. ...
... [42][43][44][45][46] In addition, the single primitive (SP, pcu net, space group Pm3m, No. 221) network also produces a relatively smaller CPB in simulations, 10,30 which can be attributed to the fact that it consists of hexavalent nodes with less local self-uniformity. 41 Considering the excellent photonic properties of these structures, we apply our method to the corresponding space groups. 8 We also consider the space groups of their correlated double-network morphologies, 16,47 However, it has been proposed that these double-network eliminate the bandgap properties. ...
Preprint
Three-dimensional (3D) photonic structures enable numerous applications through their unique ability to guide, trap, and manipulate light. Constructing new functional photonic crystals remains a significant challenge since traditional design principles based on band structure calculations require numerous time-consuming computations. Additionally, traditional design is based on enumerated structures making it difficult to find novel functional geometries. Here, we propose an ultra-fast photonic crystal performance prediction method to enable efficient structure optimization of arbitrary 3D photonic crystals even with multiple variable modulation. Our methodology combines Fourier synthesis-enabling the creation of any smooth geometry within a crystallographic space group-with deep learning, which facilitates efficient photonic characterization within the vast parameter space. Over 2 million structures can be explored within 2 hours using a mainstream desktop workstation. The ideal structures with desired band properties, such as large photonic bandgap, specific frequency ranges, etc., could be rapidly discovered. We systematically confirmed the well-documented assumption that the most significant photonic bandgaps are found in minimal surface morphologies, in which the single diamond (dia net) with Fd3m (227) symmetry reigns supreme among known photonic structures, followed by the chiral single gyroid (srs net) with I4132 (214) symmetry. Additionally, a less well-known 3D photonic crystal with lcs topology within Ia3d (230) was rediscovered to exhibit a wide complete photonic bandgap, comparable to the diamond and the gyroid net. Our method not only validates the assumed hierarchy of photonic structures but also lays the foundation for the tailored design of functional materials and offers fresh insights into the advancement of next-generation optical devices and information technology.
... Further complicating the picture, photonic bandgaps have also been observed in non-SHU disordered structures, such as jammed packings [26][27][28], and even in systems with only short-range correlations, like equilibrium hard spheres [29], which are not hyperuniform. The role of short range correlations in disordered systems is especially pronounced for vector waves, for which bandgaps have so far only been observed in networked materials, such as honeycomb or tetrahedrally bonded structures [10,17,27,30,31]. In these random networks, the emergence of bandgaps has been linked to the similarity of local geometry and topology across the network [23,27,30]. ...
... The role of short range correlations in disordered systems is especially pronounced for vector waves, for which bandgaps have so far only been observed in networked materials, such as honeycomb or tetrahedrally bonded structures [10,17,27,30,31]. In these random networks, the emergence of bandgaps has been linked to the similarity of local geometry and topology across the network [23,27,30]. This raises the question: is there a unifying feature across these systems that promotes the formation of isotropic photonic bandgaps? ...
Preprint
We introduce a new class of functional correlated disordered materials, termed Gyromorphs, which uniquely combine liquid-like translational disorder with quasi-long-range rotational order, induced by a ring of G delta peaks in their structure factor. We generate gyromorphs in 2d and 3d by spectral optimization methods, verifying that they display strong discrete rotational order but no long-range translational order, while maintaining rotational isotropy at short range for sufficiently large G. We show that they are distinct from quasicrystals, which we argue are akin to Fourier ''duals'' of gyromorphs. Using a coupled dipole approximation, we show that these structures outperform quasicrystals, stealthy hyperuniformity, and Vogel spirals in the formation of isotropic bandgaps in 2d, for both scalar and vector waves, and open complete isotropic bandgaps in 3d. Finally, we introduce ''polygyromorphs'' with several rotational symmetries at different length scales (i.e., multiple rings of delta peaks), enabling the formation of multiple bandgaps in a single structure, thereby paving the way for fine control over optical properties.
... Although this is easily achieved for periodic structures, there are many examples in nature where partially disordered morphologies give rise to structural colour, either in the feathers of birds 6,7 or in insects [8][9][10][11][12][13] . While some theories exist [14][15][16] , the detailed nature of the order-disorder interplay giving rise to structural colour is not well understood. Therefore, it is timely to study biological specimens in which evolution likely optimised this interplay [8][9][10][11][12][13] . ...
... Structurally, this arises from the absence of correlations in the bondrotation angle between neighbouring tetrahedrally connected nodes 11 . The described locally tetrahedrally correlated but longrange disordered morphology is reminiscent of a recent theoretical study 16 . Starting from a periodic, tetrahedrally interconnected strut network disorder is introduced through different bond-angle distributions. ...
Article
Full-text available
While artificially photonic materials are typically highly ordered, photonic structures in many species of birds and insects do not possess a long-range order. Studying their order-disorder interplay sheds light on...
... Further exploration of the amorphous version of the Laves graph (or srs net, which is the skeletal graph of the Gyroid), proposed by Seller et al [37] would be highly relevant. In the cubic TPMS,differences in homogeneity related to domain sizes śand the resulting packing frustrationś are key distinguishing features of the Diamond and the Gyroid surfaces [17,38,39]. ...
Preprint
Full-text available
Bicontinuous geometries, both ordered and amorphous, are commonly found in many soft matter systems. Ordered bicontinuous phases can be modelled by periodic minimal surfaces, including Schoen's (G)yroid or Schwarz' (P)rimitive and (D)iamond surfaces. By contrast, a minimal surface model for amorphous phases has been lacking. Here, we study minimal surface models for amorphous bicontinuous phases, such as sponge phases. Using the Surface Evolver with a novel topology-stabilising minimisation scheme we numerically construct amorphous minimal surfaces from both a continuous random network model for amorphous Diamond and from a randomly perforated parallel sheet model. As per Hilbert's embedding theorem, the Gaussian curvature of these surfaces cannot be constant. Our analysis of Gaussian curvature variances finds no substantial long-wavelength curvature variations in the amorphous Diamond minimal surfaces. However, their Gaussian curvature variance is substantially larger than that of the cubic P, D and G surfaces. Our work demonstrates the superior curvature homogeneity of the cubic P, D and G surfaces compared to their entropy-favoured amorphous counterparts and to other periodic minimal surfaces. This general geometric result is relevant to bicontinuous structure formation in soft matter and biology across all length scales.
... how closely the environment conforms to tetrahedrality) [45][46][47][48] . It has also been demonstrated that for two dimensional systems, the degree of local order is closely linked to its hyperuniformity, and this in turn is linked to the existence of a photonic band gap 49,50 , although the causal relationship between hyperuniformity and the photonic band gap remains unclear 51,52 . The relationship established here between the homogeneity of local environments and the electronic band structure thus further em-phasises the importance of non-periodic order in amorphous materials for their reciprocal space properties. ...
Preprint
Understanding the structural origins of the properties of amorphous materials remains one of the most important challenges in structural science. In this study we demonstrate that local 'structural simplicity', embodied by the degree to which atomic environments within a material are similar to each other, is powerful concept for rationalising the structure of canonical amorphous material amorphous silicon (a-Si). We show, by restraining a reverse Monte Carlo refinement against pair distribution function (PDF) data to be simpler, that the simplest model consistent with the PDF is a continuous random network (CRN). A further effect of producing a simple model of a-Si is the generation of a (pseudo)gap in the electronic density of states, suggesting that structural homogeneity drives electronic homogeneity. That this method produces models of a-Si that approach the state-of-the-art without the need for chemically specific restraints (beyond the assumption of homogeneity) suggests that simplicity-based refinement approaches may allow experiment-driven structural modelling techniques to be developed for the wide variety of amorphous semiconductors with strong local order.
... Light transport experiments are traditionally a fertile playground for the study of novel transport regimes, often allowing to cast intriguing parallels between the propagation dynamics in seemingly unrelated research fields [1]. In most cases, however, peculiar structural and/or optical properties must be purposely introduced in a scattering material in order to observe anomalous diffusion, such as tailored spatial correlations [2][3][4][5][6][7][8][9][10][11] or wavefront-shaped illumination conditions [12][13][14]. For instance, transient anomalous transport regimes were previously observed in structures with carefully engineered fractal heterogeneous inclusions [15][16][17][18]. ...
Article
Full-text available
Transport processes underpin a wide variety of phenomena, ranging from chemistry, to physics and ecology. Despite their pervasiveness, however, several distinctive features of these processes are still elusive, making it difficult to recognize and classify the associated transport regimes. Using light scattering as a probe to explore different propagation regimes, we report on the experimental observation of non-self-similar light transport through turbid membranes. Our results show that a breakdown of self-similarity can arise for light waves even in the presence of isotropic and homogeneous disorder, and can be tuned by varying the turbidity of the system. By introducing the concept of self-similarity for light propagation, we provide a unified framework for the classification of light transport regimes—overcoming the dichotomy between normal and anomalous diffusion—and show that non-self-similar propagation is a common and experimentally accessible phenomenon. This insight can help to understand and model other scenarios where light transport is dominated by rare propagation events, such as in nonlinear and active media, but also in other fields of research beyond optics. Published by the American Physical Society 2024
... Varying interference conditions usually make structural color iridescent, i.e., the perceived color depends on the observation and viewing angle [1]. Disorder in photonic structures can result in angle-independent colors [2][3][4] and for random structures lead to broad-band scattering, producing white without absorption [5][6][7][8]. The presence of pigments in disordered structures, in contrast, always results in angle-independent colors due to wavelength-selective absorption of light [9][10][11]. ...
Article
Full-text available
Brilliant colors in nature arise from the interference of light with periodic nanostructures resulting in structural color. While such biological photonic structures have long attracted interest in insects and plants, they are little known in other groups of organisms. Unexpected in the kingdom of Amoebozoa, which assembles unicellular organisms, structural colors were observed in myxomycetes, an evolutionary group of amoebae forming macroscopic, fungal-like structures. Previous work related the sparkling appearance of Diachea leucopodia to thin film interference. Using optical and ultrastructural characterization, we here investigated the occurrence of structural color across 22 species representing two major evolutionary clades of myxomycetes including 14 genera. All investigated species showed thin film interference at the peridium, producing colors with hues distributed throughout the visible range that were altered by pigmentary absorption. A white reflective layer of densely packed calcium-rich shells is observed in a compound peridium in Metatrichia vesparium, whose formation and function are still unknown. These results raise interesting questions on the biological relevance of thin film structural colors in myxomycetes, suggesting they may be a by-product of their reproductive cycle.
Article
In photonic crystals, the propagation of light is governed by their photonic band structure, an ensemble of propagating states grouped into bands, separated by photonic band gaps. Due to discrete symmetries in spatially strictly periodic dielectric structures their photonic band structure is intrinsically anisotropic. However, for many applications, such as manufacturing artificial structural color materials or developing photonic computing devices, but also for the fundamental understanding of light-matter interactions, it is of major interest to seek materials with long range nonperiodic dielectric structures which allow the formation of isotropic photonic band gaps. Here, we report the first ever 3D isotropic photonic band gap for an optimized disordered stealthy hyperuniform structure for microwaves. The transmission spectra are directly compared to a diamond pattern and an amorphous structure with similar node density. The band structure is measured experimentally for all three microwave structures, manufactured by 3D laser printing for metamaterials with refractive index up to n=2.1. Results agree well with finite-difference-time-domain numerical investigations and a priori calculations of the band gap for the hyperuniform structure: the diamond structure shows gaps but being anisotropic as expected, the stealthy hyperuniform pattern shows an isotropic gap of very similar magnitude, while the amorphous structure does not show a gap at all. Since they are more easily manufactured, prototyping centimeter scaled microwave structures may help optimizing structures in the technologically very interesting region of infrared.
Article
The optics of correlated disordered media is a conceptually rich research topic emerging at the interface between the physics of waves in complex media and nanophotonics. Inspired by photonic structures in nature and enabled by advances in nanofabrication processes, recent investigations have unveiled how the design of structural correlations down to the subwavelength scale could be exploited to control the scattering, transport, and localization of light in matter. From optical transparency to superdiffusive light transport to photonic gaps, the optics of correlated disordered media challenges our physical intuition and offers new perspectives for applications. This review examines the theoretical foundations, state-of-the-art experimental techniques, and major achievements in the study of light interaction with correlated disorder, covering a wide range of systems: from short-range correlated photonic liquids to Lévy glasses containing fractal heterogeneities to hyperuniform disordered photonic materials. The mechanisms underlying light scattering and transport phenomena are elucidated on the basis of rigorous theoretical arguments. Ongoing research on mesoscopic phenomena such as transport phase transitions and speckle statistics and the current development of disorder engineering for applications such as light-energy management and visual appearance design are overviewed. Finally, special efforts are made to identify the main theoretical and experimental challenges to address in the near future.
Article
Full-text available
The bright and iridescent blue color from Morpho butterfly wings has attracted worldwide attentions to explore its mysterious nature for long time. Although the physics of structural color by the nanophotonic structures built on the wing scales has been well established, replications of the wing structure by standard top-down lithography still remains a challenge. This paper reports a technical breakthrough to mimic the blue color of Morpho butterfly wings, by developing a novel nanofabrication process, based on electron beam lithography combined with alternate PMMA/LOR development/dissolution, for photonic structures with aligned lamellae multilayers in colorless polymers. The relationship between the coloration and geometric dimensions as well as shapes is systematically analyzed by solving Maxwell's Equations with a finite domain time difference simulator. Careful characterization of the mimicked blue by spectral measurements under both normal and oblique angles are carried out. Structural color in blue reflected by the fabricated wing scales, is demonstrated and further extended to green as an application exercise of the new technique. The effects of the regularity in the replicas on coloration are analyzed. In principle, this approach establishes a starting point for mimicking structural colors beyond the blue in Morpho butterfly wings.
Article
Full-text available
Great progress has been made in understanding the atomic structure of metallic glasses, but there is still no clear connection between atomic structure and glass-forming ability. Here we give new insights into perhaps the most important question in the field of amorphous metals: how can glass-forming ability be predicted from atomic structure? We give a new approach to modelling metallic glass atomic structures by solving three long-standing problems: we discover a new family of structural defects that discourage glass formation; we impose efficient local packing around all atoms simultaneously; and we enforce structural self-consistency. Fewer than a dozen binary structures satisfy these constraints, but extra degrees of freedom in structures with three or more different atom sizes significantly expand the number of relatively stable, 'bulk' metallic glasses. The present work gives a new approach towards achieving the long-sought goal of a predictive capability for bulk metallic glasses.
Article
Full-text available
Non-crystalline photonic band gap (PBG) materials have received increasing attention, and sizeable PBGs have been reported in quasi-crystalline structures and, more recently, in disordered structures. Band structure calculations for periodic structures produce accurate dispersion relations, which determine group velocities, dispersion, density of states and iso-frequency surfaces, and are used to predict a wide-range of optical phenomena including light propagation, excited-state decay rates, temporal broadening or compression of ultrashort pulses and complex refraction phenomena. However, band calculations for non-periodic structures employ large super-cells of hundreds to thousands building blocks, and provide little useful information other than the PBG central frequency and width. Using stereolithography, we construct cm-scale disordered PBG materials and perform microwave transmission measurements, as well as finite-difference time-domain (FDTD) simulations. The photonic dispersion relations are reconstructed from the measured and simulated phase data. Our results demonstrate the existence of sizeable PBGs in these disordered structures and provide detailed information of the effective band diagrams, dispersion relation, iso-frequency contours, and their angular dependence. Slow light phenomena are also observed in these structures near gap frequencies. This study introduces a powerful tool to investigate photonic properties of non-crystalline structures and provides important effective dispersion information, otherwise difficult to obtain.
Article
Full-text available
We introduce the first designs for high-Q photonic cavities in slab architectures in hyperuniform disordered solids displaying isotropic band gaps. Despite their disordered character, hyperuniform disordered structures have the ability to tightly confine the TE-polarised radiation in slab configurations that are readily fabricable. The architectures are based on carefully designed local modifications of otherwise unperturbed hyperuniform dielectric structures. We identify a wide range of confined cavity modes, which can be classified according to their approximate symmetry (monopole, dipole, quadrupole, etc.) of the confined electromagnetic wave pattern. We demonstrate that quality factors (Q) Q>109Q>10^{9} can be achieved for purely 2D structures, and that for three--dimensional finite-height photonic slabs, quality factors Q>20,000Q>20,000 can be maintained.
Article
Full-text available
Combining molecular dynamics (MD) simulation and Voronoi polyhedral analyses, we discussed the microstructure evolution in liquid and glassy silicon during cooling by focusing on the fraction of various clusters. Liquid-liquid phase transition (LLPT) is detected in supercooled liquid silicon However, freezing the high-density liquid (HDL) to the glassy state is not achieved as the quenching rate goes up to 10(14) K/s. The polyamorphism in glassy silicon is found to be mainly associated with low-density liquid (LDL).
Article
Full-text available
Whiteness arises from diffuse and broadband reflection of light typically achieved through optical scattering in randomly structured media. In contrast to structural colour due to coherent scattering, white appearance generally requires a relatively thick system comprising randomly positioned high refractive-index scattering centres. Here, we show that the exceptionally bright white appearance of Cyphochilus and Lepidiota stigma beetles arises from a remarkably optimised anisotropy of intra-scale chitin networks, which act as a dense scattering media. Using time-resolved measurements, we show that light propagating in the scales of the beetles undergoes pronounced multiple scattering that is associated with the lowest transport mean free path reported to date for low-refractive-index systems. Our light transport investigation unveil high level of optimisation that achieves high-brightness white in a thin low-mass-per-unit-area anisotropic disordered nanostructure.
Article
Full-text available
Background The colorful wing patterns of butterflies, a prime example of biodiversity, can change dramatically within closely related species. Wing pattern diversity is specifically present among papilionid butterflies. Whether a correlation between color and the evolution of these butterflies exists so far remained unsolved.ResultsWe here investigate the Cattlehearts, Parides, a small Neotropical genus of papilionid butterflies with 36 members, the wings of which are marked by distinctly colored patches. By applying various physical techniques, we investigate the coloration toolkit of the wing scales. The wing scales contain two different, wavelength-selective absorbing pigments, causing pigmentary colorations. Scale ridges with multilayered lamellae, lumen multilayers or gyroid photonic crystals in the scale lumen create structural colors that are variously combined with these pigmentary colors.Conclusions The pigmentary and structural traits strongly correlate with the taxonomical distribution of Parides species. The experimental findings add crucial insight into the evolution of butterfly wing scales and show the importance of morphological parameter mapping for butterfly phylogenetics.
Article
The photonic structures that cause the electric blue colour of some Morpho butterflies are desirable within industry. They have been reproduced on very small areas to produce prototype biomimetic structures, yet there is currently no technique to accurately manufacture these structures at a commercial level. Here, tissue and cell culture techniques are explored as a possible method of producing large quantities of the blue scales of Morpho peliedes. This represents the first attempt to culture the wing scales of a Morpho butterfly, and many problems were encountered. These could be remedied in future studies, but it is concluded that cell culture is probably not a viable technique to provide the Morpho structure for industry in the near future. However, tissue culture experiments revealed important results of scale development.
Article
Photonic integration has long been pursued, but remains immature compared with electronics. Nanophotonics is expected to change this situation. However, despite the recent success of nanophotonic devices, there has been no demonstration of large-scale integration. Here, we describe the large-scale and dense integration of optical memories in a photonic crystal chip. To achieve this, we introduce a wavelength-addressable serial integration scheme using a simple cavity-optimization rule. We fully exploit the wavelength-division-multiplexing capability, which is the most important advantage of photonics over electronics, and achieve an extremely large wavelength-channel density. This is the first demonstration of the large-scale photonic integration of nanophotonic devices coupled to waveguides in a single chip, and also the first dense wavelength-division-multiplexing nanophotonic devices other than filters. This work paves the way for optical random-access memories and for a large-scale wavelength-division-multiplexing photonic network-on-chip.