plato’s cube and the natural geometry of fragmentation

8
Plato’s cube and the natural geometry of fragmentation abor Domokos a,b , Douglas J. Jerolmack c, d,1 , Ferenc Kun e , and J ´ anos T ¨ or ¨ ok a,f a MTA-BME Morphodynamics Research Group, Budapest University of Technology and Economics, Budapest, Hungary b Department of Mechanics, Budapest University of Technology and Economics, Budapest, Hungary e Department of Theoretical Physics, University of Debrecen, Debrecen, Hungary c Department of Earth and Environmental Science, University of Pennsylvania, Philadelphia, USA d Mechanical Engineering and Applied Mechanics, University of Pennsylvania, Philadelphia, USA f Department of Theoretical Physics, Budapest University of Technology and Economics, Budapest, Hungary 1 Corresponding author: [email protected] ABSTRACT Plato envisioned Earth’s building blocks as cubes, a shape rarely found in nature. The solar system is littered, however, with distorted polyhedra — shards of rock and ice produced by ubiquitous fragmentation. We apply the theory of convex mosaics to show that the average geometry of natural 2D fragments, from mud cracks to Earth’s tectonic plates, has two attractors: “Platonic” quadrangles and “Voronoi” hexagons. In 3D the Platonic attractor is dominant: remarkably, the average shape of natural rock fragments is cuboid. When viewed through the lens of convex mosaics, natural fragments are indeed geometric shadows of Plato’s forms. Simulations show that generic binary breakup drives all mosaics toward the Platonic attractor, explaining the ubiquity of cuboid averages. Deviations from binary fracture produce more exotic patterns that are genetically linked to the formative stress field. We compute the universal pattern generator establishing this link, for 2D and 3D fragmentation. Introduction Solids are stressed to their breaking point when growing crack networks percolate through the material 1, 2 . Failure by fragmentation may be catastrophic 1, 3 (Fig. 1), but this process is also exploited in industrial applications 4 . Moreover, fragmentation of rock and ice is pervasive within planetary shells 1, 5, 6 , and creates granular materials that are literally building blocks for planetary surfaces and rings throughout the solar system 610 (Fig. 1). Plato postulated that the ide- alised form of Earth’s building blocks is a cube, the only space-filling Platonic solid 11, 12 . We now know that there is a zoo of geometrically permissible polyhedra associated with fragmentation 13 (Fig. 2). Nevertheless, observed distributions of fragment mass 1417 and shape 1821 are self-similar, and models indicate that geometry (size and dimensionality) mat- ters more than energy input or material composition 16, 22, 23 in producing these distributions. Fragmentation tiles the Earth’s surface with telltale mosaics. Jointing in rock masses forms three-dimensional (3D) mosaics of polyhedra, often revealed to the observer by 2D planes at outcrops (Fig. 2). The shape and size of these polyhedra may be highly regular, even approaching Plato’s cube, or resemble a set of random intersecting planes 24 . Alternatively, quasi-2D patterns such as columnar joints sometimes form in solidifica- tion of volcanic rocks 25 . These patterns have been reproduced in experiments of mud and corn starch cracks, model 2D frag- mentation systems, where the following have been observed: fast drying produces strong tension that drives the formation of primary (global) cracks that criss-cross the sample and make “X” junctions 2527 (Fig. 3); slow drying allows the for- mation of secondary cracks that terminate at “T” junctions 26 ; and “T” junctions rearrange into ”Y” junctions 25, 28 to either maximise energy release as cracks penetrate the bulk 2931 , or during reopening-healing cycles from wetting/drying 32 (Fig. 3). Whether in rock, ice or soil, the fracture mosaics cut into stressed landscapes (Fig. 3), form pathways for focused fluid flow, dissolution and erosion that further disintegrate these materials 33 . Experiments and simulations provide anecdotal evidence that the geometry of fracture mosaics is genetically related to the formative stress field 34 . It is difficult to determine, however, if similarities in fracture patterns among different systems are more than skin deep. First, different communi- ties use different metrics to describe fracture mosaics and fragments, inhibiting comparison among systems and scales. Second, we do not know whether different fracture patterns represent distinct universality classes, or are merely descrip- tive categories applied to a pattern continuum. Third, it is unclear if and how 2D systems map to 3D. Here we introduce the mathematical framework of convex mosaics 35 to the fragmentation problem. This approach relies on two key principles: that fragment shape can be well approx- imated by convex polytopes 24 (2D polygons, 3D polyhedra; Fig. 2a); and that these shapes must fill space without gaps, since fragments form by the disintegration of solids. Without loss of generality (SI Section 1.1), we choose a model that ignores the local texture of fracture interfaces 36, 37 . Fragments can then be regarded as the cells of a convex mosaic 35 , which arXiv:1912.04628v1 [cond-mat.soft] 10 Dec 2019

Upload: others

Post on 18-May-2022

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Plato’s cube and the natural geometry of fragmentation

Plato’s cube and the natural geometry offragmentationGabor Domokosa,b, Douglas J. Jerolmackc, d,1, Ferenc Kune, and Janos Toroka,f

aMTA-BME Morphodynamics Research Group, Budapest University of Technology and Economics, Budapest,HungarybDepartment of Mechanics, Budapest University of Technology and Economics, Budapest, HungaryeDepartment of Theoretical Physics, University of Debrecen, Debrecen, HungarycDepartment of Earth and Environmental Science, University of Pennsylvania, Philadelphia, USAdMechanical Engineering and Applied Mechanics, University of Pennsylvania, Philadelphia, USAfDepartment of Theoretical Physics, Budapest University of Technology and Economics, Budapest, Hungary1Corresponding author: [email protected]

ABSTRACT

Plato envisioned Earth’s building blocks as cubes, a shape rarely found in nature. The solar system is littered, however,with distorted polyhedra — shards of rock and ice produced by ubiquitous fragmentation. We apply the theory of convexmosaics to show that the average geometry of natural 2D fragments, from mud cracks to Earth’s tectonic plates, has twoattractors: “Platonic” quadrangles and “Voronoi” hexagons. In 3D the Platonic attractor is dominant: remarkably, the averageshape of natural rock fragments is cuboid. When viewed through the lens of convex mosaics, natural fragments are indeedgeometric shadows of Plato’s forms. Simulations show that generic binary breakup drives all mosaics toward the Platonicattractor, explaining the ubiquity of cuboid averages. Deviations from binary fracture produce more exotic patterns that aregenetically linked to the formative stress field. We compute the universal pattern generator establishing this link, for 2D and 3Dfragmentation.

IntroductionSolids are stressed to their breaking point when growingcrack networks percolate through the material1, 2. Failureby fragmentation may be catastrophic1, 3 (Fig. 1), but thisprocess is also exploited in industrial applications4. Moreover,fragmentation of rock and ice is pervasive within planetaryshells1, 5, 6, and creates granular materials that are literallybuilding blocks for planetary surfaces and rings throughoutthe solar system6–10 (Fig. 1). Plato postulated that the ide-alised form of Earth’s building blocks is a cube, the onlyspace-filling Platonic solid11, 12. We now know that there is azoo of geometrically permissible polyhedra associated withfragmentation13 (Fig. 2). Nevertheless, observed distributionsof fragment mass14–17 and shape18–21 are self-similar, andmodels indicate that geometry (size and dimensionality) mat-ters more than energy input or material composition16, 22, 23 inproducing these distributions.

Fragmentation tiles the Earth’s surface with telltale mosaics.Jointing in rock masses forms three-dimensional (3D) mosaicsof polyhedra, often revealed to the observer by 2D planes atoutcrops (Fig. 2). The shape and size of these polyhedra maybe highly regular, even approaching Plato’s cube, or resemblea set of random intersecting planes24. Alternatively, quasi-2Dpatterns such as columnar joints sometimes form in solidifica-tion of volcanic rocks25. These patterns have been reproducedin experiments of mud and corn starch cracks, model 2D frag-mentation systems, where the following have been observed:fast drying produces strong tension that drives the formationof primary (global) cracks that criss-cross the sample and

make “X” junctions25–27 (Fig. 3); slow drying allows the for-mation of secondary cracks that terminate at “T” junctions26;and “T” junctions rearrange into ”Y” junctions25, 28 to eithermaximise energy release as cracks penetrate the bulk29–31, orduring reopening-healing cycles from wetting/drying32 (Fig.3). Whether in rock, ice or soil, the fracture mosaics cut intostressed landscapes (Fig. 3), form pathways for focused fluidflow, dissolution and erosion that further disintegrate thesematerials33.

Experiments and simulations provide anecdotal evidencethat the geometry of fracture mosaics is genetically relatedto the formative stress field34. It is difficult to determine,however, if similarities in fracture patterns among differentsystems are more than skin deep. First, different communi-ties use different metrics to describe fracture mosaics andfragments, inhibiting comparison among systems and scales.Second, we do not know whether different fracture patternsrepresent distinct universality classes, or are merely descrip-tive categories applied to a pattern continuum. Third, it isunclear if and how 2D systems map to 3D.

Here we introduce the mathematical framework of convexmosaics35 to the fragmentation problem. This approach relieson two key principles: that fragment shape can be well approx-imated by convex polytopes24 (2D polygons, 3D polyhedra;Fig. 2a); and that these shapes must fill space without gaps,since fragments form by the disintegration of solids. Withoutloss of generality (SI Section 1.1), we choose a model thatignores the local texture of fracture interfaces36, 37. Fragmentscan then be regarded as the cells of a convex mosaic35, which

arX

iv:1

912.

0462

8v1

[co

nd-m

at.s

oft]

10

Dec

201

9

Page 2: Plato’s cube and the natural geometry of fragmentation

Figure 1. Fragmentation across planets and scales. Leftcolumn shows planetary surfaces and rings: (a) Saturn’s ringscomposed of ice (inset); (b) Jupiter’s moon Europa showingcracked planetary shell; (c) polygonal cracks on Pluto; and(d) surface of the asteroid Bennu. Right column showsexample processes forming fragments on Earth: (e) icebergcalving; (f) rock falls; (g) volcanic eruptions that producepyroclastic flows, forming breccia deposits (inset); and (h)mine blasting. Image credits in Table S2.

may be statistically characterised by three parameters. Celldegree (v) is the average number of vertices of the polytopes,and nodal degree (n) is the average number of polytopes meet-ing at one vertex38: we call [n, v] the symbolic plane. Wedefine the third parameter 0≤ p≡ NR/(NR +NI)≤ 1 as theregularity of the mosaic. NR is the number of regular nodesin which cell vertices only coincide with other vertices, cor-responding in 2D to “X” and “Y” junctions with n = 4 and3, respectively. NI is the number of irregular nodes wherevertices lie along edges (2D) or faces (3D) of other cells, corre-sponding in 2D to “T” junctions of n = 2 (Fig. 2b). We definea regular (irregular) mosaic as having p = 1 (p = 0). For 3Dmosaics we also introduce f as the average number of faces.In contrast to other descriptions of fracture networks24, ourframework does not delineate stochastic from deterministicmosaics; networks made from random or periodic fracturesmay have identical parameter values (Fig. 3). This theoryprovides a global chart of geometrically admissible 2D and

3D mosaics in the symbolic plane.In this paper we measure the geometry of a wide variety

of natural 2D fracture mosaics and 3D rock fragments, andfind that they form clusters within the global chart. Remark-ably, the most significant cluster corresponds to the “Platonicattractor”: fragments with cuboid averages. Discrete ElementMethod (DEM) simulations of fracture mechanics show thatcuboid averages emerge from primary fracture under the mostgeneric stress field. Geometric simulations show how sec-ondary fragmentation by binary breakup drives any initialmosaic toward cuboid averages.

1 2D mosaics in theory and in natureThe geometric theory of 2D convex mosaics is essentiallycomplete35 and is given by the formula38:

v =2n

n− p−1, (1)

which delineates the admissible domain for convex mosaicswithin the [n, v] symbolic plane (Fig. 3) — i.e., the globalchart. Boundaries on the global chart are given by: (i) thep = 1 and p = 0 lines; and (ii) the overall constraints that theminimal degree of regular nodes and cells is 3, while the min-imal degree of irregular nodes is 2. We constructed geometricsimulations of a range of stochastic and deterministic mo-saics (see SI Section 2) to illustrate the continuum of patternscontained within the global chart (Fig. 3)

We describe two important types of mosaics, which help toorganise natural 2D patterns. First are primitive mosaics, pat-terns formed by binary dissection of domains. If the dissectionis global we have regular primitive mosaics (p = 1) composedentirely of straight lines which, by definition, bisect the en-tire sample. These mosaics occupy the point [n, v] = [4,4]in the symbolic plane35. In nature, the straight lines appearas primary, global fractures. Next, we consider the situationwhere the cells of a regular primary mosaic are sequentiallybisected locally. Irregular (T-type) nodes are created resultingin a progressive decrease p→ 0 and concomitant decreasen→ 2 toward an irregular primitive mosaic. The value v = 4,however, is unchanged by this process (Fig. 3) so in the limitwe arrive at [n, v] = [2,4]. In nature these local bisectionscorrespond to secondary fracturing3, 34. Fragments producedfrom primary vs. secondary fracture are indistinguishable.Further, any initial mosaic subject to secondary splitting ofcells will, in the limit, produce fragments with v = 4 (SI Sec-tion 1). Thus, we expect primitive mosaics associated withthe line v = 4 in the global chart to be an attractor in 2D frag-mentation, as noted by39, and we expect the average angle tobe a rectangle26 (Fig. 2). We call this the Platonic attractor.As a useful aside, a planar section of a 3D primitive mosaic(e.g., a rock outcrop) is itself a 2D primitive mosaic (Fig. 2).The second important pattern is Voronoi mosaics which are,in the averaged sense, hexagonal tilings [n, v] = [3,6]. Theyoccupy the peak of the 2D global chart (Fig. 3).

2/8

Page 3: Plato’s cube and the natural geometry of fragmentation

We measured a variety of natural 2D mosaics (SI Section2) and found, encouragingly, that they all lie within the globalchart permitted by Eq. 1. Mosaics close to the Platonic (v = 4)line include patterns known to arise under primary and/orsecondary fracture: jointed rock outcrops, mud cracks, andpolygonal frozen ground. Mosaics close to Voronoi includemud cracks and, most intriguingly, Earth’s tectonic plates.Hexagonal mosaics are known to arise in the limit for systemssubject to repeated cycles of fracturing and healing25 (Fig. 3).We thus consider Voronoi mosaics to be a second importantattractor in 2D. Horizontal sections of columnar joints alsobelong to this geometric class; however, their evolution isinherently 3D, as we discuss in section 2.

It is known that Earth’s tectonic plates meet almost ex-clusively at “Y” junctions; there is debate, however, aboutwhether this “Tectonic Mosaic” formed entirely from surfacefragmentation, or contains a signature of the structure of man-tle dynamics underneath5, 40, 41. We examine the tectonic plateconfiguration41 as a 2D convex mosaic, treating the Earth’scrust as a thin shell. We find [n, v] = [3.0,5.8], numbers thatare remarkably close to a Voronoi mosaic. Indeed, the slightdeviation from [n, v] = [3,6] is because the Earth’s surface is aspherical manifold, rather than planar (SI Section 2.3). Whilethis analysis doesn’t solve the surface/mantle question, it sug-gests that the Tectonic Mosaic has evolved from episodes ofbrittle fracture and healing.

The rest of our observed natural 2D mosaics plot betweenthe Platonic and Voronoi attractors (Fig. 3). We suspectthat these fractured landscapes, which include mud cracksand permafrost, either: initially formed as regular primitivemosaics and are in various stages of evolution toward theVoronoi attractor; or were Voronoi mosaics that are evolvingvia secondary fracture towards the Platonic attractor.

2 Extension to 3D mosaicsThere is no formula for 3D convex mosaics analogous to thep = 1 line of Eq. 1 that defines the global chart. There existsa conjecture, however, with a strong mathematical basis38; atpresent this conjecture extends only to regular mosaics. Wedefine the harmonic degree as h = nv/(n+ v). The conjectureis that d < h ≤ 2d−1, where d is system dimension. For 2Dmosaics we obtain the known result35, 38 h = 2, consistentwith the p = 1 substitution in Eq. 1. In 3D the conjectureis equivalent to 3 < h ≤ 4, predicting that all regular 3Dconvex mosaics live within a narrow band in the symbolic[n, v] plane38 (Fig. 4). Plotting a variety of well-studiedperiodic and random 3D mosaics (SI Section 3), we confirmthat all of them are indeed confined to the predicted 3D globalchart (Fig. 4). Unlike the 2D case, we cannot directly measuren in most natural 3D systems. We can, however, measurethe polyhedral cells: the average numbers f , v of faces andvertices, respectively. Values for [ f , v] may be plotted inwhat we call the Euler plane, where the lines bounding thepermissible domain correspond to simple polyhedra (upper)where vertices are adjacent to three edges and three faces, and

their duals which have triangular faces (lower; Fig. 4). Simplepolyhedra arise as cells of mosaics in which the intersectionsare generic — i.e., at most three planes intersect at one point— and this does not allow for odd values of v.

As in 2D, 3D regular primitive mosaics are created by inter-secting global planes. These mosaics occupy the point [n, v] =[8,8] on the 3D global chart (Fig. 4). Cells of regular primi-tive mosaics have cuboid averages [ f , v] = [6,8]35, 38. This isthe Platonic attractor, marked by the v = 8 line in the globalchart. 3D Voronoi mosaics, similar to their 2D counterparts,are associated with the Voronoi tessellation defined by somerandom process. If the latter is a Poisson process then weobtain35 [n, v] = [4,27.07], [ f , v] = [15.51,27.07] (Fig. 4).

Prismatic mosaics are created by regarding the 2D patternas a base, that is extended in the normal direction. The pris-matic mosaic constructed from a 2D primitive mosaic hascuboid averages, and is therefore statistically equivalent to a3D primitive mosaic. The prismatic mosaic created from a 2DVoronoi base is what we call a columnar mosaic, and it hasdistinct statistical properties: [n, v] = [6,12], [ f , v] = [8,12].Thus, the three main natural extensions of the two dominant2D patterns are 3D primitive, 3D Voronoi, and columnar mo-saics.

Regular primitive mosaics appear to be the dominant 3Dpattern resulting from primary fracture of brittle materials42.Moreover, dynamic brittle fracture produces binary breakup insecondary fragmentation23, 43, driving the 3D averages [ f , v]towards the Platonic attractor. The most common example innature is fractured rock (Figs. 2, 4). The other two 3D mosaicsare more exotic, and seem to require more specialised condi-tions to form in nature. Columnar joints like the celebratedGiants Causeway, formed by the cooling of large basaltic rockmasses25, 31, 44 (Fig. 4), appear to correspond to columnarmosaics. In these systems, the hexagonal arrangement anddownward (normal) penetration of cracks arise as a conse-quence of maximising energy release29–31. The only potentialexample of 3D Voronoi mosaics that we know of is septariannodules, such as the famous Moeraki Boulders45 (Fig. 4).These enigmatic concretions have complex growth and com-paction histories, and contain internal cracks that intersect thesurface46. Similar to primitive mosaics, the intersection of 3DVoronoi mosaics with a surface is a 2D Voronoi mosaic.

3 Connecting primary fracture patterns tomechanics with simulations

We hypothesise that primary fracture patterns are geneticallylinked to distinct stress fields, in order of most generic tomost rare. In a 2D homogeneous stress field we may de-scribe the stress tensor with eigenvalues |σ1| ≥ |σ2| andcharacterise the stress state by the dimensionless parame-ter µ = σ2/σ1, whose admissible domain is µ ∈ [−1,1].In 3D this corresponds to eigenvalues |σ1| ≥ |σ2| ≥ |σ3|,stress state µ1 = σ2/σ1,µ2 = σ3/σ1, i = sgn(σ1), and do-main µ1,µ2 ∈ [−1,1], |µ1| ≥ |µ2| (and it is double covereddue to i = ±1) (Fig. 5). There is a unique map from these

3/8

Page 4: Plato’s cube and the natural geometry of fragmentation

(a1) (f,v)=(6,8)

(a2) (f,v)=(6,8)

(a3) (f,v)=(5,6)

(a4) (f,v)=(6,7)

(b1) (b2)

(a5) (f,v)=(9,14)

(a6) (f,v)=(6,8)

(a7) (f,v)=(7,9)

Figure 2. (a1-a7) Natural fragments approximated byconvex polyhedra. (b1) Granite wall showing global cracks.(b2) Approximation of fragmentation pattern by regularprimitive mosaic (black lines) and its irregular version withsecondary cracks (red lines).

stress field parameters to the location of the resultant fracturemosaics in the global chart; we call this map the mechanicalpattern generator. (Results may be equivalently cast on theFlinn diagram47 commonly used in structural geology: Fig.S10). The 2D pattern generator is described by the singlescalar function v(µ) (and n can be computed from [1]), whilethe 3D pattern generator is characterised by scalar functionsn(µ1,µ2, i), v(µ1,µ2, i), f (µ1,µ2, i). Computing the full pat-tern generator is beyond the scope of the current paper, evenin 2D. Instead, we perform generic DEM simulations48, 49 ofa range of scenarios to interpret the important primary mosaicpatterns described above (see Methods).

In 2D, we find that pure shear produces regular primitivemosaics (Fig. 3), implying v(−1) ≈ 4. This corresponds tothe Platonic attractor. In contrast, hydrostatic tension createsregular Voronoi mosaics (Fig. 3; SI Section 4 ) such thatv(1)≈ 6 — the Voronoi attractor. Both are in agreement withour expectations.

In 3D, we first conducted DEM simulations of hard materi-als at [µ1,µ2] locations corresponding to shear, uniform 2Dtension and uniform 3D tension ([−0.5,−0.25], [1,−0.2] and[1,1]), respectively (Methods, SI Section 4). The resultingmosaics displayed the expected fracture patterns for brittle ma-terials: primitive, columnar and Voronoi, respectively (Fig. 5).To obtain a global, albeit approximate, picture of the 3Dpattern generator we ran additional DEM simulations that uni-formly sampled the stress space on a 9×9 (∆µ=0.25) grid,for both i=+ 1 and i=− 1 (SI Section 4). The constructedpattern generator demarcates the boundaries in stress-statespace that separate the three primary fracture patterns (Fig.5). The vast proportion of this space is occupied by primitivemosaics, which are also the only pattern generated under neg-ative volumetric stress. Such compressive stress conditionsare pervasive in natural rocks. Columnar mosaics are a distantsecond in terms of frequency of occurrence; they occupy anarrow stripe in the stress space. Most rare are Voronoi mo-

(13)

(18)

(15)

(16) (17)

(14)

(19)

(20)(23)

(11)

ത𝑛

ҧ𝑣

6

6

5423

(2)

(3)

(4)

(5)

2D VORONOI MOSAICS

2D DELAUNAY MOSAICS

2D PRIMITIVE MOSAICS

(12)

(13)(14)

(16)

(17)(18) (19)

(20)

(21)(6) (8)

(9)

(1) (11) (22)

MECHANICS

GEOMETRY

(15)

(22)

(12)

(6)

(1)

(2)

(3)

(4)

(5)

(7)

(7)

(8)

(9)

(10)

(10) (21)

(21)

NATURE

Figure 3. Mosaics in 2D. Left: symbolic plane [n, v] withgeometrically admissible domain (defined in Eq. (1)) shadedgrey. Patterns (1-7) marked with black circles aredeterministic periodic patterns. Patterns (8-12) are geometricsimulations of random mosaics: (8) regular primitive; (9-10)advanced (irregular) primitive; (11) Poisson-Voronoi; and(12) Poisson–Delaunay. Red squares (13-21) correspond toanalysed images of natural 2D mosaics shown on the right:(13) columnar joints, Giant’s Causeway; (14) mud cracks;(15) tectonic plates; (16) Martian surface; (17) permafrost inAlaska; (18) mud cracks; (19) dolomite outcrop; (20)permafrost in Alaska; and (21) granite rock surface. Imagecredits in Table S2. Patterns (22-23) are generated by genericDEM simulation (see Methods): (22) general stress state witheigenvalues σ1 > σ2; and (23) isotropic stress state withσ1 = σ2.

saics which only occur in a single corner of the stress space(Fig. 5). Boundaries separating the three patterns shiftedsomewhat for simulations that used softer materials (Fig. S9),but the ranking did not. These primary fracture mosaics serveas initial conditions for secondary fracture. While our DEMsimulations do not model secondary fracture, we remind thereader that binary breakup drives any initial mosaic toward anirregular primitive mosaic with cuboid averages (SI Section1) — emphasising the strength of the Platonic attractor.

4 Geometry of natural 3D fragmentsBased on the pattern generator (Fig. 5) we expect that natu-ral 3D fragments should have cuboid properties on average,[ f , v] = [6,8]. To test this we collected 556 particles fromthe foot of a weathering dolomite rock outcrop (Fig. 6) andmeasured their values of f and v, plus mass and additionalshape descriptors (see Methods; SI Section 5). We find strik-ing agreement: the measured averages [ f , v] = [6.63,8.93] arewithin 12 % of the theoretical prediction, and distributions forf and v are centred around the theoretical values. Moreover,odd values for v are much less frequent than even values, illus-trating that natural fragments are well approximated by simplepolyhedra (Fig. 6). We regard these results as direct confir-mation of the hypothesis, while also recognising significant

4/8

Page 5: Plato’s cube and the natural geometry of fragmentation

0

5

10

15

20

25

30

0 5 10 15 20 25 30

0

5

10

15

20

25

30

0 5 10 15

ത𝑛

ҧ𝑣

ҧ𝑓

ҧ𝑣

PR

IMIT

IVE

VO

RO

NO

IC

OLU

MN

AR

EXPERIMENTS

(A)

(B)

(C)

CONJECTURE ON ADMISSIBLE BAND

Figure 4. Mosaics in 3D. 28 uniform honeycombs, theirduals, Poisson-Voronoi, Poisson-Delaunay and primitiverandom mosaics plotted on the parameter planes. Left panel:[n, v] plane, where continuous black line corresponds toprismatic mosaics38. Shaded grey area marks the predicteddomain based on the conjecture d < h≤ 2d−1. Right panel:[ f , v] plane, where straight black lines correspond to simplepolyhedra (top) and their duals (bottom); for the tetrahedron[ f , v] = [4,4], these two are identical. Mosaics observed inNature marked by red circles on both panels: (A) 3D Voronoimosaics observed on some special septarian nodules46; (B)columnar mosaics observed on columnar joints; and (C) 3Dprimitive mosaic observed on fractured rock.

variability in the natural data.

To better understand the full distributions of fragmentshapes, we used geometric simulations of regular and irregularprimitive mosaics. The cut model simulates regular primitivemosaics as primary fracture patterns by intersecting an initialcube with global planes (Fig.6) while the break model sim-ulates irregular primitive mosaics resulting from secondaryfragmentation processes. We fit both of these models to theshape descriptor data using three parameters: one for thecutoff in the mass distribution, and two accounting for un-certainty in experimental protocols (see Methods; SI Section5). The best fit model, which corresponds to a moderatelyirregular primitive mosaic, produced topological shape distri-butions that are very close to those of natural fragments (meanvalues [ f , v] = [6.58,8.74]). We also analysed a much larger,previously collected data set (3728 particles) containing adiversity of materials and formative conditions19. Althoughvalues for v and f were not reported, measured values forclassical shape descriptors19, 21 could be used to fit to the cutand break models (SI Section 5). We find very good agree-ment (R2>0.95), providing further evidence that natural 3Dfragments are predominantly formed by binary breakup (SI,Fig.12). Finally, we use the cut model to demonstrate how 3Dprimitive fracture mosaics converge asymptotically toward thePlatonic attractor as more fragments are produced (Fig. 6).

−1 0 1

1

02D 1D

𝜇1 = 𝜎2/𝜎1

𝜇2=𝜎3/𝜎

1

(B): COLUMNAR MOSAICS

(A) VORONOI MOSAICS

(C) PRIMITIVE MOSAICS

2D

Figure 5. Illustration of the 3D pattern generator. Left: thei = 1 leaf of the [µ1,µ2] plane. Colours indicate patternsobserved in 74 DEM simulations. Blue: primitive, green:columnar, red: Voronoi. Right: (A) 3D Voronoi-type mosaicat µ1 = µ2 = i = 1. Observe surface patterns on all planarsections agreeing with 2D Voronoi-type tessellations, andalso with surface patterns of the shown septarian nodule. (B)Columnar mosaic at µ1 = 1,µ2 = 0, i = 1. Observe surfacepattern on horizontal section agreeing with 2D Voronoi-typetessellation and parallel vertical lines on vertical sections,both matching observation on the illustrated basalt columnarjoints. (C) 3D primitive mosaic at µ1 =−0.5, µ2 =−0.25,i = 1. Observe surface patterns on all sections agreeing with2D primitive mosaics, and also corresponding to fracturepatterns on the illustrated rock. Image credits in Table S2.

5 Discussion and implications

The application and extension of the theory of convex mo-saics provides a new lens to organise all fracture mosaics —and the fragments they produce — into a geometric globalchart. There are attractors in this global chart, arising fromthe mechanics of fragmentation. The Platonic attractor pre-vails in nature because binary breakup is the most genericfragmentation mechanism, producing averages correspondingto quadrangle cells in 2D and cuboid cells in 3D. Remarkably,a geometric model of random intersecting planes can accu-rately reproduce the full shape distribution of natural rockfragments. Our findings illustrate the remarkable prescienceof Plato’s cubic Earth model. One cannot, however, directly‘see’ Plato’s cubes; rather, their shadows are seen in the statis-tical averages of many fragments. The relative rarity of othermosaic patterns in nature make them exceptions that provethe rule. Voronoi mosaics are a second important attractor in2D systems such as mud cracks, where healing of fracturesreorganises junctions to form hexagonal cells. Such healing israre in natural 3D systems. Accordingly, columnar mosaicsarise only under specific stress fields, that are consistent withiconic basalt columns experiencing contraction under direc-tional cooling. 3D Voronoi mosaics require very special stressconditions, hydrostatic tension, and may describe rare andpoorly understood concretions known as septarian nodules.

5/8

Page 6: Plato’s cube and the natural geometry of fragmentation

556 DOLOMITE FRAGMENTS COLLECTED AND MEASURED

600.000 DIGITAL FRAGMENTS PRODUCED BY N=50 INTERSECTING PLANES

(a) (b)

(c) (d) (e)

Figure 6. Natural rock fragments and geometric modelling.(a): Dolomite rock outcrop at Hármashatárhegy, Hungary,from which we sampled and measured natural fragmentsaccumulating at its base, highlighted in inset. (b): The cutmodel shown with N = 50 intersecting planes, and examplesof digital fragments (inset) drawn from the 600,000fragments produced. (c): Evolution of f and v showingconvergence toward the cuboid values of 8 and 6,respectively, with increasing N. (d) and (e): Probabilitydistributions of f and v, respectively, for natural dolomitefragments and fits of the cut and break models.

We have shown that Earth’s tectonic mosaic has a geometrythat likely arose from brittle fracture and healing, consistentwith what is known about plate tectonics5 (Fig. 3). Thisopens the possibility of inferring stress history from observedfracture mosaics. Space missions are accumulating an ever-growing catalogue of 2D and 3D fracture mosaics from di-verse planetary bodies, that challenge understanding (Fig. 1).Geometric analysis of surface mosaics may inform debateson planetery dynamics, such as whether Pluto’s polygonalsurface (Fig. 1c) is a result of brittle fracture or vigorousconvection7. Another potential application is using 2D out-crop exposures to estimate the 3D statistics of joint networksin rock masses, which may enhance prediction of rock fallhazards and fluid flow50.

6 Methods

6.1 Methods of mechanical simulationsInitial samples were randomised cubic assemblies of spheresglued together, with periodic boundary conditions in all direc-tions. The glued contact was realised by a flat elastic cylinderconnecting the two particles which was subject to deforma-tion from the relative motion of the glued particles. Forcesand torques on the particles were calculated based on the de-formation of the gluing cylinder. The connecting cylinderbroke permanently if the stress acting upon it exceeded theTresca criterion49. Stress field was implemented by slowlydeforming the underlying space. In order to avoid that thereis only one percolating crack, we have set a strong viscousfriction between the particles and the underlying space. This

acts as a homogeneous drag to the particles which ensures ahomogeneous stress field in the system.

For any given shear rate the fragment size is controlled bythe particle space viscosity and the Tresca criterion limit. Weset values that produce reasonable-sized fragments relativeto our computational domain, allowing us to characterise themosaics. Another advantage of the periodic system was thatwe could avoid any wall effect that would distort the stressfield. We note here that it is possible to slowly add a shearcomponent to the isotropic tensile shear test and obtain a struc-ture which has average values of n and v that are between theprimitive mosaics and the Voronoi case. Details of mechanicalsimulations are discussed in SI Section 4.

6.2 Methods for fitting geometric model results tofield data

In the simulation we first computed a regular primitive mosaicby dissecting the unit cube with 50 randomly chosen planes,resulting in 6×105 fragments. We refer to this simulation asthe cut model. Subsequently we further evolved the mosaic bybreaking individual fragments. We implemented a standardmodel of binary breakup14, 19 to evolve the cube by secondaryfragmentation: at each step of the sequence, fragments eitherbreak with a probability pb into two pieces, or keep their cur-rent size until the end of the process with a probability 1− pb.The cutting plane is placed in a stochastic manner by takinginto account that it is easier to break a fragment in the middleperpendicular to its largest linear extent. Inspired by similarcomputational models19, we used pb dependent on axis ratios(see SI Section 5). This computation, which we call the breakmodel, provides an approximation to an irregular primitivemosaic; this secondary fragmentation process influenced thenodal degree n, but not [v, f ].

In order to compare numerical results with the experimentaldata obtained by manual measurements, we have to take intoaccount several sampling biases. First, there is always a lowercutoff in size for the experimental samples. We implementedthis in simulations by selecting only fragments with m > m0,m0 being the cutoff threshold. Second, there is experimentaluncertainty when determining shape descriptors — especiallymarginally stable or unstable equilibria for the larger dataset(see SI Section 5). We implemented this in the computationsby letting the location of the centre of mass be a randomvariable with variation σ0 chosen to be small with respect tothe smallest diameter of the fragment. We kept only thoseequilibria which were found in 95% of the cases. Third, thereis experimental uncertainty in finding very small faces. Weimplemented this into the computations by assuming that facessmaller than A0P will not be found by experimenters, whereP denotes the smallest projected area of the fragment. Usingthe above three parameters we fitted the seven computationalhistograms to the seven experimental ones by minimising thelargest deviation, and we achieved matches with R2

max ≥ 0.95from all histograms (see SI Section 5 for details). Results forthe small data set are shown in Fig. 6.

6/8

Page 7: Plato’s cube and the natural geometry of fragmentation

AcknowledgementThis research was supported by NKFIH grants K119245 (GD)K116036 (JT) and K119967 (FK) and EMMI FIKP grantVIZ (GD, JT) The authors thank Krisztián Halmos for hisinvaluable help with field data measurements.

Author ContributionGD originated the concept of the paper, led mathematical andphilosophical components, and contributed to field data. DJJcontributed to geophysical interpretation, and led formulationof the manuscript. FK led interpretation of fragmentation me-chanics, and contributed to field data. JT performed geometricand DEM computations, and led data analysis. All authorscontributed to writing the manuscript.

References1. Turcotte, D. L. Fractals and chaos in geology and geo-

physics (Cambridge University Press, 1997).

2. Kawamura, H., Hatano, T., Kato, N., Biswas, S. &Chakrabarti, B. K. Statistical physics of fracture, friction,and earthquakes. Rev. Mod. Phys. 84, 839–884 (2012).

3. Adler, P. M. & Thovert, J.-F. Fractures and FractureNetworks (Springer, Dordrecht, 1999).

4. Prasher, C. L. Crushing and grinding process handbook(John Wiley & Sons, 1987).

5. Mallard, C., Coltice, N., Seton, M., Müller, R. D. &Tackley, P. J. Subduction controls the distribution andfragmentation of Earth’s tectonic plates. Nature 535,140–143 (2016).

6. McEwen, A. S. Tidal reorientation and the fracturing ofjupiter’s moon europa. Nature 321, 49–51 (1986).

7. Trowbridge, A., Melosh, H., Steckloff, J. & Freed, A. Vig-orous convection as the explanation for pluto’s polygonalterrain. Nature 534, 79 (2016).

8. Cuzzi, J. N. et al. An evolving view of saturn’s dynamicrings. Science 327, 1470–1475 (2010).

9. Brooker, L. et al. Clastic polygonal networks aroundlyot crater, mars: Possible formation mechanisms frommorphometric analysis. Icarus 302, 386 – 406 (2018).

10. Brilliantov, N. et al. Size distribution of particles insaturn’s rings from aggregation and fragmentation. Proc.Natl. Acad. Sci. 112, 9536–9541, DOI: 10.1073/pnas.1503957112 (2015).

11. Domokos, G. & Lángi, Z. Plato’s error and a mean fieldformula for convex mosaics. Axiomathes DOI: 10.1007/s10516-019-09455-w (2019).

12. Plato. Timaeus (translated by B. Jowett) (Aeterna Press.Original edition MacMillan and Co. London, 1892.,2015).

13. Grünbaum, B. Uniform tilings of 3-space. Geombina-torics 4, 49–56 (1994).

14. Ishii, T. & Matsushita, M. Fragmentation of long thinglass rods. J. Phys. Soc. Jpn. 61, 3474–3477 (1992).

15. Oddershede, L., Dimon, P. & Bohr, J. Self-organized crit-icality in fragmenting. Phys. Rev. Lett. 71, 3107 (1993).

16. Wittel, F. K., Kun, F., Herrmann, H. J. & Kröplin, B. H.Fragmentation of shells. Phys. Rev. Lett. 93, 035504(2004).

17. Timár, G., Blömer, J., Kun, F. & Herrmann, H. J. Newuniversality class for the fragmentation of plastic materi-als. Phys. Rev. Lett. 104, 095502 (2010).

18. Kun, F., Wittel, F. K., Herrmann, H. J., Kröplin, B. H. &Maloy, K. J. Scaling behaviour of fragment shapes. Phys.Rev. Lett. 96, 025504 (2006).

19. Domokos, G., Kun, F., Sipos, A. A. & Szabó, T. Univer-sality of fragment shapes. Sci. Reports 5, 9147 (2015).

20. Ma, G., Zhou, W., Zhang, Y., Wang, Q. & Chang, X.Fractal behavior and shape characteristics of fragmentsproduced by the impact of quasi-brittle spheres. PowderTechnol. 325, 498 – 509 (2018).

21. Szabó, T., Domokos, G., Grotzinger, J. P. & Jerolmack,D. J. Reconstructing the transport history of pebbles onMars. Nat. Commun. 6, 8366 (2015).

22. Steacy, S. & Sammis, C. An automaton for fractal patternsof fragmentation. Nature 353, 250–252 (1991).

23. Aström, J. A., Ouchterlony, F., Linna, R. P. & Timonen, J.Universal dynamic fragmentation in d dimensions. Phys.Rev. Lett. 92, 245506 (2004).

24. Dershowitz, W. & Einstein, H. Characterizing rock jointgeometry with joint system models. Rock mechanics rockengineering 21, 21–51 (1988).

25. Goehring, L. Evolving fracture patterns: columnar joints,mud cracks and polygonal terrain. Philos. TransactionsRoyal Soc. A: Math. Phys. Eng. Sci. 371, 20120353(2013).

26. Ma, X., Lowensohn, J. & Burton, J. C. Universal scalingof polygonal desiccation crack patterns. Phys. Rev. E 99,012802 (2019).

27. Goehring, L., Nakahara, A., Dutta, T., Kitsunezaki, S.& Tarafdar, S. Desiccation Cracks and their Patterns:Formation and Modelling in Science and Nature (JohnWiley & Sons, 2015).

28. Goehring, L., Conroy, R., Akhter, A., Clegg, W. J. &Routh, A. F. Evolution of mud-crack patterns duringrepeated drying cycles. Soft Matter 6, 3562–3567 (2010).

29. Aydin, A. & Degraff, J. M. Evoluton of polygonal fracturepatterns in lava flows. Science 239, 471–476 (1988).

30. Jagla, E. A. & Rojo, A. G. Sequential fragmentation: Theorigin of columnar quasihexagonal patterns. Phys. Rev. E65, 026203 (2002).

7/8

Page 8: Plato’s cube and the natural geometry of fragmentation

31. Hofmann, M., Anderssohn, R., Bahr, H.-A., Weiß, H.-J.& Nellesen, J. Why hexagonal basalt columns? Phys.Rev. Lett. 115, 154301 (2015).

32. Cho, H.-J. J., Lu, N. B., Howard, M. P., Adams, R. A. &Datta, S. S. Crack formation and self-closing in shrink-able, granular packings. Soft matter (2019).

33. Voigtländer, A. & Krautblatter, M. Breaking rocks madeeasy. blending stress control concepts to advance geomor-phology. Earth Surf. Process. Landforms 44, 381–388(2019).

34. Bahat, D., Rabinovitch, A. & Frid, V. Tensile Fracturingin Rocks (Springer-Verlag Berlin Heidelberg, 2005).

35. Schneider, R. & Weil, W. Stochastic and integral geome-try (Springer Science & Business Media, 2008).

36. Mills, K. Fractography (American Society of Metals(ASM) handbook, volume 12, 1991).

37. Bahat, D. & Rabinovitch, V., A. adn Frid. Tensile Frac-turing in Rocks (Springer, Berlin, 2005).

38. Domokos, G. & Lángi, Z. On some average propertiesof convex mosaics. Exp. Math. DOI: 10.1080/10586458.2019.1691090 (2019).

39. Bohn, S., Pauchard, L. & Couder, Y. Hierarchical crackpattern as formed by successive domain divisions. Phys.Rev. E 71, 046214 (2005).

40. Anderson, D. L. How many plates? Geology 30, 411–414(2002).

41. Bird, P. An updated digital model of plate boundaries.Geochem. Geophys. Geosystems 4 (2003).

42. Hernández, G. & Herrmann, H. J. Discrete models fortwo- and three-dimensional fragmentation. Phys. A: Stat.Mech. its Appl. 215, 420 – 430 (1995).

43. Kekalainen, P., Aström, J. A. & Timonen, J. Solution forthe fragment-size distribution in a crack-branching modelof fragmentation. Phys. Rev. E 76, 026112 (2007).

44. Lamur, A. et al. Disclosing the temperature of columnarjointing in lavas. Nat. Commun. 9, 1432 (2018).

45. Boles, J. R., Landis, C. A. & Dale, P. The MoerakiBoulders; anatomy of some septarian concretions. J.Sedimentary Res. 55, 398–406 (1985).

46. Astin, T. Septarian crack formation in carbonate con-cretions from shales and mudstones. Clay Miner. 21,617–631 (1986).

47. Flinn, D. On folding during three dimensional progressivedeformation. Quart. Journ. Geol. Soc. Lond. 118, 385–428 (1962).

48. Plimpton, S. Fast parallel algorithms for short-rangemolecular dynamics. J. computational physics 117, 1–19(1995).

49. Brendel, L., Török, J., Kirsch, R. & Bröckel, U. A con-tact model for the yielding of caked granular materials.Granul. Matter 13, 777–786 (2011).

50. Council, N. R. et al. Rock fractures and fluid flow:contemporary understanding and applications (NationalAcademies Press, 1996).

8/8