aglimpseintodiscrete differentialgeometry...and combinatorics, since the geometry of a map is...

7
COMMUNICATION A Glimpse into Discrete Differential Geometry Keenan Crane and Max Wardetzky Communicated by Joel Hass EDITOR’S NOTE. The organizers of the two-day AMS Short Course on Discrete Differential Geometry have kindly agreed to provide this introduction to the subject. The AMS Short Course runs in conjunction with the 2018 Joint Mathematics Meetings. The emerging field of discrete differential geometry (DDG) studies discrete analogues of smooth geometric objects, providing an essential link between analytical descrip- tions and computation. In recent years it has unearthed a rich variety of new perspectives on applied problems in computational anatomy/biology, computational mechan- ics, industrial design, computational architecture, and digital geometry processing at large. The basic philoso- phy of discrete differential geometry is that a discrete object like a polyhedron is not merely an approximation of a smooth one, but rather a differential geometric object in its own right. In contrast to traditional numerical anal- ysis which focuses on eliminating approximation error in the limit of refinement (e.g., by taking smaller and smaller finite differences), DDG places an emphasis on the so-called “mimetic” viewpoint, where key properties of a system are preserved exactly, independent of how large or small the elements of a mesh might be. Just as algorithms for simulating mechanical systems might seek to exactly preserve physical invariants such as total energy or momentum, structure-preserving models of Keenan Crane is assistant professor of computer science at Carnegie Mellon University. His e-mail address is kmcrane@cs .cmu.edu. Max Wardetzky is professor of mathematics at University of Göt- tingen. His e-mail address is [email protected] .de. For permission to reprint this article, please contact: [email protected]. DOI: http://dx.doi.org/10.1090/noti1578 Figure 1. Discrete differential geometry reimagines classical ideas from differential geometry without reference to differential calculus. For instance, surfaces parameterized by principal curvature lines are replaced by meshes made of circular quadrilaterals (top left), the maximum principle obeyed by harmonic functions is expressed via conditions on the geometry of a triangulation (top right), and complex-analytic functions can be replaced by so-called circle packings that preserve tangency relationships (bottom). These discrete surrogates provide a bridge between geometry and computation, while at the same time preserving important structural properties and theorems. discrete geometry seek to exactly preserve global geo- metric invariants such as total curvature. More broadly, DDG focuses on the discretization of objects that do not naturally fall under the umbrella of traditional numerical analysis. This article provides an overview of some of the themes in DDG. November 2017 Notices of the AMS 1153

Upload: others

Post on 02-Jun-2020

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: AGlimpseintoDiscrete DifferentialGeometry...and combinatorics, since the geometry of a map is determined entirely by incidence relationships.1 On the flip side, this means a different

COMMUNICATION

A Glimpse into DiscreteDifferential Geometry

Keenan Crane and Max WardetzkyCommunicated by Joel Hass

EDITOR’S NOTE. The organizers of the two-day AMSShort Course on Discrete Differential Geometry havekindly agreed to provide this introduction to thesubject. The AMS Short Course runs in conjunctionwith the 2018 Joint Mathematics Meetings.

The emerging field of discrete differential geometry (DDG)studies discrete analogues of smooth geometric objects,providing an essential link between analytical descrip-tions and computation. In recent years it has unearthed arich variety of new perspectives on applied problems incomputational anatomy/biology, computational mechan-ics, industrial design, computational architecture, anddigital geometry processing at large. The basic philoso-phy of discrete differential geometry is that a discreteobject like a polyhedron is not merely an approximationof a smooth one, but rather a differential geometric objectin its own right. In contrast to traditional numerical anal-ysis which focuses on eliminating approximation errorin the limit of refinement (e.g., by taking smaller andsmaller finite differences), DDG places an emphasis onthe so-called “mimetic” viewpoint, where key propertiesof a system are preserved exactly, independent of howlarge or small the elements of a mesh might be. Justas algorithms for simulating mechanical systems mightseek to exactly preserve physical invariants such as totalenergy or momentum, structure-preserving models of

Keenan Crane is assistant professor of computer science atCarnegie Mellon University. His e-mail address is [email protected].

Max Wardetzky is professor of mathematics at University of Göt-tingen. His e-mail address is [email protected] permission to reprint this article, please contact:[email protected]: http://dx.doi.org/10.1090/noti1578

Figure 1. Discrete differential geometry reimaginesclassical ideas from differential geometry withoutreference to differential calculus. For instance,surfaces parameterized by principal curvature linesare replaced by meshes made of circularquadrilaterals (top left), the maximum principleobeyed by harmonic functions is expressed viaconditions on the geometry of a triangulation (topright), and complex-analytic functions can bereplaced by so-called circle packings that preservetangency relationships (bottom). These discretesurrogates provide a bridge between geometry andcomputation, while at the same time preservingimportant structural properties and theorems.

discrete geometry seek to exactly preserve global geo-metric invariants such as total curvature. More broadly,DDG focuses on the discretization of objects that do notnaturally fall under the umbrella of traditional numericalanalysis. This article provides an overview of some of thethemes in DDG.

November 2017 Notices of the AMS 1153

Page 2: AGlimpseintoDiscrete DifferentialGeometry...and combinatorics, since the geometry of a map is determined entirely by incidence relationships.1 On the flip side, this means a different

COMMUNICATION

Figure 2. A given geometric quantity from the smoothsetting, like curvature 𝜅, may have several reasonabledefinitions in the discrete setting. Discretedifferential geometry seeks definitions that exactlyreplicate properties of their smooth counterparts.

The GameOur article is organized around a “game” often played indiscrete differential geometry in order to come up with adiscrete analogue of a given smooth object or theory:(1) Write down several equivalent definitions in the

smooth setting.(2) Apply each smooth definition to an object in the

discrete setting.(3) Analyze trade-offs among the resulting discrete

definitions, which are invariably inequivalent.Most often, none of the resulting discrete objects preserveall the properties of the original smooth one—a so-called no free lunch scenario. Nonetheless, the propertiesthat are preserved often prove invaluable for particularapplications and algorithms. Moreover, this activity yieldssome beautiful and unexpected consequences—such asa connection between conformal geometry and purecombinatorics, or a description of constant-curvaturesurfaces that requires no definition of curvature! Tohighlight some of the challenges and themes commonlyencountered in DDG, we first consider the simple exampleof the curvature of a plane curve.

Discrete Curvature of Planar CurvesHow do you define the curvature of a discrete curve? For asmootharc-lengthparameterized curve𝛾(𝑠) ∶ [0, 𝐿] → ℝ2,curvature 𝜅 is classically expressed in terms of secondderivatives. In particular, if 𝛾 has unit tangent 𝑇 ∶= 𝑑

𝑑𝑠𝛾and unit normal 𝑁 (obtained by rotating 𝑇 a quarter turnin the counter-clockwise direction), then

(1) 𝜅 ∶= ⟨𝑁, 𝑑2

𝑑𝑠2 𝛾⟩ = ⟨𝑁, 𝑑𝑑𝑠𝑇⟩ .

Suppose instead we have a polygonal curve with vertices𝛾1,… ,𝛾𝑛 ∈ ℝ2, as often used for numerical computation(see Figure 2, right). Here we hit upon the most elemen-tary problem of discrete differential geometry: discretegeometric objects are often not sufficiently differentiable(in the classical sense) for standard definitions to apply.For instance, our curvature definition (Equation 1) causestrouble, since at vertices our discrete curve is not twicedifferentiable, nor does it have well defined normals. The

basic approach of DDG is to find alternative characteriza-tions in the smooth setting that can be applied to discretegeometry in a natural way. With curvature, for instance,we can apply the fundamental theorem of calculus toEquation 1 to acquire a different statement: if 𝜑 is theangle from the horizontal line to the unit tangent 𝑇, then

∫𝑏

𝑎𝜅 𝑑𝑠 = 𝜑(𝑏) −𝜑(𝑎) mod 2𝜋.

Putmore simply: curvature is the rate at which the tangentturns. This characterization can be applied naturally toour polygonal curve: along any edge the change in angleis clearly zero. At a vertex it is simply the turning angle𝜃𝑖 ∶= 𝜑𝑖,𝑖+1 −𝜑𝑖−1,𝑖 between the directions 𝜑𝑖−1,𝑖,𝜑𝑖,𝑖+1of the two incident edges, yielding our first notion ofdiscrete curvature:(2) 𝜅𝐴

𝑖 ∶= 𝜃𝑖 ∈ (−𝜋,𝜋).Are there other characterizations that also lead naturallyto a discrete formulation? Yes: for instance we canconsider the motion of 𝛾 that most quickly reduces itslength. In the smooth case it is well known that thechange in length with respect to a smooth variation𝜂(𝑠) ∶ [0, 𝐿] → ℝ2 that vanishes at the endpoints of thecurve is given by integration against curvature:

𝑑𝑑𝜀|𝜀=0

length(𝛾 + 𝜀𝜂) = −∫𝐿

0⟨𝜂(𝑠), 𝜅(𝑠)𝑁(𝑠)⟩ 𝑑𝑠.

Hence, the velocity that most quickly reduces length is𝜅𝑁. For a polygonal curve, we can simply differentiatethe sum of the edge lengths 𝐿 ∶= ∑𝑛−1

𝑖=1 |𝛾𝑖+1 − 𝛾𝑖| withrespect to any vertex position. At a vertex 𝑖 we obtain

(3) 𝜕𝛾𝑖𝐿 = 𝛾𝑖 −𝛾𝑖−1|𝛾𝑖 −𝛾𝑖−1|

− 𝛾𝑖+1 −𝛾𝑖|𝛾𝑖+1 −𝛾𝑖|

=∶ 𝑇𝑖−1,𝑖 −𝑇𝑖,𝑖+1,

i.e., just a difference of unit tangent vectors 𝑇𝑖,𝑖+1 alongconsecutive edges. If 𝑁𝑖 ∈ ℝ2 is the unit angle bisector atvertex 𝑖, this difference can also be expressed as(4) 𝜅𝐵

𝑖 𝑁𝑖 ∶= 2 sin(𝜃𝑖/2)𝑁𝑖,

Figure 3. Different characterizations of curvature inthe smooth setting naturally lead to different notionsof discrete curvature. (Here we abbreviate 𝑇𝑖,𝑖+1 and𝑇𝑖−1,𝑖 by 𝑢 and 𝑣, respectively.)

1154 Notices of the AMS Volume 64, Number 10

Page 3: AGlimpseintoDiscrete DifferentialGeometry...and combinatorics, since the geometry of a map is determined entirely by incidence relationships.1 On the flip side, this means a different

COMMUNICATION

providing a discretization of the curvature normal 𝜅𝑁.A closely-related idea is to consider how the length ofa curve changes if we displace it by a small constantamount in the normal direction. As observed by Steiner,the new length of a smooth curve can be expressed as

(5) length(𝛾 + 𝜀𝑁) = length(𝛾) − 𝜀∫𝐿

0𝜅(𝑠) 𝑑𝑠.

Since this formula holds for any small piece of thecurve, it can be used to obtain a notion of curvatureat each point. How do we define normal offsets in thepolygonal case? At vertices we again encounter the issuethat we have no notion of normals. One idea is to break thecurve into individual edges which can then be translatedby 𝜀 along their respective normal directions. We can thenclose the gaps between edges in a variety of ways: using(A) a circular arc of radius 𝜀, (B) a straight line, or by(C) extending the edges until they intersect (see Figure 3,bottom left). If we then calculate the lengths for thesenew curves, we get

length𝐴 = length(𝛾) − 𝜀∑𝑛−1𝑖=2 𝜃𝑖,

length𝐵 = length(𝛾) − 𝜀∑𝑛−1𝑖=2 2 sin(𝜃𝑖/2),

length𝐶 = length(𝛾) − 𝜀∑𝑛−1𝑖=2 2 tan(𝜃𝑖/2).

Mirroring the observation in the smooth setting, we cannow say that whatever change we observe in the lengthprovides a definition for discrete curvature. The first twoare the same as ones we have seen already: the circular arcyielding the expression from Equation 2, and the straightline corresponding to Equation 4. The third one providesyet another notion of discrete curvature

𝜅𝐶𝑖 ∶= 2 tan(𝜃𝑖/2).

Finally, in the smooth case it is also well known thatcurvature has magnitude equal to the inverse of theradius of the so-called osculating circle, which agrees withthe curve up to second order. A natural way to define anosculating circle for a polygon is to take the circle passingthrough a vertex and its two neighbors. From the formulafor the radius 𝑅𝑖 of a circumcircle in terms of the sidelengths of the corresponding triangle, one easily gets adiscrete curvature that is different from the ones we sawbefore:(6) 𝜅𝐷

𝑖 ∶= 1/𝑅𝑖 = 2 sin(𝜃𝑖)/𝑤𝑖,

The fundamentalbehavior ofgeometry is

neither inherentlysmooth nordiscrete.

where 𝑤𝑖 ∶= |𝛾𝑖+1 −𝛾𝑖−1|. Apart frommerely being differentexpressions,wecanno-tice that 𝜅𝐴, 𝜅𝐵, and 𝜅𝐶

are all invariant undera uniform scaling ofthe curve, whereas 𝜅𝐷

scales like the smoothcurvature 𝜅. This sit-uation demonstratesanother common phe-nomenon in discretedifferential geometry, namely that depending on whichsmooth characterization is used as a starting point, one

may end up with pointwise or integrated quantities in thediscrete case.

As one might imagine, there are many other possi-ble starting points for obtaining a discrete analogue ofcurvature. Eventually, however, all starting points endup leading back to the same definitions, suggesting thatthere may be only so many possibilities. For example, if𝜙 ∶ ℝ2 → ℝ is the signed distance from a smooth closedcurve 𝛾, then applying the Laplacian Δ yields the curva-ture of its level curves; in particular, Δ𝜙|𝜙=0 yields thecurvature of 𝛾. Likewise, if we apply the Laplacian to thesigned distance function for a discrete curve, we recover𝜅𝐴 on one side and 𝜅𝐵 on the other. Yet another approachis the theory of normal cycles (as discussed by Morvan),related to the Steiner formula from Equation 5. Here,rather than settle on a single normal 𝑁𝑖 at each vertexwe consider all unit vectors between the unit normalsof the two incident edges, ultimately leading back to thefirst discrete curvature 𝜅𝐴. The theory of normal cyclesapplies equally well to both smooth and polygonal curves,again reinforcing the perspective that the fundamentalbehavior of geometry is neither inherently smooth nordiscrete, but can be well captured in both settings bypicking the appropriate ansatz. More broadly, the factthat equivalent characterizations in the smooth settinglead to different inequivalent definitions in the discretesetting is not special to the case of curves, but is one ofthe central themes in discrete differential geometry.

From here, a natural question arises: which discretecurvature is “best”? A traditional criterion for discrimi-nating among different discrete versions is the questionof convergence: if we consider finer and finer approxi-mating polygons, will our discrete curvatures converge tothe classical smooth one? However, convergence does notalways single out a best version: treated appropriately, allfour of our discrete curvatures will converge. We must

Figure 4. Typically, not all properties of a smoothobject can be preserved exactly at the discrete level.For curve-shortening flow, for example, 𝜅𝐴 exactlypreserves the total curvature, 𝜅𝐵 exactly preservesthe center of mass, and with 𝜅𝐷 the flow remainsstationary (up to rescaling) for any circular solution.However, no local definition of discrete curvature canprovide all three properties simultaneously.

November 2017 Notices of the AMS 1155

Page 4: AGlimpseintoDiscrete DifferentialGeometry...and combinatorics, since the geometry of a map is determined entirely by incidence relationships.1 On the flip side, this means a different

COMMUNICATION

therefore look beyond convergence, toward exact preser-vation of properties and relationships from the smoothsetting. Which properties should we try to preserve? Theanswer of course depends on what we aim to use thesecurvatures for.

As a toy example, consider the curve-shortening flow(depicted in Figure 4, top left), where a curve evolvesaccording to the velocity that most quickly reducesits length. As discussed above, this velocity is equalto the curvature normal 𝜅𝑁. A smooth, simple curveevolving under this flow exhibits several basic properties:it has at all times total curvature 2𝜋, its center ofmass remains fixed, it tends toward a circle of vanishingradius, and remains embedded for all time, i.e., no self-crossings arise (Gage-Grayson-Hamilton). Do our discretecurvatures furnish these same properties? A numericalexperiment is shown in Figure 4. Here we evolve ourpolygon by a simple time-discrete flow 𝛾𝑖 ← 𝛾𝑖 + 𝜏𝜅𝑖𝑁𝑖with a fixed time step 𝜏 > 0. For 𝜅𝐷, 𝑁𝑖 is the unit vectoralong the circumradius; otherwise it is the unit anglebisector. Not surprisingly, 𝜅𝐴 preserves total curvature(due to the fundamental theorem of calculus); 𝜅𝐵 doesnot drift (consider summing Equation 3 over all vertices);and 𝜅𝐷 has circular polygons as limit points (sinceall velocities point toward the center of a commoncircle). However, no discrete curvature satisfies all threeproperties simultaneously. Moreover, for a constant timestep 𝜏 no such flow can guarantee that new crossingsdo not occur. This situation illustrates the no free lunchidea: no matter how hard we try, we cannot find a singlediscrete object that preserves all the properties of itssmooth counterpart. Instead, we have to pick and choosethe properties best suited to the task at hand.

Suppose that instead of curvature flow, we considertwo other beautiful topics in the geometry of planecurves: the Whitney–Graustein theorem, which classifiesregular homotopy classes of curves by their total curva-ture, and Kirchhoff’s famous analogy between motionsof a spherical pendulum and elastic curves, i.e., curvesthat extremize the bending energy ∫𝐿0 𝜅2 𝑑𝑠 subject toboundary conditions. Among the curvatures discussedabove, only 𝜅𝐴 provides a discrete version of Whitney–Graustein, but does not provide an exact discrete analogueof Kirchhoff. Likewise, 𝜅𝐶 preserves the structure of theKirchhoff analogy, but not Whitney–Graustein. This kindof no free lunch situation is a characteristic feature ofDDG. A similar obstacle is encountered in the theory ofordinary differential equations, where it is known thatthere are no numerical integrators for Hamiltonian sys-tems that simultaneously conserve energy, momentum,and the symplectic form. From a computational point ofview, making judicious choices about which quantitiesto preserve for which applications goes hand-in-handwith providing formal guarantees on the reliability androbustness of algorithms.

We now give a few glimpses into recent topics andtrends in DDG.

Figure 5. What is the simplicial analogue of aconformal map? Requiring all angles to be preservedis too rigid, forcing a global similarity (left). Askingonly for preservation of so-called length cross ratiosprovides just the right amount of flexibility,maintaining much of the structure found in thesmooth setting such as invariance under Möbiustransformations (right).

Discrete Conformal GeometryA conformal map is, roughly speaking, a map that pre-serves angles (see Figure 1, bottom left). A good exampleis Mercator’s projection of the globe: even though areagets stretched out badly—making Greenland look muchbigger than Australia!—the directions “north” and “east”remain at right angles, which is very helpful if you aretrying to navigate the sea. A beautiful fact about confor-mal maps is that any surface can be conformally mappedto a space of constant curvature (“uniformization”), pro-viding it with a canonical geometry. This fact, plus thefact that conformal maps can be efficiently computed(e.g., by solving sparse linear systems), have led in recentyears to widespread development of conformal mappingalgorithms as a basic building block for digital geometryprocessing algorithms. In applications, discrete confor-mal maps are used for everything from sensor networklayout to comparative analysis of medical or anatomicaldata. Of course, to process real data one must be able tocompute conformal maps on discrete geometry.

What does it mean for a discrete map to be conformal?As with curvature, one can play the game of enumeratingseveral equivalent characterizations in the smooth setting.Consider for instance a map 𝑓 ∶ 𝑀 → 𝐷2 ⊂ ℂ from a disk-like surface 𝑀 with Riemannian metric 𝑔 to the unit disk𝐷2 in the complex plane. This map is conformal if it

too few degrees offreedom relative to

the number ofconstraints

preserves angles, ifit preserves infinitesi-mal circles, if it canbe expressed as a pairof real conjugate har-monic functions 𝑓 =𝑎+ 𝑏𝑖, if it is a criticalpoint of the Dirichletenergy ∫𝑀 |𝑑𝑓|2 𝑑𝐴, or

if it induces a new metric ̃𝑔 ∶= 𝑑𝑓 ⊗ 𝑑𝑓 that at eachpoint is a positive rescaling of the original one: ̃𝑔 = 𝑒2𝑢𝑔.Each starting point leads down a path toward different

1156 Notices of the AMS Volume 64, Number 10

Page 5: AGlimpseintoDiscrete DifferentialGeometry...and combinatorics, since the geometry of a map is determined entirely by incidence relationships.1 On the flip side, this means a different

COMMUNICATION

consequences in the discrete setting, and to algorithmswith different computational tradeoffs.

Oddly enough, the most elementary characterizationof conformal maps, angle preservation, does not translatevery well to the discrete setting (see Figure 5). Considerfor instance a simplicial map that takes a triangulateddisk 𝐾 = (𝑉,𝐸,𝐹) to a triangulation in the plane. Anymap that preserves interior angles will be a similarity oneach triangle, i.e., it can only rigidly rotate and scale. Butsince adjacent triangles share edges, the scale factor forall triangles must be identical. Hence, the only discretesurfaces that can be conformally flattened in this senseare those that are (up to global scale) developable, i.e.,that can be rigidly unfolded into the plane. This outcomeis in stark contrast to the smooth setting, where anydisk can be conformally flattened. This situation reflectsa common scenario in DDG: rigidity, or what in finiteelement analysis is sometimes called locking. There aresimply too few degrees of freedom relative to the numberof constraints: we want to match angles at all 3𝐹 corners,but have only 2𝑉 < 3𝐹 degrees of freedom. Hence, ifwe insist on angle preservation we have no chance ofcapturing the flexibility of smooth conformal maps.

Other characterizations provide greater flexibility. Oneidea is to associate each vertex of our discrete disk𝐾witha circle in the plane. A theorem of Koebe implies that onecan always arrange these circles such that two circles aretangent if they belong to a shared edge and all boundarycircles are tangent to a common circle bounding the rest.For a regular triangular lattice approximating a region𝑈 ⊂ ℂ, Thurston noticed that this map approximates asmooth conformalmap 𝑓 ∶ 𝑈 → 𝐷2 as the region isfilledbysmaller and smaller circles (see Figure 1, bottom), as laterproved by Rodin and Sullivan. Unlike a traditional finiteelement discretization, these so-called circle packingsalso preserve many of the basic structural properties ofconformal maps. For instance, composition with a Möbiustransformation of the disk yields another uniformizationmap, as in the smooth setting. More broadly, circlepackings provide an unexpected bridge between geometryand combinatorics, since the geometry of a map isdetermined entirely by incidence relationships.1 On theflip side, this means a different theory is needed toaccount for the geometry of irregular triangulations, asmore commonly used in applications.

An alternative theory starts from the idea that undera conformal map the Riemannian metric 𝑔 experiences auniform scaling at each point: ̃𝑔 = 𝑒2𝑢𝑔. In other words,vectors tangent to a given point 𝑝 ∈ 𝑀 shrink or growby a positive factor 𝑒𝑢. In the simplicial setting 𝑔 isreplaced by a piecewise Euclidean metric, i.e., a collectionof positive edge lengths ℓ ∶ 𝐸 → ℝ>0 that satisfy thetriangle inequality in each face. Two such metrics ℓ, ̃ℓare then said to be discretely conformally equivalent ifthey are related by ̃ℓ𝑖𝑗 = 𝑒(𝑢𝑖+𝑢𝑗)/2ℓ𝑖𝑗 for any collection ofdiscrete scale factors𝑢 ∶ 𝑉 → ℝ. Thoughatfirst glance this

1See “Circle Packing” in the December 2003 Notices www.ams.org/notices/200311/fea-stephenson.pdf.

relationship looks like a simple numerical approximation,it turns out to provide a complete discrete theory thatpreserves much of the structure found in the smoothsetting, with close ties to theories based on circles. Anequivalent characterization is the preservation of lengthcross ratios 𝔠𝑖𝑗𝑘𝑙 ∶= ℓ𝑖𝑗ℓ𝑘𝑙/ℓ𝑗𝑘ℓ𝑙𝑖 associated with eachedge ∈ 𝐸; for a mesh embedded in ℝ𝑛 these ratios areinvariant underMöbius transformations, againmimickingthe smooth theory. This theory also leads to efficient,convex algorithms for discrete Ricci flow, which is astarting point for many applications in digital geometryprocessing.

More broadly, discrete conformal geometry and dis-crete complex analysis is an active area of research, withelegant theories not only for triangulations but also forlattice-based discretizations, whichmake contact with thetopic of (discrete) integrable systems, discussed below.Yet basic questions about properties like convergence, ordescriptions that are compatible with extrinsic geometry,are still only starting to be understood.

Discrete Differential OperatorsDifferential geometry and in particular Riemannian man-ifolds can be studied from many different perspectives.In contrast to the purely geometric perspective (basedon, say, notions of distance or curvature), differentialoperators provide a very different point of view. Oneof the most fundamental operators in both physics andgeometry is the Laplace–Beltrami operatorΔ (or Laplacianfor short) acting on differential 𝑘-forms. It describes, forexample, heat diffusion, wave propagation, and steadystate fluid flow, and is key to the Schrödinger equationin quantum mechanics. It also provides a link betweenanalytical and topological information: for instance, onclosed Riemannian manifolds the dimension of harmonic𝑘-forms (i.e., those in the kernel of Δ) equals the di-mension of the 𝑘th cohomology—a purely topologicalquantity. The spectrum of the Laplacian (i.e., the listof eigenvalues) likewise reveals a great deal about thegeometry of the manifold. For example, the first nonzeroeigenvalue of the 0-form Laplacian provides an upper anda lower bound on optimally cutting a compact Riemannianmanifold 𝑀 into two disjoint pieces of, loosely speaking,maximal volume and minimal perimeter (Cheeger-Buser).These so-called Cheeger cuts have a wide range of ap-plications across machine learning and computer vision;more broadly, eigenvalues and eigenfunctions of Δ helpto generalize traditional Fourier-based simulation andsignal processing to more general manifolds.

These observations motivate the study of discreteLaplacians, which can be defined even in the purely com-binatorial setting of graphs. Here we briefly outline theirdefinition for orientable finite simplicial 𝑛-manifolds,such as polyhedral surfaces, without boundary. Our ex-position is similar to what has become known as discreteexterior calculus. To this end, consider the simplicialboundary operators 𝜕𝑘 ∶ 𝐶𝑘 → 𝐶𝑘−1 acting on 𝑘-chains(i.e., formal linear combinations of 𝑘-simplices). The cor-responding dual spaces (cochains) 𝐶𝑘 ∶= Hom(𝐶𝑘, ℝ) and

November 2017 Notices of the AMS 1157

Page 6: AGlimpseintoDiscrete DifferentialGeometry...and combinatorics, since the geometry of a map is determined entirely by incidence relationships.1 On the flip side, this means a different

COMMUNICATION

respective dual operators 𝛿𝑘 ∶ 𝐶𝑘 → 𝐶𝑘+1 give rise to thechain complex

{0} → 𝐶0 → 𝐶1 → … → 𝐶𝑛 → {0}.The chain property says that 𝛿𝑘∘𝛿𝑘−1 = 0, and one henceobtains simplicial cohomology 𝐻𝑘 ∶= ker(𝛿𝑘)/im(𝛿𝑘−1).To define a Laplacian in this setting, we equip each 𝐶𝑘

with a positive definite inner product (⋅, ⋅)𝑘, and let 𝛿∗𝑘 be

the adjoint operator with respect to these inner products,i.e., (𝛿𝑘𝛼,𝛽)𝑘+1 = (𝛼,𝛿∗

𝑘 𝛽)𝑘 for all 𝛼,𝛽. The Laplacian on𝑘-cochains is then defined as

Δ𝑘 ∶= 𝛿∗𝑘 𝛿𝑘 +𝛿𝑘−1𝛿∗

𝑘−1.The resulting space of harmonic 𝑘-cochains, {𝛼 ∈𝐶𝑘|Δ𝑘𝛼 = 0} is then isomorphic to 𝐻𝑘—just as inthe smooth setting. This fact is independent of thechoice of inner product, mirroring the fact that coho-mology depends only on topological structure. Likewise,for any inner product one obtains a discrete Hodgedecomposition

𝐶𝑘 = ker(Δ𝑘) ⊕ im(𝛿𝑘−1) ⊕ im(𝛿∗𝑘 ),

where here the subspaces do depend on the choice ofinner product.

At this point we return again to the game of DDG:which choice of inner product is best? A trivial innerproduct leads to purely combinatorial graph Laplacians,which do not (in general) converge to their smooth coun-terparts (e.g., when approximating a smooth manifold bya polyhedral one). Another choice is to consider linearinterpolation of 𝑘-cochains over 𝑛-dimensional simplices,resulting in what is known asWhitney elements. For 𝑛 = 2,we get the so-called cotan Laplacian (Pinkall and Polth-ier), which is widely used in digital geometry processing.Though other choices are possible, we again encountera no free lunch situation: no choice of inner productcan preserve all the properties of the smooth Laplacian.Which properties do we care about? Beyond convergence,perhaps the most desirable properties are the maximumprinciple (which ensures, for instance, proper behaviorfor heat flow), and the property that, for flat domains,linear functions are in the kernel (leading to a properdefinition of barycentric coordinates). For general un-structured meshes there are no discrete Laplacians withall of these properties. However, certain types of meshes(such asweighedDelaunay triangulations)do indeed allowfor “perfect” discrete Laplacians, offering a connectionbetween geometry and (discrete) differential operators.

Discrete Integrable SystemsAnother topic that has provided inspiration for manyideas in DDG is parameterized surface theory. Considerfor instance the problem of dressing a given surfaceby a fishnet stocking, i.e., a woven material composed ofinextensible yarns following transversal “warp” and “weft”directions (see Figure 6, right). This task corresponds todecorating a surface with a tiling where each vertexis incident to four parallelograms. Infinitesimally, such atiling is known as aweakChebyshev net (Chebyshev 1878),and locally corresponds to a regularly parameterized

surface 𝑓 ∶ 𝑈 ⊂ ℝ2 → ℝ3 where the directional derivatives𝑓𝑢 and 𝑓𝑣 along the coordinate directions satisfy |𝑓𝑢|𝑣 =|𝑓𝑣|𝑢 = 0, i.e., partial derivatives with respect to oneparameter have constant length along the parameterlines of the other parameter. The special case of rhombictilings (|𝑓𝑢| = |𝑓𝑣| = 1) are known as (strong) Chebyshevnets. Can every smooth surface be wrapped in a stocking?Locally (i.e., in a small patch around any given point)the answer is “yes.” Globally, however, there are severeobstructions to doing so, which provide some fascinatingconnections to physics.

Consider for instance the special case of so-calledK-surfaces, characterized by constant Gauß curvature𝐾 = −1. Every K-surface admits a parameterization 𝑓 ∶𝑈 ⊂ ℝ2 → ℝ3 aligned with the two transversal asymptoticdirections along which normal curvature vanishes. Hence,if 𝑁 is the unit surface normal then(7) ⟨𝑓𝑢𝑢,𝑁⟩ = ⟨𝑓𝑣𝑣,𝑁⟩ = 0.Asymptotic parameterizations are weak Chebyshev netssince(8) 𝑎 ∶= |𝑓𝑢|, 𝑏 ∶= |𝑓𝑣| satisfy 𝑎𝑣 = 𝑏𝑢 = 0.Moreover, one can show that the angle 𝜙 betweenasymptotic lines satisfies the sine-Gordon equation

(9) 𝜙𝑢𝑣 −𝑎𝑏 sin𝜙 = 0,and conversely, every solution to the sine-Gordon equa-tion describes a parameterized K-surface. Hilbert usedthis equation (and Chebyshev nets) to prove that the com-plete hyperbolic plane cannot be embedded isometricallyinto ℝ3. More generally, the sine-Gordon equation hasattracted much interest both in mathematics as an exam-ple of an infinite-dimensional integrable system, and inphysics as an example of a system that admits remarkablystable soliton solutions, akin to waves that travel uninter-rupted all the way across the ocean. Another key propertyof the sine-Gordon equation is the existence of a so-calledspectral parameter 𝜆 > 0: Equation 9 is invariant undera rescaling 𝑎 → 𝜆𝑎 and 𝑏 → 𝜆−1𝑏, giving rise to a one-parameter associated family of K-surfaces. Geometrically,the parameter 𝜆 rescales the edges of parallelogramswhile preserving the angle between asymptotic lines.

Do these properties depend critically on the smoothnature of the solutions, or can they also be faithfullycaptured in the discrete setting? Hirota derived such adiscrete version without any reference to geometry. LaterBobenko and Pinkall suggested a geometric definition ofdiscrete K-surfaces that recovers Hirota’s equation. Intheir setting, discrete K-surfaces are defined as discrete(weak) Chebyshev nets with the additional property thatall four edges incident to any vertex lie in a commonplane. The last requirement is a natural discrete analogueof Equation 7. This definition of discrete K-surfaces alsocomes with a spectral parameter 𝜆 and results in Hirota’sdiscrete sine-Gordon equation—without requiring anynotion of discrete Gauß curvature. Only recently has adiscrete version of Gauß curvature been suggested thatresults in discrete K-surfaces indeed having constantnegative Gauß curvature.

1158 Notices of the AMS Volume 64, Number 10

Page 7: AGlimpseintoDiscrete DifferentialGeometry...and combinatorics, since the geometry of a map is determined entirely by incidence relationships.1 On the flip side, this means a different

COMMUNICATION

Figure 6. Left: two discrete parameterizations of apseudosphere (constant Gauß curvature 𝐾 = −1), onewith a Chebychev net along asymptotic directions(left) and another along principal curvature lines(right). Right: a discrete Chebyshev net on a surfaceof varying curvature, resembling the weft and warpdirections of a woven material.

Figure 7. Discrete parameterized surfaces play a rolein architectural geometry, where special incidencerelationships on quadrilaterals translate tomanufacturing constraints like zero nodal torsion, oroffset surfaces of constant thickness. Here acurvature line parameterized surface discretized by aconical net is used in the design of a railway station.

For discrete K-surfaces with all equal edge lengths(i.e., the rhombic case) the four neighboring verticesof a given vertex must lie on a common circle. Byconsidering a subset of the diagonals of the quadrilaterals,one obtains another quad mesh with the property thatall quads have a circumscribed circle, resulting in so-called cK-nets (see Figure 6, left). In the discrete setting,regular networks of circular quadrilaterals play the roleof curvature line parameterized surfaces (as in Figure1, top left). This transformation therefore mimics thesmooth setting, where the angle bisectors of asymptoticlines are lines of principal curvature. More broadly, thetheory of quad nets with special incidence relationships isclosely linked to physical manufacturing considerationsin the field of architectural geometry. For example, a quadnet is conical if the four quads around each vertex aretangent to a common cone—such surfaces admit faceoffsets of constant width, making them attractive for theconstruction of (for instance) glass-paneled structures, asin Figure 7.

For further reading, see Discrete Differential Geometry(2008, Alexander Bobenko ed.).

Image CreditsFigures 1–6 courtesy of Keenan Crane and Max Wardetzky.Figure 7 courtesy of B. Schneider.Author photo of Keenan Crane courtesy of Keenan Crane.Author photo of Max Wardetzky courtesy of Max Wardetzky.

ABOUT THE AUTHORS

Keenan Crane works at the inter-face between differential geometryand geometric algorithms. His sci-entific career started out studyingPluto—back when it was still aplanet!

Keenan Crane

Max Wardetzky’s interest in dis-crete differential geometry comesfrom his outstanding and trulyinspiring teachers in differentialgeometry and his exposure tomarvelous people in computergraphics.

Max Wardetzky

November 2017 Notices of the AMS 1159