9
Deep Functional Maps: Structured Prediction for Dense Shape Correspondence Or Litany 1,2 Tal Remez 1 Emanuele Rodol` a 3,4 Alex Bronstein 2,5 Michael Bronstein 2,3 1 Tel Aviv University 2 Intel 3 USI Lugano 4 Sapienza University of Rome 5 Technion Abstract We introduce a new framework for learning dense corre- spondence between deformable 3D shapes. Existing learn- ing based approaches model shape correspondence as a la- belling problem, where each point of a query shape receives a label identifying a point on some reference domain; the correspondence is then constructed a posteriori by com- posing the label predictions of two input shapes. We pro- pose a paradigm shift and design a structured prediction model in the space of functional maps, linear operators that provide a compact representation of the correspondence. We model the learning process via a deep residual network which takes dense descriptor fields defined on two shapes as input, and outputs a soft map between the two given ob- jects. The resulting correspondence is shown to be accu- rate on several challenging benchmarks comprising multi- ple categories, synthetic models, real scans with acquisition artifacts, topological noise, and partiality. 1. Introduction 3D acquisition technology has made great progress in the last decade, and is being rapidly incorporated into com- mercial products ranging from Microsoft Kinect [43] for gaming, to LIDARs used in autonomous cars. An essen- tial building block for application design in many of these domains is to recover 3D shape correspondences in a fast and reliable way. While handling real-world scanning ar- tifacts is a challenge by itself, additional complications arise from non-rigid motions of the objects of interest (typi- cally humans or animals). Most non-rigid shape correspon- dence methods employ local descriptors that are designed to achieve robustness to noise and deformations; however, relying on such “handcrafted” descriptors can often lead to inaccurate solutions in practical settings. Partial remedy to this was brought by the recent line of works on learning shape correspondence [28, 36, 29, 8, 10, 9, 31]. A key draw- back of these methods lies in their emphasis on learning Figure 1. Correspondence results obtained by our network model on two pairs from the FAUST real scans challenge. Corresponding points are assigned the same color. The average error for the left and right pairs is 5.21cm and 2.34cm respectively. Accurate cor- respondence is obtained despite mesh “gluing” in areas of contact. a descriptor that would help in identifying corresponding points, or on learning a labeling with respect to some refer- ence domain. On the one hand, by focusing on the descrip- tor, the learning process remains agnostic to the way the fi- nal correspondence is computed, and costly post-processing steps are often necessary in order to obtain accurate solu- tions from the learned descriptors. On the other hand, meth- ods based on a label space are restricted to a fixed number of points and rely on the adoption of an intermediate reference model. Contribution. In this work we propose a task-driven ap- proach for descriptor learning, by including the computa- tion of the correspondence directly as part of the learning procedure. Perfect candidates for this task are neural net- works, due to their inherent flexibility to the addition of computational blocks. Our main contributions can be sum- marized as follows: We introduce a new structured prediction model for shape correspondence. Our framework allows end-to- 1

Deep Functional Maps: Structured Prediction for Dense ... · Deep Functional Maps: Structured Prediction for Dense Shape Correspondence Or Litany 1;2 Tal Remez Emanuele Rodola`3;4

Embed Size (px)

Citation preview

Page 1: Deep Functional Maps: Structured Prediction for Dense ... · Deep Functional Maps: Structured Prediction for Dense Shape Correspondence Or Litany 1;2 Tal Remez Emanuele Rodola`3;4

Deep Functional Maps:Structured Prediction for Dense Shape Correspondence

Or Litany1,2 Tal Remez1

Emanuele Rodola3,4 Alex Bronstein2,5 Michael Bronstein2,3

1Tel Aviv University 2Intel 3USI Lugano 4Sapienza University of Rome 5Technion

Abstract

We introduce a new framework for learning dense corre-spondence between deformable 3D shapes. Existing learn-ing based approaches model shape correspondence as a la-belling problem, where each point of a query shape receivesa label identifying a point on some reference domain; thecorrespondence is then constructed a posteriori by com-posing the label predictions of two input shapes. We pro-pose a paradigm shift and design a structured predictionmodel in the space of functional maps, linear operators thatprovide a compact representation of the correspondence.We model the learning process via a deep residual networkwhich takes dense descriptor fields defined on two shapesas input, and outputs a soft map between the two given ob-jects. The resulting correspondence is shown to be accu-rate on several challenging benchmarks comprising multi-ple categories, synthetic models, real scans with acquisitionartifacts, topological noise, and partiality.

1. Introduction3D acquisition technology has made great progress in

the last decade, and is being rapidly incorporated into com-mercial products ranging from Microsoft Kinect [43] forgaming, to LIDARs used in autonomous cars. An essen-tial building block for application design in many of thesedomains is to recover 3D shape correspondences in a fastand reliable way. While handling real-world scanning ar-tifacts is a challenge by itself, additional complicationsarise from non-rigid motions of the objects of interest (typi-cally humans or animals). Most non-rigid shape correspon-dence methods employ local descriptors that are designedto achieve robustness to noise and deformations; however,relying on such “handcrafted” descriptors can often lead toinaccurate solutions in practical settings. Partial remedy tothis was brought by the recent line of works on learningshape correspondence [28, 36, 29, 8, 10, 9, 31]. A key draw-back of these methods lies in their emphasis on learning

Figure 1. Correspondence results obtained by our network modelon two pairs from the FAUST real scans challenge. Correspondingpoints are assigned the same color. The average error for the leftand right pairs is 5.21cm and 2.34cm respectively. Accurate cor-respondence is obtained despite mesh “gluing” in areas of contact.

a descriptor that would help in identifying correspondingpoints, or on learning a labeling with respect to some refer-ence domain. On the one hand, by focusing on the descrip-tor, the learning process remains agnostic to the way the fi-nal correspondence is computed, and costly post-processingsteps are often necessary in order to obtain accurate solu-tions from the learned descriptors. On the other hand, meth-ods based on a label space are restricted to a fixed number ofpoints and rely on the adoption of an intermediate referencemodel.

Contribution. In this work we propose a task-driven ap-proach for descriptor learning, by including the computa-tion of the correspondence directly as part of the learningprocedure. Perfect candidates for this task are neural net-works, due to their inherent flexibility to the addition ofcomputational blocks. Our main contributions can be sum-marized as follows:

• We introduce a new structured prediction model forshape correspondence. Our framework allows end-to-

1

Page 2: Deep Functional Maps: Structured Prediction for Dense ... · Deep Functional Maps: Structured Prediction for Dense Shape Correspondence Or Litany 1;2 Tal Remez Emanuele Rodola`3;4

end training: it takes base descriptors as input, and re-turns matches.

• We show that our approach consistently outperformsexisting descriptor and correspondence learning meth-ods on several recent benchmarks.

2. Related work

Shape correspondence is an active area of research incomputer vision, graphics, and pattern recognition, with avariety of both classical and recent methods [11, 23, 13].Since a detailed review of the literature would be out ofscope for this paper, we refer the interested reader to therecent surveys on shape correspondence [39, 6]. Moreclosely related to our approach is the family of meth-ods based on the notion of functional maps [33], mod-eling correspondences as linear operators between spacesof functions on manifolds. In the truncated Laplacianeigenbases, such operators can be compactly encoded assmall matrices – drastically reducing the amount of vari-ables to optimize for, and leading to an increased flexibil-ity in manipulating correspondences. This representationhas been adopted and extended in several follow-up works[34, 22, 25, 3, 18, 35, 27, 26, 32], demonstrating state-of-the-art performance in multiple settings.

While being often adopted as a useful tool for the post-processing of some initial correspondence, functional mapshave rarely been employed as a building block in correspon-dence learning pipelines.

Descriptor learning. Descriptor and feature learning arekey topics in computer vision, and in recent years they havebeen actively investigated by the shape analysis community.Litman et al. [28] introduced optimal spectral descriptors, adata-driven parametrization of the spectral descriptor model(i.e., based on the eigen-decomposition of the Laplacian),demonstrating noticeable improvement upon the classicalaxiomatic constructions [37, 5]. A similar approach wassubsequently taken in [42] and more recently in [8, 10, 17].Perhaps most closely related to ours is the approach of Cor-man et al. [15], where combination weights for an input setof descriptors are learned in a supervised manner. Similarlyto our approach, they base their construction upon the func-tional map framework [33]. While their optimality criterionis defined in terms of deviation from a ground-truth func-tional map in the spectral domain, we aim at recovering anoptimal map in the spatial domain. Our structured predic-tion model will be designed with this requirement in mind.

Correspondence learning. Probably the first example oflearning correspondence for deformable 3D shapes is the“shallow” random forest approach of Rodola et al. [36].More recently, Wei et al. [41] employed a classical (ex-trinsic) CNN architecture trained on huge training sets for

learning invariance to pose changes and clothing. Convo-lutional neural networks on non-Euclidean domains (sur-faces) were first considered by Masci et al. [29] with theintroduction of the geodesic CNN model, a deep learn-ing architecture where the classical convolution operationis replaced by an intrinsic (albeit, non-shift invariant) coun-terpart. The framework was shown to produce promisingresults in descriptor learning and shape matching applica-tions, and was recently improved by Boscaini et al. [9] andgeneralized further by Monti et al. [31]. These methodsare instances of a broader recent trend of geometric deeplearning attempting to generalize successful deep learningparadigms to data with non-Euclidean underlying structuresuch as manifolds or graphs [12].

3. Background

Manifolds. We model shapes as two-dimensional Rieman-nian manifolds X (possibly with boundary ∂X ) equippedwith the standard measure dµ induced by the volume form.Throughout the paper we will consider the space of func-tions L2(X ) = {f : X → R | 〈f, f〉X < ∞}, with thestandard manifold inner product 〈f, g〉X =

∫X f · g dµ.

The positive semi-definite Laplace-Beltrami opera-tor ∆X generalizes the notion of Laplacian from Eu-clidean spaces to surfaces. It admits an eigen-decomposition ∆Xφi = λiφi (with proper boundary con-ditions if ∂X 6= ∅), where the eigenvalues form a discretespectrum 0 = λ1 ≤ λ2 ≤ . . . and the eigenfunctionsφ1, φ2, . . . form an orthonormal basis for L2(X ), allowingus to expand any function f ∈ L2(X ) as a Fourier series

f(x) =∑i≥1

〈φi, f〉Xφi(x) . (1)

Aflalo et al. [2] have recently shown that Laplacian eigen-bases are optimal for representing smooth functions onmanifolds.

Functional correspondence. In order to compactly en-code correspondences between shapes, we make use of thefunctional map representation introduced by Ovsjanikov etal. [33]. The key idea is to identify correspondences by alinear operator T : L2(X )→ L2(Y), mapping functions onX to functions on Y . This can be seen as a generalizationof classical point-to-point matching, which is a special casewhere delta functions are mapped to delta functions.

The linear operator T admits a matrix representationC = (cij) with coefficients cji = 〈ψj , Tφi〉Y , where{φi}i≥1 and {ψj}j≥1 are orthogonal bases on L2(X ) andL2(Y) respectively, leading to the expansion:

Tf =∑ij≥1

〈φi, f〉X cjiψj . (2)

Page 3: Deep Functional Maps: Structured Prediction for Dense ... · Deep Functional Maps: Structured Prediction for Dense Shape Correspondence Or Litany 1;2 Tal Remez Emanuele Rodola`3;4

A good choice for the bases {φi}, {ψj} is given by theLaplacian eigenfunctions on the two shapes [33, 2]. Thischoice is particularly convenient, since (by analogy withFourier analysis) it allows to truncate the series (2) after thefirst k coefficients – yielding a band-limited approximationof the original map. The resulting matrix C is a k× k com-pact representation of a correspondence between the twoshapes, where typically k � n (here n is the number ofpoints on each shape).

Functional correspondence problems seek a solution forthe matrix C, given a set of corresponding functions fi ∈L2(X ) and gi ∈ L2(Y), i = 1, . . . , q, on the two shapes. Inthe Fourier basis, these functions are encoded into matricesF = (〈φi, fj〉X ) and G = (〈ψi, gj〉Y), leading to the least-squares problem:

minC‖CF− G‖2F . (3)

In practice, dense q-dimensional descriptor fields (e.g.,HKS [37] or SHOT [38]) on X and Y are used as the corre-sponding functions.

Label space. Previous approaches at learning shape cor-respondence phrased the matching problem as a labellingproblem [36, 29, 10, 9, 31]. These approaches attempt tolabel each vertex of a given query shapeX with the index ofa corresponding point on some reference shape Z (usuallytaken from the training set), giving rise to a dense point-wise map TX : X → Z . The correspondence between twoqueries X and Y can then be obtained via the compositionT−1Y ◦ TX [36].

Given a training set S = {(x, π∗(x))} ⊂ X × Y ofmatches under the ground-truth map π∗ : X → Y , label-based approaches compute a descriptor FΘ(x) whose op-timal parameters are found by minimizing the multinomialregression loss:

`mr(Θ) = −∑

(x,π∗(x))∈S

〈δπ∗(x), logFΘ(x)〉Y , (4)

where δπ∗(x) is a delta function on Y at point π∗(x).Such an approach essentially treats the correspondence

problem as one of classification, where the aim is to ap-proximate as closely as possible (in a statistical sense) thecorrect label for each point. The actual construction of thefull correspondence is done a posteriori by a compositionstep with an intermediate reference domain, or by solvingthe least-squares problem (3) with the learned descriptorsas data.

Discretization. In the discrete setting, shapes are repre-sented as manifold triangular meshes with n vertices (ingeneral, different for each shape). The Laplace-Beltramioperator ∆ is discretized as a symmetric n× n matrix L =A−1W using a classical linear FEM scheme [30], where

Figure 2. Given a source and a target shape as input, our networkoutputs a soft correspondence matrix whose columns can be inter-preted as probability distributions over the target shape.

the stiffness matrix W contains the cotangent weights, andthe mass matrix A is a diagonal matrix of vertex area el-ements. The manifold inner product 〈f, g〉 is discretizedas the area-weighted dot product f>Ag, where the vec-tors f ,g ∈ Rn contain the function values of f and g ateach vertex. Note that under such discretization we haveΦ>AΦ = I, where Φ contains the Laplacian eigenfunc-tions as its columns.

4. Deep Functional MapsIn this paper we propose an alternative model to the

labelling approach described above. We aim at learningpoint-wise descriptors which, when used in a functionalmap pipeline such as (3), will induce an accurate correspon-dence. To this end, we construct a neural network whichtakes as input existing, manually designed descriptors andimproves upon those while satisfying a geometrically mean-ingful criterion. Specifically, we consider the soft error loss

`F =∑

(x,y)∈(X ,Y)

P (x, y)dY(y, π∗(x)) = ‖P ◦DY‖F , (5)

where DY is the n × n matrix of geodesic distances on Y ,◦ is the element-wise product, and

P = |ΨCΦ>A|∧ (6)

is a soft correspondence matrix, which can be interpretedas the probability of point x ∈ X mapping to point y ∈Y (see Figure 2); here, Φ,Ψ are matrices containing thefirst k eigenfunctions {φi}, {ψj} as their columns, | · | actselement-wise, and X∧ is a column-wise normalization ofX. In the formula above, the k × k matrix C represents afunctional map obtained as the least-squares solution to (3)under learned descriptors F,G.

Matrix P represents a rank-k approximation of the spa-tial correspondence between the two shapes, thus allowing

Page 4: Deep Functional Maps: Structured Prediction for Dense ... · Deep Functional Maps: Structured Prediction for Dense Shape Correspondence Or Litany 1;2 Tal Remez Emanuele Rodola`3;4

X

Φ

SHOT

Res

1

Res

2

Res

K

· · · 〈·, ·〉XFF

Θ FMC P

Softcor `F

Ψ

Y SHOT

Res

1

Res

2

Res

K

· · · 〈·, ·〉YGG

Figure 3. FMNet architecture. Input point-wise descriptors (SHOT [38] in this paper) from a pair of shapes are passed through anidentical sequence of operations (with shared weights), resulting in refined descriptors F,G. These, in turn, are projected onto theLaplacian eigenbases Φ,Ψ to produce the spectral representations F, G. The functional map (FM) and soft correspondence (Softcor)layers, implementing Equations (3) and (6) respectively, are not parametric and are used to set up the geometrically structured loss `F (5).

us to interpret the soft error (5) as a probability-weightedgeodesic distance from the ground-truth. This measure, in-troduced in [25] as an evaluation criterion for soft maps,endows our solutions with guarantees of mapping nearbypoints on X to nearby points on Y . On the contrary, theclassification cost (4), adopted by existing label-based cor-respondence learning approaches, considers equally corre-spondences that deviate from the ground-truth, no matterhow far. Further, notice that Equation (6) is asymmetric,implying that each pair of training shapes can be used twicefor training (i.e., in both directions). Also note that, differ-ently from previous approaches operating in the label space,in our setting the number of effective training examples (i.e.pairs of shapes) increases quadratically with the number ofshapes in the collection. This is a significant advantage insituations with scarce training data.

We implement descriptor learning using a Siamese resid-ual network architecture [21]. To this network, we con-catenate additional non-parametric layers implementing theleast-squares solve (3) followed by computation of the softcorrespondence according to (6). In particular, the solutionto (3) is obtained in closed form as C = GF†, where †

denotes the pseudo-inverse operation. The complete archi-tecture (named “FMNet”) is illustrated in Figure 3.

5. Implementation details

Data. For increased efficiency, we down-sample the inputshapes to 15K vertices by edge contraction [19]; in casethe input mesh has a smaller amount of vertices, it is keptat full resolution. As input feature for the network we usethe 352-dimensional SHOT descriptor [38], computed onall vertices of the remeshed shapes. The choice of the de-scriptor is mainly driven by its fast computation and its lo-

cal nature, making it a robust candidate in the presence ofmissing parts. Note that while this descriptor is not, strictlyspeaking, deformation invariant, it was shown to work wellfor the deformable setting in practice [35, 27]. Recent workon learning-based shape correspondence makes use of thesame input feature [9, 31].

Network. Our network architecture consists of 7 fully-connected residual layers as described in [21] with ex-ponential linear units (ELUs) [14] and no dimensionalityreduction, implemented in TensorFlow1 [1]. Dependingon the dataset, we used 20K (FAUST synthetic), 100K(FAUST real scans), and 1K (SHREC’16) training mini-batches, each containing ∼1K randomly chosen ground-truth matches. For the FAUST real dataset, sampling wasweighted according to the vertex area elements to preventunduly aggregation of matches to high-resolution portionsof the surface. For the three datasets, we used respectivelyk = 120, 70, 100 eigenfunctions for representing the k × kmatrix C inside the network. Training was done using theADAM optimizer [24] with a learning rate of α = 10−3,β1 = 0.9, β2 = 0.999 and ε = 10−8. The average predic-tion runtime for a pair of FAUST models is 0.25 seconds.

Upscaling. Given two down-sampled shapes X and Y , thenetwork predicts a k× k matrix C encoding the correspon-dence between the two. Since this matrix is expressed w.r.t.basis functions {φi}i, {ψj}j of the low-resolution shapes,it can not be directly used to recover a point-wise map be-tween the full-resolution counterparts X and Y . Therefore,we perform an upscaling step as follows.

Let πX : X → X be the injection mapping each pointin X to the corresponding point in the full shape X (this

1Code and data are available at https://github.com/orlitany/DeepFunctionalMaps.

Page 5: Deep Functional Maps: Structured Prediction for Dense ... · Deep Functional Maps: Structured Prediction for Dense Shape Correspondence Or Litany 1;2 Tal Remez Emanuele Rodola`3;4

1 200 400 600 800 10000

20

40

60

80

100

Best matches

Hit

rate

(%)

CMC

FMNetSHOTSiamese

0 0.02 0.04 0.06 0.08 0.1

20

40

60

80

100

Geodesic error

%C

orre

spon

denc

es

Correspondence quality

FMNetSHOTSiamese

Figure 4. Comparison between our structured prediction model(FMNet), metric learning (Siamese), and baseline SHOT in termsof CMC (left) and geodesic error (right). While the Siamese modelproduces better descriptors in terms of proximity (left), these donot necessarily induce a good functional correspondence (right).

map can be easily recovered by a simple nearest-neighborsearch in R3), and similarly for shape Y . Further, de-note by T : X → Y the point-to-point map recovered fromC using the baseline recovery approach of [33]. A mapT : X ⊃ Im(πX ) → Y is obtained via the composi-tion T = πY ◦ T ◦ π−1

X . However, while T is dense in X ,the map T is sparse in X . In order to map each point inX to a point in Y , we construct pairs of delta functionsδxi : X → {0, 1} and δT (xi) : Y → {0, 1} supported atcorresponding points (xi, T (xi)) for i = 1, . . . , |X |; notethat we have as many corresponding pairs as the number ofvertices in the low-resolution shape X . We use these corre-sponding functions to define the minimization problem:

C∗ = arg minC‖CF− G‖2,1 , (7)

where F = (〈φi, δxj 〉X ) and G = (〈ψi, δT (xj)〉Y) containthe Fourier coefficients (in the full-resolution basis) of thecorresponding delta functions, and the `2,1-norm allows todiscard potential mismatches in the data2. Problem (7) isnon-smooth and convex, and can be solved globally usingADMM-like techniques. A dense point-to-point map be-tween X and Y is finally recovered from the optimal func-tional map C∗ by the nearest-neighbor approach of [33].

6. ResultsWe performed a wide range of experiments on real and

synthetic data to demonstrate the efficacy of our method.Qualitative and quantitative comparisons were carried outwith respect to the state of the art on multiple recent bench-marks, encapsulating different matching scenarios.

2The matrix norm ‖X‖2,1 is defined as the sum of the `2 norms of thecolumns of X.

0 2 4 6 8 100

2000

4000

6000

8000

FMNetSHOTSiamese

Figure 5. Distance distributions (in descriptor space) between cor-rect matches. Since FMNet does not optimize a distribution crite-rion of this kind, it exhibits a heavy tail despite producing excellentcorrespondences.

Error measure. We measure correspondence quality ac-cording to the Princeton benchmark protocol [23]. As-sume to be given a match (x, y) ∈ X × Y , whereas theground-truth correspondence is (x, y∗). Then, we measurethe geodesic error:

ε(x) =dY(y, y∗)

area(Y)1/2, (8)

having units of normalized geodesic length on Y (ideally,zero). We plot cumulative curves showing the percent ofmatches that have error smaller than a variable threshold.

Metric learning. As a proof of concept, we start by study-ing the behavior of our framework when the functional maplayer is removed, and the soft error criterion (6) is replacedwith the siamese loss [20]:

`s(Θ) =∑

x,x+∈S

γ‖FΘ(x)− FΘ(x+)‖22

+∑

x,x−∈D

(1− γ)(µ− ‖FΘ(x)− FΘ(x−)‖2)2+ , (9)

where γ ∈ (0, 1) is a trade-off parameter, µ > 0 is the mar-gin, and (x)+ = max(0, x). Here, the sets S,D ⊂ X × Yconstitute the training data consisting of knowingly similarand dissimilar pairs of points respectively. By consideringthis loss function, we transform our structured predictionmodel into a metric learning model. The learned descriptorsFΘ(x) can be subsequently plugged into (3) to compute acorrespondence; this metric learning approach was recentlyused in a functional map pipeline in [17]. For this test weuse FAUST templates [7] as our data and SHOT [38] as aninput feature.

From the CMC curves3 of Figure 4 (left) and the distancedistributions of Figure 5 we can clearly see that the model(9) succeeds at producing descriptors that attract each other

3Cumulative error characteristic (CMC) curves evaluate the probabil-ity (y-axis) of finding the correct match within the first k best matches(x-axis), obtained as `2-nearest neighbors in descriptor space.

Page 6: Deep Functional Maps: Structured Prediction for Dense ... · Deep Functional Maps: Structured Prediction for Dense Shape Correspondence Or Litany 1;2 Tal Remez Emanuele Rodola`3;4

inter AE inter WE intra AE intra WEZuffi et al. [44] 3.13 6.68 1.57 5.58Chen et al. [13] 8.30 26.80 4.86 26.57

FMNet 4.83 9.56 2.44 26.16

Table 1. Comparison with the state of the art in terms of averageerror (AE) and worst error (WE) on the FAUST challenge with realscans. The error measure is reported in cm.

at corresponding points, while mismatches are repulsed.However, as put in evidence by Figure 4 (right), these de-scriptors do not perform well when they are used for seekinga dense correspondence via (3). Contrarily, our structuredprediction model yields descriptors that are optimized forsuch a correspondence task, leading to a noticeable gain inaccuracy.

Real scans. We carried out experiments on real 3D acqui-sitions using the recent FAUST benchmark [7]. The datasetconsists of real scans (∼200K vertices per shape) of dif-ferent people in a variety of poses, acquired with a full-body 3D stereo capture system. The benchmark is dividedinto two parts, namely the ‘intra-class’ (60 pairs of shapes,with each pair depicting different poses of the same sub-ject) and the ‘inter-class’ challenge (40 pairs of differentsubjects in different poses). The benchmark does not pro-vide ground-truth correspondence for the challenge pairs,whereas the accuracy evaluation is provided by an onlineservice. Hence, for these experiments we only comparewith methods that made their results publicly available viathe official ranking: the recent convex optimization ap-proach of Chen and Koltun [13], and the parametric methodof Zuffi and Black [44].

As training data for our approach we use the officialtraining set of 100 shapes provided with FAUST. Sinceground truth correspondences are only given between low-resolution templates registered to the scans (and not be-tween the scans themselves), during training we augmentedour data by sampling, for each template vertex, one out of10 nearest neighbors to the real scan; this step makes thenetwork more robust to noise in the form of small vertexdisplacement.

The comparison results are reported in Table 1. Read-ing these results, we see that our approach considerably im-proves upon [13] (around 50%); note that while the lattermethod is not learning-based, it relies on a pose prior wherethe shapes are put into initial alignment in 3D in order todrive the optimization to a good solution. The approachof [44] obtains slightly better results than our method, butlacks in generality: it is based on a human-specific paramet-ric model (called the ‘stitched puppet’), trained on a collec-tion of 6000 meshes in different poses from motion capturedata. Our model is trained on almost two orders of mag-nitude less data, and can be applied to any shape category

0 0.02 0.04 0.06 0.08 0.10

20

40

60

80

100

Geodesic error

%C

orre

spon

denc

es

FMNetOSD [?]GCNN [29]LSCNN [8]ADD3 [10]mADD3 [10]

Figure 6. Comparison with learning-based shape matching ap-proaches on the SCAPE dataset. Our method (FMNet) was trainedon FAUST data, demonstrating excellent generalization, while allother methods were trained on SCAPE.

0 0.02 0.04 0.06 0.08 0.10

20

40

60

80

100

Geodesic error

%C

orre

spon

denc

es

FMNetMoNet [31]ACNN [9]GCNN [29]ADD [10]RF [36]BIM [23]

Figure 7. Comparison with learning-based approaches on theFAUST humans dataset. Dashed and solid curves denote perfor-mance before and after refinement respectively. FMNet has 98%correspondences with zero error (top left corner of the plot).

(e.g., animals) as we demonstrate later in this Section.

Transfer. We demonstrate the generalization capabilitiesof FMNet by performing a series of experiments on theSCAPE dataset of human shapes [4], where our network istrained on FAUST scans data as described previously. Wecompare with state-of-the-art learning-based approaches fordeformable shape correspondence, namely optimal spectraldescriptors (OSD) [28], geodesic CNNs (GCNN) [29], lo-calized spectral CNNs (LSCNN) [8], and two variants ofanisotropic diffusion descriptors (ADD3, mADD3) [10].With the exception of FMNet, which was trained only onFAUST data, all the above methods were trained on 60shapes from the SCAPE dataset. The remaining 10 shapesare used for testing. The results are reported in Figure 6;note that for a fair comparison, we show the raw predictedcorrespondence (i.e., without post-processing) for all meth-ods.

Page 7: Deep Functional Maps: Structured Prediction for Dense ... · Deep Functional Maps: Structured Prediction for Dense Shape Correspondence Or Litany 1;2 Tal Remez Emanuele Rodola`3;4

Figure 8. Results of FMNet on the SHREC’16 Partial Correspondence benchmark. Each partial shape is matched to the full shape on theleft; the color texture is transferred via the predicted correspondence.

Synthetic shapes. For these experiments we reproduce ver-batim the experimental setup of [10, 9, 31]: the training setconsists of the first 80 shapes of the FAUST dataset; the re-maining 20 shapes are used for testing. Differently fromthe comparisons of Table 1, the shapes are now taken fromthe synthetic dataset provided with FAUST (∼7K verticesper shape), for which exact ground-truth correspondence isavailable. We compare with the most recent state of theart in learning-based shape correspondence: random forests(RF) [36], anisotropic diffusion descriptors (ADD) [10],geodesic CNNs (GCNN) [29], anisotropic CNNs (ACNN)[9], and the very recent MoNet model [31]. As a represen-tative method for the family of axiomatic techniques, weadditionally include blended intrinsic maps (BIM) [23] inthe comparison.

The results are reported in Figure 7; here, all meth-ods were post-processed with the correspondence refine-ment technique of [40]. For FMNet and MoNet (the top-performing competitor) we also report curves before re-finement. We note that the raw prediction of MoNet hasa higher accuracy at zero. This is to be expected due tothe classifier nature of this method (and all the methods inthis comparison): the logistic loss (4) aims at high point-wise accuracy, but has limited global awareness of the cor-respondence. Indeed, the task-driven nature of our approachinduces lower accuracy at zero, but better global behavior –note how the curve “saturates” at around 0.06 while MoNetnever does. As a result, FMNet significantly outperformsMoNet after refinement, producing almost ideal correspon-dence with zero error.

Partial non-human shapes. Our framework does not relyon any specific shape model, as it learns from the shapecategories represented in the training data. In particular,it does not necessarily require the objects to be completeshapes: different forms of partiality can be tackled if ade-quately represented in the training set.

We demonstrate this by running our method on the recentSHREC’16 Partial Correspondence challenge [16]. Thebenchmark consists of hundreds of shapes of multiple cate-

gories with missing parts of various forms and sizes; a train-ing set is also provided. We selected the ‘dog’ class fromthe ‘holes’ sub-challenge, being this among the hardest cat-egories in the benchmark. The dataset is officially split intojust 10 training shapes, and 26 test shapes. Qualitative ex-amples of the obtained solutions are reported in Figure 8.

7. Discussion and conclusions

We introduced a new neural network based method fordense shape correspondence, structured according to thefunctional maps framework. Building upon the recent suc-cess of end-to-end learning approaches, our network di-rectly estimates correspondences. This is in contrast to pre-vious descriptor learning techniques, that do not account forpost processing while training. We showed this methodol-ogy to be beneficial via an evaluation on a several challeng-ing benchmarks, comprising synthetic models, real scanswith acquisition artifacts, and partiality. Being model-free,we demonstrated our method can be adapted to differentshape categories such as dogs. Furthermore, we showed ourmethod is capable of generalizing between different data-sets.

Limitations. Laplacian eigenfunctions are inherently sen-sitive to topological changes. Indeed, such examples provedto be more challenging for our method. A different choiceof a basis may be useful in mitigating this issue.

As shown by recent worksaddressing partial correspondenceusing functional maps [35], spe-cial care should be taken when re-covering matrix C from the spec-tral representation of the descrip-tors. While our method was ableto recover most pairs with miss-ing parts, it failed to recover cor-respondences under extreme partiality (see inset). Thiscould be addressed by incorporating partiality priors intoour structured prediction model.

Page 8: Deep Functional Maps: Structured Prediction for Dense ... · Deep Functional Maps: Structured Prediction for Dense Shape Correspondence Or Litany 1;2 Tal Remez Emanuele Rodola`3;4

Acknowledgments

The authors wish to thank Jonathan Masci, FedericoMonti, Dan Raviv and Maks Ovsjanikov for useful discus-sions. OL, TR and AB are supported by the ERC grantno. 335491. ER and MB are supported by the ERC grantsno. 307047 and 724228, a Google Research Faculty Award,a Radcliffe Fellowship, a Nvidia equipment grant, and aTUM-IAS Rudolf Diesel Industry Fellowship.

References[1] M. Abadi, A. Agarwal, P. Barham, E. Brevdo, Z. Chen,

C. Citro, G. S. Corrado, A. Davis, J. Dean, M. Devin, et al.Tensorflow: Large-scale machine learning on heterogeneoussystems, 2015. Software available from tensorflow. org, 1,2015. 4

[2] Y. Aflalo, H. Brezis, and R. Kimmel. On the optimality ofshape and data representation in the spectral domain. SIAMJ. Imaging Sciences, 8(2):1141–1160, 2015. 2, 3

[3] Y. Aflalo, A. Dubrovina, and R. Kimmel. Spectral gen-eralized multi-dimensional scaling. IJCV, 118(3):380–392,2016. 2

[4] D. Anguelov, P. Srinivasan, D. Koller, S. Thrun, J. Rodgers,and J. Davis. Scape: Shape completion and animation ofpeople. Trans. Graphics, 24(3):408–416, 2005. 6

[5] M. Aubry, U. Schlickewei, and D. Cremers. The wave kernelsignature: A quantum mechanical approach to shape analy-sis. In Proc. ICCV, 2011. 2

[6] S. Biasotti, A. Cerri, A. Bronstein, and M. Bronstein. Recenttrends, applications, and perspectives in 3d shape similarityassessment. In Computer Graphics Forum, 2015. 2

[7] F. Bogo, J. Romero, M. Loper, and M. J. Black. FAUST:Dataset and evaluation for 3D mesh registration. In Proc.CVPR, June 2014. 5, 6

[8] D. Boscaini, J. Masci, S. Melzi, M. M. Bronstein, U. Castel-lani, and P. Vandergheynst. Learning class-specific descrip-tors for deformable shapes using localized spectral convolu-tional networks. Computer Graphics Forum, 34(5):13–23,2015. 1, 2, 6

[9] D. Boscaini, J. Masci, E. Rodola, and M. M. Bronstein.Learning shape correspondence with anisotropic convolu-tional neural networks. In Proc. NIPS, 2016. 1, 2, 3, 4,6, 7

[10] D. Boscaini, J. Masci, E. Rodola, M. M. Bronstein, andD. Cremers. Anisotropic diffusion descriptors. ComputerGraphics Forum, 35(2):431–441, 2016. 1, 2, 3, 6, 7

[11] A. M. Bronstein, M. M. Bronstein, and R. Kimmel. General-ized multidimensional scaling: a framework for isometry-invariant partial surface matching. PNAS, 103(5):1168–1172, 2006. 2

[12] M. Bronstein, J. Bruna, Y. LeCun, A. Szlam, and P. Van-dergheynst. Geometric deep learning: going beyond Eu-clidean data. arXiv:1611.08097, 2016. 2

[13] Q. Chen and V. Koltun. Robust nonrigid registration by con-vex optimization. In Proc. ICCV, 2015. 2, 6

[14] D.-A. Clevert, T. Unterthiner, and S. Hochreiter. Fast andaccurate deep network learning by exponential linear units(elus). arXiv:1511.07289, 2015. 4

[15] E. Corman, M. Ovsjanikov, and A. Chambolle. Superviseddescriptor learning for non-rigid shape matching. In Proc.ECCV, 2014. 2

[16] L. Cosmo, , et al. SHREC’16: Partial matching of de-formable shapes. In Proc. 3DOR, 2016. 7

[17] L. Cosmo, E. Rodola, J. Masci, A. Torsello, and M. Bron-stein. Matching deformable objects in clutter. In Proc. 3DV,2016. 2, 5

[18] D. Eynard, E. Rodola, K. Glashoff, and M. M. Bronstein.Coupled functional maps. In Proc. 3DV, 2016. 2

[19] M. Garland and P. S. Heckbert. Surface simplification usingquadric error metrics. In Proc. SIGGRAPH, pages 209–216,1997. 4

[20] R. Hadsell, S. Chopra, and Y. LeCun. Dimensionality reduc-tion by learning an invariant mapping. In Proc. CVPR, 2006.5

[21] K. He, X. Zhang, S. Ren, and J. Sun. Deep residual learningfor image recognition. In Proc. CVPR, 2016. 4

[22] Q. Huang, F. Wang, and L. J. Guibas. Functional map net-works for analyzing and exploring large shape collections.Trans. Graphics, 33(4):36, 2014. 2

[23] V. G. Kim, Y. Lipman, and T. A. Funkhouser. Blended in-trinsic maps. Trans. Graphics, 30(4):79, 2011. 2, 5, 6, 7

[24] D. P. Kingma and J. Ba. ADAM: A method for stochasticoptimization. In ICLR, 2015. 4

[25] A. Kovnatsky, M. M. Bronstein, X. Bresson, and P. Van-dergheynst. Functional correspondence by matrix comple-tion. In Proc. CVPR, 2015. 2, 4

[26] O. Litany, E. Rodola, A. M. Bronstein, and M. M. Bronstein.Fully spectral partial shape matching. Computer GraphicsForum, 36(2), 2017. 2

[27] O. Litany, E. Rodola, A. M. Bronstein, M. M. Bronstein, andD. Cremers. Non-rigid puzzles. Computer Graphics Forum,35(5):135–143, 2016. 2, 4

[28] R. Litman and A. M. Bronstein. Learning spectral descrip-tors for deformable shape correspondence. Trans. PAMI,36(1):170–180, 2014. 1, 2, 6

[29] J. Masci, D. Boscaini, M. M. Bronstein, and P. Van-dergheynst. Geodesic convolutional neural networks on Rie-mannian manifolds. In Proc. 3dRR, 2015. 1, 2, 3, 6, 7

[30] M. Meyer, M. Desbrun, P. Schroder, and A. H. Barr. Discretedifferential-geometry operators for triangulated 2-manifolds.Visualization&Mathematics, pages 35–57, 2003. 3

[31] F. Monti, D. Boscaini, J. Masci, E. Rodola, J. Svoboda, andM. M. Bronstein. Geometric deep learning on graphs andmanifolds using mixture model CNNs. In Proc. CVPR, 2017.1, 2, 3, 4, 6, 7

[32] D. Nogneng and M. Ovsjanikov. Informative descriptorpreservation via commutativity for shape matching,. Com-puter Graphics Forum, 36(2), 2017. 2

[33] M. Ovsjanikov, M. Ben-Chen, J. Solomon, A. Butscher, andL. Guibas. Functional maps: a flexible representation ofmaps between shapes. Trans. Graphics, 31(4):30:1–30:11,July 2012. 2, 3, 5

Page 9: Deep Functional Maps: Structured Prediction for Dense ... · Deep Functional Maps: Structured Prediction for Dense Shape Correspondence Or Litany 1;2 Tal Remez Emanuele Rodola`3;4

[34] J. Pokrass, A. M. Bronstein, M. M. Bronstein, P. Sprech-mann, and G. Sapiro. Sparse modeling of intrinsic cor-respondences. Computer Graphics Forum, 32(2):459–468,2013. 2

[35] E. Rodola, L. Cosmo, M. M. Bronstein, A. Torsello, andD. Cremers. Partial functional correspondence. ComputerGraphics Forum, 36(1):222–236, 2017. 2, 4, 7

[36] E. Rodola, S. Rota Bulo, T. Windheuser, M. Vestner, andD. Cremers. Dense non-rigid shape correspondence usingrandom forests. In Proc. CVPR, 2014. 1, 2, 3, 6, 7

[37] J. Sun, M. Ovsjanikov, and L. Guibas. A concise and prov-ably informative multi-scale signature based on heat diffu-sion. In Proc. SGP, 2009. 2, 3

[38] F. Tombari, S. Salti, and L. Di Stefano. Unique signaturesof histograms for local surface description. In Proc. ECCV,2010. 3, 4, 5

[39] O. van Kaick, H. Zhang, G. Hamarneh, and D. Cohen-Or.A survey on shape correspondence. Computer Graphics Fo-rum, 30(6):1681–1707, 2011. 2

[40] M. Vestner, R. Litman, E. Rodola, A. Bronstein, and D. Cre-mers. Product manifold filter: Non-rigid shape correspon-dence via kernel density estimation in the product space. InProc. CVPR, 2017. 7

[41] L. Wei, Q. Huang, D. Ceylan, E. Vouga, and H. Li. Densehuman body correspondences using convolutional networks.In Proc. CVPR, 2016. 2

[42] T. Windheuser, M. Vestner, E. Rodola, R. Triebel, andD. Cremers. Optimal intrinsic descriptors for non-rigid shapeanalysis. In Proc. BMVC, 2014. 2

[43] Z. Zhang. Microsoft Kinect sensor and its effect. IEEE Mul-timedia, 19(2):4–10, 2012. 1

[44] S. Zuffi and M. J. Black. The stitched puppet: A graphicalmodel of 3D human shape and pose. In Proc. CVPR, 2015.6