photometric stereo using constrained bivariate regression ... · the unknown normal vec-tor was...

8
Photometric Stereo using Constrained Bivariate Regression for General Isotropic Surfaces Satoshi Ikehata * The University of Tokyo, Japan [email protected] Kiyoharu Aizawa The University of Tokyo, Japan [email protected] Abstract This paper presents a photometric stereo method that is purely pixelwise and handles general isotropic surfaces in a stable manner. Following the recently proposed sum- of-lobes representation of the isotropic reflectance function, we constructed a constrained bivariate regression problem where the regression function is approximated by smooth, bivariate Bernstein polynomials. The unknown normal vec- tor was separated from the unknown reflectance function by considering the inverse representation of the image for- mation process, and then we could accurately compute the unknown surface normals by solving a simple and ef- ficient quadratic programming problem. Extensive evalu- ations that showed the state-of-the-art performance using both synthetic and real-world images were performed. 1. Introduction Photometric stereo estimates the surface normals of an object from appearance variations under different lighting conditions. Since Woodham [20] first introduced the photo- metric stereo for Lambertian scenes, the extension to real- world objects which exhibit diverse appearances beyond Lambertian model has drawn significant interest. Traditionally, certain parametric reflectance models are assumed to inversely solve the photometric stereo problem. One of the most popular classes assumes a basic Lambertian model but augmented with outlier detection for handling all non-Lambertian regions of the scene [21, 11]. This strategy is numerically stable and relatively robust to shadows and image noises, but complex reflections such as rough specu- larities can be highly disruptive. In contrast to the first class, a second class of methods treats non-Lambertian reflections as inliers using the nonlinear bidirectional reflectance dis- tribution function (BRDF) [9, 16]. While these methods are more capable of handling a wide variety of objects includ- * This work was supported by the Grants-in-Aid for JSPS Fellows (No. 248615). ing rough surfaces, they may suffer from numerical insta- bilities derived from the complex nonlinear optimization. Instead of explicitly modeling the parametric form of reflectance, the monotonicity property of reflectance has been recently integrated in the photometric stereo prob- lem [2, 10, 17]. Chandraker and Ramamoorthi [6] show that an isotropic BRDF consists of a sum of lobes whose contribution to the reflected intensity decreases monotoni- cally as the surface normal deviates away from the direction where the reflectance lobe is concentrated (i.e., referred to as a preferred direction). Following this remark, the surface normal has been recovered utilizing the monotonicity of the reflectance function under the assumption that the number of lobes is one and its preferred directions are known (e.g., the lighting direction in [10] and half vector in [2, 17]). While effective, these methods are highly disruptive when the assumption on the preferred direction is incorrect or the reflectance function is composed of two or more lobes. Furthermore, to our knowledge, simultaneous estimation of both azimuth and elevation angles has never been achieved by enforcing the monotonicity of a reflectance function that has a preferred direction that is different from the lighting vector (note that [2, 17] assume that the azimuth angle of the surface normal is known). This paper presents a photometric stereo algorithm for accurate estimation of surface normals of a general isotropic scene by enforcing the monotonicity of a reflectance func- tion with an unknown lobe number and preferred directions. For this purpose, the bivariate reflectance model is devel- oped in Section 2 where pixelwise appearances are well approximated by a bivariate monotonic (and therefore in- vertible) smooth function of the dot products between the surface normal and the lighting direction, and between the lighting and viewing directions. We may then consider the inverse representation of the image formation process, where the unknown normal vector is separated from the un- known monotonic inverse reflectance function. By parame- terizing the latter using the Bernstein polynomials [13], we obtain a set of constrained linear equations in both the sur- face normals and reflectance parameters, leading to a sim-

Upload: others

Post on 03-Oct-2020

7 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Photometric Stereo using Constrained Bivariate Regression ... · The unknown normal vec-tor was separated from the unknown reflectance function by considering the inverse representation

Photometric Stereo using Constrained Bivariate Regression for GeneralIsotropic Surfaces

Satoshi Ikehata∗

The University of Tokyo, [email protected]

Kiyoharu AizawaThe University of Tokyo, [email protected]

Abstract

This paper presents a photometric stereo method thatis purely pixelwise and handles general isotropic surfacesin a stable manner. Following the recently proposed sum-of-lobes representation of the isotropic reflectance function,we constructed a constrained bivariate regression problemwhere the regression function is approximated by smooth,bivariate Bernstein polynomials. The unknown normal vec-tor was separated from the unknown reflectance functionby considering the inverse representation of the image for-mation process, and then we could accurately computethe unknown surface normals by solving a simple and ef-ficient quadratic programming problem. Extensive evalu-ations that showed the state-of-the-art performance usingboth synthetic and real-world images were performed.

1. IntroductionPhotometric stereo estimates the surface normals of an

object from appearance variations under different lightingconditions. Since Woodham [20] first introduced the photo-metric stereo for Lambertian scenes, the extension to real-world objects which exhibit diverse appearances beyondLambertian model has drawn significant interest.

Traditionally, certain parametric reflectance models areassumed to inversely solve the photometric stereo problem.One of the most popular classes assumes a basic Lambertianmodel but augmented with outlier detection for handling allnon-Lambertian regions of the scene [21, 11]. This strategyis numerically stable and relatively robust to shadows andimage noises, but complex reflections such as rough specu-larities can be highly disruptive. In contrast to the first class,a second class of methods treats non-Lambertian reflectionsas inliers using the nonlinear bidirectional reflectance dis-tribution function (BRDF) [9, 16]. While these methods aremore capable of handling a wide variety of objects includ-

∗This work was supported by the Grants-in-Aid for JSPS Fellows (No.248615).

ing rough surfaces, they may suffer from numerical insta-bilities derived from the complex nonlinear optimization.

Instead of explicitly modeling the parametric form ofreflectance, the monotonicity property of reflectance hasbeen recently integrated in the photometric stereo prob-lem [2, 10, 17]. Chandraker and Ramamoorthi [6] showthat an isotropic BRDF consists of a sum of lobes whosecontribution to the reflected intensity decreases monotoni-cally as the surface normal deviates away from the directionwhere the reflectance lobe is concentrated (i.e., referred toas a preferred direction). Following this remark, the surfacenormal has been recovered utilizing the monotonicity of thereflectance function under the assumption that the numberof lobes is one and its preferred directions are known (e.g.,the lighting direction in [10] and half vector in [2, 17]).While effective, these methods are highly disruptive whenthe assumption on the preferred direction is incorrect orthe reflectance function is composed of two or more lobes.Furthermore, to our knowledge, simultaneous estimation ofboth azimuth and elevation angles has never been achievedby enforcing the monotonicity of a reflectance function thathas a preferred direction that is different from the lightingvector (note that [2, 17] assume that the azimuth angle ofthe surface normal is known).

This paper presents a photometric stereo algorithm foraccurate estimation of surface normals of a general isotropicscene by enforcing the monotonicity of a reflectance func-tion with an unknown lobe number and preferred directions.For this purpose, the bivariate reflectance model is devel-oped in Section 2 where pixelwise appearances are wellapproximated by a bivariate monotonic (and therefore in-vertible) smooth function of the dot products between thesurface normal and the lighting direction, and between thelighting and viewing directions. We may then considerthe inverse representation of the image formation process,where the unknown normal vector is separated from the un-known monotonic inverse reflectance function. By parame-terizing the latter using the Bernstein polynomials [13], weobtain a set of constrained linear equations in both the sur-face normals and reflectance parameters, leading to a sim-

Page 2: Photometric Stereo using Constrained Bivariate Regression ... · The unknown normal vec-tor was separated from the unknown reflectance function by considering the inverse representation

ple, quadratic programming problem.The proposed framework benefits from an efficient pix-

elwise optimization that is easily amenable to parallel pro-cessing and does not require typical smoothness constraintsfor both object structure and reflectance, which could dis-rupt the recovery of fine details.

2. Photometric stereo using constrained bivari-ate regression

In this section, we formulate the photometric stereo asa constrained bivariate regression problem. Henceforth werely on the following assumptions:(1) The relative position between the camera and the objectis fixed across all images.(2) The object is illuminated by a point light source at infin-ity from varying and known directions.(3) The camera view is orthographic, and the radiometricresponse function is linear.

2.1. Problem Statement

Diverse appearances of real-world objects can be en-coded by a BRDF (ρ) that relates the observed intensity I ata given point on the object to the associated surface normaln ∈ R3, the incoming lighting direction l ∈ R3, and theoutgoing viewing direction v ∈ R3 via

I = ρ(n, l,v)max (nT l, 0), (1)

where max (nT l, 0) accounts for attached shadows. Thereis a problem for photometric stereo in recovering the sur-face normaln of a scene by inversely solving Eq. (1) from acollection of m observations under different lighting condi-tions. Note that except for uncalibrated photometric stereoproblems such as [8], l and v are usually known.

Recently Chandraker and Ramamoorthi [6] have pre-sented a semiparametric model of an isotropic BRDF thatis represented as a sum of K different functions giving

ρ =

K∑k=1

ρk(nTαk). (2)

Here ρk are (unknown) nonlinear functions, and αk (i.e.,||αk|| = 1) are called preferred directions, along whichρk are concentrated. It is known that physically valid re-flectance functions satisfy the following requirements.(L1) Monotonicity: ρk′ > 0.(L2) Nonnegativity: ρk ≥ 0.(L3) Passing thorough the origin: ρk(0) = 0.

Chandraker and Ramamoorthi [6] have shown that in-versely solving Eq. (2) under known surface normals givesgood estimates of a wide variety of isotropic BRDFs with-out suffering from the curse of dimensionality. Unfortu-nately, however, solving Eq. (2) directly is prohibitively dif-

ficult in the context of the photometric stereo problem be-cause of the numerous unknown parameters, some of whichare coincident in the same term (i.e.,n,αk, ρk). There-fore, most of the conventional photometric stereo algo-rithms have assumed that the dominant preferred directionof the reflectance function is unique and known [10, 2, 17].

Instead of this approach, we only assume that the pre-ferred direction (αk) of each function (ρk) is lying on theplane spanned by the lighting and viewing directions as

αk =pkl+ qkv

||pkl+ qkv||, (3)

where pk and qk are nonnegative unknown values (i.e.,pk ≥ 0 and qk ≥ 0). The degree of freedom of αk isactually 1 because ||αk|| = 1. 1 Then, this assumptionprovides us following important result.

Theorem: Suppose there is no shadow at a surfacepoint (i.e., ∀i nT li, lTi v ≥ 0 and Ii ≥ 0, where i is theindex of the light) and ρ(n, l,v) in Eq. (1) has the formof Eq. (2), whose parameters satisfy the requirements ofa physically valid BRDF (L1)–(L3) and Eq. (3). Then,it is guaranteed that there exists at least one continuousbivariate function f(x, y) ∀x, y ∈ [0, 1], which satisfiesf ≥ 0, ∂f/∂x > 0, ∂f/∂y ≤ 0 and ∀i Ii = f(nT li, l

Ti v).

Proof : From Eq. (3), nTαk is transformed into

nTαk =pkn

T l+ qknTv√

p2k + q2k + 2pkqklTv. (4)

Here we used ||l|| = ||v|| = 1. Eq. (4) illustrates thatnTαk is non-decreasing for nT l with fixed lTv andnon-increasing for lTv with fixed nT l since p, q arenon-negative constant values and nTv is constant overdifferent lightings. From (L1), it is guaranteed that eachρk(n

Tαk) is also non-decreasing/non-increasing fornT l and lTv when either of them is fixed. Integratingthese results into Eq. (1) and Eq. (2), it is proved thatI is monotonic increasing for nT l with fixed lTv andnon-increasing for lTv with fixed nT l, which implieswe can always define continuous functions f(x, y)which satisfy f ≥ 0, ∂f/∂x > 0, ∂f/∂y ≤ 0and ∀i Ii = f(nT li, l

Ti v) since Ii ≥ 0 and

∀i 0 ≤ nT li, lTi v ≤ 1.

We illustrate this theorem in Fig. 1-(a). We note thatassuming that f(x, y) is always passing through the y-axis(i.e., f(0, y) = 0) does not limit any kind of isotropicBRDF represented by Eq. (1) because I = nT lρ.

Following this theorem, we formulate the photometricstereo as a constrained bivariate regression problem whose

1Note that this assumption does not violate most existing BRDF modelssuch as [7, 19, 15, 12, 3].

Page 3: Photometric Stereo using Constrained Bivariate Regression ... · The unknown normal vec-tor was separated from the unknown reflectance function by considering the inverse representation

𝑥

𝑧

𝑦 𝑦

𝑥

𝑧

𝑥𝑖 , 𝑦𝑖 , 𝑧𝑖 = {𝒏𝑇𝒍𝑖 , 𝒍𝑖𝑇𝒗, 𝐼𝑖}

(a) (b)

𝑧 = 𝑓(𝑥, 𝑦) 𝑥 = 𝑔(𝑦, 𝑧)

1.0 1.0 1.0 1.0

0.0 0.0

Figure 1. (a) Collections of {x, y, z} = {nT l, lTv, I} are lyingon a continuous function of z = f(x, y) which satisfies ∂f/∂x >0, ∂f/∂y ≤ 0 and f(0, y) = 0. (b) 3-d points lying on f are alsolying on a inverse function (x = g(y, z)).

goal is to recover an unknown surface normal n and a con-tinuous bivariate function f from a collection of lightingdirections li and associated appearances Ii (i = 1, . . . ,m),which satisfy the following equations and constraints:

Ii = f(nT li, lTi v) i = 1, . . . ,m. (5)

(L4) Monotonicity (x): ∂f/∂x > 0.(L5) Monotonicity (y): ∂f/∂y ≤ 0.(L6) Nonnegativity: f ≥ 0.(L7) Passing through y-axis: f(0, y) = 0.

We call Eq. (5) the forward bivariate reflectance model.The major benefit of this formulation is that we do not needto explicitly approximate the number of lobes K and theirpreferred directions αk. However, there is a critical issue,namely the coincidence of unknown parameters n and fin the same term. We overcome this difficulty by a conve-nient, inverse representation of the imaging model appliedthrough a constrained bivariate regression framework.

2.2. Inverse Bivariate Reflectance Model

Strict monotonicity of f(x, y) (L4) guarantees theunique existence of the function giving x = g(y, f(x, y)) =g(y, z), which obeys the following requirements.(L8) Monotonicity (y): ∂g/∂y ≥ 0.(L9) Monotonicity (z): ∂g/∂z > 0.(L10) Nonnegativity: g ≥ 0.(L11) Passing through the y-axis: g(y, 0) = 0.The proof, which has been omitted for brevity, is obviousby seeing Fig. 1-(b). From the definition, each 3-D pointof {x, y, z} = {nT li, lTi v, Ii}(i = 1, . . . ,m) lying on f isalso lying on g as follows.

nT li = g(lTi v, Ii) i = 1, . . . ,m. (6)

In contradiction to Eq. (5), we call Eq. (6) the inverse bivari-ate reflectance model. Our goal is now updated to recoverthe surface normal n and a continuous bivariate function g

with some shape restrictions (L8)–(L11). The fundamentaladvantage of Eq. (6) is that unknown variables of n and gare separated, which contributes to simplifying the problem.

Constraints on g limit the solution space of Eq. (6), butthere are still multiple feasible solutions of a pair of n and gbecause {lTi v, Ii} are sparsely distributed on the valid rangeof {y, z}. To reduce the inherent ambiguity of the prob-lem, we further assume a parametric model of the inversebivariate reflectance function g(y, z). Given that the left-hand side of Eq. (6) is linear in the unknown normal vectorn, for computational simplicity we would like to impose asimilar linearity on the right-hand side in our parameterizedrepresentation of g(lTv, I) (we have omitted subscripts forsimplicity). For this purpose, we then choose to expressg(lTv, I) as a summation over p known nonlinear basisfunctions gk(lTv, I) weighted by an unknown coefficientvector β , [β1, . . . , βp]

T , leading to the representation

g(lTv, I) =

p∑k=1

βkgk(lTv, I). (7)

While nonlinear in lTv and I , g(lTv, I) is clearly linearin β. We need to choose gk carefully because estimatinga multivariate regression function subject to shape restric-tions with compact support is challenging and usually verytime-consuming [18]. Here we adopt the bivariate Bern-stein polynomials [13], where the shape-restricted regres-sion function estimation is shown to be the solution of aquadratic programming problem [5, 18], making it compu-tationally attractive. Furthermore, the Bernstein polynomi-als approximation naturally selects smooth functions withlittle computational effort, unlike other nonparametric re-gression functions (e.g., smoothing spline [4]), which im-plicitly enforces the smoothness of BRDF as in [2].

Bivariate Bernstein polynomials [13] are composed ofmultiple basis functions of the form

bk1,k2(x1, x2, N1, N2) = bk1(x1, N1)bk2(x2, N2),

bki(xi, Ni) =

(Niki

)xki(1− xi)Ni−ki (i = 1, 2),

(8)

where 0 ≤ xi ≤ 1 and Ni is the order of the polynomial asfor xi, which will be chosen as a function of the sample sizem (e.g., Ni = o(mγ

i ) with γi > 0 suitably chosen via thepopular V-fold cross-validation method as shown in [18]).We transform Ii(i = 1, . . . ,m) to lie in the unit [0, 1] viaa simple linear equation as Ii ← Ii/max (I). Note thatlTi v naturally lies in [0, 1] as we only consider the casev = [0, 0, 1]T and lz > 0. Then, the bivariate Bernsteinpolynomials approximation of g is represented as

x = g(y, z) = βT bNy,Nz (y, z)

=

Ny∑ky=0

Nz∑kz=0

βky,kzbky,kz (y, z,Ny, Nz),(9)

Page 4: Photometric Stereo using Constrained Bivariate Regression ... · The unknown normal vec-tor was separated from the unknown reflectance function by considering the inverse representation

where bNy,Nz , [b0,0, . . . , bNy,Nz ]T ∈ R(Ny+1)(Nz+1)×1

and β , [β0,0, . . . , βNy,Nz ]T ∈ R(Ny+1)(Nz+1)×1.

Unlike the B-splines procedure (which may requirequadratic constraints on the coefficients) [4], shape re-strictions (e.g., monotonicity, nonnegativity) on Eq. (9)are easily encoded via linear constraints, that is Aβ ≥ 0and Cβ = 0, where A,C are shape restriction matrices.Following [18], the shape restriction matrices required forour problem are defined as follows.

(1) Monotonicity: ∂g/∂y ≥ 0 and ∂g/∂z ≥ 0. (L8), (L9)The first-order partial derivatives of g with respect to yin Eq. (9) can be represented as

∂g(y, z)/∂y (10)

= Ny

Nz∑kz=0

Ny−1∑ky=0

(βky+1,kz − βky,kz )bky,kz (y, z,Ny − 1, Nz).

From the definition in Eq. (8), it is easy to show thatall Bernstein basis polynomials are nonnegative withrespect to 0 ≤ y, z ≤ 1. Hence, the non-decreasingconstraint (i.e., ∂g/∂y ≥ 0) is simply achieved byenforcing βky+1,kz ≥ βky,kz . The non-decreasingconstraint with respect to z (i.e., ∂g/∂z ≥ 0) isalso achieved in the same manner. The restrictionmatrix for linear constraint Amonoβ ≥ 0 is repre-sented as Amono = [ATyA

Tz ]T , which is composed

of submatrices Ay ∈ RNy(Nz+1)×(Ny+1)(Nz+1) andAz ∈ RNz(Ny+1)×(Ny+1)(Nz+1), where Ar ensures themonotonicity of the function with respect to r. 2 Note thatthe strict monotonicity constraint ∂g/∂z > 0 was eased to∂g/∂z ≥ 0 for computational simplicity.

(2) Nonnegativity: g ≥ 0. (L10)The nonnegativity of g is guaranteed when ∀i βi ≥ 0.Hence, the restriction matrix for a linear constraintAnonnegβ ≥ 0 is as Anonneg , diag([1, . . . , 1]) ∈R(Ny+1)(Nz+1)×(Ny+1)(Nz+1).

(3) Passing through the y-axis: g(y, 0) = 0. (L11)From the definition in Eq. (8), bky,kz (y, 0) = 0 for all kz 6=0. Therefore g(y, 0) =

∑Ny

ky=0 βky,0bky,0(y, 0, Ny, Nz)becomes zero for all y when ∀ky βky,0 = 0.This constraint is encoded via a linear constraintCβ = 0, where C ∈ R(Ny+1)(Nz+1)×(Ny+1)(Nz+1) ,diag([1, 0, . . . , 1, 0, . . . , 1, . . .]) with Nz of 0 between 1.

2.3. Solution Method

By substituting Eq. (9) into the inverse bivariate re-flectance model, Eq. (6) becomes

nT li = βT bNy,Nz

(lTi v, Ii) i = 1, . . . ,m, (11)

2The concrete form of the shape restriction matrix is included in thesupplementary material.

where the coefficients of Bernstein polynomials (β) are re-stricted via the following equation:

Aβ =

[AmonoAnonneg

]β ≥ 0, Cβ = 0. (12)

Collecting variations of observations at the same pixelunder different lighting directions, Eq. (11) can be mergedinto following linear problem:

LTn = BTβ. (13)

Here, B , [bNy,Nz(lT1 v, I1), . . . , bNy,Nz

(lTmv, Im)] andL , [l1, . . . , lm]. By merging unknown variables (n,β),this problem is transformed into

Px = [L −B]Tx = 0, (14)

where x , [nx, ny, nz, β0,0, . . . , βNy,Nz ]T and nx, ny, nz

are the three elements of the surface normal. Without lossof generality, we may avoid the degenerate x = 0 solu-tion to Eq. (14) by constraining

∑i xi = 1, which implies

cTx = 1 where c = [1, . . . , 1]T .Given the appearance variations (I1, I2, . . . , Im) under

different known lighting conditions (l1, l2, . . . , lm), the op-timal surface normal (n) and model parameters (β) are re-covered by solving the constrained linear problem:

minx||Px||22, s.t. Ax ≥ 0 and Cx = 0, (15)

where A , [0 A] and C ,

[cT

0 C

]. Note that Eq. (15) can

be effectively solved by the general quadratic programming.

3. Handling Retroreflective MaterialsIf the reflectance of a target object obeys Eq. (2), our

method reasonably recovers the surface normal of the ob-ject by solving Eq. (15). However, one limitation of ourmethod is that this assumption is not satisfied in the pres-ence of retroreflections that are often observed on roughsurfaces because Eq. (2) does not have the ability to rep-resent this kind of reflection. In the presence of retrore-flections, our surface normal estimation fails because of theviolation of (L5) by the behavior of retroreflections in thatthe power of reflections increases as lTv increases. Whileit may limit available materials, fortunately we have foundthat our method practically handles retroreflective materi-als by simply reversing the direction of the monotonicityconstraint on lTv in Eq. (15) (i.e., use (L8)’ ∂g/∂y ≤ 0instead of (L8)) because the retroreflections do not affectthe monotonicity for nT l while they unify the direction ofthe monotonicity for lTv over the BRDF space. The prob-lem of course is that we do not know whether the mate-rial is retroreflective or not. To overcome this difficulty,

Page 5: Photometric Stereo using Constrained Bivariate Regression ... · The unknown normal vec-tor was separated from the unknown reflectance function by considering the inverse representation

we present a practical approach for handling both non- andretroreflective materials. The important observation is thatwhen we incorrectly constrain the problem, the regressionusually fails. Therefore, after we have the regression out-puts under both constraints, we can judge which constraintwas optimal by examining regression errors.

However, we have empirically found that comparing re-gression errors in Eq. (15) does not work since the flexibleBernstein polynomials are generally well fitted to observa-tions even though the constraint was not correct; instead, wecompute the following linear regression error E for choos-ing the optimal solution:

a = arg mina

m∑i=1

‖nT li − aIi‖22, (16)

E =

m∑i=1

‖nT li − aIi‖22. (17)

Here, n is a recovered surface normal by solving Eq. (15)under the monotonicity constraint for lTv in either of twodirections. We simply choose the direction for which E issmaller. This strategy is very simple but very efficient as wewill show in Section 4.

4. Experimental ResultsIn this section, we evaluate our method on synthetic and

real image data. All experiments were performed on an In-tel Core i7-2640M (2.80GHz, single thread) machine with8GB RAM and were implemented in MATLAB. For thequantitative evaluation, we generated 32-bit HDR images ofa sphere (256×256) with foreground masks under differentBRDF settings: (A) common physical or phenomenologi-cal BRDF (Cook–Torrance [7], Ward [19] , Lafortune [12],Oren–Nayar [15] and Ashikhmin–Shirley [3]) and (B) themeasured MERL BRDF database [14]. Lighting directionswere randomly selected from a hemisphere with the ob-ject placed at the center. Additionally, for the third dataset,denoted (C), we used real images for qualitatively evalu-ating our method in practical situations. For each dataset,shadows were removed via simple thresholding as in otherworks such as [16] (Thresholds for shadow removal werefixed over algorithms but varied over objects manually). Be-cause ground truth surface normals are provided in datasets(A) and (B), we quantitatively evaluated our method by theangular error between the recovered normal map and theground truth when using these datasets.

4.1. Evaluation with Synthesized BRDF

We evaluated the performance using the synthesized im-ages in dataset (A) generated under 100 different light-ings using five common BRDFs3. Here we compared

3Details of BRDF models are described in the supplementary material.

0

1

2

3

4

5

6

7

8

9

10

0 20 40 60 80 100 120 140 160 180 200 220 240

Cook-Torrance

Ward

Lafortune

Oren-Nayer

Ashikhmin-Shirley

Average

Ave. no. of nonshadowed pixels

Mea

n A

ng

ula

r E

rro

rs (

deg

ree)

Figure 3. Experimental results of dataset (A) with a varying num-ber of images. For a fair evaluation, we display the average num-ber of nonshadowed pixels that join our algorithm on the x-axisinstead of the number of images.

0

1

2

3

4

5

6

7

8

CookTorrance Ward Lafortune OrenNayer AshikhminShirley

Ours

Shi. et al [17] (0)

Shi. et al [17] (0.05)

Shi. et al [17] (0.1)

Shi. et al [17] (0.15)

Shi. et al [17] (0.20)

Shi. et al [17] (0.25)

Shi. et al [17] (0.30)

Mea

n A

ngula

r E

rro

rs (

deg

ree)

BRDF

Figure 4. Comparison with [17]. The values in parentheses arethe radian values added to the true azimuth angle map. The noisyazimuth maps were used in [17]. The sign was chosen randomly.

0

0.005

0.01

0.015

0.02

0.025

0.03

0.035

0.04

0 5 10 15 20 25

Ny=1 (m=100)

Ny=2 (m=100)

Ny=3 (m=100)

Ny=1 (m=300)

Ny=2 (m=300)

Ny=3 (m=300)

Co

mp

uta

tion

tim

e/pix

el (

sec)

𝑁𝑦 + 1 × (𝑁𝑧 + 1)

Figure 5. Evaluation of computational time. We illustrate the per-pixel computational time for each combination of the number ofBernstein basis functions and lightings.

our method with the standard Lambertian least-squares-regression-based approach [20] (LS) and a recent Lamber-tian sparse-regression-based approach implemented withthe sparse Bayesian learning [11] (SBL) (λ is fixed by

Page 6: Photometric Stereo using Constrained Bivariate Regression ... · The unknown normal vec-tor was separated from the unknown reflectance function by considering the inverse representation

0

0.5

1

1.5

2

2.5

3

3.5

10 20 30 40 50 60 70 80 90 100

LS BQ

SBL Ours

0

0.5

1

1.5

2

2.5

3

3.5

4

4.5

10 20 30 40 50 60 70 80 90 100

0

1

2

3

4

5

6

7

8

10 20 30 40 50 60 70 80 90 100

0

0.5

1

1.5

2

2.5

3

3.5

4

10 20 30 40 50 60 70 80 90 100

0

1

2

3

4

5

6

7

8

9

10 20 30 40 50 60 70 80 90 100

0

1

2

3

4

5

10 20 30 40 50 60 70 80 90 100

Cook-Torrance Ward

Lafortune Oren-Nayar

Ashikhmin-Shirley

Average over 5 BRDF

Tlow (%) Tlow (%) Tlow (%)

Tlow (%) Tlow (%) Tlow (%)

Mea

n A

ng

ula

r E

rro

rs (

deg

ree)

M

ean

An

gu

lar

Err

ors

(d

egre

e)

Mea

n A

ng

ula

r E

rro

rs (

deg

ree)

M

ean

An

gu

lar

Err

ors

(d

egre

e)

Mea

n A

ng

ula

r E

rro

rs (

deg

ree)

M

ean

An

gu

lar

Err

ors

(d

egre

e)

Figure 2. Experimental results of dataset (A) with five different BRDFs with varying frequencies in observations.

10−6). Our method was also compared with a recent para-metric non-Lambertian photometric stereo method with thebiquadratic reflectance model [16] (BQ). In this experiment,we fixed (Ny , Nz) in Eq. (11) by (1, 5) to examine the ro-bustness of our method against these parameters.Evaluation with varying frequenciesWe first evaluated our algorithm using observations withvarying frequency to present the flexibility of our modelin comparison with models used in previous works (e.g.,BQ[16]) that only work for low-frequency reflectance val-ues. Here high-frequency specularities were discarded byusing the nonshadowed pixels with intensities that wereranked below the Tlow% (10 ≤ Tlow ≤ 100).

The result is illustrated in Fig. 2. Overall, we observedthat our method outperformed other algorithms for almostall frequencies for the Lafortune, Ashikhmin–Shirley, andOren–Nayar models, and performed competitively withSBL for the Cook–Torrance and Ward models. Interest-ingly, because of our inverse reflectance model and the sim-ple lTv constraint selection strategy, our method works wellfor the Oren–Nayar model which exhibits strong retroreflec-tive reflections, and it also works well when all the frequen-cies are included unlike other methods4. Unfortunately,however, our method did not work when the number of im-ages was very small, as will be discussed below.Valid number of input imagesWe also evaluated our algorithm using a varying number ofimages to find the valid number required for effective recov-ery. The results are displayed in Fig. 3. We observed thatthe minimum number of images required to make the algo-

4The result of BQ is similar as LS since the optimization procedure ofBQ is nonlinear and strongly affected by the initial estimation by LS.

rithm work was around 20 and more than 60 were requiredfor stable reconstruction because our method could sufferfrom the over-fitting when the number of the input imageswas very small.Comparison with another monotonicity-based approachIn this section, we compared our method with the recentelevation-angle estimation algorithm [17] assuming that thedominant reflectance lobe was pointing at the half-vectordirection. We generated a ground truth azimuth anglemap and azimuth angle maps with some radian errors be-cause [17] requires an azimuth angle map as input whereasour method simultaneously recovers all elements in the nor-mal. The radian errors varied from 0.05 to 0.3 were addedto the ground truth azimuth angle values, where the sign ofthe error was chosen randomly.

The results are illustrated in Fig. 4. We observed thatwhen true azimuth angles were given, [17] outperformedour method in the Cook–Torrance and Ward datasets. How-ever, as the amount of errors increased, the differences be-came smaller and finally our method outperformed [17].As we expected, [17] did not work for the Lafortune andAshikhmin–Shirley models because these models violatetheir assumption.Evaluation of computational timeHere, we examined the computational time required for ourcomputation. We tried various combinations of Ny andNz and m in Eq. (9) and solved our optimization prob-lem Eq. (11) using the lsqlin function in MATLAB.

The evaluation results are illustrated in Fig. 5. Here wepresent a per-pixel computational time to solve one opti-mization problem. Therefore, the actual computational timewas twice that in the figure as we applied our algorithm

Page 7: Photometric Stereo using Constrained Bivariate Regression ... · The unknown normal vec-tor was separated from the unknown reflectance function by considering the inverse representation

Mea

n A

ngu

lar

Err

ors

(deg

ree)

0

5

10

15

20

25

30

35

40

LS SBL

BQ (Tlow=100) BQ (Tlow=25)

Ours w/o retro-reflection detection Ours w/ retro-reflection detection

yel

low

-mat

te-p

last

ic

wh

ite-

acry

lic

wh

ite-

dif

fuse

-bb

all

vio

let-

rub

ber

sp

ecu

lar-

ora

nge-

phen

oli

c w

hit

e-m

arb

le

dar

k-s

pec

ula

r-fa

bri

c sp

ecu

lar-

mar

oon

-phen

oli

c sp

ecu

lar-

red

-ph

enoli

c ip

swic

h-p

ine-

221

re

d-s

pec

ula

r-pla

stic

m

aroon

-pla

stic

sp

ecu

lar-

yel

low

-ph

enoli

c gra

y-p

last

ic

silv

er-p

ain

t p

ink

-pla

stic

w

hit

e-p

ain

t b

lue-

rub

ber

p

earl

-pai

nt

red

-fab

ric

silv

er-m

etal

lic-

pai

nt2

sp

ecu

lar-

vio

let-

ph

enoli

c al

um

ina-o

xid

e gold

-pai

nt

pu

rple

-pai

nt

pin

k-j

asper

yel

low

-pla

stic

yel

low

-ph

enoli

c gre

en-p

last

ic

red

-ph

enoli

c ora

nge-

pai

nt

yel

low

-pai

nt

tefl

on

sp

ecu

lar-

wh

ite-

phen

oli

c re

d-p

last

ic

ligh

t-re

d-p

ain

t gre

en-a

cryli

c p

ink

-fab

ric

aven

turn

ine

del

rin

w

hit

e-fa

bri

c d

ark

-red

-pai

nt

pvc

pin

k-f

elt

nylo

n

spec

ula

r-gre

en-p

hen

oli

c gold

-met

alli

c-p

ain

t2

poly

eth

yle

ne

vio

let-

acry

lic

pu

re-r

ub

ber

tw

o-l

ayer

-gold

re

d-f

abri

c2

gold

-met

alli

c-p

ain

t3

silv

er-m

etal

lic-

pai

nt

neo

pre

ne-

rub

ber

p

ink

-fab

ric2

gre

en-m

etal

lic-

pai

nt

blu

e-ac

ryli

c co

lor-

chan

gin

g-p

ain

t1

alu

m-b

ron

ze

gold

-met

alli

c-p

ain

t ch

rom

e co

lor-

chan

gin

g-p

ain

t3

spec

ial-

wal

nu

t-2

24

bla

ck-p

hen

oli

c tw

o-l

ayer

-sil

ver

sp

ecu

lar-

bla

ck-p

hen

oli

c si

lico

n-n

itra

de

spec

ula

r-b

lue-

ph

enoli

c re

d-m

etal

lic-

pai

nt

hem

atit

e n

ick

el

blu

e-fa

bri

c st

eel

colo

r-ch

angin

g-p

ain

t2

gre

en-l

atex

b

eige-

fab

ric

bla

ck-o

xid

ized

-ste

el

tun

gst

en-c

arbid

e ch

rom

e-st

eel

gre

en-m

etal

lic-

pai

nt2

fr

uit

wood

-24

1

blu

e-m

etal

lic-

pai

nt

blu

e-m

etal

lic-

pai

nt2

gre

en-f

abri

c d

ark

-blu

e-pai

nt

gre

ase-

cover

ed-s

teel

al

um

iniu

m

bla

ck-o

bsi

dia

n

poly

ure

than

e-fo

am

bla

ck-f

abri

c b

lack

-soft

-pla

stic

co

lon

ial-

map

le-2

23

n

atu

ral-

209

ss4

40

ch

erry

-23

5

bra

ss

wh

ite-

fab

ric2

p

ick

led

-oak

-260

ligh

t-b

row

n-f

abri

c

Figure 6. Comparison between different methods with dataset (B). Results are aligned in ascending order of the mean angular error of ourmethod with the retroreflection detection algorithm.

twice to distinguish retroreflective materials. We observedthat the computational complexity depends on the numberof basis functions rather than the number of lightings.

4.2. Evaluation with Measured BRDF

Here we evaluated the performance of our method tothe dataset (B). We generated images under 300 differentlightings for 100 different materials from the MERL BRDFdatabase [14]. In this experiment, our method was alsocompared with LS [20], SBL [11], and BQ [16]. LS,SBL, and our method used all the frequencies in observa-tions (i.e., Tlow = 100) whereas only BQ was performedwith both Tlow = 25 and Tlow = 100 as that model wasoriginally designed to represent the low-frequency observa-tions. In this experiment, we fixed Ny = 3 and Nz = 5and performed our method with/without the retroreflectiondetection algorithm described in Section 3 to verify the ef-fectiveness of this process.

The results are illustrated in Fig. 6. We observed thatour method with our efficient retroreflection detection out-performed other algorithms for most of the materials. BQ(Tlow = 25) is more effective for some materials, but weemphasize that our method was capable of handling all fre-quencies in observations because of our flexible reflectancemodel, while BQ (Tlow = 100) does not work for mostmaterials. We also observed that the angular errors ofour method without retroreflection detection were relativelylarger than that of our method with retroreflection detectionfor materials exhibiting strong retroreflections (e.g., MERLfabrics), which indicated that our retroreflection detectionalgorithm worked quite well for those materials. Finally, the

average angular errors over 100 materials were 12.5 (LS),6.2 (SBL), 13.1 (BQ, Tlow = 100), 1.7 (BQ, Tlow = 25),2.4 (our method without retroreflection detection) and 1.2(our method with retroreflection detection), respectively.

4.3. Qualitative Evaluation with Real Images

We also evaluated our algorithm using real images: (1)100 images of two-face (this dataset is from [21]), (2) 100images of doraemon, and (3) 44 images of fatguy (these twodatasets are from [11]). In this experiment, we comparedour algorithm with LS [20], SBL [11] (λwas fixed by 10−1)and BQ [16] (for Tlow = 25 and 100). Note that our methodused a retro-reflective detection scheme and we fixed (Ny ,Nz) by (1, 5). The threshold of the shadow removal for eachdataset was chosen manually but fixed over algorithms.

The experimental results are illustrated in Fig. 7. Herewe show both recovered surface normal maps and surfacemeshes reconstructed by a poisson solver [1] with a fixedscale. We observe that our method succeeded to estimatesmoother and more reasonable normal maps and surfacemeshes. We also observe that BQ (Tlow = 25) workedpoor for those datasets since shadows could not be com-pletely removed by a simple thresholding therefore the low-frequency component in the observation was not reliable.In contrast to that, our method performed well since ourmethod could account all observations without discardingthe informative high-frequency component.

5. ConclusionIn this paper, we have proposed the constrained bivari-

ate regression-based photometric stereo, which worked for

Page 8: Photometric Stereo using Constrained Bivariate Regression ... · The unknown normal vec-tor was separated from the unknown reflectance function by considering the inverse representation

(a) (b) (c) (d) (e) (f) (g) (h) (i) (j) (k)

LS

(𝑇𝑙𝑜𝑤 = 100)

SBL

(𝑇𝑙𝑜𝑤 = 100)

BQ

(𝑇𝑙𝑜𝑤 = 100)

BQ

(𝑇𝑙𝑜𝑤 = 25)

Ours

(w/ retro-reflection detection) Input

Figure 7. Experimental results using real data (two-face, doraemon and fatguy). We illustrate (a) example images of target objects, andnormal maps recovered by (b) LS, (d) SBL, (f) BQ (Tlow = 100), (h) BQ (Tlow = 25) and (j) Ours (with a retro-reflective detection). Wealso show surface meshes generated from normal maps in (c), (e), (g), (i) and (k).

various kinds of isotropic surfaces by exploiting variousconditions shared among physically valid BRDFs. Our de-tailed experimental results have shown the state-of-the-artperformance of our method for both synthetic and real data.The current limitation is that the proposed method enforcesa global monotonic constraint with regard to lTv, whichmight not be true for surfaces with both the retro and spec-ular reflections though they are rarely observed in the realworld. Another limitation is that we assumed that shadowswere discarded from images in advance, and this might beimpractical in real scenes. To ease this condition, we areinterested in incorporating the data cleansing scheme in asimilar manner as [11].

References

[1] A. Agrawal, R. Raskar, and R. Chellappa. What is the rangeof surface reconstructions from a gradient field ? In Proc.ECCV, 2006. 7

[2] N. Alldrin, T. Zickler, and D. Kriegman. Photometric stereowith non-parametric and spatially-varying reflectance. InProc. CVPR, 2008. 1, 2, 3

[3] M. Ashikhmin and P. Shirley. An an isotropic phong brdfmodel. Journal on Graphics Tools, 5(2):25–32, 2000. 2, 5

[4] K. Bollaerts, P. H. Eilers, and I. Mechelen. Simple and mul-tiple p-splines regression with shape constraints. British J.Math. Statist. Psych., 59(2):451–469, 2006. 3, 4

[5] P. Chak. Semi-nonparametric estimation with bernstein poly-nomials. Economics Letters, 89(2):153–156, 2005. 3

[6] M. Chandraker and R. Ramamoorthi. What an image revealsabout material reflectance. In Proc. ICCV, 2011. 1, 2

[7] R. Cook and K. Torrance. A reflectance model for computergraphics. ACM Trans. on Graph., 15(4):307–316, 1981. 2, 5

[8] P. Favaro and T. Papadhimitri. A closed-form solution touncalibrated photometric stereo via diffuse maxima. In Proc.CVPR, 2012. 2

[9] D. Goldman, B. Curless, A. Hertzmann, and S. Seitz. Shapeand spatially-varying brdfs from photometric stereo. In Proc.ICCV, October 2005. 1

[10] T. Higo, Y. Matsushita, and K. Ikeuchi. Consensus photo-metric stereo. In Proc. CVPR, 2010. 1, 2

[11] S. Ikehata, D. Wipf, Y. Matsushita, and K. Aizawa. Robustphotometric stereo using sparse regression. In Proc. CVPR,2012. 1, 5, 7, 8

[12] E. Lafortune, S.-C. Foo, K. Torrance, and D. Greenberg.Non-linear approximation of reflectance functions. In Proc.ACM SIGGRAPH, 1997. 2, 5

[13] G. G. Lorentz. Bernstein Polynomials. Chelsea PublishingCompany, New York, 1986. 1, 3

[14] W. Matusik, H. Pfister, M. Brand, and L. McMillan. A data-driven reflectance model. ACM Trans. on Graph., 22(3):759–769, 2003. 5, 7

[15] M. Oren and S. Nayar. Generalization of lambert’s re-flectance model. In In Proc. of the 21st annual conferenceon Computer graphics and interactive tecniques, 1994. 2, 5

[16] B. Shi, P. Tan, Y. Matsushita, and K. Ikeuchi. A biquadraticreflectance model for radiometric image analysis. In Proc.CVPR, 2012. 1, 5, 6, 7

[17] B. Shi, P. Tan, Y. Matsushita, and K. Ikeuchi. Elevation anglefrom reflectance monotonicity. In Proc. ECCV, 2012. 1, 2,5, 6

[18] J. Wang. Shape restricted nonparametric regression withBernstein polynomials. PhD thesis, North Carolina StateUniversity, 2012. 3, 4

[19] G. Ward. Measuring and modeling anisotropic reflection.Computer Graphics, 26(2):265–272, 1992. 2, 5

[20] P. Woodham. Photometric method for determining surfaceorientation from multiple images. Opt. Engg, 19(1):139–144, 1980. 1, 5, 7

[21] L. Wu, A. Ganesh, B. Shi, Y. Matsushita, Y. Wang, andY. Ma. Robust photometric stereo via low-rank matrix com-pletion and recovery. In Proc. ACCV, 2010. 1, 7