Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Eigensurfaces of eigenmirrors

Open Access Open Access

Abstract

Typically, if an observer gazes at a curved reflector, the objects in it will appear to be distorted. We show here that for some mirrors there exist surfaces that do not appear distorted when viewed from a prescribed location. We call such mirrors eigenmirrors and the surfaces eigensurfaces. We first give an analysis of the rotationally symmetric case and verify our work with simulations. In the general three-dimensional (3D) case, if the mirror is given, then one does not expect an eigensurface to exist. On the other hand, if we are given two viewpoints and a correspondence between the ray bundles emanating from each point, and we treat both the eigenmirror and the eigensurface as unknowns, then the problem reduces to solving a first-order nonlinear partial differential equation. We derive this partial differential equation in the 3D case and examine one example in detail.

© 2019 Optical Society of America

1. INTRODUCTION

If an observer gazes at a curved reflector, such as a mirrored sphere, then the scene appears to be distorted, as in Fig. 1. We ask: are there any surfaces that do not appear distorted in a given reflector? If so, then such surfaces pass unchanged through the optical system, and so we refer to them as eigensurfaces, and the corresponding reflectors as eigenmirrors.

 figure: Fig. 1.

Fig. 1. Distortion of a checkerboard pattern in a spherical mirror. This is a panoramic view of a test scene, with the camera appearing in the center of the photo. The checkers are 0.2159 m square, and the mirror, of radius 0.031 m, is 0.34 m above the floor. Two squares are labeled: square (1) is 0.10795 m from the camera and square (2) is 0.4572 m away, yet square (1) consists of disproportionately many more pixels. For the robot soccer application, we sought greater equity between the two.

Download Full Size | PDF

The motivation for this work is an aspherical reflector designed for vision-based control of soccer robots, as depicted in Fig. 2. In Ref. [1], the authors investigated the use of convex mirrors for endowing robots with panoramic vision, thus simplifying the control component of the problem. In fact, Fig. 1 depicts an image taken from such experiments. Here the spherical mirror is mounted above the sensor (no robot appears in this test case), which is pointed skyward. The checkerboard pattern surrounds the camera on the ground plane. The lens of the camera is clearly visible in the center of the image. The spherical mirrors were low cost and worked moderately well for this application, giving a panoramic view of the horizon, and even capturing some fluorescent lights on the ceiling, but the distortion was somewhat problematic. Each white square appearing in Fig. 1 was an ANSII letter-sized sheet of paper cut square, and so was 0.2159×0.2159 m. Hence, it is easy to see that close objects, such as the square labeled (1), which is 0.10795 m from the lens, are allocated many more pixels then a square such as (2), which is a mere 0.6096 m from the lens. Additionally, valuable pixels are wasted on the camera lens itself. The response to these problems was to design a convex, rotationally symmetric mirror so that when mounted above the robot and viewed along the axis of rotation, the ground plane appeared undistorted. Thus, the squares should look like squares and each should be allocated an equal number of pixels. The cross section of such a reflector, which we refer to as a rectifying mirror, appears in Fig. 3. An image of a checkerboard plane reflected in a rectifying mirror appears in Fig. 4, where the field of view is 360°×145°. Thus, we have an example of a curved reflector that does not distort everything viewed in it. The flat plane, at an appropriate distance, is an eigensurface of the rectifying mirror. To design this mirror, the authors derive a first-order differential equation for the cross section, using the fact that the slope of the mirror at each point is prescribed.

 figure: Fig. 2.

Fig. 2. Soccer robots at the UPenn GRASP Lab, circa 1997. Each robot was equipped with a spherical mirror to allow for panoramic views.

Download Full Size | PDF

 figure: Fig. 3.

Fig. 3. Cross section of a rotationally symmetric mirror designed to give a soccer robot a 360°×145° view of the playing field. This reflector has plane as its eigensurface. At the lowest point, the radius of curvature is 1.7 m.

Download Full Size | PDF

 figure: Fig. 4.

Fig. 4. Case of the rectifying mirror. This curved reflector was designed to be mounted on a soccer playing robot, 34 cm above the playing field, which here contains a checkerboard pattern.

Download Full Size | PDF

Given the above, one may ask the dual question: starting with a reflector, are there textured shapes that appear undistorted in the reflector? To address this question we must be careful about what our assumptions and definitions are.

We should point out that we view the problem under consideration as a foundational question. To keep things tractable, we work within the framework of geometric optics and assume a monocular observer that is ideal pinhole camera. Thus, for us, images are formed ultimately by a system of perfect perspective projection, and so the only aberration under consideration is distortion.

In the original robot soccer application, the control of the robots was based on the locations of objects of interest on the field, which, from the vision standpoint, were all color blobs. Focus was thus not a major concern, and indeed was not part of our model of the rectifying mirror. Hence, the blur that is evident in Fig. 4 was not a concern. Distortion was our concern mainly because of the pixel count problem mentioned above. Therefore, throughout, the reader should imagine holding up a surface such as a checkerboard, coffee mug, etc., to a chromed water faucet and observe the reflection.

Note that all of our examples are rotationally symmetric, but for our general problem statement no such assumption is needed. If a surface is not a portion of a surface of revolution, then we say that the surface is freeform. Working without any symmetry assumption typically makes a problem considerably more complex, but also adds degrees of freedom. For example, one expects that a freeform surface be modeled by a partial differential equation, rather than a more friendly differential equation. Our work is heavily informed by freeform optics, a topic that has recently seen explosive growth, and so some discussion of the development of freeform optics is warranted.

An early example of a freeform surface in optical design (possibly the first) is the 1959 progressive spectacle design by Kanolt [2]. A well-known example that was remarkable not only for its novelty of design but also for its commercial success, was the Polaroid SX-70 collapsible camera [35]. Stone and Forbes’ prescient 1992 work considered the Hamiltonian optics of freeform surfaces [6,7]. Thompson soon after began a systematic analysis of aberrations in freeform systems with a mind towards astronomical instrumentation [8]. More recent work on aberrations includes the work by Bauer and Rolland [9].

Illumination systems are a natural application area for freeform systems. Early work using aspheres in laser beam shaping in the 1970s was certainly in the freeform spirit [1012]. Classic references on illumination are Elmer [13] (which exhibits elements of freeform thinking), Welford and Winston [14], and Winston, Miñano, and Benítez [15]. The simultaneous multiple surface design method in 3D is described in Ref. [16]; see also [17] for applications.

The prescribed intensity problem for a single source, which is to illuminate a surface according to a prescribed intensity, was solved in the 1990s by Oliker et al. in Refs. [18,19]. The solution makes use of optimal transport and Monge–Ampère equations. These rich ideas have led to a considerable amount of work, such as [2023]. Source-target maps are also an approach to the problem, such as those described by Fournier, Cassarly, and Rolland [24].

The prescribed projection problem is to redirect a two-parameter bundle of rays to a target surface using a single reflector, is formulated by Hicks and Perline in Ref. [25]. With a single reflector, this problem is ill-posed in the sense of Hadamard. Croke and Hicks introduced the tool of exterior differential systems for the design of two or four freeform reflector systems [26,27]. Related work by Plakhov, Tabachnikov, and Treschev investigates the number of reflections needed to achieve an arbitrary transformation [28]. Other work on bundles and perfect imaging includes [2931].

The reader should bear in mind that the above only scratches the surface of the subject of freeform optics, and reflects the authors’ interest in the mathematical foundations of the subject. The total number of papers on applications of freeform optics far exceeds the above.

Returning to the problem at hand, we need to give a more precise meaning to the term “distortion.” Here we use it for whenever an image is not geometrically similar to one formed by a perspective (or orthographic) projection. In particular, in the case of undistorted image formation, straight lines in the scene map to straight lines in an image plane. If that is not the case, then one might imagine that some nonlinear transformation is at work, although the details of such a transformation (domain, range, etc.) are not a priori clear. Nevertheless, the association of nonlinear transformations with distortions is reasonably natural, given our everyday experience with the curved reflectors that surround us.

As we saw in the case of the rectifying mirror, even though a reflector is curved, one cannot immediately rule out the possibility that some surfaces (with recognizable texture) could appear unchanged when viewed in the reflector. (Weaker examples are abundant. For example, all rotationally symmetric systems image lines through the axis of rotation to lines in the image plane, be they curved or not.) Are there rotationally symmetric mirrors that perfectly image other surfaces?

The answer is “yes.” Our first result is that almost every rotationally symmetric reflector has an eigensurface when viewed along its axis of rotation. (At this point we should point out that in this paper we will ignore the issue of obstruction.) That is, there are surfaces that remain invariant, in a manner to be made precise, under the above-mentioned nonlinear transformation. This means that the observer gazing at the mirror sees a scaled, flipped, rotated, etc., version of what an observer in another fixed position sees.

Of course, the idea of having an optical system produce a rectified image is fundamental in imaging optics. The idea of having a single lens or mirror perform a prescribed transformation on some ray bundle could be said to have originated in the study of conics in ancient times, with more modern examples including the theory of perfect optical instruments, such as Cartesian ovals and more exotic examples such as the Luneberg lens, Maxwell’s fisheye, etc. [3234].

Below, we will compute the cross sections of rotationally symmetric eigenmirrors and eigensurfaces using differential equations, and in the general 3D case, we will model the problem using a first-order nonlinear partial differential equation. Other rotationally symmetric work of Hicks and Perline include [35], which describes reflectors for panoramic imaging that provide uniform resolution in the sense that equal solid angles are mapped to equal numbers of sensor pixels. These surfaces, in the far-field case, are shown to be rotationally symmetric surfaces of constant Gaussian curvature. Later, Roitman showed that all constant Gauss curvature surfaces have this property [36]. Hicks and Coletta derive and classify rotationally symmetric equicylindrical reflectors in Ref. [37], i.e., reflectors that give uniform resolution with respect to a cylindrical unwarping transformation in software.

2. PURELY PLANAR CASE

The problem statement alone is tricky, so we will begin with the purely planar case. Assume that we have a curve given by (some portion of) the graph of f(u), which is reflective, and two points p1 and p2 in the plane, which can be thought of as observers or sources or centers of projection; see Fig. 5. We treat p1 as the observer looking directly at the reflector (eigenmirror) f, and p2 as an observer looking at the eigensurface. Take B to be the bundle of rays that emanate from p1 and then strike the graph of f. Part of the data of the problem will be that we are given a transformation T between B and the rays that emanate from p2. It is important to get clear here some terminology and notation regarding T and these two bundles.

When working with bundles, it is of course necessary to have some sort of coordinates on them. To describe a ray sometimes we will give its footpoint (source) and the direction of the ray. Another method is to choose a reference plane that intersects the bundle in such a way that each ray crosses the reference plane at a unique point.

Therefore, given T and a ray r in B, while it is technically correct to write T(r), which we will sometimes denote as r˜, we will abuse notation sometimes and write T(V), where V is the unit direction of r. Note that T does not give a correspondence between point on r and T(r) for a given r.

It is useful to write S1(p), the unit circle at p. In the 3D case, we write the unit sphere at p as S2(p). Let r be a ray in B from p1 through [u,f(u)], i.e., it has direction

V=[u,f(u)]p1[u,f(u)]p1.
Observe that V=V(u).

This allows us to identify B with a subset of the unit sphere S1(p1). One may think of T as prescribing a certain view. The rays at p2 that correspond to those in B we will call B. In other words, B=T(B).

For the reflector f to have an “eigencurve,” the following must hold: if r is traced to f, and reflects in the direction

Out=Out(f(u),f(u),u),
along the ray r¯, then r¯ must intersect with the ray r˜ given by sp2+sT(V). (Note that T(V) depends on u and f(u).) In other words, for a solution to the following equation to exist, we must have that for each choice of u the equations
[u,f(u)]+tOut=p2+sT(V)
have a solution in s, t. We will refer to equation Eq. (3) as the fundamental equation. In three dimensions it has a very similar form.

In 2D, the fundamental equation generically does have a solution for a given u. That is, if we come to the problem thinking that the reflector f is given, then the system Eq. (3) will most likely have a solution for a given u. When s and t are solved for, they will depend on u, and so the eigencurve, Ec, can be described parametrically using t(u) as

Ec(u)=[u,f(u)]+t(u)Out(f(u),f(u),u),
or, equivalently, using s(u) gives
Ec(u)=p2+s(u)T(V(f(u),u)).
If, for example, one takes p1=[0,0], p2=[0,1], and T(w,z)=[aw+bz,cw+dz], we may write this out explicitly as
Ec(u)=1δ[2f(u)2bu+(2u2aub)f(u)u2a)f(u)2+(2bf(u)32au+2bf(u)2+(2u2b+2au)f(u)+2u3a)f(u)2f(u)2bu+(2u2a+ub)f(u)+u2a,((2du+b)f(u)2+(2cu2+au)f(u))f(u)2+(2f(u)3d2f(u)2cu+(2du2+2ub)f(u)+2cu3+2u2a)f(u)+(2dub)f(u)2+(2cu2au)f(u)],
where
δ=(f(u)2b+(au+du)f(u)+cu2)f(u)2+(2df(u))2+(2ub2cu)f(u)+2u2a)f(u)f(u)2b+(audu)f(u)cu2.
As an example, suppose f(u)=u3, with u[3/5,1], as in Fig. 6. We choose p1=[0,1/3], p2=[2,3/2], and T([X1,X2])=R([X2,X1]), where R is a rotation by 190° counterclockwise, i.e.,
R=[cos190°sin190°sin190°cos190°].
The rays in Fig. 6 are a geometrically accurate depiction of the bundle B. The action of T is then to flip a ray about the horizontal axis and rotate it by 190° counterclockwise, i.e., T([X1,X2])=R([X1,X2]).

 figure: Fig. 5.

Fig. 5. Problem in 2D, with no assumed symmetry.

Download Full Size | PDF

 figure: Fig. 6.

Fig. 6. Sample 2D problem. The eigenmirror is y=u3, with u[3/5,1], p1=[0,1/3], p2=[2,3/2], and T([X1,X2])=R([X1,X2]), where R is the rotation matrix given in Eq. (8).

Download Full Size | PDF

3. ROTATIONALLY SYMMETRIC EXAMPLE

Suppose one has a rotationally symmetric mirror that is to be viewed along its axis of symmetry. In that case, our schematic now has the form depicted in Fig. 7, which depicts a cross section of the mirror. As earlier, we have viewpoints p1 and p2, and a transformation T.

 figure: Fig. 7.

Fig. 7. 2D schematic of the cross section of a rotationally symmetric eigenmirror with eigensurface. One observer at p1 gazes at the mirror and the other, p2, looks directly at the eigensurface. Observe that what we have here is not a two reflector problem—there is no reflective relation between the ray r¯ and T(r) at their point of intersection on the eigensurface.

Download Full Size | PDF

Obviously, a parabolic mirror would be natural to analyze, so consider the case where the mirror cross section is given by the graph of

f(u)=u22+3,
with p1=[0,0] looking in the positive y direction and p2=[0,20] looking in the negative y direction.

Our association T will be between the ray r through [0, 0] of slope m and the ray r˜ through [0, 20] of slope m, i.e., T([V1,V2])=[V1,V2]. Taking In=p1[u,f(u)]=[u,1/2u23], gives

Out=reflect(In,n)=[7uu2+1,u4+9u262u2+2],
where reflect(In,n) means, of course, to reflect In about n. Thus, the ray r¯ is given parametrically by
t[u+7utu2+1,u22+3+u4t2u2+2+9u2t2u2+23tu2+1].
Equating and solving for s and t gives
[s=u4+2u2+232u4+16u2+36,t=2u413u214u4+16u2+36].
Substituting back into either r˜ or r¯ gives the parametric equations for the eigencurve (surface),
u[(u4+2u2+232)uu4+16u2+36,u6/216u4198u224u4+16u2+36].
We chose to analyze the parabolic mirror because it is a fundamental object in optics, and it would be worth knowing whether its eigensurface was some other familiar surface. Thus, the reader may wonder if there is some simplification of equations [Eq. (10)] that gives a familiar curve or pair of equations. Taking u=τ gives
τ[(τ2+2τ+232)ττ2+16τ+36,τ3/216τ2198τ24τ2+16τ+36].
Thus, putting aside the square root term, the equations are rational functions of relatively low degree, although the geometric meaning of the eigensurface in this case remains unclear.

In Fig. 8, we see a cross section the (black) parabolic eigenmirror and (red) eigensurface.

 figure: Fig. 8.

Fig. 8. System for which the reflective surface is a portion of the parabola f(u)=12u2+3. The first viewpoint, p1=[0,0], observes the (black) parabola from below, via the dashed ray. The second viewpoint, p2=[0,20] views that (red) eigensurface from above, via the dotted ray.

Download Full Size | PDF

Using the POV-Ray ray-tracer, we may simulate how an observer would see this pair, and compare the eigensurface with a similarly textured surface. A side view of the configuration appears in Fig. 9(A). Here the pair of surfaces lies in the center of a cubical test room, with checkerboard patterned walls. In particular, the floor is green and white, and the ceiling is green and black. A flat disk with a sandy texture has been placed on the back on the parabolic mirror. From above, at [0,0,20]=p2, an observer looks down and sees the sandy back of the parabolic mirror and the rings in the eigensurface. (It just so happens that they appear fairly evenly spaced.)

 figure: Fig. 9.

Fig. 9. (A) Side view of a parabolic mirror (upper) and the corresponding eigensurface (below). The entire scene lies within a room with a cubical room with checkered walls. (B) From above, at [0,0,20], an observer looking down sees the sandy back of the parabolic mirror and the rings in the eigensurface. (C) The view from p1 of the parabola. The proof of concept is that the rings look the same as in (B). (D) Here we see a flat disk reflected in the parabola. It is not, of course, an eigensurface of the parabola, and so the rings do not appear evenly spaced, as in (B) and (C).

Download Full Size | PDF

If we remove the eigensurface and replace it with a flat disk with a ringed pattern, we can view it reflected in the parabolic mirror and see that it is distorted in comparison to the above view in Fig. 9(B), i.e., it is not an eigensurface. The crucial comparison is between Fig. 9(C) and this view, where the observer’s eye p1 is at the origin, beneath the parabolic mirror. The result is that the rings are undistorted and look identical to those in Fig. 9(B). That is, we have found the eigensurface. To underscore the specialness of the eigensurface, in Fig. 9(D) we view from p1 a flat disk with evenly spaced rings, reflected in the parabola. Here we see that the rings are distorted, i.e., a flat disk is not, of course, an eigensurface of the parabola.

4. NONEXISTENCE OF AN EIGENSURFACE IN THE OFF-AXIS CASE

We have shown that an eigensurface can be constructed for any rotationally symmetric mirror with viewpoints both along the axis of symmetry. Let us now consider the case of a rotationally symmetric mirror with only p1 on the axis of symmetry. Note that in our above rotationally symmetric discussion, with the exception of a few degenerate choices for viewpoints on the axis, an eigenmirror always exists. Here we show that if one takes the viewpoint p2 to be off the axis of rotation, then the eigensurface can fail to exist.

Again, it is natural to consider a quadratic mirror, so we consider a rotationally symmetric mirror given by the parabolic surface z=u2+v2c, with c>0. Let p1=[0,0,0] and take the reference plane z=1 to label rays that emanate from p1. Let p2=[0,a,b], with a,b>0, and use the reference plane y=a+1 for labeling rays.

Our computations here will label the ray bundles, not with unit directions, but with points in chosen reference planes. That is, a plane will be associated with each viewpoint (essentially an image plane), and if a ray through the viewpoint strikes that plane at a point, then we will consider that point to be the label (or coordinates) of the ray. This makes writing T down in those coordinates easy.

Let the transformation from the bundle B at p1 to the bundle at p2 be given by T(u,v,1)=[u,a+1,bv]. Given an arbitrary point q1=[u,v,u2+v2c] on the mirror, we can calculate the line to it from p1. This line intersects the reference plane of p1 at

[uu2+v2c,vu2+v2c,1].
The corresponding point on the reference plane of p2 is
T[uu2+v2c,vu2+v2c,1]=[uu2+v2c,a+1,b+vu2+v2c].
If we reflect the incoming ray from p1 about the normal vector (2u,2v,1), we get the outgoing ray
r¯=[utu(4c1)(4u2+4v2+1)u2+v2+(u2+v2c)2,vtv(4c1)(4u2+4v2+1)u2+v2+(u2+v2c)2,u2+v2c+t·(u2+v2c)(4u2+4v2+1)+2(u2+v2+c)(4u2+4v2+1)u2+v2+(u2+v2c)2],
where t is a parameter. On the other hand, the ray r˜ is given by
r˜=[sxu2+v2c,sa,b+svu2+v2c].
In order for an eigensurface to exist, r¯ and r˜ must intersect. Otherwise, the viewpoints would not see the same image. Setting the components of r¯ and r˜ equal to each other is a system of three equations and two unknowns that does not yield a solution. Even allowing freedom of constants a, b, and c is not enough to solve the system. Thus, there is no eigensurface for the parabolic surface with these viewpoints.

5. ROTATIONALLY SYMMETRIC OFF-AXIS EXAMPLE

Next, we consider the case where the eigenmirror is unknown, but may be parametrized in terms of z and θ as

x(z,θ)=[ρ(z)cos(θ),ρ(z)sin(θ),z],
and p1=[0,0,0] and p2=[0,4,4]. We will show that no profile curve ρ exists for which an eigensurface exists.

Thus, our axis of symmetry is the z axis. We will index rays in B with the reference plane z=1, and index rays in B with the plane y=3. Thus, rays in B are labeled by points of the form [α,β,1], and rays in B are labeled by points of the form [α,1,β]. For T:BB, we take T([α,β,1])=[α,3,4β]. Remember–this means that a ray from p1 to [α,β,1] is transformed to the ray from p2 to [α,1,β].

Our only unknown is ρ(z), and we will compute the ordinary differential equation that ρ must satisfy in order for an eigensurface to exist.

Let n(z,θ) denote the normal to the surface at x(z,θ) and define Out to be the reflection about n of In=p1x(z,θ). Thus, the reflected ray at x(z,θ), r¯, parametrized by t>0, has the form, x(z,θ)+tOut(z,θ). Writing ρ˙ for dρdz, we have

r¯(t)=[rcos(θ)+(ρρ˙2+2ρ˙z+r)cos(θ)tρ˙2+1,ρsin(θ)+(ρρ˙+2z2z)ρ˙sin(θ)tρ˙2+1,z+2ρρ˙t+ztρ˙2ztρ˙2+1].
On the other hand, the ray from p1 that strikes the surface at x(z,θ) is named by the point
[ρ(z)cos(θ)z,ρ(z)sin(θ)z,1].
To find the corresponding ray in B, we apply T to this point, giving
[ρ(z)cos(θ)z,3,4+ρ(z)sin(θ)z],
and so the parametric equations for this ray, which is T(r), and we call r˜, in B are
r˜(s)[ρ(z)cos(θ)sz,s4,ρ(z)sin(θ)sz+4],
where s is the parameter along the ray. The question then is when and if
r¯=r˜.
To find this, we need to solve for {s,t}, which we can do by equating the first two components of r¯ and r˜, giving a two-by-two system in s, t. Solving gives
s=4zr(z)sin(θ)z,
t=ρ(ρ˙2+1)(ρsin(θ)z+4)(ρρ˙2+2ρ˙zρ)(ρsin(θ)z).
One may now equate the third components of r¯ and r˜, and in that equation eliminate s, t using the above, giving a differential equation, which may be algebraically manipulated into the following factored form:
(ρ˙z+ρ)(ρρ˙+z)(ρsin(θ)z)(ρsin(θ)z+4)=0.
The factors give two differential equations and two algebraic equations, but notice the occurrences of θ in the third and fourth factors. Here ρ must depend only on z, i.e., it is viable only if it is independent of θ. The latter two factors are therefore spurious. The first factor has solutions ρ(z)=Cz, a cone through the origin, so it is not physically meaningful.

The second factor also gives a spurious solution but warrants close inspection. Writing rρ˙+z=0 as rdr=zdz, we see that the solutions are r=±z2+C, so the eigenmirror is

x(z,θ)=[±Cz2cos(θ),±Cz2cos(θ),z].
This is a sphere of radius C. While this is a legitimate solution, p1=[0,0,0] is the center of the sphere, which means that any ray of B reflects back through p1. The parametric equations for the eigensurface are then
Es(z,θ)=[4cosθ6z2zsinθ6z2,
4sinθ6z2zsinθ6z2,
4zzsinθ6z2].
Not surprisingly, a simple calculation shows that this is a portion of the plane z=y+4. Thus, the solution from the second factor is also nonphysical.

Thus, one may conclude that for the data given in this case that no interesting eigensurfaces exist. This is fairly general, of course, since there is nothing special about p1 other than it should lie on the axis of rotation, and that p2 not lie on the axis of rotation. What is special to this example is the choice of T. It is not at all clear if another reasonable choice of T would lead to true solutions.

6. GENERAL 3D CASE

For the general case in 3D, much of our above notation can be recycled. The main difference is now that the mirror will be given parametrically in the form x(u,v)=[X(u,v),Y(u,v),Z(u,v)], although of course one can often use the graph form of x(u,v)=[u,v,f(u,v)].

Suppose that r is the ray from p1 to x(u,v), a point on a reflective surface x. Let VS2(p1) denote the direction of r. The ray r reflects at x(u,v), then proceeds along the ray r¯, which is parametrized by tx(u,v)+tOut, where Out is the unit direction after reflection. To have an eigenmirror means that for every u, v in the domain of x, we r¯ intersects with T(r)=r˜ (which is given by sp2+sT(V)); see Fig. 10. Then the fundamental equation describing eigensurfaces and eigenmirrors is r¯=r˜, i.e.,

x(u,v)+tOut=p2+sT(V).
Here s and t are unknowns, and the derivatives of x(u,v) appear in Out, but nowhere else. Since there are three scalar equations, one can use two of them to solve for s and t. Substituting the results into the remaining equation leaves one equation in u, v, xu, xv, and x. For a fixed reflector x, one does not typically expect these two rays to intersect, since they lie in R3. But in special cases they will. One must take care, though, to require that s,t>0, since we are working with rays, not lines; otherwise nonphysical solutions certainly can occur.

 figure: Fig. 10.

Fig. 10. Derivation for the partial differential equation in the freeform case.

Download Full Size | PDF

It is at this point that our thinking diverges from the above derivations and examples. If the eigensurface and eigenmirror are both treated as unknowns, then we have a first-order nonlinear partial differential equation for f.

As an example, take p1=[0,0,0], p2=[1,0,1], and T(V)=V. Then the fundamental partial differential equation [Eq. (18)] takes the form

vfu+(fu)fv+v=0.
The general solution to Eq. (19) is
f(u,v)=the value ofZin:2uZv2+F(u+Z)=0.
As a simple example, take F1, which gives f(u,v)=(v21)/2u. The corresponding eigensurface is parametrized by three (large) rational functions in u and v. A plot of the eigensurface, eigenmirror, and two viewpoints appear in Fig. 11. Ray-tracing simulations from the viewpoints p1 and p2 appear in Figs. 12 and 13, verifying our theory. That is, the pattern in the mirror from viewpoint p1 appears to be identical to the pattern on the surface from viewpoint p2.

 figure: Fig. 11.

Fig. 11. Eigenmirror z=1v22u and its corresponding eigensurface.

Download Full Size | PDF

 figure: Fig. 12.

Fig. 12. Simulation of the view from p1 of an eigenmirror (below) and an eigensurface (above), which has a checkerboard pattern on it. The eigenmirror reflects this pattern, and the view is identical to the view of the upper surface from viewpoint p2, as in Fig. 13. The numbers are actual objects in the simulation, placed in the back of the scene to emphasize the use of a different viewpoint in Fig. 13.

Download Full Size | PDF

 figure: Fig. 13.

Fig. 13. View of the eigensurface from viewpoint p2.

Download Full Size | PDF

To strengthen this claim, we performed the following experiment. Imagine that we had printed the two images in Figs. 12 and 13 and then cut out all of the white from the paper version of Fig. 12, creating a lattice of sorts. Then one could overlay that onto the printed version of Fig. 13 and any mismatch between the patterns would be visible. A variant of this procedure was performed using the fact that png format supports transparency. The (positive) result appears in Fig. 14.

 figure: Fig. 14.

Fig. 14. View of the eigensurface from viewpoint p2, overlaid with the view from p1. To do this, all pure white in Fig. 12 was first converted to transparency and then pasted onto the image in Fig. 13. As a result, a ghost image of the surface appears floating above the mirror. If there were a significant discrepancy between the image of the eigensurface and eigenmirror, it would appear in the lower composite, i.e., there is no “interference” from the overlaying.

Download Full Size | PDF

Remarks:

  • (i) Note that for most choices of T, the fundamental equations are large–even if T is relatively simple. An interesting question is to try to classify all possible solutions when T is an orthogonal matrix. Clearly, it should not matter where the viewpoints are, but are the choices of T fundamentally different, or is there a change of variable that will show that there really is essentially only one equation here?
  • (ii) While we have no reason to assume that the eigenmirror or surface in this example will be rotationally symmetric, they are. In fact, every solution to Eq. (19) is rotationally symmetric. In the case of the quadratic eigenmirror f(u,v)=(v21)/2u, of course it is possible to explicitly compute the axis of revolution, or one can simple plot it to see that it is a tilted hyperboloid. For the case of the eigensurface, the issue is more delicate, since it is given parametrically. Numerical experiments show that it is indeed rotationally symmetric, with the same axis as the eigenmirror.
  • (iii) We have not assumed that p1 and p2 “see the same side” of the eigensurface. Indeed, in this last section, that is not the case—p2 sees the bottom face of the eigensurface, while p1 sees the top face. Thus the eigensurface is to be thought of as an infinitely thin surface, textured the same on both sides.

7. CONCLUSION

A conventional flat mirror has a remarkable but underappreciated property: one observer can experience the view of another observer at a different location, without physically moving to that location. This application is remarkable and is used to great dramatic effect in an early scene from the film Saving Private Ryan. Pinned down by enemy fire, Tom Hanks’ character cannot safely see the enemy position, and so cannot make a plan of action. But using chewing gum, a small mirror and stick, he is able to obtain the viewpoint he wants—but without putting himself in danger. Applications of our work for the human eye are the most natural. Security mirrors, dental tools, etc., all are designed to work in conjunction with the human eye, but flat or spherical mirrors are not always optimal for these applications. (The problem we consider is not solved by flat reflectors because we ask for the ability to have the stolen viewpoint include rotations, flips, scales, etc.)

When can one “steal” another viewpoint using a curved reflector? Certainly, this is a fundamental-sounding question of geometric optics, and we argue that writing down and analyzing the ordinary and partial differential equations for this problem is sensible if we seek to understand the complex transformations induced on ray bundles by optical elements.

A classic struggle for photographers is having to photograph a wall or building, but being unable to stand at the ideal location to do so because of some physical obstruction. The result is keystone distortion, or the “convergence of verticals”—see Kingslake [38]. Classically, this problem was solved by using a camera that incorporated bellows and a “rising front/swing back.” In principle though, this problem could be framed in terms of eigenmirrors and eigensurfaces, with the building face being the eigensurface. The convergence of verticals is similar to the problem of anamorphosis that required the SX-70 mentioned above, to incorporate freeform elements. As an element of a photographic system, such a mirror might be able to image a surface from an awkward position in a precise fashion. (Of course, incorporating such a freeform element into an photographic system might require other complex optical elements to deal with the aberrations that we have ignored in our approximation.)

A summary of the technical development of our work is as follows:

  • 1. We derived the fundamental equation in the planar case, where the reflector is given and the surface is sought. We wrote these equations out explicitly in the case where T is a linear map and gave an example where the reflector was a cubic curve.
  • 2. We considered the equations in the planar case when the reflector was the cross section of a rotationally symmetric reflector. We gave an example where the reflector was a parabola, and we explicitly computed the parametric equations for the surface. The parametric equations that we obtained were quite complicated, and while we were able to simplify them a great deal, the eigensurface was not clearly a familiar one. We gave simulations as a proof of concept.
  • 3. We showed using the example of the parabolic reflector z=u2+v2c that if one moves the viewpoint p2 off-axis, then an eigensurface will not exist. (T is a fixed, chosen transformation.)
  • 4. The above led us to write down the equations in the case where the rotationally symmetric surface has the form x(z,θ)=[ρ(z)cos(θ),ρ(z)sin(θ),z]. Then, for a given T and fixed viewpoints, we solved for ρ and thus, in a limited sense, classified the eigensurfaces and eigenmirrors in the rotationally symmetric case. The solutions that were found should be considered trivial or nonphysical.
  • 5. We derived the fundamental equation in the general 3D case. We considered an example where T(V)=V, where V is a unit vector representing a ray, and showed that this gave a first-order nonlinear partial differential equation for the reflector. We solved this equation, found a family of new surfaces, and demonstrated the predicted properties in simulation.

Note that to say that one viewer sees an object in the same way as another could reasonably allow a change in of field of view, i.e., a contraction or expansion of the bundle B. We did not consider such examples, but the equation that one obtains when taking T to be practically anything nontrivial is a quite complicated first-order nonlinear partial differential equation. The study of such equations appears to be a rich and unexplored intersection among optical design, numerical analysis, and partial differential equations.

Funding

Directorate for Mathematical and Physical Sciences (MPS) (DMS-09-08299).

REFERENCES

1. R. A. Hicks and R. Bajcsy, “Reflective surfaces as computational sensors,” in Proceedings of the 2nd Workshop on Perception for Mobile Agents, CVPR (1999), pp. 82–86.

2. C. Kanolt, “Multifocal opthalmic lenses,” U.S. patent 2,878,721 (March 24, 1959).

3. E. H. Land, “Reflex camera,” U.S. patent 3,672,281 (June 27, 1972).

4. J. G. Baker, “Compact folding reflex camera,” U.S. patent 3,678,831 (July 25, 1972).

5. W. T. Plummer, “Reflective imaging apparatus,” U.S. patent 3,735,685 (May 29, 1973).

6. B. D. Stone and G. W. Forbes, “Characterization of first-order optical properties for asymmetric systems,” J. Opt. Soc. Am. A 9, 478–489 (1992). [CrossRef]  

7. B. D. Stone and G. W. Forbes, “Foundations of first-order layout for asymmetric systems: an application of Hamilton’s methods,” J. Opt. Soc. Am. A 9, 96–109 (1992). [CrossRef]  

8. K. Thompson, “Practical methods for the optical design of systems without symmetry,” Proc. SPIE 2774, 2–12 (1996). [CrossRef]  

9. A. Bauer and J. P. Rolland, “Design of a freeform electronic viewfinder coupled to aberration fields of freeform optics,” Opt. Express 23, 28141–28153 (2015). [CrossRef]  

10. J. McDermit and T. Horton, “Reflective optics for obtaining prescribed irradiative distributions from collimated sources,” Appl. Opt. 13, 1444–1450 (1974). [CrossRef]  

11. D. G. Burkhard and D. L. Shealy, “Design of reflectors which will distribute sunlight in a specified manner,” J. Sol. Energy 17, 221–227 (1975). [CrossRef]  

12. D. L. Shealy, “Theory of geometrical methods for design of laser beam shaping systems,” Proc. SPIE 4095, 1–15 (2000). [CrossRef]  

13. W. B. Elmer, “The optical design of reflectors,” in The Optical Design of Reflectors, W. B. Elmer, ed., 2nd ed., Vol. 302 of Wiley Series in Pure and Applied Optics (Wiley, 1980), p. 1.

14. W. Welford and R. Winston, The Optics of Nonimaging Concentrators: Light and Solar Energy (Academic, 1978).

15. R. Winston, J. C. Miñano, and P. G. Benítez, Nonimaging Optics (Academic, 2005).

16. P. Benítez, J. C. Miñano, J. Blen, R. Mohedano Arroyo, J. Chaves, O. Dross, M. Hernandez, and W. Falicoff, “Simultaneous multiple surface optical design method in three dimensions,” Opt. Eng. 43, 1489–1502 (2004). [CrossRef]  

17. J. C. Miñano, P. Benítez, W. Lin, J. Infante, F. Muñoz, and A. Santamaría, “An application of the SMS method for imaging designs,” Opt. Express 17, 24036–24044 (2009). [CrossRef]  

18. S. A. Kochengin and V. I. Oliker, “Determination of reflector surfaces from near-field scattering data,” Inverse Prob. 13, 363–373 (1997). [CrossRef]  

19. S. Kochengin and V. Oliker, “Determination of reflector surfaces from near-field scattering data. II. Numerical solution,” Numer. Math. 79, 553–568 (1998). [CrossRef]  

20. J. Rubinstein and G. Wolansky, “Intensity control with a free-form lens,” J. Opt. Soc. Am. A 24, 463–469 (2007). [CrossRef]  

21. L. A. Caffarelli, C. E. Gutiérrez, and Q. Huang, “On the regularity of reflector antennas,” Ann. Math. 167, 299–323 (2008). [CrossRef]  

22. J. Rubinstein and G. Wolansky, “Geometrical optics and optimal transport,” J. Opt. Soc. Am. A 34, 1817 (2017). [CrossRef]  

23. R. D. Leo, C. E. Gutiérrez, and H. Mawi, “On the numerical solution of the far field refractor problem,” Nonlinear Anal. 157, 123–145 (2017). [CrossRef]  

24. F. R. Fournier, W. J. Cassarly, and J. P. Rolland, “Fast freeform reflector generation using source-target maps,” Opt. Express 18, 5295–5304 (2010). [CrossRef]  

25. R. A. Hicks and R. Perline, “Geometric distributions and catadioptric sensor design,” in Proceedings of the 2001 IEEE Computer Society Conference on Computer Vision and Pattern Recognition CVPR (IEEE, 2001), pp. 584–589.

26. R. A. Hicks and C. Croke, “Designing coupled free-form surfaces,” J. Opt. Soc. Am. A 27, 2132–2137 (2010). [CrossRef]  

27. C. B. Croke and R. A. Hicks, “Solution to the bundle-to-bundle mapping problem of geometric optics using four freeform reflectors,” J. Opt. Soc. Am. A 31, 2097–2104 (2014). [CrossRef]  

28. A. Plakhov, S. Tabachnikov, and D. Treschev, “Billiard transformations of parallel flows: a periscope theorem,” J. Geom. Phys. 115, 157–166 (2017). [CrossRef]  

29. F. Duerr, P. Benítez, J. C. Miñano, Y. Meuret, and H. Thienpont, “Analytic design method for optimal imaging: coupling three ray sets using two free-form lens profiles,” Opt. Express 20, 5576–5585 (2012). [CrossRef]  

30. F. Duerr, P. Benítez, J. C. Minano, Y. Meuret, and H. Thienpont, “Perfect imaging of three object points with only two analytic lens surfaces in two dimensions,” Proc. SPIE 8429, 842908 (2012). [CrossRef]  

31. R. A. Hicks, S. G. Rody, and J. D. Wright, “Bundle separation, obstructions to perfect imaging, and other qualitative aspects of simultaneous multiple surface design,” Opt. Eng. 53, 031309 (2013). [CrossRef]  

32. R. Luneberg, The Mathematical Theory of Optics (University of California, 1964).

33. A. Walther, “Irreducible aberrations of a lens used for a range of magnifications,” J. Opt. Soc. Am. A 6, 415–422 (1989). [CrossRef]  

34. O. Stavroudis, The Optics of Rays, Wavefronts, and Caustics (Academic, 1972).

35. R. Hicks and R. Perline, “Equiresolution catadioptric sensors,” Appl. Opt. 44, 6108–6114 (2005). [CrossRef]  

36. P. Roitman, “Gaussian curvature, mirrors, and maps,” Am. Math. Monthly 119, 506–509 (2012). [CrossRef]  

37. R. A. Hicks and M. Coletta, “Computational photography with panoramic sensors that have uniform resolution with respect to unwarping transformations,” J. Math. Imaging Vis. 46, 121–127(2013). [CrossRef]  

38. R. Kingslake, Optics in Photography (SPIE, 1992).

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (14)

Fig. 1.
Fig. 1. Distortion of a checkerboard pattern in a spherical mirror. This is a panoramic view of a test scene, with the camera appearing in the center of the photo. The checkers are 0.2159 m square, and the mirror, of radius 0.031 m, is 0.34 m above the floor. Two squares are labeled: square (1) is 0.10795 m from the camera and square (2) is 0.4572 m away, yet square (1) consists of disproportionately many more pixels. For the robot soccer application, we sought greater equity between the two.
Fig. 2.
Fig. 2. Soccer robots at the UPenn GRASP Lab, circa 1997. Each robot was equipped with a spherical mirror to allow for panoramic views.
Fig. 3.
Fig. 3. Cross section of a rotationally symmetric mirror designed to give a soccer robot a 360 ° × 145 ° view of the playing field. This reflector has plane as its eigensurface. At the lowest point, the radius of curvature is 1.7 m.
Fig. 4.
Fig. 4. Case of the rectifying mirror. This curved reflector was designed to be mounted on a soccer playing robot, 34 cm above the playing field, which here contains a checkerboard pattern.
Fig. 5.
Fig. 5. Problem in 2D, with no assumed symmetry.
Fig. 6.
Fig. 6. Sample 2D problem. The eigenmirror is y = u 3 , with u [ 3 / 5 , 1 ] , p 1 = [ 0 , 1 / 3 ] , p 2 = [ 2 , 3 / 2 ] , and T ( [ X 1 , X 2 ] ) = R ( [ X 1 , X 2 ] ) , where R is the rotation matrix given in Eq. (8).
Fig. 7.
Fig. 7. 2D schematic of the cross section of a rotationally symmetric eigenmirror with eigensurface. One observer at p 1 gazes at the mirror and the other, p 2 , looks directly at the eigensurface. Observe that what we have here is not a two reflector problem—there is no reflective relation between the ray r ¯ and T ( r ) at their point of intersection on the eigensurface.
Fig. 8.
Fig. 8. System for which the reflective surface is a portion of the parabola f ( u ) = 1 2 u 2 + 3 . The first viewpoint, p 1 = [ 0 , 0 ] , observes the (black) parabola from below, via the dashed ray. The second viewpoint, p 2 = [ 0 , 20 ] views that (red) eigensurface from above, via the dotted ray.
Fig. 9.
Fig. 9. (A) Side view of a parabolic mirror (upper) and the corresponding eigensurface (below). The entire scene lies within a room with a cubical room with checkered walls. (B) From above, at [0,0,20], an observer looking down sees the sandy back of the parabolic mirror and the rings in the eigensurface. (C) The view from p 1 of the parabola. The proof of concept is that the rings look the same as in (B). (D) Here we see a flat disk reflected in the parabola. It is not, of course, an eigensurface of the parabola, and so the rings do not appear evenly spaced, as in (B) and (C).
Fig. 10.
Fig. 10. Derivation for the partial differential equation in the freeform case.
Fig. 11.
Fig. 11. Eigenmirror z = 1 v 2 2 u and its corresponding eigensurface.
Fig. 12.
Fig. 12. Simulation of the view from p 1 of an eigenmirror (below) and an eigensurface (above), which has a checkerboard pattern on it. The eigenmirror reflects this pattern, and the view is identical to the view of the upper surface from viewpoint p 2 , as in Fig. 13. The numbers are actual objects in the simulation, placed in the back of the scene to emphasize the use of a different viewpoint in Fig. 13.
Fig. 13.
Fig. 13. View of the eigensurface from viewpoint p 2 .
Fig. 14.
Fig. 14. View of the eigensurface from viewpoint p 2 , overlaid with the view from p 1 . To do this, all pure white in Fig. 12 was first converted to transparency and then pasted onto the image in Fig. 13. As a result, a ghost image of the surface appears floating above the mirror. If there were a significant discrepancy between the image of the eigensurface and eigenmirror, it would appear in the lower composite, i.e., there is no “interference” from the overlaying.

Equations (34)

Equations on this page are rendered with MathJax. Learn more.

V = [ u , f ( u ) ] p 1 [ u , f ( u ) ] p 1 .
Out = Out ( f ( u ) , f ( u ) , u ) ,
[ u , f ( u ) ] + t Out = p 2 + s T ( V )
E c ( u ) = [ u , f ( u ) ] + t ( u ) Out ( f ( u ) , f ( u ) , u ) ,
E c ( u ) = p 2 + s ( u ) T ( V ( f ( u ) , u ) ) .
E c ( u ) = 1 δ [ 2 f ( u ) 2 b u + ( 2 u 2 a u b ) f ( u ) u 2 a ) f ( u ) 2 + ( 2 b f ( u ) 3 2 a u + 2 b f ( u ) 2 + ( 2 u 2 b + 2 a u ) f ( u ) + 2 u 3 a ) f ( u ) 2 f ( u ) 2 b u + ( 2 u 2 a + u b ) f ( u ) + u 2 a , ( ( 2 d u + b ) f ( u ) 2 + ( 2 c u 2 + a u ) f ( u ) ) f ( u ) 2 + ( 2 f ( u ) 3 d 2 f ( u ) 2 c u + ( 2 d u 2 + 2 u b ) f ( u ) + 2 c u 3 + 2 u 2 a ) f ( u ) + ( 2 d u b ) f ( u ) 2 + ( 2 c u 2 a u ) f ( u ) ] ,
δ = ( f ( u ) 2 b + ( a u + d u ) f ( u ) + c u 2 ) f ( u ) 2 + ( 2 d f ( u ) ) 2 + ( 2 u b 2 c u ) f ( u ) + 2 u 2 a ) f ( u ) f ( u ) 2 b + ( a u d u ) f ( u ) c u 2 .
R = [ cos 190 ° sin 190 ° sin 190 ° cos 190 ° ] .
f ( u ) = u 2 2 + 3 ,
Out = reflect ( In , n ) = [ 7 u u 2 + 1 , u 4 + 9 u 2 6 2 u 2 + 2 ] ,
t [ u + 7 u t u 2 + 1 , u 2 2 + 3 + u 4 t 2 u 2 + 2 + 9 u 2 t 2 u 2 + 2 3 t u 2 + 1 ] .
[ s = u 4 + 2 u 2 + 232 u 4 + 16 u 2 + 36 , t = 2 u 4 13 u 2 14 u 4 + 16 u 2 + 36 ] .
u [ ( u 4 + 2 u 2 + 232 ) u u 4 + 16 u 2 + 36 , u 6 / 2 16 u 4 198 u 2 24 u 4 + 16 u 2 + 36 ] .
τ [ ( τ 2 + 2 τ + 232 ) τ τ 2 + 16 τ + 36 , τ 3 / 2 16 τ 2 198 τ 24 τ 2 + 16 τ + 36 ] .
[ u u 2 + v 2 c , v u 2 + v 2 c , 1 ] .
T [ u u 2 + v 2 c , v u 2 + v 2 c , 1 ] = [ u u 2 + v 2 c , a + 1 , b + v u 2 + v 2 c ] .
r ¯ = [ u t u ( 4 c 1 ) ( 4 u 2 + 4 v 2 + 1 ) u 2 + v 2 + ( u 2 + v 2 c ) 2 , v t v ( 4 c 1 ) ( 4 u 2 + 4 v 2 + 1 ) u 2 + v 2 + ( u 2 + v 2 c ) 2 , u 2 + v 2 c + t · ( u 2 + v 2 c ) ( 4 u 2 + 4 v 2 + 1 ) + 2 ( u 2 + v 2 + c ) ( 4 u 2 + 4 v 2 + 1 ) u 2 + v 2 + ( u 2 + v 2 c ) 2 ] ,
r ˜ = [ s x u 2 + v 2 c , s a , b + s v u 2 + v 2 c ] .
x ( z , θ ) = [ ρ ( z ) cos ( θ ) , ρ ( z ) sin ( θ ) , z ] ,
r ¯ ( t ) = [ r cos ( θ ) + ( ρ ρ ˙ 2 + 2 ρ ˙ z + r ) cos ( θ ) t ρ ˙ 2 + 1 , ρ sin ( θ ) + ( ρ ρ ˙ + 2 z 2 z ) ρ ˙ sin ( θ ) t ρ ˙ 2 + 1 , z + 2 ρ ρ ˙ t + z t ρ ˙ 2 z t ρ ˙ 2 + 1 ] .
[ ρ ( z ) cos ( θ ) z , ρ ( z ) sin ( θ ) z , 1 ] .
[ ρ ( z ) cos ( θ ) z , 3 , 4 + ρ ( z ) sin ( θ ) z ] ,
r ˜ ( s ) [ ρ ( z ) cos ( θ ) s z , s 4 , ρ ( z ) sin ( θ ) s z + 4 ] ,
r ¯ = r ˜ .
s = 4 z r ( z ) sin ( θ ) z ,
t = ρ ( ρ ˙ 2 + 1 ) ( ρ sin ( θ ) z + 4 ) ( ρ ρ ˙ 2 + 2 ρ ˙ z ρ ) ( ρ sin ( θ ) z ) .
( ρ ˙ z + ρ ) ( ρ ρ ˙ + z ) ( ρ sin ( θ ) z ) ( ρ sin ( θ ) z + 4 ) = 0 .
x ( z , θ ) = [ ± C z 2 cos ( θ ) , ± C z 2 cos ( θ ) , z ] .
E s ( z , θ ) = [ 4 cos θ 6 z 2 z sin θ 6 z 2 ,
4 sin θ 6 z 2 z sin θ 6 z 2 ,
4 z z sin θ 6 z 2 ] .
x ( u , v ) + t Out = p 2 + s T ( V ) .
v f u + ( f u ) f v + v = 0 .
f ( u , v ) = the value of Z in : 2 u Z v 2 + F ( u + Z ) = 0 .
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.