PreprintPDF Available

Extended depth-of-focus wavefront design from pseudo-umbilical space curves.

Authors:

Abstract and Figures

Designing extended-depth-of-focus wavefronts is required in multiple optical applications. Caustic location and structure analysis offer a powerful tool to design such wavefronts. An intrinsic limitation of designing extended-depth-of-focus wavefronts is that any smooth surface, with a non-constant mean curvature, unavoidably introduces a separation between caustic sheets, which is proportional to the ratio of change of the mean curvature along a curved embedded in the wavefront. We present a novel method to obtain extended depth-of-focus wavefronts where the mean curvature variation ratio is reduced thanks to using a long circle-involute space curve effectively filling the wavefront surface. Additionally, we present a variant of the method in which the wavefront is modified within a small tubular neighborhood of the circle involute in order to partially meet the umbilical condition along that tubular region. Finally, we provide some numerical results showing the potential of our method in an application example.
Content may be subject to copyright.
Extended depth-of-focus wavefront design from1
pseudo-umbilical space curves2
SERGIO BARBERO1* , MANUEL RITORÉ,2
3
1Instituto de Óptica (CSIC), Serrano 121, Madrid, Spain.4
2Departamento de Geometría y Topología, Facultad de Ciencias, Universidad de Granada, 180715
Granada, Spain6
*sergio.barbero@csic.es7
Abstract: Designing extended-depth-of-focus wavefronts is required in multiple optical
8
applications. Caustic location and structure analysis offer a powerful tool to design such
9
wavefronts. An intrinsic limitation of designing extended-depth-of-focus wavefronts is that
10
any smooth surface, with a non-constant mean curvature, unavoidably introduces a separation
11
between caustic sheets, which is proportional to the ratio of change of the mean curvature
12
along a curved embedded in the wavefront. We present a novel method to obtain extended
13
depth-of-focus wavefronts where the mean curvature variation ratio is reduced thanks to using a
14
long circle-involute space curve effectively filling the wavefront surface. Additionally, we present
15
a variant of the method in which the wavefront is modified within a small tubular neighborhood
16
of the circle involute in order to partially meet the umbilical condition along that tubular region.
17
Finally, we provide some numerical results showing the potential of our method in an application
18
example.19
© 2023 Optical Society of America20
1. Introduction21
Multifocal and/or extended-depth-of-focus optical beams are ubiquitous in a wide range of
22
optical engineering applications; to name just a few: ophthalmic corrections for presbyopia [1],
23
microscopy [2], or laser micromachining [3]. Although relatively novel, the analysis or
24
prescription of the caustic structure of such optical beams offers an appealing way of designing
25
them [4–8].26
Here, we restrict our analysis to optical systems that generate smooth, in the sense described by
27
analytical functions, wavefronts; a desirable property in some applications such as visual optics.
28
Then, we exclude the classical cone-shaped-axicon, which contains derivatives discontinuities
29
at the center. Recently, it has been proposed a method to design axially symmetric wavefronts
30
under certain caustic prescriptions [8]. The method is purely within a geometrical optics
31
framework, which is a reasonably good starting point. Once the incoming and outgoing
32
wavefronts are prescribed, it is possible to provide their required coupling by means of, for
33
instance, a refractive/reflective optical surface [9].34
This kind of wavefront provides an extended depth-of-focus by means of a varying mean
35
curvature through the spatial domain. However, any smooth surface with a non-constant mean
36
curvature unavoidable introduces a difference in principal curvatures, which implies the generation
37
of two separated caustic sheets; hence preventing the formation of a perfect geometrical focal
38
line. The amount of cylinder (difference between principal curvatures) is proportional to the
39
ratio of change of one of the principal curvatures along a principal line of curvature [10]. This
40
last property points out one of the drawbacks of axially symmetric wavefronts. Since the
41
change of mean curvature, on these surfaces, is produced along meridians, and meridians are the
42
shortest path from the optical center to the edge, the ratio of mean curvature is maximal, and, in
43
consequence, they presumably induce a non-optimal amount of cylinder. This fact was analyzed
44
in [8].45
Therefore, a natural alternative to axially symmetric wavefronts is to set the mean curvature
46
variation not along meridians but, for instance, along spiral-type curves. Indeed, spiral or helical
47
wavefronts (or phase functions) are intensively used in many different optical applications [11
13].
48
Breaking axial symmetry implies that the equations relating caustics with wavefront geometry
49
are no longer ordinary differential equations but partial differential equations (PDE). Therefore,
50
a possible approach would be to handle those PDEs. However, to avoid the complexities and
51
non-uniqueness issues associated with solving PDEs, we propose here a different approach based
52
on differential geometry reasoning.53
Caustics, or focal surfaces as they are known in differential geometry, contain some special
54
curves called focal curves [14]. Given a space curve
𝛾
embedded in a surface, with its associated
55
focal surface, the focal curve
𝐶𝛾
is such that it meets that the points of
𝐶𝛾
are the centers of the
56
osculating spheres of
𝛾
[14]. Setting focal curves and recovering their generating space curves,
57
as we will see, is a way of reducing the dimensionality of the much more complicated problem of
58
recovering surfaces from their focal surfaces. The crux of such simplification is to realize that
59
some space curves can effectively fill in a surface in Euclidean space. Particularly, we considered
60
circle involutes. They are curves optimally sampling a surface, in the sense that the overall length
61
is minimum given the following restriction: the distance of any point on that surface to the circle
62
involute is less than a given magnitude [15].63
A classic result in differential geometry of spatial curves is that curvature and torsion (second
64
curvature) determine the curve up to translation and rotation [16]. Then, the second curvature of
65
the circle involute filling the wavefront can be conferred with an extra convenient optical property;
66
namely, to be close to the umbilical, hence reducing the light dispersion generated by caustic
67
sheet separation. This approach is admissible because, generically, any portion of a smooth
68
surface (at least
𝐶2
) providing spatially varying mean curvature can only contain umbilical lines
69
or isolated umbilical points [10, 17]. In principle, umbilicity or pseudo-umbilicity (close to) are
70
attainable if those curves are not very close to each other on the embedded surface. This is an
71
issue that we have explored experimentally in this paper.72
Inspired by the aforementioned differential geometry properties we propose a method based
73
on several sequential steps. 1) We construct an axial symmetric asphere meeting a sagittal
74
caustic prescription (section 2.1); 2) We build up a uniform arc-length spiral-type circle involute
75
optimally filling that asphere (section 2.2); 3) Those curves are modified such that one of the
76
curvatures meets the imposed extended depth-of-focus target; 4) Small tubular neighborhoods
77
of those curves are constructed such that the umbilical condition (zero astigmatism) is, at least,
78
partially met (subsection 2.3), and, finally, 5) the cloud points are approximated with a surface
79
using Zernike polynomial least squares fitting (subsection 2.4). Once the surface is described
80
analytically by means of a Zernike expansion, caustic sheets can be computed as explained
81
in subsection 2.5. Overall, the method offers an efficient solution to the problem of extended
82
depth-of-focus multifocal wavefront design under some caustic prescriptions.83
2. Method84
In this section, we describe step by step the design method. For a better understanding, we will
85
exemplify it with an example of application; namely, a wavefront aimed at providing prescribed
86
intensities (linear, exponential, or any other type [20]) over a depth of field along a segment of the
87
optical axis, which can be achieved using an axicon or gaxicon [21]. Particularly, we designed a
88
wavefront providing constant intensity distribution. For this purpose, a one-to-one mapping must
89
be established between the generating curve
𝛾
of the wavefront and its associated focal curve
𝐶𝛾
;
90
both curves sampled with points at equal arc-length distances. In doing so, what is ensured is that
91
equal areas of the wavefront are focused on approximately equal-length segments of the optical
92
axis. The example case used in this paper [8] is applicable in visual optics problems, where it
93
has been shown the benefits of including curvature variation not only along the radial coordinate
94
but also along the angular one [1, 27
29]. We assumed an eye output wavefront, with a pupil size
95
of
𝜉=3
mm providing 60 D for far vision and 63 D for near vision. This implies a far and near
96
focal distance of 𝑧𝑐=16.67 mm and 𝑧𝑒=15.87 mm, respectively.97
2.1. Axial symmetric wavefront with prescribed sagittal caustic98
Caustic surfaces are the locus of centers of principal curvatures of the wavefront. They can
99
be obtained from the wavefront surface by means of computing distances along the wavefront
100
normals, whose magnitudes are the inverse of the principal curvatures [25]. Each point of the
101
wavefront, except if it is umbilical, generates two caustic points associated with each of the two
102
principal curvatures. Therefore, the caustic manifold comprises two surfaces or sheets. If the
103
wavefront poses axial symmetry, one of the caustic sheets (called sagittal) degenerates into the
104
segment of a straight line along the symmetry axis. Our starting point is an axially symmetric
105
wavefront generating a uniform light concentration along the sagittal caustic. Such wavefront can
106
be obtained by solving the Cauchy problem of a first-order non-linear differential equation [15]:
107
𝑧(𝑥)=𝑥
𝑔(𝑥) 𝑧(𝑥),(1)
𝑔(𝑥)=(𝑧𝑒𝑧𝑐)𝑥
𝜉+𝑧𝑐.
We solved Eq. 1 numerically with the help of ode45 MATLAB function and
𝑧(0)=0
as the
108
initial condition. The numerical solver provides sag data (
𝑧𝑖
) of the wavefront for a set of radial
109
coordinates (
𝑥𝑖
). For subsequent steps, we need to fit those data with an analytical function. For
110
the example case, we experimentally found that the optimal function is an asphere described by a
111
fourth-order power series polynomial:112
𝑧(𝑥)=𝑐1𝑥2+𝑐2𝑥4+𝑐3𝑥6+𝑐4𝑥8.(2)
The least squares coefficients were found using a simplex search optimization method, as
113
implemented with fminsearch MATLAB function. The absolute errors of the fitting, as a function
114
of radial coordinate, are shown in Fig. 1.115
0 0.5 1 1.5 2 2.5 3
Radial coordinate (mm)
-0.02
-0.015
-0.01
-0.005
0
0.005
0.01
0.015
Fitting error ( m)
Fig. 1. Fitting error (absolute errors) of wavefront data when using Eq. 2 in the example
case.
2.2. Circle involutes embedded in an axial symmetric wavefront116
We consider a regular curve
𝛾:𝐽R2
defined on some interval
𝐽R
, where by regular we
117
mean
¤𝛾1(𝑠)2+ ¤𝛾2(𝑠)20
for every
𝑠𝐽
. In this section, the dot indicates the derivative with
118
respect to the parameter
𝑠
. We assume that the coordinates in
R2
are
(𝑥, 𝑧)
, that the curve can be
119
written as 𝛾(𝑠)=(𝛾1(𝑠), 𝛾2(𝑠)) for 𝑠𝐽, and that 𝛾1>0at least in the interior of 𝐽.120
We also consider the parameterization
𝜓:𝐽× [0,2𝜋] R3
of the associated wavefront of
121
revolution (given by Eq. 2) obtained by rotating the curve 𝛾around the 𝑧-axis:122
𝜓(𝑠, 𝜃)=(𝛾1(𝑠)cos 𝜃, 𝛾1(𝑠)sin 𝜃 , 𝛾2(𝑠)).(3)
The tangent plane of the wavefront, at every point (𝑠, 𝜃), is generated by the vectors:
( ¤𝛾1(𝑠)cos 𝜃, ¤𝛾1(𝑠)sin 𝜃 , ¤𝛾2(𝑠)),(𝛾1(𝑠)sin 𝜃, 𝛾1(𝑠)cos 𝜃, 0).
Hence the Riemannian metric induced in the surface by the Euclidean one is given, in
(𝑠, 𝜃)123
coordinates, by124
©«
𝑓20
0𝑔2ª®¬:=©«
( ¤𝛾1)2+ ( ¤𝛾2)20
0𝛾2
1ª®¬.(4)
A geodesic in this metric is a curve
𝛿(𝑡):=(𝑠(𝑡), 𝜃(𝑡))
satisfying the system of second order
125
ordinary differential equations [26]126
𝑠′′ +¤
𝑓
𝑓(𝑠)2𝑔¤𝑔
𝑓2(𝜃)2=0,
𝜃′′ +2¤𝑔
𝑔𝑠𝜃=0.
(5)
In these formulas, the symbol
denotes the derivative with respect to
𝑡
. A geodesic is uniquely
127
determined by their initial values
𝛿(0)=(𝑠(0), 𝜃(0))
,
𝛿(0)=(𝑠(0), 𝜃(0))
. It is easily checked
128
that if 𝛿(𝑡)=(𝑠(𝑡), 𝜃 (𝑡)) is a geodesic, then129
𝛿𝜃0(𝑡)=(𝑠(𝑡), 𝜃(𝑡) + 𝜃0)(6)
is also a geodesic since it satisfies Eqs.
(5)
. The curve
𝛿𝜃0
is identified with the geodesic obtained
130
by rotating
𝛿
by an angle
𝜃0
. The initial conditions of the geodesic
𝛿𝜃0
are
𝛿𝜃0(0)=𝛿(0) + (0, 𝜃0)131
and 𝛿
𝜃0(0)=𝛿(0).132
Let us fix some
𝑠0𝐽
. Then the curve
𝜃↦→ (𝑠0, 𝜃 )
is sent by the parameterization
𝜓
to the
133
circle of revolution134
𝜃↦→ (𝛾1(𝑠0)cos 𝜃, 𝛾1(𝑠0)sin 𝜃, 𝛾2(𝑠0)) (7)
of length
2𝜋𝛾1(𝑠0)=2𝜋𝑔(𝑠0)
. The curve
𝜃↦→ (𝑠0, 𝜃 )
can be parameterized by arc-length
135
defining 𝑐:[0,2𝜋𝑔 (𝑠0)] R2by:136
𝑐(𝑡):=(𝑠0, 𝑡/𝑔(𝑠0)).(8)
To compute the involute around the curve
𝑐
we consider the geodesics
𝛿𝑡
in the plane with
137
coordinates (𝑠, 𝜃)satisfying the initial conditions 𝛿𝑡(0)=𝑐(𝑡),𝛿
𝑡(0)=𝑐(𝑡). That is138
𝛿𝑡(0)=(𝑠0, 𝑡/𝑔(𝑠0)), 𝛿
𝑡(0)=(0,𝑔(𝑠0)1).(9)
We fix the geodesic
𝛿0(𝑟)=(𝑠(𝑟), 𝜃(𝑟))
corresponding to the value
𝑡=0
. As observed after
Eqs.
(5)
, for any value
𝜃0
, the curve
(𝛿0)𝜃0
defined in Eq.
(6)
is a geodesic. Taking
𝜃0=𝑡/𝑔(𝑠0)
,
the initial conditions of (𝛿0)𝑡/𝑔(𝑠0)are
(𝛿0)𝑡/𝑔(𝑠0)(0)=𝛿0(0) + 0,𝑡
𝑔(𝑠0)=𝑠0,𝑡
𝑔(𝑠0)=𝛿𝑡(0),
(𝛿0)
𝑡/𝑔(𝑠0)(0)=𝛿
0(0)=0,1
𝑔(𝑠0)=𝛿
𝑡(0).
By the uniqueness of geodesics with the same initial conditions we get (𝛿0)𝑡/𝑔(𝑠0)=𝛿𝑡and so139
𝛿𝑡(𝑟)=(𝛿0)𝑡/𝑔(𝑠0)(𝑟)=𝑠(𝑟), 𝜃(𝑟) + 𝑡
𝑔(𝑠0).(10)
Hence the equation of the involute about the geodesic circle (8) is given by140
𝑡↦→ 𝛿𝑡(𝑡)=𝛿0(𝑡) + 0,𝑡
𝑔(𝑠0)=𝑠(𝑡), 𝜃(𝑡) + 𝑡
𝑔(𝑠0).(11)
Therefore it is enough to compute the geodesic
𝛿0(𝑟)=(𝑠(𝑟), 𝜃(𝑟))
to obtain the involute. The
141
corresponding curve in the wavefront is given by142
𝑡↦− 𝜓𝑠(𝑡), 𝜃 (𝑡) + 𝑡
𝑔(𝑠0).
The surface associated with Eq. 2 is:143
𝛾1(𝑠)=𝑠, (12)
𝛾2(𝑠)=𝑐1𝑠2+𝑐2𝑠4+𝑐3𝑠6+𝑐4𝑠8.
Going back to the geodesic (circle involute) curve, eqs. 5 were solved numerically using a
144
built-in MATLAB function: ode45.m, which generates a non-uniform distribution of points:
145
(𝑥𝑖, 𝑦𝑖, 𝑧 𝑗)𝑖=1, ...𝑛
. Next, we interpolate those points imposing equal arc-length separation.
146
The arc-length separation is computed by:147
𝛿𝑖=(𝑥𝑖+1𝑥𝑖)2+ (𝑦𝑖+1𝑦𝑖)2+ (𝑧𝑖+1𝑧𝑖)2.(13)
The total length is
𝐿=Í𝑛
𝑖=1𝛿𝑖
. Then, a uniform arc-length step is given as
𝛿𝑢=𝐿/(𝑛+1)
.
148
We obtained the equal arc-length points of the curve (
(𝑥𝑢, 𝑦𝑢, 𝑧𝑢)𝑢=1, .., 𝑚
) by separately
149
interpolating their
𝑠
and
𝜃
parameters using spline interpolation. Specifically, we used the interp1
150
MATLAB function with the splines option. Finally, the curve points were obtained by applying
151
again Eq. 3.152
For illustrative purposes, Fig. 2 shows a uniform equally arc-length circle involute within the
153
asphere of Eq. 2, as computed with the aforementioned procedure.154
2.3. Pseudo-umbilical tubular neighborhood of the circle involute155
Once we have built the set of points of the circle involute
𝛾:=(𝑥𝑢, 𝑦𝑢, 𝑧𝑢)
, the next step is to
156
perturb them so that the modified curve (say
𝛽
) has a prescribed focal curve. We show a scheme
157
of this procedure in Fig.3. We set that the focal curve is a line along the optical axis at the image
158
space: say, the set of points
𝐶𝛽:=(0,0, 𝜁𝑖)
. This intrinsically prescribes a radius of curvature
159
function given by 𝜁, and discretized by points 𝜁𝑖.160
Then we establish a one-to-one mapping between
𝛽
(red solid line in Fig.3) and
𝐶𝛽
(points
𝜁1
161
and
𝜁2
in Fig.3). For
𝐶𝛽
to be a focal curve of
𝛽
, the points of the former must be the centers of
162
the osculating spheres of the latter (dashed circles in Fig.3). Then we impose that the distance
163
from 𝛽to 𝐶𝛽is given by 𝜁. We transform 𝑐into 𝛽by perturbations along the normal to 𝛾.164
Fig. 2. Equal arc-length spaced circle involute points (blue dots) embedded in asphere
provided by Eq. 2 (colormap yellow-red). Axes scales are not equal for better visualizing
how the circle-involute fills in the surface.
The new curve
𝛽
is not necessarily an umbilical curve embedded in the wavefront. However,
165
we can at least approximate it to umbilicity, building up neighboring points to
𝛽
with some
166
properties. Parametrizing
𝛽
with arc length, the Darboux frame defines a positively oriented
167
orthonormal basis attached to each point
𝑝𝛽
along wavefront
𝑆
, which vectors are: 1) the unit
168
tangent:
T(𝑠)=𝛽(𝑠)
; 2) the unit normal to
𝑆
at
𝑝
:
N(𝑠)
; and 3) the tangent normal given by
169
cross product:
B(𝑠)=T(𝑠) × N(𝑠)
. Now, if
𝛽
is a curve of umbilical points, then it is possible to
170
define a local differentiable
(𝐶)
chart in a small tubular neighborhood of
𝛽
as a second-order
171
Taylor series expansion [18,19]:172
𝑊(𝑠, 𝑣)=𝛽(𝑠) + 𝑣B(𝑠) + 𝑘(𝑠)
2𝑣2N(𝑠),(14)
where
𝑘(𝑠)
is the normal curvature at
𝑝
, to which is assigned the value of
𝜁
.
𝑊(𝑠, 𝑣)
provides
173
a parametrization of that small tubular neighborhood close to the point
𝑝𝛽
. We computed
T(𝑠)174
by numerically differentiating
𝛾
with respect to arc length (by means of finite differences).
N(𝑠)175
is prescribed as the unitary version of the vector joining each pair of points of the one-to-one
176
mapping between
𝛽
and
𝐶𝛽
. Finally,
B(𝑠)
is obtained, as already mentioned, by the cross product
177
of the previous unitary vectors. Particularly, we computed one point of this chart for each point
178
of 𝛽.179
2.4. Zernike polynomials least squares fitting of the wavefront surface180
The merging of the points of the previous steps, namely the points of
𝛽
and its neighborhood points
181
in the tubular chart, provides a discretization of the wavefront:
𝑆:=(𝑥𝑗, 𝑦 𝑗, 𝑧 𝑗)𝑗=1, ...𝑛
.
182
Next, we fit that point cloud with a polynomial surface representation. Specifically, we used
183
Zernike polynomials, as they are classically used in optics; we followed the polynomial standard
184
Fig. 3. Transformation of circle involute within asphere (
𝛾
depicted with black solid
line) into a space curve with a prescribed focal curve (𝛽depicted with red solid line).
reported in [30]. The fitting scheme we used was the least squares approximation, which sets the
185
following linear system of equations:186
𝑧𝑗=
𝑗
𝐶𝑖𝑍𝑖(𝑥𝑗, 𝑦 𝑗)𝑗=1, ...𝑛, 𝑖 =1, ..𝑚, (15)
where
𝑍𝑖
denotes a Zernike polynomial,
𝐶𝑖
its associated coefficient and the index
𝑚
is the
187
number of Zernike polynomials used in the fitting. Eq. 15 can be expressed in matrix form as:188
z=𝐴c,(16)
where
z
is column vector containing
𝑧𝑗𝑗=1, ...𝑛
, and
𝑐
is a column vector containing the
𝑚189
Zernike coefficients. Matrix
𝐴
(dimension:
𝑛 𝑥 𝑚
) is, in general, overdetermined. We computed
190
𝑐
by a two steps procedure. First, we applied the pseudoinverse (Moore–Penrose inverse) to
191
matrix A, i.e. the least squares solution to the system of linear equations given by Eq. 16. The
192
pseudoinverse was computed using a built-in MATLAB function: pinv.m. Second, we refine the
193
solution by a simplex search optimization method as implemented with fminsearch MATLAB
194
function.195
2.5. Caustic construction from Zernike expansion196
We denote the caustic sheets with
𝜒
(sagittal) and
𝜓
(tangential). They can be computed applying
197
the following sequence [16].198
1.
Compute the coefficients of first (
𝐸, 𝐹 , 𝐺
) and second (
𝐿, 𝑀 , 𝑁
) fundamental forms of the
199
Zernike polynomial wavefront with cartesian parametric representation, using first, second
200
derivatives and normals (N) of the surface with respect x and y [16].201
2.
Compute the mean (
𝐻
) and Gaussian (
𝐾
) curvatures from
𝐸, 𝐹 , 𝐺
and
𝐿, 𝑀 , 𝑁
using
202
equations:203
𝐻=𝐸 𝑁 2𝐹 𝑀 +𝐺 𝐿
𝐸𝐺 𝐹2(17)
𝐾=𝐿𝑁 𝑀2
𝐸𝐺 𝐹2
3. Obtain the two principal curvatures (𝑘 𝑛) from 𝐻and 𝐾applying:204
𝑘1=𝐻+𝐻2𝐾(18)
𝑘2=𝐻𝐻2𝐾
4. Finally, the equations providing caustic curves, are:205
𝜒(𝑥, 𝑦, 𝑧)=𝑊(𝑥, 𝑦, 𝑧) + 𝑘1(𝑥, 𝑦, 𝑧)N(𝑥, 𝑦, 𝑧)(19)
𝜉(𝑥, 𝑦, 𝑧)=𝑊(𝑥, 𝑦, 𝑧) + 𝑘2(𝑥, 𝑦, 𝑧)N(𝑥, 𝑦, 𝑧)
To characterize how the final wavefront deviates from being umbilical at all points, i.e. one
206
provided a single caustic spike, we computed the difference between the two sheet caustic locations,
207
which is itself a surface that from now on we refer to as caustic astigmatism. Additionally, as an
208
average metric evaluation of that condition, we used the root-mean-square of caustic astigmatism,
209
defined as:210
𝑅𝑀𝑆𝑎=Í𝑛
𝑖=1(𝜒𝑥
𝑖𝜉𝑥
𝑖)2+ ( 𝜒𝑦
𝑖𝜉𝑦
𝑖)2+ ( 𝜒𝑧
𝑖𝜉𝑥
𝑖)2
𝑛,(20)
where 𝜒and 𝜉super-indexes denote the 𝑥, 𝑦, 𝑧 components of the caustic location.211
Additionally, we also tracked the difference between the sagittal caustic (
𝜒
) –the caustic sheet
212
closer to the optical axis and the nominal prescribed depth-of-focus line; this surface is from
213
now on referred to as sagittal error. And, as before, we used the root-mean-square as an average
214
metric:215
𝑅𝑀𝑆𝑥=Í𝑛
𝑖=1(𝜒𝑥
𝑖)2+ ( 𝜒𝑦
𝑖)2+ ( 𝜒𝑧
𝑖𝑔(𝑥))2
𝑛,(21)
where 𝑔(𝑥)is the z-coordinate of the prescribed focal line as given by Eq. 1.216
3. Simulations and numerical results217
We generated two different wavefronts associated with the example case described in section
218
2. First, a depth-of-focus wavefront, where we applied all steps of the proposed method except
219
the one pursuing the pseudo-umbilical condition (subsection 2.3). Second, a depth-of-focus
220
wavefront where we fully applied the proposed method including the step trying to attain the
221
pseudo-umbilical condition along the embedded circle involute. For reasons of economy of
222
notation, from now on we will refer to the former as the circle-involute generated wavefront
223
(CIGE), and the latter as the circle-involute pseudo-umbilical generated wavefront (CIPU).224
We compared the optical performance of both, algorithms and their output wavefronts (CIGE
225
and CIPU), in terms of caustic astigmatism and sagittal error and their associated metrics:
𝑅𝑀𝑆𝑎
226
and
𝑅𝑀𝑆𝑥
. We did that in order to analyze the always-present trade-off in multifocal wavefront
227
design between astigmatism correction and axial focal prescription (see discussion at [8]). The
228
results depend on several numerical sampling patterns, being the most relevant one the size of
229
the circle generating the circle-involute curve, i.e. the parameter
𝑠0
, which controls how dense
230
the wavefront surface is sampled by the involute circle.231
We empirically found that the most optimal results are obtained for
𝑠0
ranging between
0.7232
and
1.1𝜇𝑚
. We also found that the optimal number of Zernike least-squares polynomial fitting
233
is (fifth-order expansion). This may be due to the fact that using a high number of polynomials
234
sometimes generates ill-conditioned reconstructed matrices, a typical problem in least-squares
235
reconstruction.236
Fig. 4. Caustic astigmatism (mm) of (a-b) CIGE and (c-d) CIPU wavefronts when
𝑠0=0.7𝜇𝑚
. Lateral projections on an (a-c) xz-plane and (b-d) yx-plane. Colormaps
of the surfaces are proportional to the 𝑅 𝑀 𝑆 𝑎of caustic astigmatism.
Table 1 shows comparing results of
𝑅𝑀𝑆𝑎
and
𝑅𝑀𝑆𝑥
for the two wavefronts under test: CIGE
237
and CIPU. The data in the table reveals that while for the CIPU wavefront, both
𝑅𝑀𝑆𝑎
and
238
𝑅𝑀𝑆𝑥are highly dependent on 𝑠0, those values are more stable for the CIGE wavefront.239
Reducing the value of
𝑠0
substantially reduces
𝑅𝑀𝑆𝑎
in the CIPU wavefront, both in absolute
240
terms –a
56.46%
from
𝑠0=1.1
to
𝑠0=0.7𝜇𝑚
and in comparison to the CIGE wavefront: at
241
𝑠0=1.1
the
𝑅𝑀𝑆𝑎
is
16.04%
smaller but at
𝑠0=0.7
this reduction is improved up to
63.01%
.
242
However, that
𝑅𝑀𝑆𝑎
reduction when reducing
𝑠0
is achieved at the expense of a slight increase
243
in the value of
𝑅𝑀𝑆𝑥
; a
13.3%
from
𝑠0=1.1
to
𝑠0=0.7𝜇𝑚
. On the other hand, for the CIGE
244
there is only a small difference in
𝑅𝑀𝑆𝑎
:
1.15%
between the two limit
𝑠0
values. Significantly,
245
although for CIGE
𝑅𝑀𝑆𝑥
is always bellow
𝑅𝑀𝑆𝑎
, this does not occur with CIPU, where
𝑅𝑀𝑆𝑎
246
is substantially smaller (52.61%) than 𝑅𝑀 𝑆𝑥at 𝑠0=0.7𝜇𝑚.247
𝑅𝑀𝑆𝑥
is always smaller for CIGE with respect to CIPU: at
𝑠0=1.1
the relative difference, in
248
percentage, is 14.7%, and, at 𝑠0=0.7this difference is increased up to 25.15%.249
We analyzed the surface structure of caustic astigmatism and sagittal error when
𝑠0=0.7𝜇𝑚
.
250
The lateral projections on an xz-plane and y-x plane of these surfaces are shown in figures 4
251
and 5, respectively. The colormap mappings of caustic astigmatism and sagittal surfaces are
252
proportional to the 𝑅 𝑀 𝑆𝑎and 𝑅 𝑀 𝑆 𝑥, respectively.253
Caustic astigmatism contains a cusp point (singular point) where both caustic sheets meet
254
(
𝑅𝑀𝑆𝑎=0
). From that point, caustic astigmatism evolves forming a horn-shaped surface.
255
Comparing CIGE and CIPU caustic astigmatism, there is a considerable reduction in the surface
256
size but the horn-shaped morphology is preserved.257
In both, CIGE and CIPU, the separation of the sagittal error from the optical axis is small;
258
𝑠0(𝜇𝑚) 0.7 0.8 0.9 1 1.1
𝑅𝑀𝑆𝑎
(CIGE) 513.72 513.11 512.44 512.99 519.71
𝑅𝑀𝑆𝑎
(CIPU) 190.00 242.20 302.11 365.8 436.34
𝑅𝑀𝑆𝑥
(CIGE) 300.09 300.85 301.27 300.95 301.81
𝑅𝑀𝑆𝑥
(CIPU) 400.91 387.84 375.25 363.95 353.83
Table 1.
𝑅𝑀 𝑆𝑎
and
𝑅𝑀 𝑆𝑜
optical performance values as function of
𝑠0
for CIGE and
CIPU multifocal wavefronts.
always below
1.5𝜇𝑚
. Most of the error is concentrated in differences along the z-coordinate
259
as clearly shown in 5(a-c). In both cases, the sagittal error is highly oscillatory, but the higher
260
values for CIPU, in comparison to CIPU, are due to a higher z-coordinate extension of sagittal
261
error.262
4. Discussion263
We have presented a design method, based on differential geometry, aiming at designing depth-of-
264
focus wavefronts where the optical power variation –associated to mean curvature variation– is
265
produced, not along a meridian as occurs in classical axial axicons, but along a spiral-type circle
266
involute filling the surface. A classical axicon, i.e. a conical surface, has two major drawbacks:
267
first, not equal areas of the refractive surface provide uniform intensity, and, second, the axial
268
symmetry makes the design sensitive to pupil variations. Our proposal overcomes both issues.269
Our method can also, as an alternative, control the separation between caustic sheets (hence
270
improving the energy concentration) by means of pursuing pseudo-umbilicity along the circle
271
involute space curve. The radius of the circle generating the involute is a critical factor to control
272
the separation between caustic sheets. As seen in section 3, reducing its value helps to decrease
273
that difference. However, this has a theoretical threshold, because when the distance between
274
the circle involute loops is too close the umbilicity condition of the tubular neighbors starts
275
to interfere and, as mentioned in section 1, the impossibility of a smooth surface to contain
276
an extended region of umbilical points becomes apparent. It is important to point out that
277
the reduction of the separation between caustic sheets obtained by our method faces a design
278
trade-off; namely, the amount of concentration of light along the prescribed axis segment and the
279
separation between caustics. This trade-off, already discussed in axially symmetric multifocal
280
wavefronts [8], is intrinsic to multifocality because of the general laws of differential geometry in
281
smooth surfaces [10]. However, we comment that the error in sagittal, concentrated along the
282
z-coordinate as analyzed in section 3, could be, in practice, corrected a posteriori by means of
283
readjusting the mean curvature at the apex of the wavefront.284
5. Disclosures285
The authors declare no conflicts of interest.286
6. Data Availability Statement287
Data availability. No data were generated or analyzed in the presented research.288
Fig. 5. Sagittal errors (
𝜇𝑚
) of (a-b) CIGE and (c-d) CIPU wavefronts when
𝑠0=0.7
𝜇𝑚
. Lateral projections on an (a-c) xz-plane and (b-d) yx-plane. Colormaps of the
surfaces are proportional to the 𝑅 𝑀 𝑆 𝑥of sagittal errors.
References289
1.
S. Barbero, “Smooth multifocal wavefronts with prescribed mean curvature for visual optics applications, Appl. Opt.
290
60(21), 6147–6154 (2021).291
2.
T. Zhao, T. Mauger, and G. Li, “Optimization of wavefront-coded infinity-corrected microscope systems with
292
extended depth of field,” Biomed. Opt. Express 4(8), 1464–1471 (2013).293
3.
F. Courvoisier, J. Zhang, M.K. Bhuyan, M. Jacquot, and J.M. Dudley, Applications of femtosecond Bessel beams to
294
laser ablation,” Appl. Phys. A 112, 29-–34 (2013).295
4. D. McGloin, and K. Dholakia, “Bessel beams: Diffraction in a new light, Contem. Phys., 46(1), 15–28 (2005).296
5.
L. Froehly, F. Courvoisier, A. Mathis, M. Jacquot, L. Furfaro, R. Giust, P.A. Lacourt, and J.M. Dudley, Arbitrary
297
accelerating micron-scale caustic beams in two and three dimensions, Opt. Express. 19, 16455–16465 (2011).298
6.
A. Zannotti, C. Denz, M.A. Alonso, “Shaping caustics into propagation-invariant light, Nat. Commun. 11, 3597
299
(2020).300
7.
E. Espíndola-Ramos, G. Silva-Ortigoza, C.T. Sosa-Sánchez, I. Julián-Macías, O. Cabrera-Rosas, P. Ortega-Vidals,
301
A. González-Juárez, R. Silva-Ortigoza, M.P. Velázquez-Quesada, and G.F. Torres del Castillo, “Paraxial optical fields
302
whose intensity pattern skeletons are stable caustics,” J. Opt. Soc. Am. A 36, 1820–1828 (2019).303
8.
S. Barbero, “Multifocal wavefronts with prescribed caustics in axially symmetric optical systems, Opt. Express 30,
304
14274–14286 (2022).305
9.
R. Vila, J. Portilla, and S. Barbero, “Robust numerical solution to the Levi-Civita wavefront coupling problem via
306
level set computation of the point characteristic function, J. Opt. Soc. Am. A 40, 277–284 (2023).307
10.
S. Barbero, and M.M. Gonzalez, “Admissible surfaces in progressive addition lenses,” Opt. Lett. 45, 5656–5659
308
(2020).309
11.
M.W. Beijersbergen, R.P.C. Coerwinkel, M. Kristensen, J.P. Woerdman, “Helical-wavefront laser beams produced
310
with a spiral phaseplate,” Optics Communications, 112(5-6), 321-327 (1994).311
12.
G. Mikula, Z. Jaroszewicz, A. Kolodziejczyk, K. Petelczyc, and M. Sypek, “Imaging with extended focal depth by
312
means of lenses with radial and angular modulation,” Optics Express, 15(15), 9184–9193 (2007).313
13.
Y. Shen, X. Wang, Z. Xie, C. Min, X. Fu, Q. Liu, M. Gong, and X. Yuan, “Optical vortices 30 years on: OAM
314
manipulation from topological charge to multiple singularities,” Light. Sci. Appl. 890 (2019).315
14.
R.. Uribe-Vargas, “On vertices, focal curvatures and differential geometry of space curves, Bull Braz Math Soc, New
316
Series 36, 285—307 (2005).317
15.
S. Barbero, and M. Ritore “Circle involute as an optimal scan path, minimizing acquisition time, in surface topography,”
318
Surf. Topogr. Metrol. Prop. 7(21), 035010 (2019).319
16. J. J. Stoker, Differential geometry (Wiley-Interscience, New York, 1969).320
17.
S. Izumiya, M. C. Romero, M.A. Soares, F. Tari, Differential geometry from a singularity theory viewpoint (World
321
Scientific Publishing, Singapore, 2016).322
18.
C. Gutierrez, and J. Sotomayor, “Structural stable configurations of lines of principal curvature,” Asterisque 9899,
323
185–215 (1982).324
19.
R. Garcia, and J. Sotomayor, “On the patterns of principal curvature lines around a curve of umbilic points,” Anais
325
da Academia Brasileira de Ciências, 771, 13–24 (2005)326
20.
B. Chebbi and I. Golub, “Development of spot size and lateral intensity distribution generated by exponential,
327
logarithmic, and linear axicons," ,” J. Opt. Soc. Am. A 31, 2447–2452 (2014).328
21.
R.G. Gonzalez-Acuña, and J.C. Guitiérrez-Vega, “Generalization of the axicon shape: the gaxicon,” J. Opt. Soc. Am.
329
A35, 1915–1918 (2018).330
22. O.N. Stavroudis, The optics of rays, wavefronts, and caustics (New York, Academic, 1972).331
23.
I.A. Porteous, Geometric differentiation. For the Intelligence of curves and surfaces. Second Ed. (Cambridge,
332
Cambridge University, 2001).333
24.
L.N. Thibos, R.A. Applegate, J.T. Schwiegerling, and R. Webb R, “Standards for reporting the optical aberrations of
334
eyes, J. Refract. Surg. 18(5), S652–60 (2002).335
25. O.N. Stavroudis, The optics of rays, wavefronts, and caustics (New York, Academic, 1972).336
26. D.R. Struik, Lectures on classical differential geometry, (Dover, 2003).337
27.
P. Gracia, C. Dorronsoro, and S. Marcos, “Multiple zone multifocal phase designs,” Opt. Lett. 38(18), 3526-3529
338
(2013).339
28.
M. Vinas, C. Dorronsoro, V. Gonzalez, D. Cortes, A. Radhakrishnan, and S. Marcos, “Testing vision with angular
340
and radial multifocal designs using Adaptive Optics,” Vision. Res. 132, 85–96 (2017).341
29.
L.A. Romero, A.G. Marrugo, M.S. Millán, “Trade-Off asymmetric profile for extended-depth-of-focus ocular lens,”
342
Photonics. 9(2), 119 (2022).343
30.
L.N. Thibos, R.A. Applegate, J.T. Schwiegerling, and R. Webb R, “Standards for reporting the optical aberrations of
344
eyes, J. Refract. Surg. 18(5), S652–60 (2002).345
ResearchGate has not been able to resolve any citations for this publication.
Article
Full-text available
The Levi-Civita wavefront coupling problem consists of, given two prescribed wavefronts, obtaining a refractive or reflective surface coupling them. We propose a robust numerical method to solve Levi-Civita’s problem, whose rationale is to consider that Levi-Civita’s solutions are level surfaces of the point characteristic function established between points of incoming and outgoing wavefronts. The method obtains both surface data points and their normals, enabling a more robust surface reconstruction. We carry out a detailed error analysis of our method by means of comparing the surface data estimation with nominal surfaces obtained in reference tests offering analytical solutions to Levi–Civita’s problem. The method offers, in computer simulations, highly accurate results with moderate computational cost.
Article
Full-text available
Multifocal and/or extended depth-of-focus designs are widely used in many optical applications. In most of them, the optical configuration has axial symmetry. A usual design strategy consists of exploring the optimal wavefronts that emerging out of the optical system would provide the desired multifocal properties. Those properties are closely related to light concentration on caustic surfaces. We present a systematic analysis of how to obtain those multifocal wavefronts given some prescriptions on the locations of caustics. In particular, we derive several multifocal wavefronts under archetypical prescriptions in the sagittal caustic alone, or combined with the tangential one at certain points, with some emphasis on visual optics applications.
Article
Full-text available
We explore the possibility of extending the depth of focus of an imaging lens with an asymmetric quartic phase-mask, while keeping the aberration within a relatively low level. This can be intended, for instance, for ophthalmic applications, where no further digital processing can take place, relying instead on the patient’s neural adaptation to their own aberrations. We propose a computational optimization method to derive the design-strength factor of the asymmetric profile. The numerical and experimental results are shown. The optical experiment was conducted by means of a modulo-2π phase-only spatial light modulator. The proposed combination of the asymmetric mask and the lens can be implemented in a single refractive element. An exemplary case of an extended-depth-of focus intraocular lens based on the proposed element is described and demonstrated with a numerical experiment.
Article
Full-text available
Multifocal lenses comprising progressive power surfaces are commonly used in contact and intraocular lens designs. Given a visual performance metric, a wavefront engineering approach to design such lenses is based on searching for the optimal wavefront at the exit pupil of the eye. Multifocal wavefronts distribute the energy along the different foci thanks to having a varying mean curvature. Therefore, a fundamental step in the wavefront engineering approach is to generate the wavefront from a prescribed mean curvature function. Conventionally, such a thing is done by superimposing spherical wavefront patches and maybe adding a certain component of spherical aberration to each spherical patch in order to increase the depth-of-field associated with each focus. However, such a procedure does not lead to smooth wavefront solutions and also restricts the type of available multifocal wavefronts. We derive a new, to the best of our knowledge, mathematical method to uniquely construct multifocal wavefronts from mean curvature functions (depending on radial and angular coordinates) under certain numerically justified approximations and restrictions. Additionally, our procedure leads to a particular family of wavefronts (line-umbilical multifocal wavefronts) described by 2 conditions: (1) to be smooth multiplicative separable functions in the radial and angular coordinates; (2) to be umbilical along a specific segment connecting the circle center with its edge. We provide several examples of multifocal wavefronts belonging to this family, including a smooth variant of the so-called light sword element.
Article
Full-text available
Progressive addition lenses (PALs) contain a surface of spatially varying curvature, which supplies variable optical power for different viewing areas over the lens. We derive complete compatibility equations providing the exact magnitude of a cylinder along lines of curvature on any arbitrary PAL smooth surface. These equations reveal that, contrary to current knowledge, the cylinder and its derivative depend not only on the principal curvature and its derivatives along the principal line but also on the geodesic curvature and its derivatives along the line orthogonal to the principal line. We quantify the relevance of the geodesic curvature through numerical computations. We also derive an extended and exact Minkwitz theorem restricted only to be applied along lines of curvature, but excluding umbilical points.
Article
Full-text available
Structured light has revolutionized optical particle manipulation, nano-scaled material processing, and high-resolution imaging. In particular, propagation-invariant light fields such as Bessel, Airy, or Mathieu beams show high robustness and have a self-healing nature. To generalize such beneficial features, these light fields can be understood in terms of caustics. However, only simple caustics have found applications in material processing, optical trapping, or cell microscopy. Thus, these technologies would greatly benefit from methods to engineer arbitrary intensity shapes well beyond the standard families of caustics. We introduce a general approach to arbitrarily shape propagation-invariant beams by smart beam design based on caustics. We develop two complementary methods, and demonstrate various propagation-invariant beams experimentally, ranging from simple geometric shapes to complex image configurations such as words. Our approach generalizes caustic light from the currently known small subset to a complete set of tailored propagation-invariant caustics with intensities concentrated around any desired curve.
Article
Full-text available
We construct exact solutions to the paraxial wave equation in free space characterized by stable caustics. First, we show that any solution of the paraxial wave equation can be written as the superposition of plane waves determined by both the Hamilton–Jacobi and Laplace equations in free space. Then using the five elementary stable catastrophes, we construct solutions of the Hamilton–Jacobi and Laplace equations, and the corresponding exact solutions of the paraxial wave equation. Therefore, the evolution of the intensity patterns is governed by the paraxial wave equation and that of the corresponding caustic by the Hamilton–Jacobi equation.
Article
Full-text available
Thirty years ago, Coullet et al. proposed that a special optical field exists in laser cavities bearing some analogy with the superfluid vortex. Since then, optical vortices have been widely studied, inspired by the hydrodynamics sharing similar mathematics. Akin to a fluid vortex with a central flow singularity, an optical vortex beam has a phase singularity with a certain topological charge, giving rise to a hollow intensity distribution. Such a beam with helical phase fronts and orbital angular momentum reveals a subtle connection between macroscopic physical optics and microscopic quantum optics. These amazing properties provide a new understanding of a wide range of optical and physical phenomena, including twisting photons, spin-orbital interactions, Bose-Einstein condensates, etc., while the associated technologies for manipulating optical vortices have become increasingly tunable and flexible. Hitherto, owing to these salient properties and optical manipulation technologies, tunable vortex beams have engendered tremendous advanced applications such as optical tweezers, high-order quantum entanglement, and nonlinear optics. This article reviews the recent progress in tunable vortex technologies along with their advanced applications.
Article
Full-text available
We generalize the shape of the traditional axicon by analytically finding the function of the output surface when the input surface is not flat but an arbitrary continuous function that possesses rotational symmetry. Several illustrative examples are presented and tested using ray tracing techniques without the paraxial approximation.