ArticlePDF Available

General formula for bi-aspheric singlet lens design free of spherical aberration

Optica Publishing Group
Applied Optics
Authors:

Abstract and Figures

In this paper, we present a rigorous analytical solution for the bi-aspheric singlet lens design problem. The input of the general formula presented here is the first surface of the singlet lens; this surface must be continuous and such that the rays inside the lens do not cross each other. The output is the correcting second surface of the singlet; the second surface is such that the singlet is free of spherical aberration.
This content is subject to copyright. Terms and conditions apply.
General formula for bi-aspheric singlet lens
design free of spherical aberration
RAFAEL G. GONZÁLEZ-ACUÑA1,*AND HÉCTOR A. CHAPARRO-ROMO2,3
1Optics Center, Physics Department, Tecnológico de Monterrey Garza Sada 2501, Monterrey, N.L. 64849, Mexico
2Av. Universidad No. 3000, Universidad Nacional Autónoma de México, C.U., Mexico City, Mexico
3e-mail: moxaika@live.com
*Corresponding author: rafael123.90@hotmail.com
Received 24 August 2018; revised 3 October 2018; accepted 3 October 2018; posted 5 October 2018 (Doc. ID 343158); published 25 October 2018
In this paper, we present a rigorous analytical solution for the bi-aspheric singlet lens design problem. The input
of the general formula presented here is the first surface of the singlet lens; this surface must be continuous and
such that the rays inside the lens do not cross each other. The output is the correcting second surface of the singlet;
the second surface is such that the singlet is free of spherical aberration. © 2018 Optical Society of America
https://doi.org/10.1364/AO.57.009341
1. INTRODUCTION
The design of optical systems with aspheric surface has the goal
to strongly reduce spherical aberration. Spherical aberration on
lenses has been extensively studied by Ref. [1]. Luneberg [2]
established a method for computing the shape of the second
surface from an initial first surface that introduces spherical
aberration, which he described just for special cases. Many
authors proposed a lens design with two aspheric surfaces to
correct spherical aberration [3,4].
The problem of the design of a singlet free of spherical aber-
ration with two aspheric surfaces is also known as the
Wasserman and Wolf problem [5]. The problem has been
solved with a numerical approach by Ref. [6]. Recently, Ref. [7]
has shown a rigorous analytical solution of a singlet lens free of
spherical aberration for the special case when the first surface is
flat or conical. Since its publication, several works inspired by
its solution have emerged [712], all of them free of spherical
aberration. The solution has six different signs; therefore, it is a
set of 2664 possible solutions, where only one is right. We
test the formula provided by those in Ref. [7], when the first
surface is not flat or conic, and the equation system does not
give correct answers.
In this paper, we present a rigorous analytical solution for
the design of lenses free of spherical aberration. The solution
presented here has just one sign; therefore, it is a set of just two
possible solutions. Our solution is robust because the set of
solutions is valid for negative and positive refraction indices.
The model allows use of continuous functions, such that
the rays inside the lens do not cross each other. This model
will compute the second surface in order to correct the spherical
aberration produced by the first surface.
2. MATHEMATICAL MODEL
The goal is to determine the shape of the second surface
rb,zb, given a first surface ra,za, in order to correct the
spherical aberration generated by the first surface. Therefore
the objective is to find rb,zbgiven ra,za, where rais the
only independent variable, and zb,rb, and zaare functions of
ra. The origin of the coordinate system is located at the center
of the input surface za00.
We assume that the singlet lens has refraction index nand is
radially symmetric. At the center, the singlet lens has thickness
t. The distance from the object to the first surface is ta. The
distance from the second surface to the image is tb, as can be
seen in Fig. 1.
The first fundamental equation for this model is the vector
form of the Snells law:
v21
nna×na×v1 naffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1
1
n2na×v1·na×v1
r,
(1)
where v1is the unitary vector of the incident ray, v2is the uni-
tary vector of the refracted ray, and nais the normal vector of
the first surface, as seen in Fig. 2.
The unitary vectors written of the first surface are
8
>
>
>
>
>
<
>
>
>
>
>
:
v1ra,zata,0
ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
r2
azata2
p,
v2rbra,zbza,0
ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
rbra2zbza2
p,
naz0
a,1,0
ffiffiffiffiffiffiffiffi
1z02
a
p,
(2)
where z0
ais the derivative with respect to raof the sagitta of the
first surface.
Research Article Vol. 57, No. 31 / 1 November 2018 / Applied Optics 9341
1559-128X/18/319341-05 Journal © 2018 Optical Society of America
Replacing Eq. (2) in Eq. (1) and separating the Cartesian
components, we get the following expressions for the direction
cosines ri,ziof the vector:
8
>
>
>
>
>
>
>
>
>
>
>
<
>
>
>
>
>
>
>
>
>
>
>
:
rirbra
Ψzataz0
ara
nffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
r2
ataza2
p1z02
a
z0
aΦ,
zizbza
Ψrazataz0
az0
a
nffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
r2
ataza2
p1z02
aΦ,
with,
Ψffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
zbza2rbra2
p
Φffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1razataz0
a2
n2r2
ataza21z02
a
qffiffiffiffiffiffiffiffi
1z02
a
p
(3)
and, evidently, z2
ir2
i1. Relations (3) come from the ap-
plication of the Snell law for an arbitrary ray striking the singlet
lens at point ra,za. Note that the expressions in the right sides
of Eq. (3) are fully expressed in terms of the coordinates of the
input surface, i.e., zizira,zaand so on.
Now let us focus on the Fermats principle, which is the sec-
ond fundamental equation of our model. We assume for a
spherical aberration-free singlet lens, the optical path of any
non-central ray must be equal to the optical path of the axial ray:
tant tbsgntaffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
r2
azata2
p
nffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
rbra2zbza2
p
sgntbffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
r2
bzbttb2
q,(4)
where sgntaand sgntbare the sign functions of the variable ta
or tb, respectively.
Now we have a system of three equations: two components
of vector form of Snells law, Eq. (3), and Fermats principle,
Eq. (4). Furthermore, we have two unknowns, rband zb.
The solution of the system is
8
>
>
>
>
>
<
>
>
>
>
>
:
rbrizbza
zira,
zb
h0s1ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
z2
i2nfizizatbtzi1raritr2
i
rizatbtrazi2f2
ih1n2
s
1n2
,
(5)
where s1is the only sign; it comes from the fact that when the
refraction index is positive, the rays are refracted to the opposite
direction when the refraction index is negative. Also, we define
the following auxiliary variables:
8
>
>
>
>
<
>
>
>
>
:
fisgntaffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
r2
ataza2
ptatb,
h0nfizin2tzizar2
izarariziz2
ittb,
h1r2
a2rarittbza2t2r2
i1zi2
2ttbza1zi:
(6)
Equation (5) is the most important result in this work. It could
look cumbersome, but it is notorious that it could be expressed
in closed form for an almost arbitrary freeform input surface.
The condition for the validity of Eq. (5) is that the surface nor-
mal should be perpendicular to the tangent plane to the input
surface at the origin, and the rays do not cross each other inside
the lens.
From a topological point of view, since the singlet lens is a
homogeneous optical element, the input and output surfaces
are simply connected sets on R2that can be defined as
Ωafra,zaR2jza<z
bg,
Ωbfrb,zbR2jzb>z
ag,(7)
where Ωais homeomorphic to Ωb, and both surfaces are topo-
logically equivalent. Thus, there exists a continuous and bijec-
tive function f, such that f:ΩaΩb, and whose inverse f1
is continuous.
There are many functions fthat map both sets, but there is
only one that is physically valid and corresponds to the one that
satisfies the variational Fermat principle of minimum optical
length, which is constant. In our case, fis given by Eq. (5).
The uniqueness of fhas as consequence that the Snell law is
automatically fulfilled at the second interface zbas well.
Now, since fis continuous, it means that fmaps open balls
from Ωato Ωb, then the ray neighborhoods are preserved.
Therefore, the validity of Eq. (5) also requires that the rays
do not intersect each other inside the lens because, in the case
when the rays inside cross each other, Ωboverlaps itself, and
then Ωbis no longer a simple connected set; it is not homeo-
morphic with respect to Ωa, the vicinity of the neighborhoods
is not preserved, and finally we do not have a homogeneous
optical element.
-20 -10 10 20 30
-10
-5
5
-30 40
-z z
r
-r
OI
o
Fig. 1. Diagram of a singlet lens free of spherical aberration. The
first surface is given by ra,za, and the second surface is given by
rb,zb. The distance between the first surface and the object is ta,
the thickness at the center of the lens is t, and the distance between
the second surface and the image is tb.
r
z
r
z
Fig. 2. Left: Zoom at point Pof Fig. (1); three unitary vectors can
be seen: v1is the unitary vector of the incident ray, v2is the unitary
vector of the refracted ray, and nais the normal vector of the first sur-
face. Right: Zoom at point Qof Fig. (1); unitary vectors of the second
surface can be seen: v2is the unitary vector of the incited ray, v3is the
unitary vector of the refracted ray, and nbis the normal vector of the
second surface.
9342 Vol. 57, No. 31 / 1 November 2018 / Applied Optics Research Article
3. RESULTS AND EXAMPLES
In this section, we report some illustrative results of Eq. (5), in
six different contexts, with the objective to show the capacity of
our model. We compare our results with Ref. [7], then we in-
ject the model with functions that have never been evaluated,
and we observe that they satisfy the optical path conditions; we
load the system with a fixed input function and change the
distance between the image and the second surface, in order
to show that the model adjusts the second surface and corrects
the spherical aberration. Then, we show an example of negative
refraction index, and an example when the object is at minus
infinity, ta, and finally we compute the efficiency
of Eq. (5).
A. Comparison Between Solutions
Table 1shows the results of the comparison of the solution
given in Ref. [7] with ours, for plane, spherical, parabolic,
and cosine cases. The sagitta for each surface corresponds to
flat za0, parabolic-aspheric case zar2
a200, spheric-
aspheric case za100 ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1002r2
a
p, and cosine-aspheric case
zacosra3. In addition, the error difference between both
solutions is reported.
The difference is computed by the following formula:
RS Error X
N
zbzbc2
N,(8)
where Nis the number of samples, zbis the sagitta of Eq. (5),
and zbc is the sagitta of the solution of Ref. [7]. We evaluate the
values of the first quadrant because the solutions provided by
Ref. [7] are not symmetric and fail in the second quadrant. For
the evaluation, we use a sample of N500.
With the data in Table 1, we can argue that our solution in
the first quadrant and for conic surfaces is equal to the solution
of Ref. [7]. Also, it is important to mention that all this
computation was made without the point zbc 0, because in
the method of Ref. [7], that point diverges and for our method
at the same point, we have zb0tza0.
B. Gallery of Lenses with New Solution
This time, the first surfaces are not conic. The model is robust
enough that it can admit nonstandard functions as first surfa-
ces; the sagittae of the first functions of Fig. 3are
8
>
>
>
>
>
>
<
>
>
>
>
>
>
:
azaexpraJ0rawith,ra<0
expraJ0rawith,ra0,
bza1
18 r2
acosra
2,
cza5 cosra
5,
dza1
18 r2
acosra
2,
(9)
where J0is the Bessel function of order zero. Therefore, the case
when the first surface is the first term of Eq. (9) is the plot (a) in
Fig. 3, the case when the first surface is the second term of
Eq. (9) is the plot (b) in Fig. 3and so on.
The shapes presented in Fig. 3are illustrative examples of
the many examples computed in [13]. The general formula still
gives a correct second surface in order to get a singlet lens free of
spherical aberration. Please see Appendix A.
C. Performance of the Formula
The same surface is evaluated for different object distances, and
the system shows the capacity to generate the solution such as
the singlet is free of spherical aberration in the image point. In
Fig. 4, we change tband keep fixed the sagitta zaJ0ra.We
find that the maximum diameter of the lens is proportional
to tb.
Table 1. Benchmark of Resolver Power Between
Analytical Solutions to Design of Singlet Lenses Free of
Spherical Aberration
Solution According
to J.C.V.E
Solution According to
R.G.G.A. and H.A.C.R.
Basic configuration for all cases: n1.5and s11for (1) flat, (2) spherical,
(3) paraboloid, and (4) cosine; we have: (1) ta70 mm,t10 mm,
tb75 mm,RS 1.20 ×1029. (2) ta70 mm,t8mm,
tb80 mm,RS 4.54 ×1028. (3) ta80 mm,t10 mm,
tb100 mm,RS 1.75 ×1028. (4) ta70 mm,t8mm,
tb80 mm,RS=.
rr
O
z
r
I
I
I
I
O
O
O
z
z
z
0
(a)
(b)
(c)
(d)
Fig. 3. Gallery of singlet lenses bi-aspherical free of spherical aber-
ration. For all cases, the configuration is ta70 mm,t8mm,
tb75 mm, and n1.5.
Research Article Vol. 57, No. 31 / 1 November 2018 / Applied Optics 9343
D. Example for Negative Refraction Index
In Fig. 5, the ray tracing can be seen for a lens with negative
refraction index n1.5, for zaJ0ra2; therefore, we
take the sign as s11. It can be seen that for a negative re-
fraction index, both sets Ωa,Ωbdo not cross each other,
ΩaΩb; therefore maximum radius tends to infinity.
E. Example When the Object Is at Minus Infinity
Now, when the object is far away, ta, we need to com-
pute the limit when tafor fi,ri,zi, which can be
written as
8
>
>
>
>
>
>
>
>
>
<
>
>
>
>
>
>
>
>
>
:
lim
ta
fizatb,
lim
ta
ziffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
n21z02
an2
n2z02
a1
qffiffiffiffiffiffiffiffi
z02
a1
pz02
a
nz 02
an,
lim
ta
ri
z0
anffiffiffiffiffiffiffiffi
z02
a1
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
n21z02
an2
n2z02
a1
q1
nz02
a1
:(10)
Then, we can use Eq. (5) with the parameters of Eq. (10)
and test it. For illustration proposes, we choose za
cosra2r2200 and plot the ray tracing in Fig. 6. From
the figure, it is clear that the output rays converge to the image
point despite ta.
F. Efficiency
To validate the efficiency of Eq. (5) we compare two vectors, v3
and v
3;v3comes from the image to the second surface, and v
3
is computed using Snelllaw with the normal vector of the sec-
ond surface, nb, and the unitary vector, v2(Fig. 2); therefore,
nb,v3,v
3are written as
8
>
>
>
>
>
<
>
>
>
>
>
:
nbz0
b,r0
b,0
ffiffiffiffiffiffiffiffiffiffi
z02
br02
b
p,
v3rb,zbttb,0
ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
r2
bzbttb2
p,
v
3nnb×nb×v2 nbffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1n2nb×v2·nb×v2
p
:
(11)
The percentage efficiency of ray is measured how close ends in
the image position, thus percentage efficiency is given by
E100%
v
3v3
v3
×100%:(12)
We compute the efficiency for 550 rays for all the examples
presented in the paper, and the average of all the examples
is 99.999999999999860%100%. We believe that the error
is not zero because the equations are very large, and when
evaluated, we get computational errors such as truncation.
Please notice that for all examples, the singlets are free of
spherical aberration even when the incident angles are very large;
this happens because we do not use any paraxial approximation.
4. CONCLUSION
In this paper, we generalized the second surface in a lens in
order to get an image free of spherical aberration for a given
arbitrary first surface. When the first surface is such as the rays
-10 -5 5 10
80
O
I
z
60
40
20
-20
60
z
40
20
-20
5
-5
-10 10
O
I
20
z
r
O
I
-20
-5 5
(a) (b) (c) (d)
Fig. 4. (a) tb75 mm, (b) tb55 mm, (c) tb35 mm,
(d) tb15 mm. The constant configuration of cases (a)(d) is
ta30 mm,t5mm,n1.5.
Fig. 5. Configuration of the singlet lens with negative refraction
index: ta20 mm,t3mm,tb30 mm,n1.5.
Fig. 6. Configuration of the singlet lens with object at infinity:
t3.5mm,tb50 mm,n1.5.
9344 Vol. 57, No. 31 / 1 November 2018 / Applied Optics Research Article
inside the lens do not cross each other. We exhaustively test the
formula by ray tracing, and we find that the solution presented
works for positive and negative refractive indices. For a negative
refraction index, we take s11.
This general formula expands the variety of lenses free of
spherical aberration. The efficiency of the method allows to de-
sign robust optical systems whose first surfaces are not restricted
to the conical family functions.
APPENDIX A
We share the program code in Mathematica language to graph
surfaces free of spherical aberration. The program is composed
of three sections: the first one declares Eq. (5); in the second
section, constants of the optical system are defined; finally, the
graphing process is performed for the functions of the first sur-
face and the second surface simultaneously.
Singlet free of spherical aberration.
1
2 (* Clean register. *)
3Clear[Global*]
4
5 (* Define first surface function *)
6 za[ra_] := Module[{h}, h = Cos[ra 0.5]];
7
8 (* Define variables of Eq. (5)*)
9 Phi[ra_] := Module[{h},
10 h = Sqrt[1 (ra + (ta + za[ra]) Derivative[1][za][ra])^2/
11 (n^2 (ra^2 + (ta za[ra])^2) (1 + (Derivative[1][za][ra])^2))]/
12 Sqrt[1 + (Derivative[1][za][ra])^2]]
13
14 ri[ra_] := Module[{h},
15 h = (ra + (ta + za[ra]) Derivative[1][za][ra])/(
16 n Sqrt[ra^2+(taza[ra])^2] (1 + (Derivative[1][za][ra])^2))
17 Derivative[1][za][ra] Phi[ra]]
18
19 zi[ra_] := Module[{h},
20 h = (Derivative[1][za][
21 ra] (ra + (ta + za[ra]) Derivative[1][za][ra]))/(
22 n Sqrt[ra^2 + (ta
23 za[ra])^2] (1 + (Derivative[1][za][ra])^2)) +Phi[
24 ra]];
25
26 fi[ra_] := Module[{h}, h = ta tb Sign[ta] Sqrt[ra^2 + (ta za
[ra])^2]];
27
28 h0[ra_] := Module[{h},
29 h = ri[ra]^2 za[ra] + fi[ra] n zi[ra]
30 ra ri[ra] zi[ra] + (t + tb) zi[ra]^2 n^2 (za[ra] + t zi[ra])];
31
32 h1[ra_] := Module[{h},
33 h = ra^2 + 2 ra ri[ra] t + (tb za[ra])^2 +
34 t^2 (ri[ra]^2 + (1 + zi[ra])^2)
35 2 t (tb za[ra]) (1 + zi[ra])];
36
37 zb[ra_] := Module[{h},
38 h = (h0[ra] +
39 s1 Sqrt[zi[ra]^2 (fi[ra]^2
40 2 fi[ra] n (ra ri[ra] + ri[ra]^2 t +
41 zi[ra] (t (zi[ra] 1) tb + za[ra])) +
42 h1[ra] n^2 (ra zi[ra] +
43 ri[ra] (t + tb za[ra]))^2)])/(n^2 + 1)];
44
45 rb[ra_] := Module[{h}, h = ra + (ri[ra] (za[ra] + zb[ra]))/zi[ra]];
46
47 (* Define system configurations*)
48 s1 = 1;
49 n = 1.5;
50 t = 8;
51 ta = 60;
52 tb = 70;
53 basez = ta;
54 topez = tb + t;
55 rmax = 16.8; (* Iterative value *)
56
57 (* System plot *)
58 g1 = Plot[{za[ra], topez, basez}, {ra, rmax, rmax},
59 PlotStyle >{Black, Transparent, Transparent},
60 AspectRatio >Automatic,AxesLabel > {r, z}];
61 g2 = ParametricPlot[{rb[ra], zb[ra]}, {ra, rmax, rmax},
62 AspectRatio > 1/1, AxesOrigin > {0, 0}, PlotStyle >
{Black},
63 Mesh >None];
64 Show[g1, g2]
Please take care of the copy-paste process to avoid character
coding errors between the pdf reader format and the mathema-
tica work environment. We show the Code 1, Ref. [13].
Funding. Consejo Nacional de Ciencia y Tecnología
(Conacyt) (593740).
REFERENCES
1. D. Malacara-Hernández and Z. Malacara-Hernández, Handbook of
Optical Design (CRC Press, 2016).
2. R. K. Luneburg and M. Herzberger, Mathematical Theory of Optics
(University of California, 1964).
3. D. Malacara, Two lenses to collimate red laser light,Appl. Opt. 4,
16521654 (1965).
4. E. M. Vaskas, Note on the WassermanWolf method for designing
aspheric surfaces,J. Opt. Soc. Am. 47, 669670 (1957).
5. G. Wassermann and E. Wolf, On the theory of aplanatic aspheric
systems,Proc. Phys. Soc. London Sect. B 62,28 (1949).
6. P. D. Lin and C.-Y. Tsai, Determination of unit normal vectors of
aspherical surfaces given unit directional vectors of incoming and
outgoing rays: reply,J. Opt. Soc. Am. A 29, 1358 (2012).
7. J. C. Valencia-Estrada, R. B. Flores-Hernández, and D. Malacara-
Hernández, Singlet lenses free of all orders of spherical aberration,
Proc. R. Soc. A 471, 20140608 (2015).
8. J. C. Valencia-Estrada, M. V. Pereira-Ghirghi, Z. Malacara-Hernández,
and H. A. Chaparro-Romo, Aspheric coefficients of deformation for a
Cartesian oval surface,J. Opt. 46,100107 (2017).
9. M. Avendaño-Alejo, E. Román-Hernández, L. Castañeda, and V. I.
Moreno-Oliva, Analytic conic constants to reduce the spherical
aberration of a single lens used in collimated light,Appl. Opt. 56,
62446254 (2017).
10. G. Castillo-Santiago, M. Avendaño-Alejo, R. Daz-Uribe, and L.
Castañeda, Analytic aspheric coefficients to reduce the spherical
aberration of lens elements used in collimated light,Appl. Opt. 53,
49394946 (2014).
11. N. D. C. Lozano-Rincón and J. C. Valencia-Estrada, Paraboloid-
aspheric lenses free of spherical aberration,J. Modern Opt. 64,
11461157 (2017).
12. J. C. Valencia-Estrada, J. Garca-Marquez, L. Chassagne, and S.
Topsu, Catadioptric interfaces for designing VLC antennae,Appl.
Opt. 56, 75597566 (2017).
13. R. G. González-Acuña and H. A. Chaparro-Romo, Singlet free of spheri-
cal aberration,2018, https://doi.org/10.6084/m9.figshare.7163357.
Research Article Vol. 57, No. 31 / 1 November 2018 / Applied Optics 9345
... Various solutions focus on reducing the spherical aberration of single plano-convex or convex-plano lenses by using caustic formulas [2,3] and expansion in Taylor's series [4]. The polar equation of an aspherical surface of a lens free from primary-and higherorder spherical aberration can be obtained in series form [5]. Ray tracing is used to obtain the desired parameters of the surfaces of thick lenses to correct the marginal spherical aberration [6,7]. Analytical solutions deduced by Schwarzschild's formula correcting all orders of spherical aberration for singlet lenses has been proposed [8]. ...
Article
Full-text available
It is still challenging to find a spherical-aberration-free singlet lens with well corrected coma due to an undesired and complicated residual high-order coma. In this paper, we present a spherical-aberration-free singlet lens with reduced coma containing high-order coma contribution. This design algorithm is to deduce the front aspherical surface parameters from the back spherical surface using meridional ray tracing to find the proper values of the back focal length and the back spherical radius to reduce the coma. The exemplary lens demonstrates an excellent well-balanced and diffraction-limited performance at the field angle ranging from 0.0° to 2.5° with a working F# equal to 1.65.
... Several approaches for correcting aberrations in Digital Holography Microscopy have been developed and applied. Many of these methods are numerical techniques, limited to measuring thin and sparse objects such as living cells or thin lenses, where object phases are considered negligible in the reconstructed unwrapped phase map, or using a beam with a different wavelength (physical methods) [8,[11][12][13][14][15]. To our knowledge, there are no reports on the use of our proposed optical method, which utilize a single beam and capture to address aberration issues. ...
Article
Full-text available
This work utilizes a Gabor Holographic Optical Scheme integrated with a microscope objective and a thin convex plane lens. This bi-telecentric lens system corrects spherical aberration from the objective, maintains consistent magnification across various reconstruction distances, and ensures a plane incidence on CMOS. Depending on the focal lengths of the objective and lens, the final image can be enlarged or reduced compared to the classic Gabor system, resulting in high-quality reconstructed phase images without spherical aberration. This setup was employed to capture phase distribution and intensity images of planktonic objects, such as copepods, achieving superior image quality.
... In a 1949, Wasserman and Wolf formulated the problem how to design a lens without spherical aberration in an analytical way, and it has since been known as the 'Wasserman-Wolf problem' [16]. In 2018, Rafael G. González-Acuña and Héctor A. Chaparro-Romo, finally found a closed formula for a lens surface that eliminates spherical aberration [2]. Their equation can be applied to specify a shape for one surface of a lens, where the other surface has any given shape. ...
... producing a smooth bijective mapping). If only a few rays cross each other inside the lens, the method will result in an inhomogeneous and nonphysical optical element [32,33]. ...
Article
Full-text available
We propose a method to design the exact phase profile of at least one metasurface in a stigmatic singlet that can be made to implement a desired ray mapping. Following the generalized vector law of refraction and Fermat’s principle, we can obtain exact solutions for the required lens shape and phase profile of a phase gradient metasurface to respect particular ray conditions (e.g., Abbe sine) as if it were a freeform refractive element. To do so, the method requires solving an implicit ordinary differential equation. We present comparisons with Zemax simulations of illustrative designed lenses to confirm the anticipated optical behaviour.
... where d is the center lens thickness and R 1 and R 2 are the radius of curvature parameters for the front and back surfaces of the lens, respectively. While the previously-discussed quasioptical system [42] relied on a separate penalty function to optimize the lenses for reducing aberrations at the focus, the focal length for each design in our simplex was used as an input in a spherical aberration correction model based on Fermat's principle [53] to adjust the hyperbolic face of the lens, with an example shown in the block diagram (see Fig. 3). The ray-tracing simulation for evaluating each design is separated into several stages, starting at the back focus of the first aspheric lens with rays propagating towards its hyperbolic surface. ...
Article
Full-text available
Terahertz (THz) time-domain spectroscopy has been investigated for assessment of the hydration levels in the cornea, intraocular pressure, and changes in corneal topography. Previous efforts at THz imaging of the cornea have employed off-axis parabolic mirrors to achieve normal incidence along the spherical surface. However, this comes at the cost of an asymmetric field-of-view (FOV) and a long scan time because it requires raster-scanning of the collimated beam across the large mirror diameter. This paper proposes a solution by designing a pair of aspheric lenses that can provide a larger symmetric spherical FOV (9.6 mm) and reduce the scan time by two orders of magnitude using a novel beam-steering approach. A hyperbolic-elliptical lens was designed and optimized to achieve normal incidence and phase-front matching between the focused THz beam and the target curvature. The lenses were machined from a slab of high-density polyethylene and characterized in comparison to ray-tracing simulations by imaging several targets of similar sizes to the cornea. Our experimental results showed excellent agreement in the increased symmetric FOV and confirmed the reduction in scan time to about 3-4 seconds. In the future, this lens design process can be extended for imaging the sclera of the eye and other curved biological surfaces, such as the nose and fingers.
... where is the center lens thickness and 1 and 2 are the radius of curvature parameters for the front and back surfaces of the lens, respectively. The focal length for each design was used as an input in a spherical aberration correction model based on Fermat's principle [51] to adjust the hyperbolic face of the lens, with an example shown in the block diagram (see Fig. 3). ...
Preprint
Full-text available
Terahertz (THz) time-domain spectroscopy has been investigated for assessment of the hydration levels in the cornea, and changes in corneal topography and the intraocular pressure. Previous efforts at THz imaging of the cornea have employed off-axis parabolic mirrors to achieve normal incidence along the spherical surface. However, this comes at the cost of an asymmetric field-of-view (FOV) and a long scan time because it requires raster-scanning the collimated beam across the large mirror diameter. In this paper, we propose a solution by designing a pair of aspheric lenses that can provide a larger symmetric spherical FOV (9.6 mm) and reduce the scan time by two orders of magnitude using a novel beam-steering approach. Using Nelder-Mead optimization, a hyperbolic-elliptical lens was designed to achieve normal incidence and phase-front matching between the focused THz beam and the target curvature. The lenses were machined from a slab of high-density polyethylene and characterized in comparison to ray-tracing simulations by imaging several targets of similar sizes to the cornea. Our experimental results showed excellent agreement in the increased symmetric FOV and confirmed the reduction in scan time to about 3-4 seconds. In the future, this lens design process can be extended for imaging the sclera of the eye and other curved biological surfaces, such as nose and fingers.
Conference Paper
Single-pixel reflection imaging of curved surfaces requires normal incidence and phase-front matching over a large FOV. Hyperbolic-Elliptical lenses were designed for a collocated THz-TDS setup to achieve imaging of samples such as the cornea in under 2 seconds.
Article
Full-text available
This paper presents a model to design bi-aspherical catadioptric lenses with limited image diffraction. A first refractive Cartesian oval surface that does not introduce any spherical aberration is used. When total internal reflection occurs, this surface can also be simultaneously used as a mirror. The reflective characteristics of Cartesian ovals are also well described in this paper. The theoretical work described here can considerably reduce computing time in optical system design. This model is applied to examples of antennae design for visible light communications (VLC).
Article
Full-text available
We study the formation of caustic surfaces produced by conic lenses, considering a plane wavefront propagating parallel to the optical axis. The shape of the caustic can be modified by changing the parameters of the lens in such a way that if we are able to vanish the caustic, the optical system produces the sharpest diffraction-limited images. Alternatively, caustic surfaces with a large area can be applied to the design of non-imaging optical systems, with potential applications such as diffusers of light for illumination or solar concentrators. Here, we provide analytic equations for the conic constants, principal surfaces, and caustic surfaces, and also approximations at the third and fifth orders formed by conic lenses, in order to reduce the spherical aberration at these orders.
Article
Full-text available
This paper describes a method to design families of singlet lenses free of all orders of spherical aberration. These lenses can be mass produced according to Schwarzschild's formula and therefore one can find many practical applications. The main feature of this work is the application of an analysis that can be extended to grazing or maximum incidence on the first surface. Also, here, the authors present some developments that corroborate geometrical optics results, along with the axial thick lensmaker's formula, which can be applicable to any pair of finite conjugate planes for any lens shape (bending) and can be used instead of the classical thick lensmaker's formula, which always assumes that the object is at infinity, to attain better accuracy.
Article
Full-text available
We provide closed-form formulas for aspheric terms for either plano–convex or convex–plano aspheric lenses as functions of the paraxial parameters involved in the process of refraction. These formulas are obtained through an expansion in Taylor’s series from the exact caustic equation produced by aspheric lenses considering a plane wavefront propagating parallel to the optical axis and impinging on the refracting surface. A comparison of the aspheric coefficients obtained through our analytic formulas and commercial optical design software is presented, showing good agreement. This is useful in reducing spherical aberration.
Article
Full-text available
We reply to the comments on our paper previous paper. While the results obtained are the same as ours, we hold that, by using homogeneous coordinate notation, our method enables first-order and second-order derivatives of non-axially symmetrical systems to be computed numerically (such as [J. Opt. Soc. Am. A 28, 747 (2011)]), which are necessary for automatic optical design [Appl. Opt. 2, 1209 (1963)].
Article
A method to design singlet paraboloid-aspheric lenses free of all orders of spherical aberration with maximum aperture is described. This work includes all parametric formulas to describe paraboloid-aspheric or aspheric-paraboloid lenses for any finite conjugated planes. It also includes the Schwarzchilds approximations (which can be used to calculate one rigorous propagation of light waves in physic optics) to design convex paraboloid-aspheric lenses for imaging an object at infinity, with explicit formulas to calculate thicknesses easily. The results were verified with software through ray tracing.
Article
In this paper we propose a mathematical model as a theoretical and practical contribution to the theory of aspheric surfaces, that allows the geometrical calculation of an aspherical refractive surface in cylindrical coordinates \((r,z)\), that correspond to a non-degenerated Cartesian oval of revolution, according to standard formulae [ISO-10110-12 (2007)], including: vertex curvature C, conical constant K and aspheric deformation coefficients \(A_4,\,A_6,\,A_8,\,A_{10}\) and \(A_{14}\). The results are shown for an object and its respective image located at finite distances from the vertex (origin of coordinates) on the optical axis.
Article
Two lenses to collimate the light from a He-Ne laser, with focal ratios ƒ/4 and ƒ/2.65 are described. These systems have good spherical correction, and the coma is small.
Article
Methods are described for the design of two aspheric surfaces for any given centred system, so as to achieve exact aplanatism. The practical application of the methods is illustrated by the design of a reflecting microscope.