Open Access
28 April 2021 Straightforward path to Zernike polynomials
Author Affiliations +
Abstract

Starting from Weierstrass’ approximation theorem, Zernike polynomials are obtained by a few straightforward steps involving only the recast of the aberration function as a double sum in the polar coordinates followed by the weighted orthogonalization of a power series. The origin of the name Fringe Zernike polynomials is also explained.

Zernike polynomials are used extensively in microlithography to characterize the imaging optics and in evaluating the resulting images. Yet few lithographers have questioned how these polynomials are obtained. Frits Zernike invented the eponymous circle polynomials as solutions of a self-adjoint differential equation subject to circular boundary conditions.13 The angular parts of his solutions are simply cosmφ and sinmφ with m0, but the general expression for the radial part of the solution for 0r1 looks quite daunting at first:

Eq. (1)

Rnm(r)=rm(nm2)!(dd(r2))(nm)/2{rn+m(r21)(nm)/2}=s=0(nm)/2(1)s(ns)!s!(n+m2s)!(nm2s)!rn2s.

An alternate derivation of the above formulae is described in Born and Wolf.4 To understand fully both derivations, the reader has to be familiar with some specialized topics in mathematical physics. It is the purpose of this letter to lessen the complexity and demonstrate that the Zernike polynomials can be obtained using straight-forward mathematics involving three steps described below.

Let W(x,y) be an aberration function or any function that is continuous within and on the unit circle. According to Weierstrass’ approximation theorem, W(x,y) may be expressed in a polynomial to arbitrary degree of accuracy:

Eq. (2)

W(x,y)=limNp,q=0NApqxpyq,
where p and q are integers.5 Anticipating the split of the polynomials into radial and angular parts, our first step is to express xpyq in polar corrdinates. We accomplish this by letting x=rcosθ and y=rsinθ and making use of Euler’s formula and the binomial theorem
xpyq=(rcosθ)p(rsinθ)q=rp+q(eiθ+eiθ2)p(eiθeiθ2i)q=rp+q2p+qiqs=0p(ps)(eiθ)ps(eiθ)st=0q(qt)(eiθ)qt(eiθ)t=rp+q2p+qiqs=0pt=0q(ps)(qt)(1)t(eiθ)p+q(s+t)(eiθ)s+t=rp+q2p+qiqei(p+q)θl=0p+qCle2ilθ,
where we have let l=s+t and combined all the exponential terms with fixed l into a single term with coefficient Cl. We then insert the above expression into Eq. (2) and get

Eq. (3)

W(x,y)=W(r,θ)=m=0rml=0mClmei(m2l)θ,
where we have let m=p+q and, just like before, combined all the exponential terms with fixed m into a single term with coefficient Clm.

Our second step is to re-arrange the terms in Eq. (3). To start, let us write down its first few terms, say from m=0 to m=4. They are

C00+r(C01eiθ+C11eiθ)+r2(C02ei2θ+C12+C22ei2θ)+r3(C03ei3θ+C13eiθ+C23eiθ+C33ei3θ)+r4(C04ei4θ+C14ei2θ+C24+C34ei2θ+C44ei4θ).

These terms can be rearranged so that the following pattern can be seen:

{C00+r(C01eiθ+C11eiθ)+r2(C02ei2θ+C22ei2θ)+  r3(C03ei3θ+C33ei3θ)+r4(C04ei4θ+C44ei4θ)}+{r2C12+r3(C13eiθ+C23eiθ)+r4(C14ei2θ+C34ei2θ)}+{r4C24}.

Continuing this process and making use of Euler’s formula, Eq. (3) can be expressed as

W(r,θ)=m=0rm(Amcosmθ+Bmsinmθ)+m=0r2rm(Amcosmθ+Bmsinmθ)+m=0r4rm(Amcosmθ+Bmsinmθ)+

Grouping all the cosine and sine terms together, we have

Eq. (4)

W(r,θ)=m=0cosmθ  rm(Am+Amr2+Amr4+)+m=0sinmθ  rm(Bm+Bmr2+Bmr4+)=m=0cosmθ  rmk=0Amkr2k+m=0sinmθrmk=0Bmkr2k.

In reaching the above expression, no requirement of rotational symmetry about an axis had to be imposed.

Our third and last step involves expressing r2k as a linear combination of orthogonal polynomials satisfying the orthogonal relation over the interval [0, 1]. Once this is accomplished, both summations over k can be expressed as linear combinations of these polynomials. Therefore, the first thing to do is to obtain these orthogonal radial polynomials (actually the Zernike radial polynomials) by orthogonalizing the set {1,r2,r4r2k}. We do this by first letting r2=u so that the orthogonalization process becomes for the set {1,u,u2,uk}. We may then associate the orthogonalization of this set with shifted Legendre polynomials Pk(u). Pk(u)’s can be obtained through the Gram-Schmidt orthogonalization process on the above set or simply by making use of the formula6

Pk(u)=1k!dkduk{uk(u1)k}.
(See also Ref. 5, pp. 233–239. The Legendre polynomials discussed in the text are defined on the interval [1,1] and the associated formula is called Rodrigues’ formula.) The first three shifted Legendre polynomials are P0(u)=1, P1(u)=2u1, P2(u)=6u26u+1. Therefore, we may express 1 as P0(u), u as P1(u)2+P0(u)2, u2 as P2(u)6+P1(u)2+P0(u)3, and so on. Hence any linear combination of powers of u can be expressed as a linear combination of Pk(u)’s. There is only one catch, however. We have to include the common factor rm=um/2 in Eq. (4) in the orthogonalization process, so if Gkm(u)’s are the resulting polynomials, our orthogonalization relation has to be, instead of 01Pk(u)·Pk(u)du=Const.δkk,

Eq. (5)

01um/2Gkm(u)·um/2Gkm(u)du=01Gkm(u)Gkm(u)umdu=Const.δkk.

The second integral in Eq. (5) suggests that the presence of the factor um/2 may be regarded as orthogonalizing the set {1,u,u2,uk} with the weight um. The formula for the polynomials obtained by orthogonalizing the set {1,u,u2,} with the weight equal to um instead of 1 (which would result in shifted Legendre polynomials) is given as

Eq. (6)

Gkm(u)=1k!1umdkduk{umuk(u1)k}.

The validity of Eq. (6) can be established as follows. First, the polynomial so generated is of order k because the term of the highest power inside the brackets to be differentiated is um+2k. Second, the following integral is valid (Ref. 6, p. 324):

01Gkm(u)ukumdu=0,0k<k.

Since Gkm(u) is a linear combination of powers of u with uk being the term of the highest power, Eq. (5) therefore stands.

Now with powers of u represented by Gkm(u)’s, Eq. (4) can be recast as

W(r,θ)=m,k=0CkmrmGkm(r2)cosmθ+m,k=0DkmrmGkm(r2)sinmθ,
where Ckm and Dkm are the new coefficients.

The Zernike polynomial is simply Zkm(r,θ)=rmGkm(r2){cosmθsinmθ=Rkm(r){cosmθsinmθ, where Rkm(r)=rmGkm(r2) is called the Zernike radial polynomial. Since the angular parts are already orthogonal, as

02πcosmθsinmθ  dθ=0,02πcosmθcosmθ  dθ=π(1+δm0)δmm,02πsinmθsinmθ  dθ=π(1δm0)δmm,
and since Rkm(r)’s satisfy
01Rkm(r)Rkm(r)dr2=01Gkm(r2)Gkm(r2)r2  mdr2=Const.δkk
because of Eq. (5), Zkm(r,θ)’s therefore satisfy the orthogonal relation over an area bounded by the unit circle, as
0102πZkm(r,θ)Zkm(r,θ)rdrdθ=Const.δkkδmm.

The explicit expression for the Zernike radial polynomials can now be written down immediately as

Rkm(r)=rmGkm(r2)=rmk!dkd(r2)k{r2(m+k)(r21)k}.
Defining n=m+2k brings us to Eq. (1) put forth originally by Frits Zernike.

Incidentally, using k instead of n to index the Zernike polynomials is not a bad thing. One advantage is that k is independent of m. The ordering sequence of the Zernike polynomials used by Zeiss and ASML is a modified version of the indexing scheme originated at the University of Arizona. We can learn the origin of this Fringe indexing scheme from Katherine Creath and Robert E. Parks’ article:7 “The first program for analyzing interferograms was written by Jim Rancourt, PhD 1974 (Fig. 11),[19]… Later, Loomis, PhD 1980, wrote a FRINGE MANUAL, and updated the program to output the 37 “FRINGE” Zernike polynomials,[20] and the beginning of the confusion about whose numbering of the polynomials one might be using.” Citation [19] in their article is: Optical Sciences Center, “FRINGE Software Program,” OSC Newsletter 8(12), 29 (1974). Citation [20] refers to John S. Loomis, FRINGE User’s Manual, Optical Sciences Center, University of Arizona, Tucson, AZ, November 1976. Hence we believe that it was John Loomis who invented this indexing scheme in conjunction with the wavefront-fitting program called FRINGE, originally written by Jim Rancourt. It is therefore a gross misnomer that the Zernike polynomials we lithographers use are often referred to as Fringe Zernike polynomials, as if there are various sets of such polynomials; it is the “Fringe” indexing scheme of the one and only set of Zernike polynomials!

The indexing scheme used by Zeiss and ASML is shown in Fig. 1. As one can see, rows are arranged by the ascending order of m+k. Since the power of every radial polynomial is n=m+2k and since (m+2k)+m=2(m+k) is fixed for every row, the rightmost entry of every row, with m=0, has the highest power. Table 1 lists explicitly the Zernike polynomials according to this indexing scheme.

Fig. 1

Indexing scheme of Zernike polynomials used by Zeiss and ASML. These plots were originally generated by Marco Moers.

JM3_20_2_020501_f001.png

Table 1

Explicit expressions of the first 36 Zernike polynomials.

IndexMathematical expressionNamem (period)m+kn=m+2k (power)
11Piston000
2rcosθTilt x111
3rsinθTilt y111
42r21Focus012
5r2cos2θAstigmatism x222
6r2sin2θAstigmatism y222
7(3r32r)cosθComa x123
8(3r32r)sinθComa y123
96r46r2+1Spherical aberration024
10r3cos3θThree-fold x333
11r3sin3θThree-fold y333
12(4r43r2)cos2θAstigmatism x234
13(4r43r2)sin2θAstigmatism y234
14(10r512r3+3r)cosθComa x135
15(10r512r3+3r)sinθComa y135
1620r630r4+12r21Spherical aberration036
17r4cos4θFour-fold x444
18r4sin4θFour-fold y444
19(5r54r3)cos3θThree-fold x345
20(5r54r3)sin3θThree-fold y345
21(15r620r4+6r2)cos2θAstigmatism x246
22(15r620r4+6r2)sin2θAstigmatism y246
23(35r760r5+30r34r)cosθComa x147
24(35r760r5+30r34r)sinθComa y147
2570r8140r6+90r420r2+1Spherical aberration048
26r5cos5θFive-fold x555
27r5sin5θFive-fold y555
28(6r65r4)cos4θFour-fold x456
29(6r65r4)sin4θFour-fold y456
30(21r730r5+10r3)cos3θThree-fold x357
31(21r730r5+10r3)sin3θThree-fold y357
32(56r8105r6+60r410r2)cos2θAstigmatism x258
33(56r8105r6+60r410r2)sin2θAstigmatism y258
34(126r9280r7+210r560r3+5r)cosθComa x159
35(126r9280r7+210r560r3+5r)sinθComa y159
36252r10630r8+560r6210r4+30r21Spherical aberration0510

If the pupil function is rather roughly behaved, it may be necessary to include Zernike polynomials of very high orders. For numerical computations involving Zernike radial polynomials of n40, Janssen and Dirksen suggested an alternate form of Eq. (1) with advantages in computation time, accuracy and ease of implementation.8 Based on Janssen and Dirksen’s integral expression, Shakibaei and Paramesran found a concise recursive relation for Rnm(r) leading to a reduction in computational complexity.9

Acknowledgments

The author wishes to thank A. J. E. M. Janssen for helpful suggestions and for clarifications regarding the derivation of an equation in Ref. 8 and the derivation of Zernike polynomials in Born and Wolf.10 He also thanks Bernd Geh for helpful suggestions and extensive discussions on the broader topic associated with Zernike polynomials.

References

1. 

F. Zernike, “Beugungstheorie des schneidenver-fahrens und seiner verbesserten form, der phasenkontrastmethode,” Physica, 1 689 –704 (1934). https://doi.org/10.1016/S0031-8914(34)80259-5 Google Scholar

2. 

F. Zernike, “Diffraction theory of the knife-edge test and its improved form, the phase-contrast method,” J. Micro/Nanolithogr. MEMS MOEMS, 1 (2), 87 –94 (2002). https://doi.org/10.1117/1.1488608 Google Scholar

3. 

B. R. A. Nijboer, “The diffraction theory of aberrations,” (1942). Google Scholar

4. 

M. Born and E. Wolf, Principle of Optics, 7th ed.Cambridge University Press, Cambridge (1999). Google Scholar

5. 

F. W. Byron and R. W. Fuller, Mathematics of Classical and Quantum Physics, 239 Dover Publications, New York (1992). Google Scholar

6. 

A. Yen and S.-S. Yu, Optical Physics for Nanolithography, 316 –320 SPIE Press, Bellingham, Washington, DC (2018). Google Scholar

7. 

K. Creath and R. E. Parks, “Optical metrology at the optical sciences center: a historical review,” Proc. SPIE, 9186 91860T (2014). https://doi.org/10.1117/12.2064376 PSISDG 0277-786X Google Scholar

8. 

A. J. E. M. Janssen and P. Dirksen, “Computing Zernike polynomials of arbitrary degree using the discrete Fourier transform,” J. Eur. Opt. Soc. Rapid Publ., 2 07012 (2007). https://doi.org/10.2971/jeos.2007.07012 Google Scholar

9. 

B. H. Shakibaei and R. Paramesran, “Recursive formula to compute Zernike radial polynomials,” Opt. Lett., 38 2487 (2013). https://doi.org/10.1364/OL.38.002487 OPLEDP 0146-9592 Google Scholar

10. 

J. Braat and P. Török, Imaging Optics, 888 –890 Cambridge University Press, Cambridge (2019). Google Scholar
© 2021 Society of Photo-Optical Instrumentation Engineers (SPIE) 1932-5150/2021/$28.00 © 2021 SPIE
Anthony Yen "Straightforward path to Zernike polynomials," Journal of Micro/Nanopatterning, Materials, and Metrology 20(2), 020501 (28 April 2021). https://doi.org/10.1117/1.JMM.20.2.020501
Received: 13 February 2021; Accepted: 8 April 2021; Published: 28 April 2021
Lens.org Logo
CITATIONS
Cited by 2 scholarly publications.
Advertisement
Advertisement
RIGHTS & PERMISSIONS
Get copyright permission  Get copyright permission on Copyright Marketplace
KEYWORDS
Zernike polynomials

Monochromatic aberrations

Differential equations

Mathematics

Optical lithography

Physics

Spherical lenses

Back to Top