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

Nijboer–Zernike’s aberration theory: computational achievements via Tchebychev’s polynomials approximation theory

Open Access Open Access

Abstract

Nijboer–Zernike’s diffraction theory of aberration is a nearly abandoned jewel of physical optics. The present paper constitutes an attempt to extend its practical feasibility. It is found that such a task can be achieved by using what is probably the most important property of  Tchebychev’s polynomials.

© 2022 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. INTRODUCTION

I first met Emil Wolf in 1997 during a meeting on laser beam characterization at the University of Laval in Quebec City, Canada. Since then, my career, similarly to that of all scientists who started studying optics in the early nineties, has considerably been influenced by Wolf’s charisma as well as by the huge deal of work Wolf and his co-workers produced over four decades, especially in the field of classical coherence theory, of which he was universally considered the father. Wolf will also be remembered as one of the authors of the Optics Bible, namely, the monumental Principle of Optics (Principles henceforth), written together with Born [1]. A story about Principles Emil Wolf enjoyed telling his guests (I was not an exception during a lunch at Rochester’s Institute of Optics) was the following [2]:

In 1956, Born was already in retirement. I was on a visiting appointment at New York University, still working on our book [2]. One day, I received a letter from Born in which he asked me why the manuscript was not yet finished. I wrote back saying that the manuscript is almost completed, except for a chapter on partial coherence on which I was still working. Born replied at once saying, “Wolf, who apart from you, is interested in coherence? Leave the chapter out and send the manuscript to the printers.” I finished the chapter anyway, and our book was published in 1959, only a few months before the invention of the laser, and many of the reviews of our book, which were then appearing, stressed that Principles of  Optics contained an account of coherence theory, which had become of crucial importance to the understanding of some features of laser light.

I had the privilege of coauthoring a couple of works with Wolf, the first of which saw the light in the tea room of the cafeteria of the Department of Physics and Astronomy in Rochester, not far from a photographic portrait Olivia Newton-John, bearing the dedication:

To Emil, with love—Olivia

Not many know that Olivia Newton-John was Max Born’s granddaughter.

The present paper has been conceived as a tribute to Wolf’s memory and legacy. It deals with a classical topic of classical optics, namely, the Nijboer–Zernike diffraction theory of aberrations [37]. In 1951, Wolf published an important review paper on the subject [8], which later formed the core of Ch. IX in Principles.

Despite its formal elegance, Nijboer’s theory did not receive the attention it deserved. The main reasons have partly been indicated by Nijboer himself but were clarified 20 years ago by Janssen and his co-workers [912], who developed an important extension of the Nijober formalism aimed at overcoming three main drawbacks that, for nearly 80 years, have confined Nijboer–Zernike’s theory as a beautiful but nearly useless practical tool: (i) entity of the phase distortion; (ii) entity of defocusing; and (iii) amplitude uniformity of the wavefront. From a mathematical perspective, one of the main technical problems is the evaluation of several paraxial 1D diffraction integrals containing Bessel functions of the first kind. Such integrals are the computational bricks with which Nijboer’s perturbative aberration theory was built up. When the diffracted field is required across the focal plane, such integrals can exactly be evaluated in terms of suitable linear combinations of Bessel functions of the first kind [46,8]. In the presence of defocusing, a quadratic phase term appears inside the integrals and makes their computation a dramatically more demanding task. It should be stressed that, as far as the practical solution of several problems in lithography, microscopy, and astronomy, Janssen’s extension of the Nijboer–Zernike aberration theory constitutes an optimal answer. Interested readers are encouraged to consult the important review published on the subject in 2013 [13].

The scope of the present paper is different. Here, we want to show that the “old” Nijboer–Zernike theory has probably been abandoned too early, and that new computational strategies can be conceived in order to extend its practical feasibility at least to deal with the presence of arbitrary defocusing. To this end, a short mathematical tour will first be presented within the realm of Tchebychev’s polynomials. Unique among several families of orthogonal polynomials is their capability to provide excellent low-order polynomial approximants of continuous functions. In Section 2, the basics of Tchebychev’s polynomials approximation properties are briefly resumed and employed on the zeroth-order Bessel function of the first kind. Section 3 constitutes the core of the paper: Nijboer–Zernike’s theory is numerically reformulated in such a way that a quantitative study of primary optical aberrations, namely, spherical aberration, coma, and astigmatism, can be carried out as well as for arbitrarily high defocusing. To this end, all Nijboer’s diffraction integrals are converted into quickly converging series expansions whose evaluation can be effectively implemented through simple recursive algorithms. One of the most important practical consequences is that the numerical effort needed to evaluate the diffracted wavefield turns out to be practically independent of the defocusing amount, a fact that alone should render such an approach particularly attractive. In order to validate our algorithm, several numerical checks will be carried out using numerous previously published results by Nijboer himself [1,46,8] as well as by Janssen and co-workers [9,10,12,13].

2. TCHEBYCHEV’S EXPANSION OF BESSEL FUNCTIONS

A. Approximation Problem Involving Tchebychev’s Polynomials

Tchebychev polynomials ${T_n}(x)$ constitute a well-known family of orthogonal functions across the real interval $[- 1,1]$, being ([14], Formula 8.949.3)

$$\begin{array}{*{20}{l}}{\int_{- 1}^1 \frac{{{T_n}(x) {T_m}(x)}}{{\sqrt {1 - {x^2}}}} {\rm d}x = \pi \!\left\{{\begin{array}{ll}{\frac{{{\delta _{n,m}}}}{2}}&\;\;{m \ne 0,n \ne 0 ,}\\[3pt]{1}&\;\;{m = n = 0 ,}\end{array}} \right.}\end{array}$$
where ${\delta _{n,m}}$ denotes Kronecker’s symbol and the function $1/\sqrt {1 - {x^2}}$ represents the weight function, which is, incidentally, the primitive of function $\arcsin x$. This is related to the classical “trigonometric” definition of ${T_n}$:
$${{T_n}(\cos \theta) = \cos n\theta ,\quad n \ge 0 .}$$
The most important property of Tchebychev's polynomials states the following:

Property. Among all reduced polynomials of degree $n$, ${T_n}$ is the polynomial that deviates least from zero within the interval $[- 1,1]$.

By “reduced,” it should be mean a polynomial of the form ${x^n} + {a_{n - 1}}{x^{n - 1}} + \ldots + {a_0}$, i.e., such that the coefficient of the leading term be unitary. A simple proof of such property can be found, for instance, in [15] (Sec. 8.5.4). The “least deviation” property of Tchebychev’s polynomial is connected to its capability to build up excellent low-order polynomial approximations of continuous functions within the interval $[- 1,1]$. For the scope of the present paper, it is worth briefly illustrating such capabilities.

To this aim, consider the following approximation problem: given a continuous function $f(x)$, with $x \in [- 1,1]$, to find its best $n$ th-order polynomial approximation ${Q_n}(x)$ such that, given any other polynomial of degree $n$, say ${p_n}(x) \ne {Q_n}(x)$, the following inequality holds:

$${{\omega _n} = \mathop {{\rm max}}\limits_{x \in [- 1,1]} |f(x) - {Q_n}(x)| \lt \mathop {{\rm max}}\limits_{x \in [- 1,1]} |f(x) - {p_n}(x)| ,}$$
where the symbol ${\omega _n}$ has been introduced to denote the ${L_\infty}$ norm of the error associated with the best approximation. Equation (3) simply states that ${Q_n}(x)$ is, among all infinite $n$th-order polynomials, the one that “least deviates” from $f(x)$ within the interval $[- 1,1]$. It should be noted that finding, in closed form, the best polynomial ${Q_n}(x)$ for a typical function $f$ is an almost impossible task (except for a few particular cases). On the other hand, several other polynomial approximations of $f$ could be found in a considerably easier way, e.g., by truncating orthogonal expansions. As a consequence, it would be desirable, for practical reasons, to find one of such approximations, which could be sufficiently close, in somewhat sense, to the best one ${Q_n}(x)$. From such perspective, Tchebychev’s family plays a role of pivotal importance. More precisely, consider the series expansion representation of the function $f(x)$ in terms of Tchebychev’s polynomials $\{{T_n}\}$, i.e.,
$${f(x) = \sum\limits_{k = 0}^\infty {a_k} {T_k}(x) ,\quad x \in [- 1,1] ,}$$
with the expanding coefficients $\{{a_k}\}$ being given, for intance, by the inner product defined through Eq. (1). Consider now the $n$ th-order polynomial approximation, say ${f_n}(x)$, obtained by truncating the series in Eq. (4) up to the $n$th-order term,
$${{f_n}(x) = \sum\limits_{k = 0}^n {a_k} {T_k}(x) .}$$

From what we’ve said here, the approximation provided by ${f_n}(x)$ will certainly be worse than that of the best one ${Q_n}(x)$, since ${\omega _n} \lt {{\cal S}_n}$, where ${{\cal S}_n}$ is defined by

$${{{\cal S}_n} = \mathop {{\rm max}}\limits_{x \in [- 1,1]} |f(x) - {f_n}(x)| .}$$
Nevertheless, the following theorem will play an important role in the rest of the paper ([15], Sec. 8.5.3):

Theorem. The ${L_\infty}$ norms ${\omega _n}$ and ${S_n}$ satisfy the following inequalities:

$${{{\cal A}_n} \le {\omega _n} \le {{\cal S}_n} \le {\cal S}_n^ \star ,}$$
where quantities ${\cal S}_n^ \star$ and ${{\cal A}_n}$ are given by
$${{\cal S}_n^ \star = \sum\limits_{k = n + 1}^\infty |{a_k}| ,}$$
and
$${{\cal A}_n} = \frac{1}{{\sqrt 2}} {\left(\,{\sum\limits_{k = n + 1}^\infty a_k^2} \right)^{1/2}} ,$$
respectively. In the next section, the inequalities in Eq. (7) will be applied to the zeroth-order Bessel function of the first kind.

B. Tchebychev’s Expansion of ${J_0}$ Function

Bessel functions of the first kind play a dominant role within paraxial optics. In the Nijboer–Zernike aberration theory, the optical field generated by aberrated systems is represented via superposition of diffraction integrals whose integrands are products of Bessel functions, Zernike polynomials, and quadratic phase factors, the integration domain being (in suitable units) the interval [0, 1]. To overcome the difficulties of evaluating such highly oscillating (especially in the presence of nonnegligible defocusing) integrals, suitable representations of Bessel functions like ${J_\ell}(vx)$ (with $\ell \ge 0,v \ge 0$) for $x \in [0, 1]$ will be employed. To illustrate the main idea, consider first the zeroth-order Bessel function ${J_0}(vx)$, for which the following expansion holds {[16], Formula 9.3.6(3)}:

$${{J_0}(v x) = \sum\limits_{k = 0}^\infty {\epsilon _k} {{(- 1)}^k} J_k^2(v/2) {T_{2k}}(x) ,\quad x \in [- 1,1] ,}$$
where ${\epsilon _k} = 2 - {\delta _{k,0}}$. Equation (10) is a direct consequence of Graf’s addition theorem of Bessel functions and of the trigonometric definition of Tchebychev polynomials in Eq. (2). Due to the evenness of ${J_0}$, only Tchebychev polynomials of even order are involved. For such reason, even values of the truncation order $n$ will be considered. Since in the following only positive values of the Bessel function argument vx will be considered, it could appear to be redundant to carry out the subsequent analysis within the $x$ interval $[- 1,1]$ rather than the interval [0, 1]. Actually, it could be possible, in principle, to employ the so-called shifted Tchebychev polynomials $T_n^{\,*}(x) = {T_n}(2x - 1)$, but the involved math would be more cumbersome. On the other hand, due to the parity features of first-kind Bessel functions, all subsequent approximation formulas, as well as norm inequalitites, will keep their validity also in the restricted domain [0, 1]. The $n$ th-order polynomial approximant of ${J_0}(vx)$, say ${f_n}(v, x)$, is then introduced as follows:
$${{f_n}(v, x) = \sum\limits_{k = 0}^{n/2} {\epsilon _k} {{(- 1)}^k} J_k^2(v/2) {T_{2k}}(x) ,\quad x \in [- 1,1] ,}$$
where, for the sake of clarity, the explicit dependance on $v$ has been explitated.

Equations (10) and (11) have the form in Eqs. (4) and (5), respectively, and, in this particular case, it is possible to analytically evaluate all quantities into Eq. (7) except, of course, ${\omega _n}$. This implies that it should be possible to quantitatively compare the performances of Tchebychev’s approximant into Eq. (11) and the (unknown) best approximant ${Q_n}(x)$. The upper limit ${\cal S}_n^ \star$, defined by Eq. (8), is first evaluated, yielding

$${{\cal S}_n^ \star = 1 - \sum\limits_{k = 0}^{n/2} {\epsilon _k} J_k^2(v/2) ,}$$
which, in the present case, also coincides with ${S_n}$. As far as the quantity ${{\cal A}_n}$ is concerned, from Eqs. (9) and (11) it is not difficult to prove that (see Appendix A)
$$\begin{array}{*{20}{l}}{{{\cal A}_n} = {{\left[{_2{F_3}\left({\begin{array}{*{20}{c}}{1/2,1/2; - {v^2}}\\{1,1,1}\end{array}} \right) - \sum\limits_{k = 0}^{n/2} {\epsilon _k}J_k^4(v/2)} \right]}^{1/2}} ,}\end{array}$$
where $_2{F_3}$ denotes a generalized hypergeometric function.

In Fig. 1, an example of applications of the above inequalities is illustrated, also to give readers an idea about the powerfulness of Tchebychev’s approximation. In particular, in the left column, the behaviors of the quantities ${{\cal A}_n}$ and ${\cal S}_n^ \star$ are shown, as a function of $v$, as solid and dashed curves, respectively, for (a) $n = 4$, (c) $n = 8$, and (e) $n = 12$. The thin vertical line is located at the abscissa $v = 10$. It can be appreciated how, for a given $v$, on sufficiently increasing $n$ the curves representing ${A_n}$ and ${\cal S}_n^ \star$ becomes practically indistinguishable. According to Eq. (7), this implies that the $n$th-order polynomial approximant ${f_n}(v;x)$ given in Eq. (11) is expected to be close to the best $n$th-order polynomial approximant ${Q_n}(x)$ of the Bessel function ${J_0}(vx)$.

 figure: Fig. 1.

Fig. 1. Left column: behaviors of the quantities ${{\cal A}_n}$ (solid curve) and ${\cal S}_n^ \star$ (dashed curve), as a function of $v$, for (a) $n = 4$, (c) $n = 8$, and (e) ${n = 12}$. The thin vertical line is located at the abscissa $v = 10$. Right column: plots of polynomial approximations of the function ${J_0}(10x)$ (dashed curves) within the interval $x \in [0 ,1 ]$, (b) $n = 4$, (d) $n = 8$, and (f) $n = 12$. The solid curve denotes the exact profile of the Bessel function.

Download Full Size | PDF

This is confirmed, at least at a visual level, by the right column, where plots of different polynomial approximations of the function ${J_0}(10x)$ are shown (dashed curves) for the same values of the truncation $n$, namely, (b) $n = 4$, (d) $n = 8$, and (f) $n = 12$. In the next section, such a capability of Tchebychev’s polynomials to generate excellent low-order approximations will be employed to numerically revisit and improving the Nijboer–Zernike aberration perturbative theory.

3. APPLICATION TO NIJBOER–ZERNIKE’S DIFFRACTION THEORY OF ABERRATION

A. Brief Résumé of Nijboer’s Formalism

In a perfect optical system, the light waves, which proceed from each element of the object, emerge in the image space as convergent waves spherical in form. Such an instrument represents an idealization that cannot be realized in practice; in the image space of an actual instrument, the wave surfaces will, as a rule, be of a more complicated form.

These are the opening words of an important historical review of the subject of aberration theory, written in a 1951 paper by Wolf [8]. The paper formed the core of Ch. IX in Principles, one of the most beautiful of the whole opera. In the rest of paper, the Nijboer–Zernike diffraction aberration theory [46] will be quantitatively revisited in the light of that reported in the previous section. To this end, readers are encouraged to first go through ([1], Ch. IX) or even through [46,8] in order to familiarize with the notations as well as the meaning of the symbols, which will be employed in the rest of the paper.

Given a centered imaging optical system, the deformation from the perfectly spherical convergent wavefront in the region of the exit pupil, measured in reduced wavelength units $\rlap{--} \lambda = \lambda /2\pi$, will be denoted $\Phi (\tau ,\vartheta)$, where $(\tau ,\vartheta)$ represent polar coordinates upon the exit pupil plane, normalized so that $\tau$ ranges from 0 to 1.

Further, the image space will be parametrized by dimensionless cylindrical coordinates $(u,v,\varphi)$: the variable $u$, related to the Fresnel number, represents a dimensionless measure of the position of the transverse observation plane, with respect to the focal one (corresponding to $u = 0$). The couple $(v,\varphi)$ represents (dimensionless) polar coordinates across the typical plane $u = {\rm const}$. The diffracted wavefield, say $\Psi (u,v,\varphi)$, will be expressed, within paraxial approximation and apart from unessential overall phase and amplitude factors, by the following integral [8]:

$$\begin{split}{\Psi (u,v,\varphi) }&= \frac{1}{\pi}\int_0^1 \!\!\int_0^{2\pi} \exp \left[{\rm i}\frac{u}{2}{\tau ^2} + {\rm i}v\tau \cos (\vartheta - \varphi)\right.\\&\quad - \left.{\rm i} \Phi (\tau ,\vartheta) \vphantom{\frac{u}{2}}\right]\tau {\rm d}\tau {\rm d}\vartheta .\end{split}$$
Note that the sign choice into Eq. (14) follows that in [5,8], which is the opposite of that in [1].

For an optical system free of any aberration, i.e., when $\Phi \equiv 0$, the integral in Eq. (14) turns out to be equal to $2{\cal L}_0^0(u,v)$, where [1]

$${{\cal L}_0^0(u,v) = \int_0^1 \exp \left({{\rm i}\frac{u}{2}{\tau ^2}} \right) {J_0}(v \tau) \tau {\rm d}\tau }$$
will be referred to as von Lommel’s integral. Suitable generalizations of Eq. (15) will be considered in the following, and precisely
$${{\cal L}_\ell ^m(u,v) = \int_0^1 \exp \left({{\rm i}\frac{u}{2}{\tau ^2}} \right) {J_\ell}(v \tau) {\tau ^{m + 1}} {\rm d}\tau ,}$$
where $m$ and $\ell$ denote nonnegative integer parameters.

The perturbative Nijboer approach to the diffraction theory of aberrations was conceived to deal with “small” values of $|\Phi |$, typically of the order of unity. To this end, the aberration function $\Phi$ is first written in the following form:

$${\Phi (\tau ,\vartheta) = \beta R_n^m(\tau) \cos m\vartheta ,}$$
with $\beta$ being a nonnegative, dimensionless parameter, which gives an account of the aberration strength, and $R_n^m(\cdot)$ denotes the Zernike polynomial [1], which, within the interval $[0 , 1]$, satisfies the inequality $|R_n^m(\tau)| \le 1$. The subsequent step consists in substituting from Eq. (17) into Eq. (14) then expanding the result in powers of $\beta$. The nontrivial and delicate treatment, which is fully detailed in Ch. IX in Principles, eventually leads to the following expansion [8]:
$${\Psi (u,v,\varphi) = \sum\limits_{k = 0}^\infty \frac{{{{({\rm i}\beta)}^k}}}{{k!}} {{\cal C}_k}(u,v;n,m) ,}$$
where functions ${{\cal C}_k}(u,v;n,m)$ can be expressed in terms of suitable 1D integrals, which will be reduced to linear combinations of generalized Lommel integrals defined through Eq. (16). For $|\beta | \le 1$, the expansion in Eq. (18) can safely be truncated at $k = 4$, with fifth and higher powers of $\beta$ being neglected. Nijboer also found the explicit integral representations of the whole sequence $\{{{\cal C}_k}(u,v;n,m)\} _{k = 0}^4$, which are reported, for reader’s convenience, in Appendix B.

B. Numerical Computation of Nijboer’s Diffraction Integrals

The numerical computation of the contributions into the series expansion in Eq. (18) is the central practical problem of such perturbative approach to aberration theory. Nijboer developed a clever analytical treatment based on basic properties of Zernike’s polynomials in order to represent each integral in Eq. (B1) through a suitable series of Bessel functions. In particular, such series terminate when $u = 0$, i.e., across the focal plane, thus allowing the sequence $\{{{\cal C}_k}(0,v;n,m)\} _{k = 0}^4$ to be exactly found. The same is no longer true when the computation is required out of the focal plane, i.e., for $u \ne 0$.

In order for the practical feasibility of Nijboer’s theory to be extended for dealing with any couple $(u,v)$, the following idea will be pursued: given a primary aberration described by Eq. (17), each power of the Zernike polynomial $R_m^n(\tau)$ into the diffraction integrals of Eq. (B1) will be expanded as powers of $\tau$. In this way, each term of the sequence $\{{{\cal C}_k}(u,v;n,m)\} _{k = 0}^4$ will be recast as a linear combination of generalized Lommel integrals, as shown in Eq. (16). The latter can be converted into rapidly converging series by using the generalization of Eq. (10) to expand Bessel functions ${J_\ell}(v\tau)$, with $\ell \ne 0$, in terms of Tchebychev polynomials. In particular, the following expansions will be employed ([16], Sec. 9.3.6):

$${{J_{2p}}(v \tau) = \sum\limits_{k = 0}^\infty {\epsilon _k} {J_{p + k}}(v/2) {J_{p - k}}(v/2) {T_{2k}}(\tau) ,}$$
which is valid for even values of the Bessel function order ${\ell = 2p}$, and
$${{J_{2p + 1}}(v \tau) = 2\sum\limits_{k = 0}^\infty {J_{p + k + 1}}(v/2) {J_{p - k}}(v/2) {T_{2k + 1}}(\tau) ,}$$
which is valid for odd $\ell = 2p + 1$. It should be noted, as pointed out in Section 2, that both expansions are formally valid within the symmetric interval $\tau \in [- 1,1]$. However, due to the parity properties of Bessel functions of integer orders ${J_\ell}$, the above expansions will be employed only for $\tau \in [ 0, 1]$.

For $\ell \ne 0$, the lower bound ${{\cal A}_n}$ of the maximum deviation ${{\cal S}_n}$, defined in Eq. (9), can still be exactly computed for any $\ell$, as detailed in Appendix A, which yields

$$\begin{split}\mathcal{A}_{n}^{2}(\ell ,v) & =\frac{{{v}^{2{\ell} }}\Gamma {{\left( \ell +\frac{1}{2} \right)}^{2}}}{\pi }{{}_{2}}{{{\tilde{F}}}_{3}}\left( \begin{array}{c} \ell +\frac{1}{2},\ell +\frac{1}{2};-{{v}^{2}} \\[3pt] \ell +1,\ell +1,2\ell +1 \end{array} \right)\\[-4pt]&\quad-\sum\limits_{k=0}^{\lfloor n/2 \rfloor}{}a_{k}^{2}(\ell ,v), \end{split}$$
where the symbol $_2{\tilde F_3}(\cdot)$ denotes a regularized hypergeometric function and
$${{a}_{k}}(\ell ,v)=\left\{\begin{array}{*{20}{l}} {{\epsilon }_{k}}{{J}_{\ell /2+k}}(v/2){{J}_{\ell /2-k}}(v/2), &\;\; {\ell} \; {\rm even}, \\[3pt] 2{{J}_{\lfloor\ell /2 \rfloor + 1 + k}}(v/2){{J}_{\lfloor\ell /2\rfloor - k}}(v/2), & \;\;{\ell} \;{\rm odd}. \\ \end{array} \right.$$
On the contrary, it seems no longer possible to achieve a closed-form expression also for the upper bound ${\cal S}_n^ \star (\ell ,v)$, differently from what we did in Eq. (12) for $\ell = 0$. However, its numerical evaluation does not present any problems, due to the rapid convergence of the single terms $|{a_k}(\ell ,v)|$ to zero for $k \gg \lfloor n/2 \rfloor$. Accordingly, the following inequality holds:
$${{{\cal A}_n}(\ell ,v) \le \mathop {{\rm max}}\limits_{x \in [- 1,1]} |{J_\ell}(vx) - {f_n}(v,x)| \le {\cal S}_n^ \star (\ell ,v) .}$$

In the following, each generalized Lommel integral in Eq. (16) will be replaced by its $N$th-order approximant, defined by

$${\cal L}_{\ell }^{m}(u,v)\simeq \left\{ \begin{array}{*{20}{l}} \sum\limits_{k=0}^{\lfloor N/2 \rfloor}{}{{a}_{k}}(\ell ,v)\mathcal{T}_{2k}^{m}(u/2), & {\ell} \;{\rm even}, \\[8pt] \sum\limits_{k=0}^{\lfloor N/2 \rfloor}{{a}_{k}}(\ell ,v)\mathcal{T}_{2k+1}^{m}(u/2), & {\ell} \; {\rm odd}, \end{array} \right.$$
where functions ${\cal T}_k^m(\cdot)$ are defined by
$${{\cal T}_k^m(\alpha) = \int_0^1 \exp ({\rm i}\alpha {\tau ^2}) {T_k}(\tau) {\tau ^{m + 1}} {\rm d}\tau ,}$$
with $m,k \ge 0$, and where $\alpha$ can take, in principle, any complex value. Differently from ${\cal L}_\ell ^m(u,v)$, functions ${\cal T}_k^m(\alpha)$ can exactly be evaluated through the following recursive algorithm:
$${{\cal T}_{k + 1}^m(\alpha) = 2 {\cal T}_k^{m + 1}(\alpha) - {\cal T}_{k - 1}^m(\alpha) ,\quad k \ge 1 ,m \ge - 1 ,}$$
with the initial conditions
$$\begin{split}&{{\cal T}_0^m(\alpha) = \frac{1}{{2(- {\rm i}\alpha {)^{1 + m/2}}}} \left[{\Gamma \left({\frac{m}{2} + 1} \right) - \Gamma \left({\frac{m}{2} + 1, - {\rm i}\alpha} \right)} \right] ,}\\&{{\cal T}_1^m(\alpha) = {\cal T}_0^{m + 1}(\alpha) .}\end{split}$$
Here, symbols $\Gamma (\cdot)$ and $\Gamma (\cdot , \cdot)$ denote the gamma function and the incomplete gamma function, respectively.

C. Preliminary Numerical Checks

Since the central idea is to replace, into Eq. (16), the Bessel function ${J_\ell}(v\tau)$ via its $N$ th-order approximant ${f_N}(v,\tau)$, we introduce the absolute error, say ${\cal E}$, i.e., the difference between the exact value in Eq. (16) and the approximated one in Eq. (24), which can be recast as follows:

 figure: Fig. 2.

Fig. 2. Behavior of the absolute error in Eq. (28) obtained by evaluating Lommel’s function ${\cal L}_0^0(0,v)$ via Eq. (24) as a function of the truncation order $N$ (dots), for (a) $v = 10$, (b) $v = 20$, (c) $v = 30$, and (d) $v = 40$. The open circles represent the values of the upper bound given by Eq. (29).

Download Full Size | PDF

$${{\cal E} = \int_0^1 \exp \left({{\rm i}\frac{u}{2} {\tau ^2}} \right) \left[{{J_\ell}(v\tau) - {f_N}(v,\tau)} \right] {\tau ^{m + 1}} {\rm d}\tau ,}$$
so that, on taking into account the conjecture in Eq. (23), the following upper limit can straightforwardly be given by
$${|{\cal E}| \le \int_0^1 \left| {{J_\ell}(v\tau) - {f_N}(v,\tau)} \right| {\tau ^{m + 1}} {\rm d}\tau \lt \frac{{{\cal S}_n^ \star (\ell ,v)}}{{m + 2}} .}$$
Although Eq. (29) could appear, in principle, quite attractive from a mathematical point of view, its practical usefulness is, unfortunately, limited. To show why, it is worth checking the predictive capabilities of our algorithm on test cases for which analytical solutions are already available.

For instance, when $u = 0$, each of the integrals into Eq. (B1) have been evaluated in closed form by Nijboer in the case of primary aberrations, e.g., spherical aberration, coma, astigmatism. It is then rather natural to use Nijboer’s own results to carry out preliminary checks about the numerical efforts required by the present method.

The simplest test consists in evaluating Lommel’s integral ${\cal L}_0^0(u,v)$, which, across the focal plane, $u = 0$ reduces to

$${{\cal L}_0^0(0,v) = \int_0^1 {J_0}(v \tau) \tau {\rm d}\tau = \frac{{{J_1}(v)}}{v} .}$$
In Fig. 2, the behavior of the absolute error in Eq. (28) obtained by evaluating Lommel’s function ${\cal L}_0^0(0,v)$ via Eq. (24) is plotted against the truncation order $N$ (dots), for (a) $v = 10$, (b) $v = 20$, (c) $v = 20$, and (d) $v = 40$. The open circles represent the values of the upper bound given by Eq. (29).

Another example of exact evaluation of Lommel’s function ${\cal L}_0^0$ is when $u = v$, which yields [17]

$${{\cal L}_0^0(v,v) = \frac{1}{{2{\rm i}v}} \left[{\exp \left({{\rm i}\frac{v}{2}} \right) {J_0}(u) - \exp \left({- {\rm i}\frac{v}{2}} \right)} \right] ,}$$
which occurs, for example, when the field diffracted by an aberration-free imaging optical system has to be evaluated along the boundary of the geometrical shadow [1].

Similarly, as done in Fig. 2, in Fig. 3 the behavior of the absolute error in Eq. (28) obtained by evaluating Lommel’s function ${\cal L}_0^0(v,v)$ via Eq. (24) is plotted against the truncation order $N$ (dots) for (a) $v = 10$, (b) $v = 20$, (c) $v = 30$, and (d) $v = 40$. The open circles represent again the values of the upper bound given by Eq. (29).

 figure: Fig. 3.

Fig. 3. Behavior of the absolute error in Eq. (28) obtained by evaluating Lommel’s function ${\cal L}_0^0(v,v)$ via Eq. (24) as a function of the truncation order $N$ (dots), for (a) $v = 10$, (b) $v = 20$, (c) $v = 30$, and (d) $v = 40$. The open circles represent the values of the upper bound given by Eq. (29).

Download Full Size | PDF

Per Figs. 2 and 3 (and for several others’ error behaviors not show here), it is possible to understand why the upper bound in Eq. (29) is not numerically significative to guarantee an efficient choice of the truncation order $N$.

Nevertheless, it can be appreciated, within the behavior of the absolute error, a sort of plateau, which extends for values of the truncation orders up to, roughly, $v/2$. When $N \gt v/2$, the error reduction considerably increases, according to the exponential law represented in figures by thick black lines. The following rule for the truncation error choice will then be adopted:

$${N = \left\lceil {\frac{v}{2} + K} \right\rceil ,}$$
where the values of the integer $K$ have to be chosen depending on the particular case under test.

D. Recurrence Relation for von Lommel’s Integrals

The numerical evaluation of the generalized von Lommel integrals ${\cal L}_\ell ^m(u,v)$ can further be improved by deriving a simple but computationally important recurrence relation, based on a well-known property of Bessel functions. To this end, consider Eq. (16) written for $\ell + 1$, i.e.,

$${{\cal L}_{\ell + 1}^m(u,v) = \int_0^1 \exp \left({{\rm i}\frac{u}{2}{\tau ^2}} \right) {J_{\ell + 1}}(v \tau) {\tau ^{m + 1}} {\rm d}\tau ,}$$
where the Bessel function ${J_{\ell + 1}}$ can be recast as follows:
$${{J_{\ell + 1}}(x) = \frac{{2\ell}}{x} {J_\ell}(x) - {J_{\ell - 1}}(x) ,}$$
so that, after simple algebra, Eq. (33) takes on the following form:
$$\begin{array}{*{20}{l}}{{\cal L}_{\ell + 1}^m(u,v) = \frac{{2\ell}}{v} {\cal L}_\ell ^{m - 1}(u,v) - {\cal L}_{\ell - 1}^m(u,v) ,\quad \begin{array}{*{20}{c}}{\ell = 1, 2, \ldots ,}\\{m = 0, 1, \ldots}\end{array}}\end{array}.$$
The above recurrence relation has to be employed starting from the initial values ${\{{\cal L}_\ell ^{- 1}(u,v)\} _{\ell \ge 0}}$, ${\{{\cal L}_0^m(u,v)\} _{m \ge 0}}$, and ${\{{\cal L}_1^m(u,v)\} _{m \ge 0}}$. In this way, the evaluation of a 2D matrix of integrals $\{{\cal L}_\ell ^m(u,v)\} _{\ell \ge 0}^{m \ge - 1}$ is reduced to the evaluation of only three 1D vectors of integrals, as sketched into Fig. 4.
 figure: Fig. 4.

Fig. 4. Recurrence relation in Eq. (35).

Download Full Size | PDF

It is worth giving readers at least an idea about the computational times of the present method, compared to those pertinent to standard integration packages of Mathematica. To this end, the evaluation of the quantity ${\cal L}_{25}^{- 1}(60,2\pi)$ has been carried out on using both Eq. (24) and the Mathematica command NIntegrate. The result of the computational experiment is reported in Table 1. The first column contains the value of the parameter $K$ to be inserted into Eq. (32) in such a way a relative error of the order of magnitude given in the fifth column can be reached. The third column contains the value of the corresponding WorkingPrecision parameter to be set into the NIntegrate command in order for the same relative error order to be achieved. The second and the fourth columns contain the computational times (expressed in seconds). The relative errors have been computed with respect to an “exact” value of ${\cal L}_{25}^{- 1}(60,2\pi)$ evaluated again via Mathematica NIntegrate command, but on setting WorkingPrecision to 200. From the table it is possible to realize that the computational speed of the algorithm of Eq. (24) turns out to be, for a given relative error order, about one order of magnitude greater than the speed of the standard Mathematica integration package, regardless of the level of accuracy required. The results of several other numerical experiments, not reported here, fully confirmed such a speed difference.

Tables Icon

Table 1. Evaluation of the Quantity ${\cal L}_{25}^{- 1}(60,2\pi)$a

 figure: Fig. 5.

Fig. 5. Behavior of the relative error obtained by evaluating the diffraction integral $V_n^m(u,v)$ defined through Eq. (36) via the expansions given in Eq. (37) (solid curve) and in Eq. (40) (dots). The chosen values of the parameters are $n = 25$, $u = 60$, and $v = 2\pi$, while (a) $m = 1$, (b) $m = 5$, (c) $m = 15$, and (d) $m = 25$. The “exact” value has been evaluated through the standard integration package of Mathematica.

Download Full Size | PDF

E. Numerical Comparison with Janssen’s Extended Nijboer–Zernike Theory

The present section will be devoted to present numerical comparisons between our algorithm and Janssen’s extended Nijboer–Zernike [9,10,13]. Although the evaluation of generalized von Lommel’s integrals in Eq. (16) was also considered in [10], here we prefer to employ our algorithm to evaluate directly the following integral:

$$V_n^m(u,v) = \int_0^1 R_n^m(\tau) \exp \left({{\rm i} \frac{u}{2} {\tau ^2}} \right) {J_m}(v \tau) \tau {\rm d} \tau ,$$
which, with a slight change of notation, was posed as the main computational brick of Janssen’s approach, where in the following $n$ and $m$ will be chosen such a way that $n - m$ is even. In [9,12], the above integral representation of $V_n^m(u,v)$ has been converted into a notable power series expansion, which will be referred to as the first Janssen formula, namely [9,10,13],
$$V_n^m(u,v) = \exp \left({{\rm i} \frac{u}{2}} \right) \sum\limits_{l = 1}^\infty \frac{{{{(- {\rm i}u/v)}^{l - 1}}}}{l} \sum\limits_{j = 0}^p {{\cal V}_{\textit{lj}}} \frac{{{J_{m + l + 2j}}(v)}}{v} ,$$
where $p = \frac{{n - m}}{2}$ and $q = \frac{{n + m}}{2}$, and the expanding coefficients $\{{{\cal V}_{\textit{lj}}}\}$ are defined by
$$\begin{split} {{{\cal V}_{\textit{lj}}}} &= {{(- 1)}^p} (m + l + 2j)\left[\!\vphantom{\left({\begin{array}{*{20}{c}}{m + j + l - 1}\\{l - 1}\end{array}} \right)^1}\left({\begin{array}{*{20}{c}}{m + j + l - 1}\\{l - 1}\end{array}} \right) \left({\begin{array}{*{20}{c}}{j + l - 1}\\{l - 1}\end{array}} \right)\right.\\&\quad\times\left.\left({\begin{array}{*{20}{c}}{l - 1}\\{p - j}\end{array}} \right){{\left({\begin{array}{*{20}{c}}{q + l + j}\\l\end{array}} \right)}^{\!- 1}} \right] ,\end{split}$$
with $l = 1,2, \ldots$ and $j = 0,1, \ldots ,p$.

In [13], Haver and Janssen pointed out how the convergence of the series into Eq. (37) becomes numerically problematic when $u$ becomes greater than about 50. This is due in part to the fact that the truncation order of Janssen’ series grows with the modulus of $u$. To overcome such a problem, Janssen and co-authors gave the series in Eq. (37) new, computationally more efficient forms, which will not be considered here [12,13]. We shall limit ourselves to work on a single example, namely, the computation of $V_n^m(u,v)$, with $n = 25$, $u = 60$, and $v = 2\pi$, and for different (odd) values of $m$.

To this end, it is worth recalling the explicit definition of Zernike polynomial’s $R_n^m(\tau)$, namely,

$$R_n^m(\tau) = \sum\limits_{k = 0}^p {(- 1)^k} \left({\begin{array}{*{20}{c}}{n - k}\\k\end{array}} \right) \left({\begin{array}{*{20}{c}}{n - 2k}\\{p - k}\end{array}} \right){\tau ^{n - 2k}} ,$$
which, once substituted into Eq. (36), gives at once
$$V_n^m(u,v) = \sum\limits_{k = 0}^p {(- 1)^k} \left({\begin{array}{*{20}{c}}{n - k}\\k\end{array}} \right) \left({\begin{array}{*{20}{c}}{n - 2k}\\{p - k}\end{array}} \right){\cal L}_m^{n - 2k}(u,v) .$$

In Fig. 5, the behavior of the relative error, against the truncation orders, obtained by evaluating the diffraction integral $V_{25}^m(60,2\pi)$ via the expansions given in Eq. (37) (solid curve) and in Eq. (40) (dots), is plotted for (a) $m = 1$, (b) $m = 5$, (c) $m = 15$, and (d) $m = 25$. The “exact” value has again been evaluated through the standard integration package of Mathematica. It should be appreciated how the numerical effort required to Eq. (40) to evaluate the integral is substantially independent of $m$, being the truncation order given by Eq. (32), with $K$ not exceeding 23. Moreover, the error reduction is monotonically decreasing on growing the truncation order, i.e., the parameter $K$. The above are the main differences with respect Janssen’ series, whose truncation order displays a clear dependance on $m$, and the partial sums into Eq. (37) do not monotonically decrease on increasing their order.

 figure: Fig. 6.

Fig. 6. Isophotes in the presence of third-order spherical aberration of the amount $\beta = 1/2$. The figure should be visually compared with Fig. 3 in [8].

Download Full Size | PDF

 figure: Fig. 7.

Fig. 7. Same as in Fig. 6, but for a fifth-order spherical aberration. The figure should be visually compared with Fig. 4 in [8].

Download Full Size | PDF

F. Third- and Fifth-Spherical Aberration

The first example of application of our algorithm to Nijboer–Zernike’s theory concerns spherical aberration, characterized by the following phase function inside the integral in Eq. (14):

$${{\Phi _{{\rm sph}}}(\tau ,\vartheta) = \beta R_n^0(\tau) .}$$
Since the phase is independent of the angular variable $\vartheta$, the resulting diffracted field, say ${\Psi _{{\rm sph}}}$, turns out to be radial, too. Moreover, it is not difficult to prove that all contributions in Eq. (B1) take on the simpler form
$${{{\cal C}_k}(u,v;n,0) = 2 {{(- 1)}^k} \int_0^1 \exp \left({{\rm i}\frac{u}{2}{\tau ^2}} \right){{[R_n^0(\tau)]}^k}{J_0}(v\tau)\tau {\rm d}\tau ,}$$
for $k = 0, 1, 2, \ldots$, which allows the numerical evaluation of the field to be easily achieved in the whole space $(u,v)$.

Consider now third- and fifth-order spherical aberrations, corresponding to set into Eq. (41) $n = 4$ and $n = 6$, respectively. The parameter $\beta$ will be set to 1/2, in order to reproduce Figs. 3 and 4 in [8]. To this end, it is sufficient to expand the factors ${[R_n^0(\tau)]^k}$ into the integrals in Eq. (42) and then to apply the definition in Eq. (16) to recast them as a linear combination of generalized Lommel integrals ${\cal L}_0^m(u,v)$. For space reasons, the explicit expressions of ${{\cal C}_k}$’s will be not reported here. Each function ${\cal L}_0^m(u,v)$ is then numerically evaluated by suitably choosing the truncation order $N$ via Eq. (32). Results are shown in Figs. 6 and 7, which are aimed at reproducing the isophotes of Figs. 3 and 4 in [8], respectively. The visual agreement is perfect, and several numerical tests (not shown here) were performed directly on the exact analytical expressions of the focal field, for further checking purposes.

An important comment, of a numerical nature, in regard to the above figure: once the truncation order $N$ has been fixed, our algorithmic efforts are independent of the variabile $u$, i.e., of the position of the observation plane. This represents, in our opinion, one of the most important features of the present approach, which is basically due to the marriage between the simplicity and elegance of Nijboer’s analytical treatment and the extraordinary efficiency of Bessel function Tchebychev’s expansions, which allows detailed maps of the diffracted field to be produced in a few minutes. To give an idea, the time needed to generate Figs. 6 and 7, both containing about 20,000 sampled points, was about 7 min, working with an unoptimized Mathematica code running onto a commercial laptop equipped with a 3 GHz Intel Core i7 processor and 16 GB RAM.

 figure: Fig. 8.

Fig. 8. Isophotes at the plane $u = 0$ in presence of a coma in Eq. (43) of amount (a) $\beta = 1$ and (b) $\beta = 3$. The figures should be compared with Figs. 5 and 6 in [8], respectively.

Download Full Size | PDF

G. Primary Coma

The coma aberration phase is [1,5]

$${{\Phi _{{\rm coma}}}(\tau ,\vartheta) = \beta R_3^1(\tau) \cos \vartheta ,}$$
where $R_3^1(\tau) = 3{\tau ^3} - 2\tau$. In this case, it is also not difficult, after substituting the explicit definition of $R_3^1$ into Eq. (B1), to recast all contributions ${{\cal C}_k}(u,v;3,1)$ in terms of generalized von Lommel’s functions, as shown in Eq. (16). A preliminary check has first been carried out with the exact results provided by Nijboer across the focal plane $u = 0$ [5]. A visual result is given in Fig. 8, where isophotes calculated across the focal plane are shown for (a) $\beta = 1$ and (b) $\beta = 3$.

In Fig. 9, a slightly modified version of the same code has been used to reproduce the numerical results presented in [9]. In particular, to adhere as much as possible with Janssen’s simulations, the aberration phase in Eq. (43) has been changed into ${\Phi _{{\rm coma}}} = \beta {\tau ^3}\cos \vartheta$, with $\beta = 1$. Isophotes of the diffracted field have been evaluated across the transverse planes (a) $u = 0$, (b) $u = \pi /2$, (c) $u = 2\pi$, and (d) $u = 4\pi$. Each figure contains a matrix of about $300 \times 200$ sampled points, with computational times of about 30 s per figure. Also in that case, the visual agreement with the results obtained by Janssen is excellent, once the considerably different densities of sampled points of the figures have been taken into account.

H. Primary Astigmatism

The last example of application deals with astigmatism, whose phase model ${\Phi _{{\rm ast}}}$ is

$${{\Phi _{{\rm ast}}}(\tau ,\vartheta) = \beta R_2^2(\tau) \cos 2\vartheta = \beta {\tau ^2} \cos 2\vartheta .}$$
The expressions of all five contributions ${{\cal C}_k}(u,v;2,2)$ are considerably easier with respect to spherical aberration and coma; in Fig. 10, the isophotes are plotted at the focal plane $u = 0$ for (a) $\beta = 1$ and (b) $\beta = 3$, in order to reproduce Figs. 7 and 8 of [8], respectively.
 figure: Fig. 9.

Fig. 9. Isophotes in presence of coma ${\Phi _{{\rm coma}}} = 3\beta {\tau ^3}\cos \vartheta$, for $\beta = 1/3$ at different transverse planes: (a) $u = 0$, (b) $u = \pi /2$, (c) $u = 2\pi$, and (d) $u = 4\pi$. The figure should be compared with Fig. 2 in [9].

Download Full Size | PDF

 figure: Fig. 10.

Fig. 10. Isophotes at the plane $u = 0$ in the presence of an astigmatism in Eq. (44) of amount (a) $\beta = 1$ and (b) $\beta = 3$. The figures should be compared with Figs. 7 and 8 in [8], respectively.

Download Full Size | PDF

Finally, Fig. 11 shows examples of field computation at different transverse planes as shown for $\beta = 1$. Isophotes of the diffracted field have been produced for (a) $u = 0$, (b) $u = \pi /2$, (c) $u = 2\pi$, and (d) $u = 4\pi$. Each figure contains a matrix of about $300 \times 200$ sampled points, with computational times of about 30 s per figure.

 figure: Fig. 11.

Fig. 11. Isophotes in the presence of astigmatism in Eq. (44) of amount $\beta = 1$ at different transverse planes: (a) $u = 0$, (b) $u = \pi /2$, (c) $u = 2\pi$, and (d) $u = 4\pi$.

Download Full Size | PDF

4. CONCLUSION

The present paper has been conceived as a tribute to Emil Wolf’s memory and legacy in respect to his authorship of Principle of Optics. The paper’s content was also intended to honor Wolf as a mathematician (he earned his PhD in math from Bristol University in 1948).

What has been celebrated is the marriage of two mathematical jewels: the Nijboer–Zernike aberration theory and the Tchebychev polynomial approximation theory. The result is a powerful computational tool able to evaluate, up to the desired degree of accuracy, all terms of classical perturbative treatment of primary aberrations for, in principle, any values of the physical parameters in the observation space. In particular, the computational effort required to achieve such a task turns out to be independent of the defocusing, a fact that could greatly improve the practical applicability of Nijboer–Zernike’s theory to a variety of scenarios, from vectorial approaches to aberration theory [11], to the numerical evaluation of Fresnel transforms [18,19]. We believe that the present paper could also open mathematical questions worthy of consideration, e.g., the use of different families of polynomials than Tchebychev to approximate continuous functions into the interval $[- 1,1]$, for instance, Legendre polynomials [20]. Another concerns the summability of the complete Nijboer’s perturbative series, in order to also deal with nonsmall values of $\beta$. To this end, an (symbolic language) algorithm for generating functions ${{\cal C}_k}(u,v;n,m)$ of arbitrary orders $k$ should first be conceived to also grasp information about their asymptotic behavior for $k \gg 1$. Then, suitable convergence acceleration algorithms, for instance, based on Padé approximants or on Levin-type nonlinear sequence transformations, could be used to practically sum Nijboer’s series.

In conclusion, no words would be better:

In our time of ever-increasing specialization, there is a tendency to concern ourselves with relatively narrow scientific problems. The broad foundations of our present-day scientific knowledge and its historical development tend to be forgotten too often. This is an unfortunate trend, not only because our horizon becomes rather limited and our perspective somewhat distorted, but also because there are many valuable lessons to be learned in looking back over the years during which the basic concepts and the fundamental laws of a particular scientific discipline were first formulated.

(Emil Wolf)

APPENDIX A: EXACT EVALUATION OF ${\omega _n}$’s LOWER BOUND ${A_n}$ DEFINED THROUGH EQ. (9)

Consider first the case of even-order Bessel functions as in the expansion in Eq. (19), for which only even-order expanding coefficients ${a_{2k}}$ will be considered, as we did for ${J_0}$. In particular, only even values of the approximant order $n$ will be chosen. In this way, we have, from Eq. (9),

$$\begin{split}A_n^2 = 2 \sum\limits_{k = 1}^\infty J_{p + k}^2(v/2) J_{p - k}^2(v/2) - 2\sum\limits_{k = 1}^{n/2} J_{p + k}^2(v/2) J_{p - k}^2(v/2) ,\end{split}$$
which, on using Formula (6.8.4.8) in [21], can be recast as follows:
$$A_n^2 = \frac{2}{\pi} \int_0^1 \frac{{J_{2p}^2(vx)}}{{\sqrt {1 - {x^2}}}} {\rm d}x - \sum\limits_{k = 0}^{n/2} {\epsilon _k} J_{p + k}^2(v/2) J_{p - k}^2(v/2) .$$
For odd-order Bessel functions ${J_{2p + 1}}$, it is still possible, by using Formula (6.8.4.9) in [21], to prove that
$$\begin{split}A_n^2 = \frac{2}{\pi} \int_0^1 \!\frac{{J_{2p + 1}^2(vx)}}{{\sqrt {1 - {x^2}}}} {\rm d}x - 2\sum\limits_{k = 0}^{(n - 1)/2}\! J_{p + 1 + k}^2(v/2) J_{p - k}^2(v/2) .\end{split}$$

Both integrals in Eqs. (A2) and (A3) can be expressed in closed form via hypergeometric functions through the following notable formula:

$$\begin{array}{*{20}{l}}{\frac{2}{\pi} \int_0^1 \frac{{J_\ell ^2(vx)}}{{\sqrt {1 - {x^2}}}} {\rm d}x = \frac{{{v^{2\ell}}\Gamma {{\left({\ell + \frac{1}{2}} \right)}^2}}}{\pi}{_2}{{\tilde F}_3}\left({\begin{array}{*{20}{c}}{\ell + \frac{1}{2},\ell + \frac{1}{2}; - {v^2}}\\{\ell + 1,\ell + 1,2\ell + 1}\end{array}} \right) ,}\end{array}$$
where the symbol $_2{\tilde F_3}(\cdot)$ denotes the regularized hypergeometric function.

APPENDIX B: EXPLICIT EXPRESSION OF ${{\cal C}_k}(u,v;n,m)$

Following Nijboer [5], the whole sequence $\{{{\cal C}_k}(u,v;\def\LDeqbreak{}n,m)\} _{k = 0}^4$ is

$$\begin{array}{rl}{{\cal C}_0}(u,v;n,m) &= 2\int_0^1 \exp \left({{\rm i}\frac{u}{2}{\tau ^2}} \right){J_0}(v\tau)\tau {\rm d}\tau ,\\[6pt]{{\cal C}_1}(u,v;n,m)& = - 2{{\rm i}^m} \cos m\varphi \int_0^1 \exp \left({{\rm i}\frac{u}{2}{\tau ^2}} \right)\\[6pt]&\quad\times\, R_n^m(\tau){J_m}(v\tau)\tau {\rm d}\tau \\[6pt]{{\cal C}_2}(u,v;n,m)& = \int_0^1 \exp \left({{\rm i}\frac{u}{2}{\tau ^2}} \right){{[R_n^m(\tau)]}^2}{J_0}(v\tau)\tau {\rm d}\tau \\[6pt]&\quad +\, {{\rm i}^{2m}} \cos 2m\varphi \int_0^1 \exp \left({{\rm i}\frac{u}{2}{\tau ^2}} \right){{[R_n^m(\tau)]}^2}\\&\quad\times\,{J_{2m}}(v\tau)\tau {\rm d}\tau ,\\[6pt]{{\cal C}_3}(u,v;n,m)& = - \frac{1}{2} \left\{3{{\rm i}^m} \cos m\varphi \int_0^1 \exp \left({{\rm i}\frac{u}{2}{\tau ^2}} \right){{[R_n^m(\tau)]}^3}\right.\\[6pt]&\quad\times\,{J_m}(v\tau)\tau {\rm d}\tau + {{\rm i}^{3m}} \cos 3m\varphi \int_0^1 \exp \left({{\rm i}\frac{u}{2}{\tau ^2}} \right)\\[6pt]&\quad\times\left.{{[R_n^m(\tau)]}^3}{J_{3m}}(v\tau)\tau {\rm d}\tau \vphantom{\left({{\rm i}\frac{u}{2}{\tau ^2}} \right)\int^1_0}\right\} ,\\[6pt]{{\cal C}_4}(u,v;n,m)& = \frac{1}{4} \left\{3\int_0^1 \exp \left({{\rm i}\frac{u}{2}{\tau ^2}} \right){{[R_n^m(\tau)]}^4}{J_0}(v\tau)\tau {\rm d}\tau \right.\\[6pt]&\quad+\, 4{{\rm i}^{2m}} \cos 2m\varphi \int_0^1 \exp \left({{\rm i}\frac{u}{2}{\tau ^2}} \right){{[R_n^m(\tau)]}^4}\\&\quad\times\,{J_{2m}}(v\tau)\tau {\rm d}\tau+ {{\rm i}^{4m}} \cos 4m\varphi \int_0^1 \exp \left({{\rm i}\frac{u}{2}{\tau ^2}} \right)\\&\quad\times{{[R_n^m(\tau)]}^4}\left.{J_{4m}}(v\tau)\tau {\rm d}\tau \vphantom{\left({{\rm i}\frac{u}{2}{\tau ^2}} \right)^1}\right\} .\end{array}$$

Acknowledgment

I am grateful to the reviewers for their criticisms and suggestions, always aimed at improving the quality of the present work. I also wish to thank T. M. Spinozzi for his invaluable help during the preparation of the manuscript. The Grant of Excellence Departments, MIUR (art. 1, 432 cc. 314-317, Lex 232/2016) to the Dipartimento di Ingegneria, Università Roma Tre, is gratefully acknowledged.

Disclosures

The author declares no conflicts of interest.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the author upon reasonable request.

REFERENCES AND NOTE

1. M. Born and E. Wolf, Principles of Optics, 7th ed. (Cambridge University, 1999).

2. http://www2.optics.rochester.edu/~stroud/BookHTML/ChapIV_pdf/IV_29.pdf.

3. B. R. A. Nijboer, “The diffraction theory of aberrations,” Ph.D. dissertation (University of Groningen, 1942).

4. B. R. A. Nijboer, “The diffraction theory of optical aberrations. Part I: general discussion of the geometrical aberrations,” Physica 10, 679–692 (1943). [CrossRef]  

5. B. R. A. Nijboer, “The diffraction theory of optical aberrations. Part II: diffraction pattern in the presence of small aberrations,” Physica 13, 605–620 (1947). [CrossRef]  

6. K. Nienhuis and B. R. A. Nijboer, “The diffraction theory of optical aberrations. Part III: general formulae for small aberrations: experimental verification of the theoretical results,” Physica 14, 605–608 (1949). [CrossRef]  

7. F. Zernike and B. R. A. Nijboer, Contribution in La Théorie des Images Optiques (Editions de la Revue d’Optique, 1949), p. 227.

8. E. Wolf, “The diffraction theory of aberrations,” Rep. Prog. Phys. 14, 95–120 (1951). [CrossRef]  

9. A. Janssen, “Extended Nijboer-Zernike approach for the computation of optical point-spread functions,” J. Opt. Soc. Am. A 19, 849–857 (2002). [CrossRef]  

10. J. Braat, P. Dirksen, and A. Janssen, “Assessment of an extended Nijboer-Zernike approach for the computation of optical point-spread functions,” J. Opt. Soc. Am. A 19, 858–870 (2002). [CrossRef]  

11. J. Braat, P. Dirksen, A. Janssen, and A. van de Nes, “Extended Nijboer-Zernike representation of the vector field in the focal region of an aberrated high-aperture optical system,” J. Opt. Soc. Am. A 20, 2281–2292 (2003). [CrossRef]  

12. A. Janssen, J. Braatz, and P. Dirksen, “On the computation of the Nijboer-Zernike aberration integrals at arbitrary defocus,” J. Mod. Opt. 51, 687–703 (2004). [CrossRef]  

13. S. van Haver and A. J. E. M. Janssen, “Advanced analytic treatment and efficient computation of the diffraction integrals in the extended Nijboer-Zernike theory,” J. Eur. Opt. Soc. Rap. Publ. 8, 13044 (2013). [CrossRef]  

14. I. S. Gardshteyn and I. M. Ryzhik, Table of Integrals, Series, and Products (Academic, 2007).

15. Y. L. Luke, The Special Functions and Their Approximations (Academic, 1969), Vol. I.

16. Y. L. Luke, The Special Functions and Their Approximations (Academic, 1969), Vol. II.

17. E. H. Linfoot and E. Wolf, “Phase distribution near focus in an aberration free diffraction image,” Proc. Phys. Soc. B 69, 823–832 (1956). See also ([1], Sec. 8.8.2). [CrossRef]  

18. Y. Wu and D. Kelly, “Paraxial light distribution in the focal region of a lens: a comparison of several analytical solutions and a numerical result,” J. Mod. Opt. 61, S57–S67 (2014). [CrossRef]  

19. Y. Wu, M. Hillenbrand, L. Zhao, S. Sinzinger, and D. Kelly, “Fresnel transform as a projection onto a Nijboer–Zernike basis set,” Opt. Lett. 40, 3275–3472 (2015). [CrossRef]  

20. J. P. Boyd and R. Petschek, “The relationships between Chebyshev, Legendre and Jacobi Polynomials: the generic superiority of Chebyshev polynomials and three important exceptions,” J. Sci. Comput. 59, 1–27 (2014). [CrossRef]  

21. Y. Brychov, Handbook of Special Functions (Taylor and Francis, 2008).

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the author upon reasonable request.

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 (11)

Fig. 1.
Fig. 1. Left column: behaviors of the quantities ${{\cal A}_n}$ (solid curve) and ${\cal S}_n^ \star$ (dashed curve), as a function of $v$ , for (a)  $n = 4$ , (c)  $n = 8$ , and (e)  ${n = 12}$ . The thin vertical line is located at the abscissa $v = 10$ . Right column: plots of polynomial approximations of the function ${J_0}(10x)$ (dashed curves) within the interval $x \in [0 ,1 ]$ , (b)  $n = 4$ , (d)  $n = 8$ , and (f)  $n = 12$ . The solid curve denotes the exact profile of the Bessel function.
Fig. 2.
Fig. 2. Behavior of the absolute error in Eq. (28) obtained by evaluating Lommel’s function ${\cal L}_0^0(0,v)$ via Eq. (24) as a function of the truncation order $N$ (dots), for (a)  $v = 10$ , (b)  $v = 20$ , (c)  $v = 30$ , and (d)  $v = 40$ . The open circles represent the values of the upper bound given by Eq. (29).
Fig. 3.
Fig. 3. Behavior of the absolute error in Eq. (28) obtained by evaluating Lommel’s function ${\cal L}_0^0(v,v)$ via Eq. (24) as a function of the truncation order $N$ (dots), for (a)  $v = 10$ , (b)  $v = 20$ , (c)  $v = 30$ , and (d)  $v = 40$ . The open circles represent the values of the upper bound given by Eq. (29).
Fig. 4.
Fig. 4. Recurrence relation in Eq. (35).
Fig. 5.
Fig. 5. Behavior of the relative error obtained by evaluating the diffraction integral $V_n^m(u,v)$ defined through Eq. (36) via the expansions given in Eq. (37) (solid curve) and in Eq. (40) (dots). The chosen values of the parameters are $n = 25$ , $u = 60$ , and $v = 2\pi$ , while (a)  $m = 1$ , (b)  $m = 5$ , (c)  $m = 15$ , and (d)  $m = 25$ . The “exact” value has been evaluated through the standard integration package of Mathematica.
Fig. 6.
Fig. 6. Isophotes in the presence of third-order spherical aberration of the amount $\beta = 1/2$ . The figure should be visually compared with Fig. 3 in [8].
Fig. 7.
Fig. 7. Same as in Fig. 6, but for a fifth-order spherical aberration. The figure should be visually compared with Fig. 4 in [8].
Fig. 8.
Fig. 8. Isophotes at the plane $u = 0$ in presence of a coma in Eq. (43) of amount (a)  $\beta = 1$ and (b)  $\beta = 3$ . The figures should be compared with Figs. 5 and 6 in [8], respectively.
Fig. 9.
Fig. 9. Isophotes in presence of coma ${\Phi _{{\rm coma}}} = 3\beta {\tau ^3}\cos \vartheta$ , for $\beta = 1/3$ at different transverse planes: (a)  $u = 0$ , (b)  $u = \pi /2$ , (c)  $u = 2\pi$ , and (d)  $u = 4\pi$ . The figure should be compared with Fig. 2 in [9].
Fig. 10.
Fig. 10. Isophotes at the plane $u = 0$ in the presence of an astigmatism in Eq. (44) of amount (a)  $\beta = 1$ and (b)  $\beta = 3$ . The figures should be compared with Figs. 7 and 8 in [8], respectively.
Fig. 11.
Fig. 11. Isophotes in the presence of astigmatism in Eq. (44) of amount $\beta = 1$ at different transverse planes: (a)  $u = 0$ , (b)  $u = \pi /2$ , (c)  $u = 2\pi$ , and (d)  $u = 4\pi$ .

Tables (1)

Tables Icon

Table 1. Evaluation of the Quantity L 25 1 ( 60 , 2 π ) a

Equations (49)

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

1 1 T n ( x ) T m ( x ) 1 x 2 d x = π { δ n , m 2 m 0 , n 0 , 1 m = n = 0 ,
T n ( cos θ ) = cos n θ , n 0 .
ω n = m a x x [ 1 , 1 ] | f ( x ) Q n ( x ) | < m a x x [ 1 , 1 ] | f ( x ) p n ( x ) | ,
f ( x ) = k = 0 a k T k ( x ) , x [ 1 , 1 ] ,
f n ( x ) = k = 0 n a k T k ( x ) .
S n = m a x x [ 1 , 1 ] | f ( x ) f n ( x ) | .
A n ω n S n S n ,
S n = k = n + 1 | a k | ,
A n = 1 2 ( k = n + 1 a k 2 ) 1 / 2 ,
J 0 ( v x ) = k = 0 ϵ k ( 1 ) k J k 2 ( v / 2 ) T 2 k ( x ) , x [ 1 , 1 ] ,
f n ( v , x ) = k = 0 n / 2 ϵ k ( 1 ) k J k 2 ( v / 2 ) T 2 k ( x ) , x [ 1 , 1 ] ,
S n = 1 k = 0 n / 2 ϵ k J k 2 ( v / 2 ) ,
A n = [ 2 F 3 ( 1 / 2 , 1 / 2 ; v 2 1 , 1 , 1 ) k = 0 n / 2 ϵ k J k 4 ( v / 2 ) ] 1 / 2 ,
Ψ ( u , v , φ ) = 1 π 0 1 0 2 π exp [ i u 2 τ 2 + i v τ cos ( ϑ φ ) i Φ ( τ , ϑ ) u 2 ] τ d τ d ϑ .
L 0 0 ( u , v ) = 0 1 exp ( i u 2 τ 2 ) J 0 ( v τ ) τ d τ
L m ( u , v ) = 0 1 exp ( i u 2 τ 2 ) J ( v τ ) τ m + 1 d τ ,
Φ ( τ , ϑ ) = β R n m ( τ ) cos m ϑ ,
Ψ ( u , v , φ ) = k = 0 ( i β ) k k ! C k ( u , v ; n , m ) ,
J 2 p ( v τ ) = k = 0 ϵ k J p + k ( v / 2 ) J p k ( v / 2 ) T 2 k ( τ ) ,
J 2 p + 1 ( v τ ) = 2 k = 0 J p + k + 1 ( v / 2 ) J p k ( v / 2 ) T 2 k + 1 ( τ ) ,
A n 2 ( , v ) = v 2 Γ ( + 1 2 ) 2 π 2 F ~ 3 ( + 1 2 , + 1 2 ; v 2 + 1 , + 1 , 2 + 1 ) k = 0 n / 2 a k 2 ( , v ) ,
a k ( , v ) = { ϵ k J / 2 + k ( v / 2 ) J / 2 k ( v / 2 ) , e v e n , 2 J / 2 + 1 + k ( v / 2 ) J / 2 k ( v / 2 ) , o d d .
A n ( , v ) m a x x [ 1 , 1 ] | J ( v x ) f n ( v , x ) | S n ( , v ) .
L m ( u , v ) { k = 0 N / 2 a k ( , v ) T 2 k m ( u / 2 ) , e v e n , k = 0 N / 2 a k ( , v ) T 2 k + 1 m ( u / 2 ) , o d d ,
T k m ( α ) = 0 1 exp ( i α τ 2 ) T k ( τ ) τ m + 1 d τ ,
T k + 1 m ( α ) = 2 T k m + 1 ( α ) T k 1 m ( α ) , k 1 , m 1 ,
T 0 m ( α ) = 1 2 ( i α ) 1 + m / 2 [ Γ ( m 2 + 1 ) Γ ( m 2 + 1 , i α ) ] , T 1 m ( α ) = T 0 m + 1 ( α ) .
E = 0 1 exp ( i u 2 τ 2 ) [ J ( v τ ) f N ( v , τ ) ] τ m + 1 d τ ,
| E | 0 1 | J ( v τ ) f N ( v , τ ) | τ m + 1 d τ < S n ( , v ) m + 2 .
L 0 0 ( 0 , v ) = 0 1 J 0 ( v τ ) τ d τ = J 1 ( v ) v .
L 0 0 ( v , v ) = 1 2 i v [ exp ( i v 2 ) J 0 ( u ) exp ( i v 2 ) ] ,
N = v 2 + K ,
L + 1 m ( u , v ) = 0 1 exp ( i u 2 τ 2 ) J + 1 ( v τ ) τ m + 1 d τ ,
J + 1 ( x ) = 2 x J ( x ) J 1 ( x ) ,
L + 1 m ( u , v ) = 2 v L m 1 ( u , v ) L 1 m ( u , v ) , = 1 , 2 , , m = 0 , 1 , .
V n m ( u , v ) = 0 1 R n m ( τ ) exp ( i u 2 τ 2 ) J m ( v τ ) τ d τ ,
V n m ( u , v ) = exp ( i u 2 ) l = 1 ( i u / v ) l 1 l j = 0 p V lj J m + l + 2 j ( v ) v ,
V lj = ( 1 ) p ( m + l + 2 j ) [ ( m + j + l 1 l 1 ) 1 ( m + j + l 1 l 1 ) ( j + l 1 l 1 ) × ( l 1 p j ) ( q + l + j l ) 1 ] ,
R n m ( τ ) = k = 0 p ( 1 ) k ( n k k ) ( n 2 k p k ) τ n 2 k ,
V n m ( u , v ) = k = 0 p ( 1 ) k ( n k k ) ( n 2 k p k ) L m n 2 k ( u , v ) .
Φ s p h ( τ , ϑ ) = β R n 0 ( τ ) .
C k ( u , v ; n , 0 ) = 2 ( 1 ) k 0 1 exp ( i u 2 τ 2 ) [ R n 0 ( τ ) ] k J 0 ( v τ ) τ d τ ,
Φ c o m a ( τ , ϑ ) = β R 3 1 ( τ ) cos ϑ ,
Φ a s t ( τ , ϑ ) = β R 2 2 ( τ ) cos 2 ϑ = β τ 2 cos 2 ϑ .
A n 2 = 2 k = 1 J p + k 2 ( v / 2 ) J p k 2 ( v / 2 ) 2 k = 1 n / 2 J p + k 2 ( v / 2 ) J p k 2 ( v / 2 ) ,
A n 2 = 2 π 0 1 J 2 p 2 ( v x ) 1 x 2 d x k = 0 n / 2 ϵ k J p + k 2 ( v / 2 ) J p k 2 ( v / 2 ) .
A n 2 = 2 π 0 1 J 2 p + 1 2 ( v x ) 1 x 2 d x 2 k = 0 ( n 1 ) / 2 J p + 1 + k 2 ( v / 2 ) J p k 2 ( v / 2 ) .
2 π 0 1 J 2 ( v x ) 1 x 2 d x = v 2 Γ ( + 1 2 ) 2 π 2 F ~ 3 ( + 1 2 , + 1 2 ; v 2 + 1 , + 1 , 2 + 1 ) ,
C 0 ( u , v ; n , m ) = 2 0 1 exp ( i u 2 τ 2 ) J 0 ( v τ ) τ d τ , C 1 ( u , v ; n , m ) = 2 i m cos m φ 0 1 exp ( i u 2 τ 2 ) × R n m ( τ ) J m ( v τ ) τ d τ C 2 ( u , v ; n , m ) = 0 1 exp ( i u 2 τ 2 ) [ R n m ( τ ) ] 2 J 0 ( v τ ) τ d τ + i 2 m cos 2 m φ 0 1 exp ( i u 2 τ 2 ) [ R n m ( τ ) ] 2 × J 2 m ( v τ ) τ d τ , C 3 ( u , v ; n , m ) = 1 2 { 3 i m cos m φ 0 1 exp ( i u 2 τ 2 ) [ R n m ( τ ) ] 3 × J m ( v τ ) τ d τ + i 3 m cos 3 m φ 0 1 exp ( i u 2 τ 2 ) × [ R n m ( τ ) ] 3 J 3 m ( v τ ) τ d τ ( i u 2 τ 2 ) 0 1 } , C 4 ( u , v ; n , m ) = 1 4 { 3 0 1 exp ( i u 2 τ 2 ) [ R n m ( τ ) ] 4 J 0 ( v τ ) τ d τ + 4 i 2 m cos 2 m φ 0 1 exp ( i u 2 τ 2 ) [ R n m ( τ ) ] 4 × J 2 m ( v τ ) τ d τ + i 4 m cos 4 m φ 0 1 exp ( i u 2 τ 2 ) × [ R n m ( τ ) ] 4 J 4 m ( v τ ) τ d τ ( i u 2 τ 2 ) 1 } .
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.