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

Steerable photonic jet for super-resolution microscopy

Open Access Open Access

Abstract

A promising technique in optical super-resolution microscopy is the illumination of the sample by a highly localized beam, a photonic jet (also called photonic nanojet). We propose a method of computation of incident field amplitude and phase profiles that produce photonic jets at desired locations in the near field after interaction with a fixed micro-scale dielectric lens. We also describe a practical way of obtaining the incident field profiles using spatial light modulators. We expect our photonic jet design method to work for a wide range of lens shapes, and we demonstrate its application numerically using two-dimensional micro-lenses of circular and square cross-sections. We furthermore offer a theoretical analysis of the resolution of photonic jet design, predicting among other that a larger lens can produce a narrower photonic jet. Finally, we give both theoretical and numerical evidence that the waist width of the achieved designed jets is increasing linearly and slowly over a large interval of radial distances. With uniform plane wave illumination, the circular two-dimensional micro-lens produces a similar-sized jet at a fixed radial distance, while the square lens does not form a jet at all. We expect our steerable optical photonic jet probe to enable highly localized adaptive real-time measurements and drive advances in super-resolution optical microscopy and scatterometry, as well as fluorescence and Raman microscopy. Our relatively weak peak jet intensity allows application in biology and health sciences, which require high resolution imaging without damaging the sample bio-molecules.

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

1. Introduction

Several far-field and near-field optical microscopy techniques overcome the Abbe diffraction limit [1] and achieve super-resolution imaging of particles and features. In a recent review article [1], Huszka and Gijs give a good introduction to the most common techniques, one of which is the use of photonic jets (also called photonic nanojets or PNJs [2,3]; see Fig. 1(b) for definitions). Here, laser or white light illumination of a dielectric micro-lens (cylindrical, hemispherical, spherical) produces a highly focused beam, a PNJ, in the shadow region of the near field exterior to the lens.

 figure: Fig. 1.

Fig. 1. Photonic nanojet (PNJ) (a) design by amplitude- and phase-modulated illumination of a dielectric micro-lens with general shape and (b) features of a PNJ, its radial distance, waist width, and decay length, as in [5]. The full width at half maximum (FWHM) amplitude contour is with respect to the amplitude $|E^{\rm tot}|$ of the total electric field.

Download Full Size | PDF

PNJ formation was observed experimentally in 2006 [4], and it was since shown to depend on the micro-lens radius, shape, and refractive index relative to the surrounding medium, as well as on the illumination wavelength. [2,5] Several works exploit this to create PNJs of various shapes and sizes using lens shaping and plane wave or gaussian illumination [6,7]. The high field intensity and the small width of the PNJ focus, relative to the illuminating laser wavelength, allows highly local measurements and applications such as detection, manipulation and trapping of nanoparticles [810], nanopatterning and nanolithography [11,12] as well as optical imaging methods [1315] and other optical metrology techniques like fluorescence [16] and Raman [17] microscopy, in addition to scatterometric measurements [18,19]. However, the highly localized nature of PNJs also makes it difficult to find the area of interest. It is therefore desirable to realize a steerable PNJ that can quickly scan the local region without any need for lens adjustment or sample movement. Advances in this direction have been obtained using optically trapped levitating nanoparticles [12], shaped optical fiber tips [20] and metalic masks [21,22]. Precise and quick steering of PNJs could potentially allow rapid selective photo-switching of closely spaced sites along a biomolecule [16,17], or unlock the full potential of the emerging label-free microsphere-assisted optical super-resolution microscopy; where a sub-classical sample is placed near or inside a PNJ, enabling the localization, characterization (e.g., sizing) [3,4,18,19] and even magnified optical image creation [1315,23,24] of the sample.

We propose a method for fast and precise PNJ steering using a computed spatially inhomogeneous incident wave and a simple fixed micro-lens. The method enables rapid focus point movement without any adjustment or translation of the lens; it is not confined to the traditional spherical (3D) or circular (2D) micro-lenses; and it results in a consistently narrow PNJ and a large range of achievable PNJ positions. The results presented in this work have direct applications in particle manipulation and improvement in optical nanopattering and nanolithography. Furthermore, due to the robustness of the method, this technique combined with an adequate measuring strategy is of direct relevance for super-resolution microscopy.

The mathematical foundation of the method is described in Section 2, and a practical way to realize the computed incident fields is proposed in Section 3. We estimate theoretically the achievable PNJ waist width as a function of the radial distance in Section 4, and present numerical results for 2D circular and square cross section dielectric lenses in Section 5.

2. Tailoring the incident field

The procedure for tailoring the incident field that produces the desired PNJ is based on the solution of a Lippmann-Schwinger equation, such as [25, Eq. (8.13), p. 309], which we here derive explicitly for completeness (Eq. (8) below). Figure 1(a) shows the setup of the problem.

We consider the time-harmonic, two-dimensional transverse electric (2D TE) case with the electric field pointing out of the plane, and with the time-dependence factor $\exp (j\omega t)$ implied and suppressed; here $j$ is the imaginary unit. The operating free-space wavelength and wavenumber are $\lambda _0$ and $k_0=2\pi /\lambda _0$, respectively. A cylindrical dielectric lens with a constant refractive index $n_L>1$ and a cross-section profile $L$ is immersed in vacuum and illuminated with an incident field $E^{\rm inc}(x,y)=A(x,y)\exp (jP(x,y))$. Here a Cartesian coordinate system $(x,y,z)$ is introduced with the $z$-axis aligned with the lens axis, as in Fig. 1. Note that the problem geometry, the material parameters, and all the fields are invariant along the $z$-axis. The real-valued functions $A(x,y)$ and $P(x,y)$ give the amplitude and the phase profile of the incident field, respectively. Define the ’lens contrast’ $\alpha =k_0^{2}(n_L^{2}-1)$, the characteristic function

$$\chi_L(\mathbf{x})=\begin{cases}1,\quad & {\mathbf{x}=(x,y)}\in L,\\0\quad & {\rm otherwise},\end{cases}$$
and the piecewise constant wavenumber
$$k(\mathbf{x})=k_0[1+\chi_L(\mathbf{x})(n_L-1)],\quad\mathbf{x}\in\mathbf{R}^{2}.$$
We assume the ’desired total field’ ${E^{\rm tot}}$ solves the Helmholtz problem
$$\left.\begin{array}{rcll}(\Delta+k(\mathbf{x})^{2}){E^{\rm tot}}(\mathbf{x}) & = & { f(\mathbf{x})}, & \mathbf{x}\in{\mathbf{R}^{2}},\\ {E^{\rm tot}}(\mathbf{x}) & = & \xi(\mathbf{x}), & \mathbf{x}\in C{.}\end{array}\right\}$$
The curve $C$ and the function $\xi$ together define the desired near-field pattern which we wish to achieve by a computed structured illumination of the lens. For example, $C$ can indicate where we want a high total field intensity, and $\xi$ can be a large real constant. As usual, decompose the total field ${E^\textrm {tot}}$ outside $L$ into the sum ${E^\textrm {tot}}={E^\textrm {inc}}+{E^\textrm {sca}}$ of an incident and a scattered field, where the incident field satisfies
$$(\Delta+k_0^{2}){E^{\rm inc}}(\mathbf{x})=f(\mathbf{x}),\quad\mathbf{x}\in\mathbf{R}^{2}.$$
Here $f$ is the source of the incident field radiating in free space, possibly located at infinity. We readily check that
$$\begin{aligned} (\Delta+k_0^{2}){E^{\rm tot}}(\mathbf{x})&=f(\mathbf{x})+(k_0^{2}-k(\mathbf{x})^{2}){E^{\rm tot}}(\mathbf{x})\\ &=(\Delta+k_0^{2}){E^{\rm inc}}-\alpha\chi_L(\mathbf{x}){E^{\rm tot}}(\mathbf{x}),\,\,\mathbf{x}\in\mathbf{R}^{2}, \end{aligned}$$
which leads to
$$(\Delta+k_0^{2}){E^{\rm sca}}(\mathbf{x})={-}\alpha\chi_L(\mathbf{x}){E^{\rm tot}}(\mathbf{x}),\quad\mathbf{x}\in\mathbf{R}^{2}.$$
Since the right-hand member of 6 is compactly supported, and since our scattered field must satisfy the Sommerfeld radiation condition in the plane, we thus have
$${E^{\rm sca}}(\mathbf{x})={-}\alpha\Phi_0\ast(\chi_LE^{\rm tot})(\mathbf{x})={-}\alpha\int_{\mathbf{y}\in L}\Phi_0(\mathbf{x}-\mathbf{y}){E^{\rm tot}}(\mathbf{y})d\mathbf{y},\quad\mathbf{x}\in {\mathbf{R}^{2}},$$
and finally
$${E^{\rm inc}}(\mathbf{x})={E^{\rm tot}}(\mathbf{x})-{E^{\rm sca}}(\mathbf{x})={E^{\rm tot}}(\mathbf{x})+\alpha\int_{\mathbf{y}\in L}\Phi_0(\mathbf{x}-\mathbf{y}){E^{\rm tot}}(\mathbf{y})d\mathbf{y},\quad\mathbf{x}\in {\mathbf{R}^{2}}.$$

Here $\Phi _0(\mathbf {x})=(j/4)H_0^{(2)}(k_0|\mathbf {x}|)$ is the outgoing fundamental solution of the Helmholtz operator in the plane, and $H_0^{(2)}$ is the Hankel function of order zero and of the second kind. This suggests the following procedure for producing a PNJ at a desired location in the near field:

  • 1. Pick a curve $C$ and a function $\xi$ on $C$ that represent the desired PNJ well via the second condition in 3. Also, pick a curve $\Gamma$ at which the tailored incident field is to be computed.
  • 2. Solve the system in 3 (numerically) for the desired ${E^\textrm {tot}}$ in a neighborhood of the lens that includes the curves $C$ and $\Gamma$. Prepare a program that approximates the numerical values of ${E^\textrm {tot}}$ in $L$ and at $\Gamma$, for example by interpolation.
  • 3. Compute the incident field ${E^\textrm {inc}}$ at $\Gamma$ using 8 and the program from the previous step.
  • 4. Illuminate the micro-lens using a source $f$ that radiates a field approximating ${E^\textrm {inc}}$ at $\Gamma$.

If the realized lens illumination produces an unsatisfactory approximation to the desired PNJ then one may repeat step 4 above using a better approximation to ${E^\textrm {inc}}$ at $\Gamma$, or one may start over and modify the curve $C$ and function $\xi$. For the numerical results presented in this paper, we always choose the curve $C$ to be a short straight linear piece located at the desired PNJ position and pointing radially away from the origin. Our numerical experimentation has shown that more complicated choices of the geometry of $C$, such as multi-component curves, closed curves, or non-radial curves, can result in involved interference patterns in the near field. While this should certainly be studied systematically and exploited for advanced near-field design, we feel it falls outside of the scope of this paper.

The curve $\Gamma$ at which the incident field profile is tailored need not be a linear piece. However, for simplicity, in this paper we always choose $\Gamma$ as illustrated in Fig. 1(a), that is, a linear piece parallel with the $x$-axis and placed just above the micro-lens. We did not investigate the practical capabilities and performance of our PNJ design method for more complicated choices of $\Gamma$.

3. Realizing the incident field profiles using spatial light modulators

The spatial profile of an optical beam as computed in Section 2 could be achieved in practice using diffractive optical elements such as phase filters or computer generated holograms (CGHs) implemented with digital micromirror devices (DMDs) or liquid crystal spatial light modulators (SLMs). These mechanisms allow to generate complex fields by effectively varying the diffraction efficiency with phase modulations [26,27] or by transferring phase modulations onto amplitude modulations using polarizing optics [28]. We here propose a mechanism, illustrated in Fig. 2, to generate the complex field required to probe the micro-lens, which leads to the generation of the nanojet based on the latter technique. We consider a linearly polarized light beam of spatial profile $u_0({ \bf x})$ propagating along the $z$ direction.

 figure: Fig. 2.

Fig. 2. Probe beam preparation through amplitude and phase modulation of a beam $u_0(\textbf {x})$. Using two spatial light modulators (SLMs) and polarization optics we generate an arbitrary complex field $u(\textbf {x}) = \Psi (\textbf {x}) \exp {i\Phi (\textbf {x})}$ by imprinting the phases $\varphi _1(\textbf {x})$ and $\varphi _2(\textbf {x})$, respectively with the SLMs. The dimensions of the beam are adjusted with a demagnifying telescopic system.

Download Full Size | PDF

Without losing generality, let us consider a Gaussian beam $u_0(\textbf {x}) = A_0 \exp (-|\textbf {x}|^{2}/w_0^{2})$ with horizontal polarization along the horizontal direction. With the use of a half-wave plate (HWP) the polarization of the beam is rotated 45${\circ }$ so that it has equal components of magnitude $A_0/\sqrt {2}$ in the horizontal ($H$) and vertical ($V$) direction. The beam is transmitted (or reflected) by a first spatial light modulator (SLM$_1$) imprinting a phase $\varphi _1(\textbf {x})$ to the horizontal component of the beam in the following way:

$$u({\bf x}) = \frac{1}{\sqrt{2}}u_0({\bf x}) ( e^{i \varphi_1 ({\bf x})} {H} + {V}).$$
With the use of a linear polarizer (LP) at 45$^{\circ }$ followed by a HWP the amplitude of the beam becomes $A_0(\exp (i\phi _1(\textbf {x}))+1)/2$ with the polarization state set to horizontal, then a second spatial light modulator (SLM$_2$) imprints a phase $\varphi _2(\textbf {x})$, and the resulting beam is
$$ u({\bf x}) = u_0({\bf x}) \cos{\Big(\frac{\varphi_1({\bf x})}{2}\Big)}\exp{i\Big( \frac{\varphi_1({\bf x}) +2 \varphi_2({\bf x})}{2}\Big)}.$$
The amplitude and phase of the beam after the SLM$_2$ can be engineered by setting the required phases $\varphi _1$ and $\varphi _2$. Note that an imaging system with unit magnification is required to prevent spurious phase delays due to propagation between the first and the second SLMs, for instance a $4f$ imaging system consisting of two lenses of equal focal lengths ($f$) located at a distance $2f$ between them in a way that the distance from the first (second) lens to the first (second) SLM is $f$ while the distance to the other SLM is $3f$. To generate an arbitrary beam with electric field $u(\textbf {x}) =\psi (\textbf {x}) \exp {i\phi (\textbf {x})}$ one would need to induce $\varphi _1(\textbf {x}) = 2 \arccos (\psi (\textbf {x})/u_0(\textbf {x}))$ to SLM$_1$ and $\varphi _2(\textbf {x})= (2\phi (\textbf {x}) - \varphi _1(\textbf {x}))/2$ to SLM$_2$.

We next address the achievable resolution of the SLM-produced incident field profiles. In general, SLMs can generate pixelated 8-bit and 10-bit phase-maps (such as $\varphi _1(\textbf {x})$ and $\varphi _2(\textbf {x})$ above) with an arbitrary phase distribution of values discretized in steps of 2$\pi$ rad$/255\approx 24.6$ mrad ($\approx$1.4$^{\circ }$) and $2\pi$ rad$/1023\approx 6.1$ mrad ($\approx 0.35^{\circ }$) respectively. The actual performance depends on the model of the SLM and on the operating wavelength.

For the "SLM-generated" part of the numerical results in Section 5, we select the phase map resolution of 24.6 mrad, and Gaussian illumination sufficiently large to cover the area of interest of the SLM, but just enough not to generate diffraction effects due to the borders of the SLM. Since we choose to modulate the theoretical incident field profiles in Section 5 in 902 steps of 26.6 nm over a domain (the curve $\Gamma$ in Fig. 1(a)) of length 24 $\mu$m, we here start with an illumination beam of diameter (FWHM) equal to 902 pixels of the SLM display. This way, each pixel of the SLM corresponds to a pixel in the domain $\Gamma$. For an SLM with pixel size of 8 $\mu$m (such as Thorlabs EXULUS HD2), the required beam diameter is 7.2 mm and the beam waist 5.2 mm. After passing through the two SLMs, the beam will have the desired amplitude and phase profiles, however at a larger scale.

To reduce the size of the beam to 24 $\mu$m without modifying its spatial profile, we require an imager with magnification $M=300$. This can be realized with consecutive telescopic systems such that the product of their magnifications $M_i$ equals $M$, for instance two telescopes: A) & B) $M_{a,b}=10$ $\rightarrow$ $f_1$ = 1000 mm, $f_2$ = 100 mm, and a third one C) $M_{c}=3$ $\rightarrow$ $f_5$ = 300 mm, $f_6 = f_2$. With these three consecutive systems (A, B and C), the desired field is obtained at the final focal plane of the last system, which corresponds to the reference space domain $\Gamma$ above the micro-lens. The imager requires ultra-precise alignment and position placement of the demagnifying lenses and of the micro-lens, which can be found by rigorous computation. We can compensate for small diffraction effects caused by misalignment, cross-talk and aberrations by introducing some feedback to the SLM phases $\varphi _1(\textbf {x})$ and $\varphi _2(\textbf {x})$.

4. Waist-width analysis of designed photonic jets

In this section we construct a simplified theoretical model of PNJ formation, and then estimate the qualitative behavior of PNJ waist width as a function of the radial distance $\varrho \ge 0$ of the PNJ. We consider the 2D TE case (see Section 2 and Fig. 1(b)) with a circular micro-lens of cross-section radius $R_L$ (see Fig. 3.1). This analysis is based on the stable bandwidth estimates for the singular spectrum of the forward operator $F$ for the Helmholtz equation in the plane, established in [29]. The reader interested only in the performance of the PNJ design algorithm proposed in sections 2 and 3 can skip the current section entirely.

 figure: Fig. 3.

Fig. 3. A schematic of the 5-step theoretical procedure for estimating the full width at half maximum (FWHM) of a PNJ with a desired profile.

Download Full Size | PDF

We shall write $F$ for the linear operator that maps a source $s$ radiating in free space to the radiated field $F(s)$ measured at a circle of radius $R_L+\varrho$ and centered at the origin. More precisely, the forward operator $F:L^{2}(B_{R_L})\mapsto L^{2}(\partial B_{R_L+\varrho }),$

$$(Fs)(\mathbf{x})=\frac{j}{4}\int_{y\in B_{R_L}}H_0^{(2)}(k_0|\mathbf{x}-\mathbf{y}|)s(\mathbf{y})d\mathbf{y},\quad \mathbf{x}\in\partial B_{R_L+\varrho},$$
maps any square-integrable ($L^{2}$), time-harmonic, 2D TE electromagnetic source $s$ supported in the disk $B_{R_L}$ of radius $R_L$ and radiating in free space, to the trace of its radiated electric field at the boundary of the disk $B_{R_L+\varrho }$ with radius $R_L+\varrho$. Indeed, in view of Eq.  (7), we can express the field scattered by the micro-lens in terms of the field radiated in free space by a source $s$ supported in the closure $\overline {B}_{R_L}$ of the disk with radius $R_L$. The forward problem associated with $F$ is to find the radiated field $F(s)$ at $\partial B_{R_L+\varrho }$ given a source $s$ in $L^{2}(B_{R_L})$. The ’inverse source problem’ (ISP) associated with $F$ is to determine a free-space source $s$ in $L^{2}(B_{R_L})$ which radiates the desired (or measured) near-field pattern at $\partial B_{R_L+\varrho }$.

Given a desired radial distance $\varrho$, we

  • 1. construct a desired PNJ profile ${E^\textrm {tot}}_\textrm {PNJ}$ at $\partial B_{R_L+\varrho }$,
  • 2. interpret ${E^\textrm {tot}}_\textrm {PNJ}$ as a field boundary datum at $\partial B_{R_L+\varrho }$ for a single-frequency inverse source problem for the Helmholtz equation in free space, with the source supported in $\overline {B}_{R_L}$,
  • 3. solve the inverse source problem and compute the optimal (in a precise sense given below) square-integrable source $s$ supported in $\overline {B}_{R_L}$ that minimizes the $L^{2}$ norm of $F(s)-{E^\textrm {tot}}_\textrm {PNJ}$ at $\partial B_{R_L+\varrho }$,
  • 4. compute the field $E^\textrm {sca}_s=F(s)$ radiated by $s$ at $\partial B_{R_L+\varrho }$ and interpret $E^\textrm {sca}_s$ as an approximation to the physically realizable field closest to the desired ${E^\textrm {tot}}_\textrm {PNJ}$ at $\partial B_{R_L+\varrho }$, and
  • 5. find the FWHM of the field $E^\textrm {sca}_s$ at $\partial B_{R_L+\varrho }$ and interpret it as the actually achievable PNJ waist width, given the radial distance $\varrho$ and the desired profile $E^\textrm {tot}_\textrm {PNJ}$.
We stress that our analysis is based on a simplified model. Indeed, in step 1 we shall use a Gaussian amplitude and constant zero phase, rather than a true Maxwellian field, to model the trace of the PNJ field at $\partial B_{R_L+\varrho }$. Also, in step 4, we assume the incident field is insignificant at $\partial B_{R_L+\varrho }$ compared to the scattered field radiated by the source $s$. We do this because we have no a priori knowledge of how an exact total field, and the pertaining incident field, should look given an arbitrary position of the PNJ. While more accurate pulse-like functions could have been used instead of Gaussian profiles, the latter are simple, exhibit a highly localized peak, and vanish quickly with increasing argument, thus modeling the qualitative features of a PNJ. We estimate the smallest achievable FWHM using the above procedure (steps 1–5), by letting the desired FWHM in $E^\textrm {tot}_\textrm {PNJ}$ approach zero and examining the corresponding asymptotic behavior of the realized FWHM in $E^\textrm {sca}_s$. A schematic of the procedure is shown in Fig. 3. Let $R\ge R_L$ and assume a PNJ is centered at the intersection of the boundary $\partial B_R$ and the negative $y$-axis in Fig. 1(a). The radial distance of this PNJ is $\varrho =R-R_L$. In step 1 of the above procedure, we model the desired profile of the PNJ at $\partial B_{R_L+\varrho }$ as the Gaussian
$${E^{\rm tot}}_{\rm PNJ}(\theta)=\exp(-(\theta+\pi/2)^{2}/d^{2}),$$
where $\theta$ is the angle coordinate in the plane. Given a desired FWHM of the PNJ profile in fractions of $\lambda _0$, it can be shown that the corresponding $d$ in 10 is found by
$$d(\text{FWHM},R) =\frac{{\rm FWHM}\cdot\lambda_0}{2R\sqrt{\ln 2}}.$$
Fig. 4 shows four Gaussians, each modelling a desired PNJ profile with $\textrm {FWHM}=0.5\lambda _0$, for $R_L$ from 2 $\mu$m to 8 $\mu$m. The radial distance of the PNJ is $2\lambda _0$ for all plots in this section. In step 3 of the above procedure, we solve the inverse source problem using a truncated singular value decomposition (TSVD) of the forward operator $F$. The singular value decomposition essentially expresses an operator in terms of eigenvectors and eigenvalues, so that a (pseudo-) inverse operator is expressed in terms of the same eigenvectors but the inverse of the eigenvalues. With $(\sigma _m,\psi _m,\phi _m)$ the singular system of $F$ described in [29, Section 2.1], we can write the action of $F$ on a $L^{2}(B_{R_L})$ (square integrable) source $s$ supported in $\overline {B}_{R_L}$ as
$$F(s)\approx\sum_{\substack{m\in\pmb{Z}\\|m|\le\mathcal{B}}}\sigma_m\langle s,\psi_m\rangle_{L^{2}(B_{R_L})}\phi_m,$$
where $\mathcal {B}$ is the stable spectral bandwidth of $F$, estimated in 2D in [29] in terms of $\mathcal {B}_-\le \mathcal {B}\le \mathcal {B}_+$ with
$$\mathcal{B}_-{=}\underset{m\ge 0}{\rm argmin}\{j_{m,1}\ge k_0R_L\},$$
$$\mathcal{B}_+{=}\underset{m\ge 0}{\rm argmin}\{y_{m,1}\ge k_0R_L\}.$$
Here $j_{m,1}$ and $y_{m,1}$ are the first positive zeros of the Bessel function $J_m(x)$ and the Neumann function (Bessel function of the second kind) $Y_m(x)$, respectively. Thus, we solve $F(s)=E^\textrm {tot}_\textrm {PNJ}$ in terms of
$$s^{{\pm}}_{\rm TSVD}=\sum_{\substack{m\in\pmb{Z}\\|m|\le\mathcal{B}_{{\pm}}}}\sigma_m^{{-}1}\langle E^{\rm tot}_{\rm PNJ},\phi_m\rangle_{L^{2}(\partial B_{R})}\psi_m,$$
and in step 4 of the above procedure we find $E^\textrm {sca}_s$ via
$$\begin{aligned} E^{{\rm sca}\pm}_s&=F(s^{{\pm}}_{\rm TSVD})\approx\sum_{\substack{m\in\pmb{Z}\\|m|\le\mathcal{B}_{{\pm}}}}\sigma_m\langle s^{{\pm}}_{\rm TSVD},\psi_m\rangle_{L^{2}(B_{R_L})}\phi_m\\ &=\sum_{\substack{m\in\pmb{Z}\\|m|\le\mathcal{B}_{{\pm}}}}\langle E^{\rm tot}_{\rm PNJ},\phi_m\rangle_{L^{2}(\partial B_{R})}\phi_m. \end{aligned}$$
We here use the fact that the right singular vectors $\psi _m$ of $F$ constitute an orthonormal system. The computed $s^{\pm }_\textrm {TSVD}$ is optimal in the sense that it is the fullest projection of the true solution $s$ onto the domain of $F$ (without the nullspace of $F$) that is based on the stable (noise-robust) component in the data $E^\textrm {tot}_\textrm {PNJ}$; see [29] for more details. Since each left singular vector $\phi _m(\theta )$ is proportional to $\exp (jm\theta )$ [29, Lemma 1], the computation in 16 amounts to a band-limited projection of the desired profile $E^\textrm {tot}_\textrm {PNJ}$ onto the Fourier basis at the unit circle. It is the sharp spectral cut-off exhibited by the map $F$ at the bandwidth index $|m|=\mathcal {B}$, in conjunction with the lens radius $R_L$ and the radial distance $R-R_L$, that ultimately determine the smallest achievable PNJ waist width. From 13 and 14 it is evident that the bandwidth $\mathcal {B}$ increases with increasing lens radius $R_L$. This is because $j_{m,1}$ and $y_{m,1}$ are monotonically increasing with $|m|$. Thus a larger lens will be able to produce a narrower PNJ, since the corresponding $F$ will have a larger bandwidth, and higher spatial frequency components will be present in the projection 16. The analysis in this section is directly extendible to 3D with the help of our singular spectrum results for $F$ from [30].

 figure: Fig. 4.

Fig. 4. Desired PNJ profiles $|E^\textrm {tot}_\textrm {PNJ}(\theta )|$ at $\partial B_{R_L+\varrho }$ ($\times 10^{5}$ V/m) for different lens radii $R_L$.

Download Full Size | PDF

We now fix the operating wavelength at $\lambda _0=532$ nm (common green laser). Figure 5 shows the physically viable field profiles produced using the 5-step procedure outlined above, taking $\mathcal {B}=\mathcal {B}_-$, and starting with the desired profiles from Fig. 4. We see that the achieved FWHM of the PNJ profile gets slightly closer to the desired FWHM of $0.5\lambda _0$ as we increase the lens radius $R_L$ and thereby the bandwidth $\mathcal {B}$. Figure 6 shows more desired and achieved FWHM for the four lens radii $R_L$. We again use $\mathcal {B}=\mathcal {B}_-$, in order to arrive at an upper bound for the achievable FWHM. The curves representing larger lens radii, and therefore larger bandwidths, are closer to the desired FWHM. The plot further suggests that the bandwidth of the forward operator $F$ is indeed the limiting factor for how narrow a PNJ can get. Figure 7 shows the lower and upper bounds on the achievable FWHM (based on $\mathcal {B}=\mathcal {B}_+$ and $\mathcal {B}=\mathcal {B}_-$ respectively) for the four examined lens radii. These bounds are the asymptotic values of the achieved FWHM as the desired FWHM approaches zero. We finally examine the achievable FWHM as function of the desired radial distance, again using the 5-step procedure outlined above. Figure 8 shows our numerical results; here we try to produce a PNJ with a FWHM waist-width of $0.5\lambda _0$ (as in Fig. 5). The achieved FWHM waist-width of the PNJ seems to be approximately linearly increasing with increasing radial distance. The rate of increase is lower for larger lenses, which implies that, theoretically, we can achieve higher angular resolution at large distances from the lens by increasing the lens radius.

 figure: Fig. 5.

Fig. 5. Resulting physically viable PNJ profiles $|E^\textrm {sca}_s(\theta )|$ at $\partial B_{R_L+\varrho }$ ($\times 10^{5}$ V/m) for different lens radii $R_L$.

Download Full Size | PDF

 figure: Fig. 6.

Fig. 6. Waist-width analysis of PNJ profile for four different lens radii with a radial distance of $2\lambda _0$. The black dots are the results from Fig. 5.

Download Full Size | PDF

 figure: Fig. 7.

Fig. 7. Angular PNJ resolution analysis using the projection from 16 with both bandwidth estimates $\mathcal {B}_{\pm }$ from [29].

Download Full Size | PDF

 figure: Fig. 8.

Fig. 8. Waist-width analysis with varying radial distance of a single PNJ with desired FWHM of $0.5\lambda _0$, for four different lens radii $R_L$. The black dots are the results from Figs. 4 and 5.

Download Full Size | PDF

5. Achieved PNJ design

We demonstrate our method of PNJ design numerically using a 2D circular cross-section lens of radius 4 $\mu$m and a 2D square cross-section lens of side length 8 $\mu$m. Both lenses have refractive index $n_L=1.4607$ (SiO$_2$ at 532 nm), and the operating wavelength is $\lambda _0=532$ nm (common green laser). The illumination is time-harmonic and 2D TE, that is, the electric field vector points out of the lens cross-section plane, and the coordinate system origin is at the center of the lens cross-section.

The curve $C$, defining the desired location of the PNJ in 3, is a straight, 10 nm-linear piece pointing radially away from the origin, and we choose $\xi (\textbf {x})=1$ for $\textbf {x}\in C$. The curve $\Gamma$ where the incident field is tailored is the straight linear piece defined by $x\in [-12\ \mu$m$,12\ \mu$m$]$, $y=4\ \mu$m. Thus $\Gamma$ is always ’just above’ the micro-lens, and parallel with the $x$-axis. We solve 3 for ${E^\textrm {tot}}$ using the Finite Element method implemented by COMSOL, although other near-field numerical methods are also applicable [31]. Furthermore, we solve 8 for ${E^\textrm {inc}}$ at $\Gamma$ using numerical quadrature and interpolation in a Matlab script. Finally, we input the computed ${E^\textrm {inc}}$ at $\Gamma$ to a second COMSOL model to find the resulting total field.

The left-most and right-most columns in Fig. 9 demonstrate the realized PNJ design using the theoretical (Section 2) and the simulated SLM-produced (Section 3) incident fields, respectively. For practical purposes, there seems little difference between the theoretical and the SLM-produced near fields. The SLM angles $\varphi _1(\textbf {x})$ and $\varphi _2(\textbf {x})$ are here discretized with the resolution of 24.6 mrad. We conjecture that the lack of symmetry in some of the profiles of the tailored incident field in the first four cases in Fig. 9 is a numerical artifact due to the finite field sampling in the steps 2 and 3 described in Section 2. Evidently, the solution compensates for the asymmetry, and correct symmetric PNJs are ultimately achieved as desired in these cases.

 figure: Fig. 9.

Fig. 9. Left column: PNJ scanning achieved at the single optical wavelength $\lambda _0=532$ nm (common green laser). A 2D SiO$_2$ micro-lens with a circular cross-section of radius 4 $\mu$m, or a square cross-section of side length $8\ \mu$m, is illuminated along the negative $y$-axis by a computed structured incident field (see Section 2). The plots show the amplitude (in V/m) of the resulting total near field, normalized to maximum intensity of 1. Middle column: The computed amplitude and phase profiles of the incident field that produce the desired total near field. The desired PNJ locations in $\mu$m are, from top to bottom: $(x,y)=(0,-4.532)$, $(x,y)=(0,-9.32)$, $(x,y)=(0,-4.532)$, $(x,y)=(0,-9.32)$ (radial distance $1\lambda _0$ or $10\lambda _0$), and $(x,y)=(3,-9)$. Right column: PNJ scanning achieved with the incident field profiles simulated as being produced by an SLM with phase steps of 24.6 mrad in the angles $\varphi _1(\textbf {x})$ and $\varphi _2(\textbf {x})$. See Section 3 for details.

Download Full Size | PDF

The numerical results indicate that our method can design the PNJ position well and over a large domain in the near field. Specifically, Fig. 10 shows the achieved radial distances for the two lens geometries over large ranges of desired radial distances. The relative error in the achieved radial distances in that figure is on average $9.9\%$ in the circular lens case (for both the theoretical and the SLM-produced incident fields), and on average $1.5\%$ (theoretical) and $2.7\%$ (SLM) in the square lens case.

 figure: Fig. 10.

Fig. 10. Desired vs. achieved radial distance. a) circular lens, b) square lens.

Download Full Size | PDF

In Fig. 11 we observe that the waist width and the decay length of the realized PNJs are increasing with increasing radial distance, and for both lens geometries. The linear nature of the waist width increase was predicted in Section 4, but overestimated, especially for large radial distances. We stress that the analysis in Section 4 uses a simple Gaussian profile to model the field traces. In conclusion, both the radial and the lateral dimension of our near-field PNJ design increase with increasing distance from the lens, with the lateral dimension growing much slower and to the point of being practically constant over large intervals of radial distances.

 figure: Fig. 11.

Fig. 11. Achieved waist width and decay length vs. achieved radial distance. a) circular lens, b) square lens.

Download Full Size | PDF

Figure 12 shows, for the theoretical incident field profiles, the maximal PNJ electric field intensity relative to the maximal tailored incident electric field intensity, as function of the achieved radial distance in the PNJ. For the circular lens, this quantity fluctuates about 30$\%$ to 35$\%$, while for the square lens it decays significantly with increasing radial distance, and seems to stabilize at about 40$\%$. The fact that we achieve PNJs with maximum amplitudes comparable to those of the tailored incident fields suggests that the phenomenon underlying our PNJ formation is not primarily focusing, but rather an interference effect.

 figure: Fig. 12.

Fig. 12. Achieved PNJ field amplitude maximum vs. radial distance. a) circular lens, b) square lens.

Download Full Size | PDF

Table 1 shows mean values of achieved PNJ parameters, for the realizations of the theoretical incident field profiles treated in Figs. 1012. For comparison, the simulated SLM-produced incident field profiles yield a mean waist width of $0.88\lambda _0$ (circular lens) and $0.70\lambda _0$ (square lens), as well as a mean decay length of $2.24\lambda _0$ (circular lens) and $1.60\lambda _0$ (square lens). For the theoretical incident field profiles, the predicted FWHM of the achieved PNJs in the example in Fig. 8 is between approximately $0.75\lambda _0$ and $1.00\lambda _0$ for $R_L=4\ \mu$m and for radial distances between approximately zero and $3\lambda _0$. This corresponds well with the numerically obtained waist widths in Fig. 11 for the circular lens and for achieved radial distances from approximately zero to approximately $3\lambda _0$. (We stress that our PNJ design procedure from Section 2 does not include specifying a desired PNJ waist width.) We next note that the PNJs produced using a square lens seem not only more precisely steerable, but also smaller than those produced using a circular lens. Their peak intensity relative to the incident field maximum intensity is also on average greater than in the circular-lens case, albeit this intensity ratio seems to drop quickly with the radial distance of the PNJ.

Tables Icon

Table 1. Mean values of the achieved PNJ parameters in Figs. 1012, for theoretical incident field profiles (Section 2).

Finally, for reference, we investigate the near-field localization occurring when a simple $-\widehat {y}$-directed uniform plane wave at wavelength $\lambda _0=532$ nm illuminates the micro-lenses considered in this section. Figure 13 shows the resulting near fields. The circular cross-section lens produces a fixed PNJ with radial distance $2.22\lambda _0$, waist width $0.84\lambda _0$, decay length $3.21\lambda _0$, and maximum PNJ field intensity (relative to maximum incident electric field intensity) $18.92$. The waist width is $0.04\lambda _0\approx 21$ nm smaller than its mean counterpart in our steerable PNJs in the circular-lens case (but larger than that for our steerable PNJs in the square-lens case), while the decay length is almost one full wavelength greater than the mean decay length in the corresponding steerable PNJs. The main difference is the fact that the steerable PNJs exhibit 60 times smaller peak intensity relative to the maximum incident field intensity, and hence probably need a more powerful illumination in order to achieve comparable maximum field strengths. Alternatively, the relatively weak PNJs are well-suited as steerable optical probes in the investigation of bio-molecules and other sensitive samples.

 figure: Fig. 13.

Fig. 13. Near fields, normalized to maximum intensity of 1, obtained using uniform plane wave illumination. a) circular lens, b) square lens.

Download Full Size | PDF

The near-field structure in Fig. 13 of the square lens illuminated by a uniform plane wave seems too far off an admissible PNJ structure for analysis.

6. Conclusion

We give a theoretical foundation for a method of spatial steering of photonic jets, for a wide range of micro-scale dielectric lens shapes used together with computed illumination profiles. We furthermore provide numerical validation of the method in the case of circular and square cross-section 2D micro-lenses. Our method produces photonic jets at the desired locations in the near field of the dielectric lens. The novelty is that we take an inverse problem approach: starting with a desired total near field ${E^\textrm {tot}}$ for a fixed micro-lens geometry, we compute an incident field ${E^\textrm {inc}}$ that produces an approximation of ${E^\textrm {tot}}$, and that specifically produces a PNJ at the desired position. We also describe a practical way of obtaining the incident field profiles using spatial light modulators (SLMs). Our work paves the way towards fast SLM-based imaging that requires no mechanical movement of the sample or of the micro-lens. We see many applications, one of which is improving the resolution and speed of confocal microscopy by replacing the scanning mirrors with SLMs and doing $x,y$, and $z$ scan with our steerable PNJ. Other applications could be nanoparticle manipulation, direct lithography writing, scatterometry, fluorescence microscopy and Raman microscopy, which all benefit from the fact that we have a highly localized steerable optical probe with a subwavelength waist width. Since we see nearly zero defocusing of the PNJ over radial distances of at least up to 15$\lambda _0$ (almost 8 $\mu$m), the technique is ideal for the abovementioned and other applications. We thus expect our steerable optical photonic jet probe to drive advances within several fields.

The observed small relative peak field strength makes our steerable PNJ probe unfavorable in certain applications. However, in other cases a steerable PNJ may be of more interest than one that is narrower but spatially fixed. We plan to explore in future work how the relative field strength can be increased. Our analysis in Section 4 leads us to expect that using larger micro-lenses will decrease the waist width of the steerable PNJ. Further, the relative peak field strength may increase with increasing size of the micro-lens.

We have not yet explored the general validity of the results in Figs. 6 and 7, that is, for other lens geometries than the circular case. Our analytical method of Section 4 is limited to circular-shaped lenses, and can therefore only make (qualitative) predictions for that case. We leave a more general analytic and numerical treatment to future work.

Finally, we did not investigate the sensitivity of our method to local minima in the optimization problem of finding the incident field profiles. It would certainly be desirable to be able to obtain a similar level of PNJ steerability using simpler incident fields. However, it is currently unclear whether such simplification is even possible.

Funding

H2020 Marie Skłodowska-Curie Actions (754558); Innovationsfonden (E13241-NanoMeas); European Metrology Programme for Innovation and Research (20FUN02/f10 POLight); Villum Fonden (25893).

Disclosures

The authors declare 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 authors upon reasonable request.

References

1. G. Huszka and M. A. M. Gijs, “Super-resolution optical imaging: A comparison,” Micro Nano Eng. 2, 7–28 (2019). [CrossRef]  

2. A. Darafsheh, “Photonic nanojets and their applications,” JPhys Photonics 3(2), 022001 (2021). [CrossRef]  

3. S. Lecler, S. Perrin, A. Leong-Hoi, and P. Montgomery, “Photonic Jet Lens,” Sci. Rep. 9(1), 4725 (2019). [CrossRef]  

4. A. Heifetz, K. Huang, A. V. Sahakian, X. Li, A. Taflove, and B. Vadim, “Experimental confirmation of backscattering enhancement induced by a photonic jet,” Appl. Phys. Lett. 89(22), 221118 (2006). [CrossRef]  

5. A. V. Itagi and W. A. Challener, “Optics of photonic nanojets,” J. Opt. Soc. Am. A 22(12), 2847 (2005). [CrossRef]  

6. J. Zhu and L. L. Goddard, “Spatial control of photonic nanojets,” Opt. Express 24(26), 30444 (2016). [CrossRef]  

7. A. Paganini, S. Sargheini, R. Hiptmair, and C. Hafner, “Shape optimization of microlenses,” Opt. Express 23(10), 13099 (2015). [CrossRef]  

8. A. A. R Neves, “Photonic nanojets in optical tweezers,” J. Quant. Spectrosc. Radiat. Transfer 162, 122–132 (2015). [CrossRef]  

9. Y. Li, H. Xin, H. Lei, L. Liu, Y. Li, Y. Zhang, and B.-J. Li, “Manipulation and detection of single nanoparticles and biomolecules by a photonic nanojet,” Light: Sci. Appl. 5(12), e16176 (2016). [CrossRef]  

10. Y. Li, H. Xin, X. Liu, Y. Zhang, H. Lei, and B. Li, “Trapping and detection of nanoparticles and cells using a parallel photonic nanojet array,” ACS Nano 10(6), 5800–5808 (2016). [CrossRef]  

11. X. A. Zhang, J. Elek, and C. H. Chang, “Three-Dimensional Nanolithography Using Light Scattering from Colloidal Particles,” ACS Nano 7(7), 6212–6218 (2013). [CrossRef]  

12. E Mcleod and C. B. Arnold, “Subwavelength direct-write nanopatterning using optically trapped microspheres,” Nat. Nanotechnol. 3(7), 413–417 (2008). [CrossRef]  

13. Z. Wang, W. Guo, L. Li, B. Luk’yanchuk, A. Khan, Z. Liu, Z. Chen, and M. Hong, “Optical virtual imaging at 50 nm lateral resolution with a white-light nanoscope,” Nat. Commun. 2(1), 218 (2011). [CrossRef]  

14. R. Ye, Y-H. Ye, H. F. Ma, L. Cao, J. Ma, F. Wyrowski, R. Shi, and J.-Y. Zhang, “Experimental imaging properties of immersion microscale spherical lenses,” Sci. Rep. 4(1), 3769 (2015). [CrossRef]  

15. H. S. S. Lai, F. Wang, Y. Li, B. Jia, L. Liu, and W. J. Li, “Super-Resolution Real Imaging in Microsphere-Assisted Microscopy,” PLoS One 11(10), e0165194 (2016). [CrossRef]  

16. K. A. Sergeeva, M. V. Tutov, S. S. Voznesenskiy, N. I. Shamich, A. Yu. Mironenko, and A. A. Sergeeva, “Highly-sensitive fluorescent detection of chemical compounds via photonic nanojet excitation,” Sens. Actuators, B 305, 127354 (2020). [CrossRef]  

17. I. S. Ruzankina and G. Ferrini, “Enhancement of Raman signal by the use of BaTiO3 Microspheres,” Proceedings – International Conference Laser Optics 2020, Iclo 2020, 9285882 (2020).

18. Z. Chen and A. Taflove, “Photonic nanojet enhancement of backscattering of light by nanoparticles: a potential novel visible-light ultramicroscopy technique,” Opt. Express 12(7), 1214 (2004). [CrossRef]  

19. X. Li, Z. Chen, A. Taflove, and V. Backman, “Optical analysis of nanoparticles via enhanced backscattering facilitated by 3-d photonic nanojets,” Opt. Express 13(2), 526 (2005). [CrossRef]  

20. J. Zelgowski, A. Abdurrochman, F. Mermet, P. Pfeiffer, J. Fontaine, and S. Lecler., “Photonic jet sub-wavelength etching using shaped optical fiber tip,” Opt. Lett. 41(9), 2073 (2016). [CrossRef]  

21. I. V. Minin, O. V. Minin, C. Y. Liu, H. D. Wei, Y. E. Geints, and A. Karabchevsky., “Experimental Demonstration of a Tunable Photonic Hook by a Partially Illuminated Dielectric Microcylinder,” Opt. Lett. 45(17), 4899 (2020). [CrossRef]  

22. C. Y. Liu, W. Y. Chen, Y. E. Geints, O. V. Minin, and I. V. Minin, “Simulation and Experimental Observations of Axial Position Control of a Photonic Nanojet by a Dielectric Cube with a Metal Screen,” Opt. Lett. 46(17), 4292 (2021). [CrossRef]  

23. I. Kassamakov, S. Lecler, A. Nolvi, A. Leong-Hoï, P. Montgomery, and E. Haeggström, “3D Super-Resolution Optical Profiling Using Microsphere Enhanced Mirau Interferometry,” Sci. Rep. 7(1), 3683 (2017). [CrossRef]  

24. A. Darafsheh, G. F. Walsh, L. Dal Negro, and V. N. Astratov, “Optical super-resolution by high- index liquid-immersed microspheres,” Appl. Phys. Lett. 101(14), 141128 (2012). [CrossRef]  

25. D. Colton and R. Kress, Inverse Acoustic and Electromagnetic Scattering Theory, 4 Edition (Springer, 2019).

26. J. A. Davis, D. M. Cottrell, J. Campos, M. J. Yzuel, and I. Moreno, “Encoding amplitude information onto phase-only filters,” Appl. Opt. 38(23), 5004 (1999). [CrossRef]  

27. V. Arrizón, U. Ruiz, R. Carrada, and L. A. González, “Pixelated phase computer holograms for the accurate encoding of scalar complex fields,” J. Opt. Soc. Am. A 24(11), 3500 (2007). [CrossRef]  

28. L. Zhu and J. Wang, “Arbitrary manipulation of spatial amplitude and phase using phase-only spatial light modulators,” Sci. Rep. 4(1), 4728 (2014). [CrossRef]  

29. M. Karamehmedović, “Explicit tight bounds on the stably recoverable information for the inverse source problem,” J. Phys. Commun. 2(9), 095021 (2018). [CrossRef]  

30. A. Kirkeby, M. T. R. Henriksen, and M. Karamehmedović, “Stability of the inverse source problem for the Helmholtz equation in R3,” Inverse Probl. 36(5), 055007 (2020). [CrossRef]  

31. M. Karamehmedović, R. Schuh, V. Schmidt, T. Wriedt, C. Matyssek, W. Hergert, A. Stalmashonak, G. Seifert, and O. Stranik, “Comparison of numerical methods in near-field computation for metallic nanoparticles,” Opt. Express 19(9), 8939 (2011). [CrossRef]  

Data Availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors 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 (13)

Fig. 1.
Fig. 1. Photonic nanojet (PNJ) (a) design by amplitude- and phase-modulated illumination of a dielectric micro-lens with general shape and (b) features of a PNJ, its radial distance, waist width, and decay length, as in [5]. The full width at half maximum (FWHM) amplitude contour is with respect to the amplitude $|E^{\rm tot}|$ of the total electric field.
Fig. 2.
Fig. 2. Probe beam preparation through amplitude and phase modulation of a beam $u_0(\textbf {x})$. Using two spatial light modulators (SLMs) and polarization optics we generate an arbitrary complex field $u(\textbf {x}) = \Psi (\textbf {x}) \exp {i\Phi (\textbf {x})}$ by imprinting the phases $\varphi _1(\textbf {x})$ and $\varphi _2(\textbf {x})$, respectively with the SLMs. The dimensions of the beam are adjusted with a demagnifying telescopic system.
Fig. 3.
Fig. 3. A schematic of the 5-step theoretical procedure for estimating the full width at half maximum (FWHM) of a PNJ with a desired profile.
Fig. 4.
Fig. 4. Desired PNJ profiles $|E^\textrm {tot}_\textrm {PNJ}(\theta )|$ at $\partial B_{R_L+\varrho }$ ($\times 10^{5}$ V/m) for different lens radii $R_L$.
Fig. 5.
Fig. 5. Resulting physically viable PNJ profiles $|E^\textrm {sca}_s(\theta )|$ at $\partial B_{R_L+\varrho }$ ($\times 10^{5}$ V/m) for different lens radii $R_L$.
Fig. 6.
Fig. 6. Waist-width analysis of PNJ profile for four different lens radii with a radial distance of $2\lambda _0$. The black dots are the results from Fig. 5.
Fig. 7.
Fig. 7. Angular PNJ resolution analysis using the projection from 16 with both bandwidth estimates $\mathcal {B}_{\pm }$ from [29].
Fig. 8.
Fig. 8. Waist-width analysis with varying radial distance of a single PNJ with desired FWHM of $0.5\lambda _0$, for four different lens radii $R_L$. The black dots are the results from Figs. 4 and 5.
Fig. 9.
Fig. 9. Left column: PNJ scanning achieved at the single optical wavelength $\lambda _0=532$ nm (common green laser). A 2D SiO$_2$ micro-lens with a circular cross-section of radius 4 $\mu$m, or a square cross-section of side length $8\ \mu$m, is illuminated along the negative $y$-axis by a computed structured incident field (see Section 2). The plots show the amplitude (in V/m) of the resulting total near field, normalized to maximum intensity of 1. Middle column: The computed amplitude and phase profiles of the incident field that produce the desired total near field. The desired PNJ locations in $\mu$m are, from top to bottom: $(x,y)=(0,-4.532)$, $(x,y)=(0,-9.32)$, $(x,y)=(0,-4.532)$, $(x,y)=(0,-9.32)$ (radial distance $1\lambda _0$ or $10\lambda _0$), and $(x,y)=(3,-9)$. Right column: PNJ scanning achieved with the incident field profiles simulated as being produced by an SLM with phase steps of 24.6 mrad in the angles $\varphi _1(\textbf {x})$ and $\varphi _2(\textbf {x})$. See Section 3 for details.
Fig. 10.
Fig. 10. Desired vs. achieved radial distance. a) circular lens, b) square lens.
Fig. 11.
Fig. 11. Achieved waist width and decay length vs. achieved radial distance. a) circular lens, b) square lens.
Fig. 12.
Fig. 12. Achieved PNJ field amplitude maximum vs. radial distance. a) circular lens, b) square lens.
Fig. 13.
Fig. 13. Near fields, normalized to maximum intensity of 1, obtained using uniform plane wave illumination. a) circular lens, b) square lens.

Tables (1)

Tables Icon

Table 1. Mean values of the achieved PNJ parameters in Figs. 1012, for theoretical incident field profiles (Section 2).

Equations (18)

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

χ L ( x ) = { 1 , x = ( x , y ) L , 0 o t h e r w i s e ,
k ( x ) = k 0 [ 1 + χ L ( x ) ( n L 1 ) ] , x R 2 .
( Δ + k ( x ) 2 ) E t o t ( x ) = f ( x ) , x R 2 , E t o t ( x ) = ξ ( x ) , x C . }
( Δ + k 0 2 ) E i n c ( x ) = f ( x ) , x R 2 .
( Δ + k 0 2 ) E t o t ( x ) = f ( x ) + ( k 0 2 k ( x ) 2 ) E t o t ( x ) = ( Δ + k 0 2 ) E i n c α χ L ( x ) E t o t ( x ) , x R 2 ,
( Δ + k 0 2 ) E s c a ( x ) = α χ L ( x ) E t o t ( x ) , x R 2 .
E s c a ( x ) = α Φ 0 ( χ L E t o t ) ( x ) = α y L Φ 0 ( x y ) E t o t ( y ) d y , x R 2 ,
E i n c ( x ) = E t o t ( x ) E s c a ( x ) = E t o t ( x ) + α y L Φ 0 ( x y ) E t o t ( y ) d y , x R 2 .
u ( x ) = 1 2 u 0 ( x ) ( e i φ 1 ( x ) H + V ) .
u ( x ) = u 0 ( x ) cos ( φ 1 ( x ) 2 ) exp i ( φ 1 ( x ) + 2 φ 2 ( x ) 2 ) .
( F s ) ( x ) = j 4 y B R L H 0 ( 2 ) ( k 0 | x y | ) s ( y ) d y , x B R L + ϱ ,
E t o t P N J ( θ ) = exp ( ( θ + π / 2 ) 2 / d 2 ) ,
d ( FWHM , R ) = F W H M λ 0 2 R ln 2 .
F ( s ) m Z Z | m | B σ m s , ψ m L 2 ( B R L ) ϕ m ,
B = a r g m i n m 0 { j m , 1 k 0 R L } ,
B + = a r g m i n m 0 { y m , 1 k 0 R L } .
s T S V D ± = m Z Z | m | B ± σ m 1 E P N J t o t , ϕ m L 2 ( B R ) ψ m ,
E s s c a ± = F ( s T S V D ± ) m Z Z | m | B ± σ m s T S V D ± , ψ m L 2 ( B R L ) ϕ m = m Z Z | m | B ± E P N J t o t , ϕ m L 2 ( B R ) ϕ m .
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.