You are on page 1of 17

Article

Fraxicon for optical applications with aperture ∼ 1 mm:


characterisation study
Haoran Mu1 , Daniel Smith1 , Soon Hock Ng1,2,∗ , Vijayakumar Anand1,3,∗ , Nguyen Hoai An Le1 , Raghu
Dharmavarapu1 , Zahra Khajehsaeidimahabadi1 , Rachael T. Richardson4,5 , Patrick Ruther6 , Paul R. Stoddart1 ,
Henrikas Gricius7 , Tomas Baravykas8 , Darius Gailevičius7 , Gediminas Seniutinas1,2 , Tomas Katkus1 , Saulius
Juodkazis1,7,9

1 Optical Sciences Centre, ARC Training Centre in Surface Engineering for Advanced Materials (SEAM),
Swinburne University of Technology, Hawthorn, Victoria 3122, Australia; haoranmu@swin.edu.au;
danielsmith@swin.edu.au; ale@swin.edu.au; rdharmavarapu@swin.edu.au;
zkhajehsaeidimahabad@swin.edu.au; pstoddart@swin.edu.au; gseniutinas@swin.edu.au;
tkatkus@swin.edu.au; sjuodkazis@swin.edu.au
2 Melbourne Centre for Nanofabrication, Australian National Fabrication Facility, Clayton, 3168, Australia
3 Institute of Physics, University of Tartu, W. Ostwaldi 1, 50411 Tartu, Estonia; vijayakumar.anand@ut.ee
4 Bionics Institute, East Melbourne, Victoria 3002, Australia; rrichardson@bionicsinstitute.org
5 University of Melbourne, Medical Bionics Department, Fitzroy, Victoria, 3065, Australia
6 Department of Microsystems Engineering (IMTEK), University of Freiburg, Germany; ruther@imtek.de
7 Laser Research Center, Physics Faculty, Vilnius University, Sauletekio Ave. 10, 10223 Vilnius, Lithuania;
henrikas.gricius@outlook.com Darius.Gailevicius@ff.vu.lt
8 Femtika Ltd., Keramiku˛ Str. 2, LT-10233 Vilnius, Lithuania; tomas.baravykas@femtika.com
9 WRH Program International Research Frontiers Initiative (IRFI) Tokyo Institute of Technology, Nagatsuta-cho,
Midori-ku, Yokohama, Kanagawa 226-8503 Japan;
* Correspondence: vijayakumar.anand@ut.ee (V.A.); soonhockng@swin.edu.au (S.H.N.)

Abstract: Emerging applications of optical technologies are driving the development of miniaturised 1

light sources, which in turn require the fabrication of matching micro-optical elements with sub-1 mm 2

cross sections and high optical quality. This is particularly challenging for spatially-constrained 3

biomedical applications where reduced dimensionality is required, such as endoscopy, optogenetics, 4

or optical implants. Planarisation of a lens by the Fresnel lens approach was adapted for a conical 5

lens (axicon) and was made by direct femtosecond 780 nm/100 fs laser writing in SZ2080™polymer 6

with photo-initiator. Optical characterisation of the positive and negative fraxicons is presented. 7

Numerical modeling of fraxicon optical performance under illumination by incoherent and spatially 8

extended light sources is compared with the ideal case of plane wave illumination. Considering the 9

potential for rapid replication in soft polymers and resists, this approach holds great promise for the 10

most demanding technological applications. 11

Keywords: fraxicon; micro-optics; RGB; SZ2080™resist, direct-laser-writing 12

Citation: Title. Journal Not Specified


2023, 1, 0. https://doi.org/
1. Introduction 13
Received:
Revised:
Ultrafast laser assisted 3D micro-/nano-fabrication (printing) using additive [1–7], sub- 14

Accepted: tractive [8–10], and patterning [11–18] modes of material structuring is becoming popular 15

Published: for a wide range of technological tasks and applications [19–23] with a good understand- 16

ing of underlying mechanisms of energy deposition and light-matter interactions [24–26]. 17


Copyright: © 2024 by the authors.
One of the most promising trends is the rapid prototyping and manufacturing of various 18
Submitted to Journal Not Specified
for possible open access publication
micro-optical components merging the refractive, diffractive, waveguiding and polarisation 19

under the terms and conditions


control or even combined functionalities [27–31]. Also inscription of waveguides in glasses 20

of the Creative Commons Attri- and crystals [32,33] as well as formation of optical vortex generators and optical memory 21

bution (CC BY) license (https:// structures via form birefringence of self-organised nano-gratings [34]. Fs-laser fabricated 22

creativecommons.org/licenses/by/ optical elements are useful for beam collimation [35], shaping, imaging [36], telecommu- 23

4.0/). nications [37], sensing [38] with expanding range of functionalities and applications due 24

Version February 21, 2024 submitted to Journal Not Specified https://www.mdpi.com/journal/notspecified


Version February 21, 2024 submitted to Journal Not Specified 2 of 17

to possibility of miniaturization and efficient fabrication by direct laser writing [39]. Ap- 25

plications of 3D polymerisation are recently reviewed for micro-mechanical applications 26

triggered by different stimuli: light, temperature, pH [40]. Use of specialised (undisclosed 27

composition) two-photon absorbing photoresists, hydrogels, glass composits developed 28

for commercial 3D printers based on fast scanning and high repetition rate fs-oscillators is 29

a fast-growing application field [41] with a vision of 3D printing applications of computer- 30

designed complex optical elements [42]. A new pathway of 3D formations out of silica 31

with resolution down to ∼ 120 nm using photo-polymerisable resist with a 2%wt. Irgacure 32

369 photoinitor and subsequent calcination was demonstrated recently [43]. Similarly, 33

3D silica structures can be produced from hydrogen silsesquioxane (HSQ) without any 34

photo-initiator by direct write with fs-laser at very different exposure conditions: low 35

repetition rate (∼ 10 kHz) and long ∼ 300 fs pulses [44,45] as well as at high (∼ 80 MHz) 36

repetition rate and short ∼ 120 fs pulses [46]. These demonstrations of 3D SiO2 structuring 37

down to nanoscale resolutions are recently extended to high-refractive index ZrO2 resists 38

and, moreover, demonstrated at writing speeds approaching 10 m/s using fast polygon 39

and stepping scanners [47]. Another strategy for a high throughput 3D printing is the use 40

of multi-focus arrays [48,49]. 41

High resolution 3D printing over large areas remains a formidable challenge, especially 42

for optical applications at shorter wavelengths. Maintaining low surface roughness of one- 43

tenth of the wavelength λ/10 or less adds to the challenge. Making flat micro-optical 44

elements for further miniaturisation and compaction of micro-optical solutions is currently 45

trending, but this is even more challenging for fabrication of optical micro-lenses and 46

functional structures. For controlled phase patterns in diffractive optical elements this is 47

of particular importance and a phase step should be defined over the most narrow lateral 48

width (a step-like height change). 49

Control of focusing from tight (< 10λ) to loose (> 10λ) is dependent on the curvature 50

and diameter of the lens D and its focal length f , which defines the f-number F# = f /D, 51

i.e. the numerical aperture N A ≈ 1/2F# , and the imaging resolution of the lens. More 52

demanding precision is required for larger micro-optics with a large N A. For flat optical 53

elements, e.g., Fresnel lenses, the 2π phase height (along the light propagation direction) is 54

defined over the height of wavelength ∼ λ, which approaches a comparable lateral width 55

for the most off-center phase rings. This is a challenging 3D laser polymerisation task 56

demanding the most high resolution laser printing. 57

High resolution structures can be made via photo-initiator-free laser writing at high 58

pulse intensity I p ∼ (1 − 10) × 1012 W/cm2 or (1-10) TW/cm2 [50]. At such high intensi- 59

ties, the photon energy hν ≈ 1.24/(λ [µm]) [eV] approaches the ponderomotive energy 60

(potential) of an electron, i.e. the cycle-averaged quiver energy of a free electron in an elec- 61

tromagnetic field of light: U p [eV]= 9.33 × (λ [µm])2 × I p [1014 W/cm2 ]. For λ ∼ 1 µm and 62

I p ≈ 10 TW/cm2 , the electron quiver energy during one optical cycle reaches U p ≈ 0.93 eV, 63

which is comparable to the photon energy hν ≈ 1.24 eV. Hence, photo-ionisation of the 64

polymer matrix (∼ 99 wt.%) can take place without a photo-initiator, which is usually 65

doped at below 1 wt% for the wavelength-specific two-photon absorption. Nonlinear 66

or defect-based absorption provides free electrons which promote further ionisation and 67

chemical bond breaking via the ponderomotive channel and avalanche ionisation. 68

Another effective polymerisation pathway is via high-MHz repetition rate exposure 69

of photo-resist, which facilitates thermal accumulation and cross-linking even with a 70

very small initial temperature augmentation due to low sub-1 nJ pulse energy. Low 71

thermal diffusivity of the glass substrate and resist DT = χ/(c p ρ) ≈ 7 × 10−7 [m2 /s] [51] 72

enhance the local temperature rise (here χ ≈ 1 [W·m−1 ·K−1 ] is the thermal conductivity, 73

ρ ≈ 2.2 [g/cm−3 ] is the mass density, and c p ≈ 700 [J/(kg·K)] is heat capacity at constant 74

pressure). High repetition rate laser writing was therefore used in this study. Ionisation of 75

the photo-resist modifies the real and imaginary parts of the refractive index ñ = n + iκ, i.e. 76

permittivity ε ≡ ñ2 , which defines energy deposition. When the real part of the permittivity 77

ϵre ≡ (n2 − κ 2 ) → 0 (epsilon-near-zero ENZ) enters 0 < ϵre < 1, the most efficient energy 78
Version February 21, 2024 submitted to Journal Not Specified 3 of 17

deposition into resist takes place [24]. It is noteworthy that the condition of ε re = 0 (or 79

n = κ) defines a runaway process of dielectric breakdown at the focus, which should be 80

avoided for high precision and resolution of 3D polymerised structures. 81

Designing and manufacturing 3D polymerized structures for beam modulation is more 82

challenging in an integrated optics framework. In most research reports, including dis- 83

tributed Bragg reflector (DBR) laser-based optics where the periodicity of grating elements 84

is parallel to the beam propagation [52], and cases where the periodicity is perpendicular 85

to the beam propagation [53], binary elements are manufactured. This is because the short 86

integration distance often demands short period grating elements, and achieving multiple 87

levels within that short period is challenging. 88

Here, we demonstrate 3D laser printing of a 0.2-mm-diameter fraxicon [54] (flat conical 89

lens) with a triangular phase profile (2π over ∼ 1 µm height) at 5 µm width in SZ2080 90

resist for integration with a micro light-emitting diode (micro-LED). The lateral step-width 91

was defined within ∼ 1 µm. Characterization of the fraxicon performance was carried 92

out with optical microscopy and several optical numerical modeling methods: ray-tracing, 93

analytical solution for Gaussian input, Rayleigh–Sommerfeld (RS) diffraction integral, and 94

holographic simulation. 95

2. Materials and Methods 96

2.1. Resist SZ2080TM 97

The popular hybrid organic-inorganic silica-zirconia composit SZ2080TM was used in 98

this study. Its composition is open (in spite of being a commercial product) with known 99

photo-initiators and their concentrations. This makes a straightforward determination of 100

the linear and nonlinear portions of the absorbed energy contributing to the final polymer- 101

ization at the defined pulse intensity as shown in ref. [24]. The instantaneous permittivity, 102

square of the complex refractive index, ε(t) ≡ (ñ(t))2 defines the amount of the absorbed 103

pulse energy as ε(t) changes during the pulse and follows the intensity envelope I p (t). 104

(a) centre (b) (c)

Fraxicon SEM
Blue Ø = 200 µm
µLED

20 µm

AFM
µLED 250
100 µm
Height (nm)

-250
fraxicon
µLED

γ
-500
1 µm 5 µm
air

0 10 20 30 40
PDMS 20 µm
Center cross section (µm)

Figure 1. (a) Blue micro-LED assembled on a polyimide substrate chip with 460 nm emission and
concept design of a flat-fraxicon for endoscopy applications made out of silicone (polydimethyl-
siloxane - PDMS). (b) Optical microscope image of a fraxicon made by direct laser writing at
780 nm/100 fs/100 MHz (C-Fiber 780 Erbium Laser, MenloSystems) in SZ2080™resist. Fraxicon has
Λ = 1 µm period. (c) Structural characterisation of fraxicon by SEM and AFM, showing blazed 2π
steps with period Λ = 5 µm.
Version February 21, 2024 submitted to Journal Not Specified 4 of 17

SZ2080TM has several properties contributing to its wide use in very different ap- 105

plications: it has a solid (gel) state, high accuracy and resolution of 3D laser printing at 106

nanoscale, ultra-low shrinking, small change in refractive index during laser exposure, 107

a high mechanical stability [55]. It is used for producing nano-photonic structures [55], 108

cell scaffolds [56], bio-medical applications [57], micro-optical elements [28,35,38], and 109

functional structures for micro-fluidics [58]. Among other virtues, also useful for this study, 110

are the glass-matching refractive index [55,59], mechanical stability [60], chemical inert- 111

ness [61], high-resilience for laser-induced damage [62], possibility of chemical doping [63] 112

or simple fictionalization of the fabricated 3D surfaces [64]. Recently, it has been used in 113

thermal post-processing for down-scaling [65] and material-morphing [61]. Interestingly, 114

besides all the aforementioned benefits it also allows the possibility to convert the mate- 115

rial into an inorganic substance which enables the realization of 3D printing of glass at a 116

nano-scale [61]. 117

2.2. Laser printing resist 118

3D laser printing of fraxicons was made by 780 nm/100 fs (C-Fiber 780 Erbium Laser, 119

MenloSystems) tightly focused conditions using an objective lens of numerical aperture 120

N A = 1.4 with a beam diameter saturating the input aperture for optimal resolution. The 121

radius at focus is r = 0.61λ/N A ≈ 340 nm. The pulse repetition rate was 100 MHz. A 122

combination of fast galvanometric scanners and synchronized precision positioning stages 123

were used, similar to the system described in [66]. The scan velocity varied along the 124

structure and depended on the distance to the geometric center. The writing strategy 125

was to scan the structure concentrically in closed loops and iterate each next loop by a 126

displacement of relative radius and height of ∆r = 50 nm (see discussion below on thermal 127

accumulation) and ∆z = 300 nm. In the center, the beam’s travel speed was vsc = 102 µm/s 128

and increased linearly to the edge, reaching 103 µm/s. Also, kinematic commands known 129

as rapid jumps (G0), were not used to perform all movements at the same accelerations. 130

This strategy minimizes the kinematic fabrication error, where smaller radius loops result in 131

greater excentrical acceleration and scan deviations. For simplicity, we did not account for 132

the cumulative dose variation; therefore, the structure features a slight spherical ramp along 133

the radial direction. Each loop had a ramp-up and ramp-down segment of 30 degrees. A 134

previous study showed that the difference in exposure dose affects the final refractive index 135

and optical performance of a micro-lens, which is better described by wave optics [67], 136

however, such intricate control of the index by polymerisation was not investigated for the 137

fraxicon fabricated in this current study. 138

Commercial SZ2080™resist was used for 3D printing with 2-benzyl-2-dimethylamino- 139

1-(4-morpholinophenyl)- butanone-1 (IRG369, Sigma Aldrich) as the photoinitiator dis- 140

solved in the initial pre-polymer at 1% wt. The peak of IRG369 absorbance (in SZ2080™) 141

was at 390 nm with emission at 400 nm as determined by photoluminescence excita- 142

tion spectrosopy [50]; for pure SZ2080™absorption/emission was at 350 nm/400 nm. 143

To improve the resolution, pure SZ2080™is preferable; however, resist with a photoini- 144

tiator is used for a larger laser processing window. The energy of a single pulse, with 145

transmission losses accounted for at the focal point, was E p ≈ 96 pJ, corresponding to 146

fluence per pulse Fp = 0.0275 J/cm2 and intensity I p = 0.275 TW/cm2 (average). Con- 147

sequently, the pulses generated a negligible ponderomotive potential. The dwell time 148

required for the beam to cross the focal diameter 2r is tdw = v2rsc = 6.8 ms and, at the 149

repetition rate R, defines the number of accumulated pulses over the focal spot is large 150

N = tdw R = 680 × 103 . Thermal spread (cooling) of the laser-heated focal volume is 151

defined by the time tth = (2r )2 /DT = 660 ns, while a time separation between pulses is 152

only 1/R = 10 ns. Hence, a very strong thermal accumulation takes place at the used direct 153

laser writing. An average temperature drop at the arrival of the next pulse occurs due 154
Version February 21, 2024 submitted to Journal Not Specified 5 of 17

(a) (b)
Confocal: 1
Optical:

Intensity
Z = 30 µm Z = 120 µm
0
1
Red

0
1 Z = 60 µm Z = 150 µm
Green
extract:

0
1
Blue
20 µm

Y Y Z = 90 µm Z = 180 µm

20 µm 20 µm
Z X
0

Figure 2. Optical characterisation of fraxicon shown in Fig. 1(b) by optical microscope with white
condenser illumination. (a) Confocal intensity distribution and its RGB color content along the focus
(a “ non-diffracting” part of the axial intensity). (b) Optical images at different axial positions along
white light propagation (along z-axis.)

to heat transfer to surrounding cold material. The temperature accumulation TN can be 155

explicitly calculated for the N pulses, when single pulse temperature jump is T1 [68]: 156

1 − βN
TN = T1 (1 + β + β2 + ... + β N ) ≡ T1 , (1)
1−β
q
tth
where β = tth +1/R is the constant which defines heat accumulation; β → 1 at high 157

repetition rate R → ∞. For the used experimental conditions, β ≈ 0.9925. The first 158

N = 10 pulses cause significant temperature jump TN = 9.67T1 (N = 10). Considering 159

the exothermic character of polymerization, a minute temperature rise at the focal region 160

causes a guided thermal polymerisation [69]. 161

The samples were prepared by drop-casting the liquid resin on a standard microscope 162

cover slip and pre-condensating at 50◦ C for 24 hours. After exposure, the samples were 163

developed in methyl-isobutyl-ketone for 30 min, then rinsed with pure-developer and 164

air dried under normal room conditions. The refractive index of the resist at visible 165

wavelengths is approximately nSZ ≈ 1.5. While the exact definition of the 2π phase height 166

was experimentally challenging for smaller period fraxicons, the height of polymerised 167

phase ramps normalised by the wavelength (or 2π in phase) is h × nSZ /λ ∼ 2, which 168

corresponds to the second-order (2 × 2π) phase steps. This strategy was used for fabrication 169

because of the more straightforward definition of the exact required geometry with a 170

focused laser pulse, which occupies a defined volume. 171

For more widespread practical implementation of the proposed 3D laser printing of 172

micro-optical elements, the fabrication conditions were optimized to complete the entire 173

laser writing step within 100 min for all 200 µm-diameter fraxicons in this study. 174

2.3. Characterisation 175

Micro-LEDs (CREE C460TR2227, CREE, Durham, USA) were used for the design 176

and prototyping of a low-profile implantable device. The footprint of the µLED was 177
Version February 21, 2024 submitted to Journal Not Specified 6 of 17

0.27 × 0.22 mm2 , with an emitter area comparable to the D = 0.2 mm diameter fraxicon. 178

The micro-LEDs were assembled on a 10 µm thick polyimide substrate [70,71], which was 179

formed by spin-coating first a 5-µm-thin polyimide layer on a silicon wafer (diameter 100 180

mm), followed by the sputter deposition of a metallic thin film. Interconnecting tracks 181

were then patterned using lift-off technology and the track thickness was increased by 182

electroplating 1 µm of gold to reduce the electrical line resistance. A second polyimide 183

layer was subsequently deposited to insulate the metal tracks. 184

To access the metal tracks, small openings were formed in the top polyimide layer by 185

reactive ion etching (RIE) with oxygen plasma. A second metallisation and electroplating 186

step was used to define "bonding pads" for the micro-LED chips and zero-insertion force 187

(ZIF) connector pads for wire bonding the test structure to a printed circuit board. Finally, 188

the shape of the polyimide substrate was defined by trenching the stack of polyimide 189

layers down to the silicon substrate with a second RIE process step. The substrates could 190

then be peeled from the silicon wafer using tweezers and the micro-LED chips were 191

assembled on the pads of the polyimide substrate by flip-chip bonding [70,71]. They 192

were subsequently underfilled with a biocompatible adhesive (EPO-TEK 301-2, Epoxy 193

Technology, Inc., Billerica, USA) to electrically insulate the pads located at the interface 194

between micro-LED chips and polyimide substrate. Structural characterisation of fraxicon 195

was carried out by optical microscopy, scanning electron microscopy (SEM) and atomic 196

force microscopy (AFM). Typical results are shown in Fig. 1. 197

2.4. Fraxicon: basic properties and design 198

The positive fraxicon was designed with a diameter D = 0.2 mm and featured 199

20 blazed rings. The thickness profile (t f ra ) of the fraxicon as a function of the radial 200
p
coordinate (r = x2 + y2 ) is given as Eqn. 2 201

   
h
t f ra = h − mod r ,h (2)
Λ

where h is the height of the fraxicon corresponding to 2π phase retardation (h = (n λ−1) = 202
SZ
1 µm, λ is the incident wavelength, and nSZ ≈ 1.5 is the refractive index of SZ2080™), 203

Λ = 5 µm is the period of gratings (rings), and mod is a function of remainder after division 204

(modulo operation). 205

The axial intensity AI (z) distribution (along the z-axis) of an axicon/fraxicon depends 206

on the radial intensity at the input and can be found from a relation based on Snell’s law 207

r = z(n − 1)α , here n is the refractive index of axicon and α is the base angle of axicon 208

(also, the angle


 required
 to the full π angle at the tip of the axicon). For the Gaussian beam 209
2
Iin = I0 exp − 2r
w2o
, where wo is the beam waist: 210

2z2 (n − 1)2 α2
 
AI (z) = I0 exp − × [2πz(n − 1)2 α2 ]. (3)
w2o

This equation can be generalized and the input intensity profile Iin (r ) can be expressed in 211

terms of r-to-z mapping AI (z) as Iin (r ) = M × AI [z = r/(n − 1)α]/r, where M = 2π (n1−1)α 212

is the axicon geometry defined constant. This is valid for the on-axis intensity in 0 < z < 213

DOF region with depth-of-focus DOF = w/[(n − 1)α] defined by the radius of beam w. 214

3. Results 215

3.1. Theory: fraxicon illuminated by an incoherent and extended source 216

The Gauss-Bessel beam generation by fraxicon is mostly discussed using spatially 217

coherent illumination such as laser beams. The timing of photon emission from the LED 218

is however disorganized. To describe beam generation by fraxicon for a spatially in- 219

coherent illumination, such as light from an LED, an incoherent imaging framework 220

is needed. The LED needs to be considered as a collection of points and the gener- 221
Version February 21, 2024 submitted to Journal Not Specified 7 of 17

(a) (b)

1 Y

0.9
Z 100 µm

Normalised Intensity
0.8

0.7

0.6

0.5 Sum
Red
Green
0.4
Blue
20 µm

200 400 600 800 1000


z (μm)

Figure 3. (a) Optical image of fraxicon with 1.5 times larger width of the 2π-steps. Image taken under
white light condenser illumination. (b) Axial intensity profile calculated from the lateral image stacks
(same as in Fig. 2). The inset shows the confocal profile of intensity.

ated beam is formed by the summation of intensities of the beam generated for every 222

point. While it is difficult to discriminate beams generated for coherent and incoher- 223

ent illuminations, the generation approach is quite different from one another. We con- 224

sider√a point in the LED as a Delta function p emitting a spherical wavefront with inten- 225

sity Is given as S(1/zs ) = exp[ j(2π/λ) x2 + y2 + z2s ]. The phase of the fraxicon is 226

given as ϕ = exp[ j(2π/λ)t f ra ]. The complex amplitude after the fraxicon is, therefore, 227

ψ = Is C1 L(rs /zs )S(1/zs )ϕ, where L is a linear phase and C1 is a complex constant. The in- 228

tensity distribution at a distance zr for a Delta-function is given as IDelta ≈ |ψ Q(1/zr )|2 ,


N
229

where Q(1/zr ) = exp[ j(π/λzr )( x2 + y2 )] and ’ ’ is a 2D convolutional operator. The inten-


N
230

sity distribution obtained for the entire LED active area can be given as ILED ≈ | IDelta O|2 ,
N
231

where O is the LED active area in a square shape filled with ones and zeros around it. It 232

must be noted that the above summation is not a complex summation but an addition of 233

intensities as the phase information is not present. The above expression is an approximate 234

one as Fresnel approximation was used for propagation between fraxicon and camera. To 235

understand the beam generation deeper, let us consider the Delta function with no linear 236

phase attached to it, i.e., the one at the center of the LED active area. This Delta function 237

generates a spherical wave that interacts with the fraxicon. We already established that 238

fraxicon and axicon consist of lens functions with different focal lengths multiplexed in 239

the transverse direction [72]. At the camera plane, one of the lens functions satisfies the 240

imaging condition generating a sharp Delta-like function. The other lens functions cause 241

ring patterns around this sharp Delta-like function typical of a squared Bessel function. 242

The other points in the LED active area with linear phases attached to them create off-axis 243

squared Bessel functions on the camera. The recorded intensity distribution is the sum of all 244

the contributions from the points of the LED active area. This is different from coherent illu- 245

mination as it would depict the behaviour of light emission from a single point. Simulation 246

results for intensity obtained for a point in an LED, intensity obtained for a circular region 247

in the LED by incoherent superposition due to spatial incoherence and intensity obtained 248

for the same circular region but with coherent superposition are shown in Figs. 4(a), (b) 249

and (c), respectively. The incoherent superposition do not generate distinct rings around 250

the central maxima as expected of a Bessel distribution due to the lack of phase relations 251

for light emitted from every point. The temporal coherence length for a Gaussian fit of LED 252
2
4 ln 2 λ0
emission spectra is given by Ltc = π ∆λ ,
where λ0 is the central emission wavelength 253

and ∆λ is its width or Full-Width-at-Half-Maximum (FWHM), and it coincides well with 254

experimentally measured values Ltc ≈ 2 µm using Mach-Zehnder interferometry [73]. 255


Version February 21, 2024 submitted to Journal Not Specified 8 of 17

Figure 4. (a) Intensity distribution obtained for a single point of the LED. (b) Intensity distribution
obtained for a circular region of the LED by incoherent superposition. (c) Intensity distribution
obtained for the same circular region as (b) but with coherent superposition.

3.2. Experimental characterisation of light intensity 256

An axicon or a conical lens is a very useful optical element that forms an elongated 257

axial intensity distribution when illuminated by a Gaussian beam. The formed Bessel- 258

Gaussian beam has a diameter defined by the first minimum of the Bessel function J0 (k ⊥ r ), 259

where r is the radial coordinate in the lateral plane, and the perpendicular component of 260

the propagation wavevector k ⊥ is defined by k ⊥ ≡ k sin γ with γ being the half-cone angle 261

of the beam with an optical axis, wavevector k = ω/c ≡ 2π/λ and k2⊥ + k2∥ = (ω/c)2 ; 262

ω, c are the cyclic frequency and speed of light, respectively. The diameter of the axially 263

extended focus is d B = 4.816/k ⊥ and the length depends on the diameter of the incident 264

beam D (lens diameter) as Zmax = D/(2 sin γ). Large D and small γ facilitates to have a 265

long, so-called, non-diffracting region of intensity on the optical axis. Axicon can be made 266

flat in the same way as a Fresnel lens is made from concentric segments. 267

A fraxicon lens of D = 300 µm, consisting of 30 circular rings of ∼ 5 µm width and 268

0.7 µm height were polymerised in SZ2080™using fs-laser direct write (Fig. 1). The step 269

between adjacent phase-ramps was within ∼ 1 µm. Fraxicons are especially promising 270

for optical devices which have strong requirements for spatial constraints. If fraxicons are 271

used for optical focusing of micro-light sources, e.g., LED, it is usual that illumination of 272

the aperture of the fraxicon will take place from an extended, non-collimated light source 273

with possible intensity inhomogeneities of tens-of-micrometer scale (see powered µLED 274

image in Fig. 1(a)). Such a situation can be modeled using an optical microscope under 275

condenser illumination of a fraxicon on a sample plane discussed next. 276

3.2.1. Confocal intensity mapping 277

Figure 2(a) shows axial intensity distribution calculated from the axial stacks as shown 278

in (b) using a standard microscope (Nikon, Optiphot) under white-light condenser illumina- 279

tion in transmission mode. Images at every ∆z = 1 µm step were recorded and 3D intensity 280

distribution was calculated using own Matlab code. The resulting confocal axial intensity 281

distribution was separated into basic red, green, blue (RGB) color channels. The width of 282

the focal region has d B ≈ 20 µm, hence, k sin γ = 4.816/d B , or γ ≈ 1.1◦ for λ ≈ 0.5 µm. 283
( R) ( B)
As expected, the width for R-red axial distribution dB
was larger than for B-blue, dB . 284

Strong widening of the intensity profile for the blue channel is most probably defined by 285

the comparatively high-NA of 0.9 of the imaging microcope lens, since the conical angle 286

γ ≈ 1.1◦ is small. 287


Version February 21, 2024 submitted to Journal Not Specified 9 of 17

The length of the most uniform intensity region is wavelength dependent since the 288

height of the phase ramps is fixed. The longest non-diffrating region was for the blue 289

wavelength. This is also clearly seen in the lateral cross sections at z > 150 µm (Fig. 2(b)). 290

Figure 3 shows fraxicon and its axial intensity cross sections under condenser illu- 291

mination in the microscope. The period of concentric phase ramps was 1.5 times larger, 292

hence ∼ 7.5 µm while the diameter and the height of the phase ramps were the same 293

as for fraxicon in Figs. 1(b) and 2. This results


√ in a longer non-diffracting region. The 294

intensity decreased along propagation as z and RGB colors had an effective different 295

length. The exact axial intensity profile depends on the wavevector spread for different 296

colors in illumination light. The ideal performance of fraxicons was modeled numerically 297

and discussed next. 298

3.3. Numerical predictions by ray and wave simulations 299

For numerical simulations we used the ideal case of triangular phase steps as 20 rings 300

with each of them having 5 µm width and 1 µm height. Material of the phase steps has 301

a refractive index of 1.5 (dispersion-free). This makes a D = 0.2 mm diameter fraxicon, 302

which was illuminated with a plane wave (a singe wavenumber) at different RGB colors. 303

3.3.1. Wave optics 304

The Rayleigh–Sommerfeld (RS) diffraction integral can be used for exact prediction 305

on axial intensity profile, which is consistent with the wave-optical approach; it is used 306

for thin graphene micro-lenses [74]. The electric field distributions U2 (r2 , z) in the axial 307

plane at different axial z positions were calculated using MATLAB program based on the 308

RS diffraction integral expressed as: 309

−i e−ikr
ZZ
U2 (r2 , θ2 , z) = U1′ (r1 , θ1 ) · · cos(n, r )dr1 dθ1 , (4)
λ r

where λ is the incident light wavelength (R=700 nm, G=546.1 nm and B=435.8 nm), k = 2π λ 310

is the wave vector, (r1 , θ1 ) and (r2 , θ2 ) are the polar coordinates in the diffraction plane (the 311

plane immediately behind the fraxicon) and q observation plane (the focal plane), respectively; 312
p
r = (z + ( x2 − x1 ) + (y2 − y1 ) = (z2 + r12 + r22 − 2r1 r2 cos(θ1 − θ2 ), cos(n, r ) is de-
2 2 2 313

fined as the cosine of the angle between the unit normal vector n of the diffraction plane 314

and the position vector r from point (r1 , θ1 ) to point (r2 , θ2 ), and U1′ (r1 , θ1 ) is the E-field 315

immediately behind the fraxicon. The incident wave U1 (r1 , θ1 ) is diffracted by the fraxicon 316

(a) (b)
Intensity Distribution (Axial plane): Normalised Intensity
1

0.8

0.6
X (µm)

0.4

0.2
20 µm

100 µm
0
Z (µm)

Figure 5. (a) Simulation by the Rayleigh–Sommerfeld (RS) diffraction integral for non-polarised plane
wave with RGB wavelengths: R=700 nm, G=546.1 nm and B=435.8 nm (top-down). The calculated
intensity cross section is given along propagation. (b) Central intensity cross section for the RGB
colors; inset shows geometry of simulated positive fraxicon with Λ = 5 µm period.
Version February 21, 2024 submitted to Journal Not Specified 10 of 17

(a) (b) (c) (d)


α B
glass
BGR G
γ R

screen
1 mm
1 mm 1 mm
B
G
fraxicon R

Figure 6. Ray tracing through (fr)axicons and their pairs (OSLO, Lambda Research Co.). Positive (a)
and negative (b) fraxicon and axicon. Pairs of (fr)axicons (i.e. collimating telescopes): two positive (c)
and a negative-positive pair (d) for the bulk and flat-optics realisations. Illumination with RGB color
light shown by arrow. Location of the planarised (fraxicon) region is shown by shaded markers. The
base angle of axicon α and γ is the half-cone angle of the Bessel beam. The refractive index of glass
was n = 1.52 and the base angle α = ±40◦ , where the positive sign is for real focus (a) and negative
for virtual focus (b); a large angle was chosen for visualisation and a strong angular dispersion of
RGB rays.

through amplitude and phase modulations, and the electric field modified by the fraxicon 317

U1′ (r1 , θ1 ) can be expressed by Eqn. 5: 318

q
U1′ (r1 ) = U1 (r1 ) · T (r1 ) · e−ik·Φ(r1 ) , (5)

where T (r1 ) is the transmission distribution (amplitude modulation) of the fraxicon, and 319

Φ(r1 ) = nSZ · t f ra (nSZ is the refractive index of SZ2080™, and t f ra is the thickness profile 320

of the fraxicon) is the phase modulation provided by the fraxicon. Consequently, the light 321

intensity distributions in the axial plane can be calculated by squaring the electric field: 322

I (r2 , z) = |U2 (r2 , z)|2 (Fig. 5). 323

Apparent differences between modeling of ideal plane wave illumination of a fraxicon 324

(Fig. 5) and experimental imaging using condenser illumination of the microscope (Fig. 3) 325

is due to presence of different k components at the same wavelength. Such a situation is 326

expected in real applications where the fraxicon is placed in front of a µLED (Fig. 1(a)). 327

However, the basic√ features of the oscillatory nature of intensity along the propagation axis, 328

its decay as ∝ z and sub-1 mm long extent of high-intensity section are consistent. Exact 329

intensity distributions can be well controlled using tailored illumination of an axicon [75], 330

hence, fraxicon as well. 331

3.3.2. Ray optics 332

The Optical Software for Layout and Optimisation (OSLO, Lambda Research Co.) is 333

a ray tracing tool used to model light propagation through (fr)axicons. Figure 6 shows 334

RGB beam propagation through positive and negative (fr)axicons and their pairs, which 335

collimates the beam. Change from an axicon to fraxicon is conveniently made by tab- 336

selection in the OSLO input; the fraxicon surface appears flat in the viewer, however, 337

it encodes fraxicon layout of 2π phase ramps. As expected by the shape of axicon, a 338

conical prism, light dispersion is evident in RGB ray tracing (Fig. 6). It manifests as color 339

aberration along the focal region. A slight color appearance of fraxicon when imaged with a 340

microscope is evident in Fig. 3(a) with red-center and blue outside edges of the phase ramps. 341

Figure 6 clearly shows the compactness of micro-optical constructions using fraxicons and 342

less dispersion in a pair of the flat optical elements (c,d) for the same diameter of the input 343

beam; alignment in Fig. 6 is at the plane of screen. 344

Such compact fraxicons have the potential for application in space telescope tech- 345

nologies. Recently, multiple-order diffractive engineered surface (MODE) lenses, which 346
Version February 21, 2024 submitted to Journal Not Specified 11 of 17

comprise a front-surface multiple-order diffractive lens (MOD) and a rear-surface diffractive 347

Fresnel lens (DFL), are planned to aid in the search for Earth-like planets and exo-planets 348

in the universe as part of the upcoming telescope array known as the Nautilus Observa- 349

tory [76–78]. 350

4. Discussion 351

The circular grating pattern makes an efficient collection of illumination at different 352

wavelengths onto the optical axis and is useful for light delivery and collection to spot 353

sizes of tens-of-µm. The long focal extension is helpful for light coupling into optical fibers 354

and laser machining using increasingly miniaturised laser sources. Photonic crystal (PhC) 355

lasers now deliver tens of Watts output powers at near-IR wavelengths from sub-1 mm 356

apertures at very low divergence angle [79]. Micro-optics based on a high damage threshold 357

SZ2080™photo-resist is a promising solution [80]. 358

Gauss-Bessel beams formed by phase profiles closely matching those of a fraxicon 359

were made with a spatial light modulator for record high ∼ 104 aspect ratio modification 360

of dielectrics and semiconductors for scribing and dicing with nanoscale resolution of 361

tens-of-nm [81]. For such material modification multiple ionisation locations along the 362

non-diffracting intensity distribution are essential and their connection occurs via back- 363

scattering under multi-pulse irradiation. Hence oscillatory nature of Bessel beam intensity 364

on the axis (Fig. 5) is beneficial for such material modification. 365

Replication of polymerised fraxicons can be made using Ni-shim (plasma coating 366

with subsequent electrochemical deposition of Ni), which replicates structures with 10 nm 367

feature sizes, e.g., nano-needles of black-Si [82]. 368

The emergence of a new generation of optical cochlear implants has highlighted the 369

need for better miniaturisation and integration of optical elements. It has been shown 370

that optical neuromodulation with visible light, facilitated by optogenetics, can confer 371

higher spatial precision of neural activation compared to traditional electrical stimulation 372

methods [83]. For cochlear implants, broad current spread from electrical devices reduces 373

the number of independent stimulating channels [84,85]. Higher spatial precision of 374

activation could greatly increase the number of independent stimulating channels and 375

enable simultaneous channel stimulation which would greatly enrich the sound quality 376

experienced by cochlear implant recipients [86,87]. Strategies to deliver focussed electrical 377

stimulation, including tripolar and focussed multipolar stimulation strategies, have failed 378

to deliver significant clinical benefit [88,89]. The emerging development of optical arrays 379

could provide an alternative solution. 380

In rodents, such as mice, rats, and gerbils, auditory neurons were modified to express 381

photosensitive ion channels. The spread of activation during optical stimulation was 382

significantly lower compared to electrical stimulation, resulting in near-physiological 383

spectral resolution when using light emitters that are in close proximity to the neural 384

tissue [71,90,91]. Furthermore, during two-channel simultaneous optical stimulation in 385

the mouse cochlea using micro-LEDs with a pitch of just 0.52 mm, channel interaction 386

was 13-15-fold lower than simultaneous electrical channel stimulation [92]. In the human 387

cochlea, where there is a greater distance between the emitter and the neural tissue, optical 388

arrays of micro-LEDs are still predicted to significantly reduce the spread of activation in 389

the cochlea to 0.4-1.0 octaves, up to 4-fold lower than electrical stimulation [93,94]. While 390

modelling data suggests waveguides could provide even greater spectral selectivity [95], 391

the fraxicon technology presented here could potentially be used to focus the emission cone 392

and improve the spectral resolution provided by LEDs in the human cochlea. 393

Antireflection coating could be used to increase transmittance of fraxicon by coating 394



a film with thickness λ/4 of a refractive index nout n f rax ≈ 1.25, where nout = 1 (air) is 395

the refractive index at the focal region and n f rax = 1.5. MgF2 is a good candidate for the 396

antireflection coating over visible spectral range. The atomic layer deposition (ALD) is a 397

candidate for conformal coating of 3D surfaces as well as magnetron sputtering. 398


Version February 21, 2024 submitted to Journal Not Specified 12 of 17

Fraxicons with ∼ 1 − µm-tall phase ramps can be made using a scanning thermal tip 399

(nano-cantilever) method based on the AFM principle (NanoFrazor; Suppl. A). By using 400

Polyphthalaldehyde (PPA) 4% resist for a gray scale mask, e.g., 3000 rpm spin coating for a 401

100 nm film, a sacrificial PPA mask can be made with high nanoscale control and resolution 402

down to 10 nm. Such mask then is used for transfer of the 3D pattern onto a substrate by 403

reactive ion etching (RIE). For Si etch (Fig. A1), etching contrast was ∼ 2.6 and translates to 404

260 nm deep structures when PPA resist is ∼ 100 nm. Different patterns and metasurface 405

structures can be easily designed for patterning into resist using open source toolbox [96]. It 406

is based on the industry standart GDSII, a binary database file format for Electronic Design 407

Automation data exchange of integrated circuit layout. 408

5. Conclusions and outlook 409

3D printing of 0.2-mm-diameter flat fraxicon lenses with 2π phase step defined over 1 410
and 5 µm lateral widths and ∼ 0.7 µm height (axial length) was made in SZ2080™(with 411
1%wt. IRG) resist. 3D polymerisation was carried out by tightly focused 780 nm/100 fs/100 MHz 412

laser irradiation. High repetition rate and strong overlap of laser pulses at the focal diame- 413
ter of 680 nm with a linear scan-step of 50 nm between the adjacent pulses determined a 414
strong thermal accumulation of 3D laser printing/polymerisation with only ∼ 0.1 nJ pulses. 415
The used tight focusing with depth-of-focus approximately 3-4 times longer than the focal 416
diameter, i.e., 2.5 − 3 µm, was significantly shorter than the axial pulse length ct p ≈ 30 µm. 417
The entire fraxicon was printed within 1.5 hours. It was tested for illumination by 418
an extended light source (condenser of a microscope) to simulate its performance for 419
µLED illumination in endoscopy and opto-probes. The RGB color analysis revealed an 420
axial color separation along light propagation which is significant for extended white 421
light source. Different axial intensity distributions are predicted from the analysis of a 422
fraxicon illumination by incoherent and coherent light sources. Fraxicons with wider 423
sections of phase ramps have lesser diffraction-related effects as compared with flat Fresnel 424
lenses, which have increasingly narrower phase ramps at larger diameters, consequently, a 425
stronger diffraction. These aspects of fraxicon use in micro-optical applications have to be 426
considered. 427

Author Contributions: Conceptualization, PRS, RTR, PR, SJ; methodology, VA, SHN, TK, TB; valida- 428

tion, HM, DS, NHAL, RD, ZK, GS; investigation, HG, TB, DG; data curation, VA, HM, DS, NHAL, RD, 429

ZK, GS; writing—original draft preparation, HM, VA, DS, PRS, DG, SJ; writing—review and editing, 430

all the authors; visualization, VA, HM, DS, SJ; supervision, VA SHN, PRS, SJ; funding acquisition, 431

SHN, SJ. All authors have read and agreed to the published version of the manuscript. 432

Funding: The project was funded by ARC LP190100505 grant, European Union’s Horizon 2020 433

research and innovation program grant agreement No. 857627 (CIPHR), ARC LP190100505, and 434

NHMRC 2002523 grants. The thermal scanning probe lithography tool NanoFrazor was acquired via 435

ARC LE160100124 grant. 436

Data Availability Statement: Data can be made available upon a reasonable request. 437

Acknowledgments: We are grateful for process development of fraxicons at Femtika. We acknowl- 438

edge a material transfer agreement with Kyoto University for testing micro-optical elements using 439

PhC laser. This work was performed in part at the Melbourne Centre for Nanofabrication (MCN) in 440

the Victorian Node of the Australian National Fabrication Facility (ANFF). 441

Conflicts of Interest: The authors declare no conflict of interest 442

Appendix A. Thermal Scanning Probe Lithography (tSPL) 443

Scanning thermal tip nanolithography (NanoFrazor, SwissLitho) has 10 nm resolution 444

(defined by the tip) and possibility to sublimate/evaporate solid linear Polyphthalaldehyde 445

(PPA; anionic from Sigma-Aldrich) resist upon fast and localised heating. The height control 446

of resist removal is adjustable; hence, 3D surface can be defined for a sacrificial mask in 447

RIE-ICP plasma etch. 448


Version February 21, 2024 submitted to Journal Not Specified 13 of 17

(a) (b)
thermal tip PPA mask

100 nm
PPA mask
PPA mask
5 µm

etched
Si Si
9 nm

10 µm -81 nm
PPA mask

Airy beam fraxicon grating

radians
PPA mask

100 nm
Si

10 µm
(c)

Figure A1. Sacrificial mask etch of micro-optical elements into Si. (a) A tilted-view SEM image of a
binary Fresnel lens ( f = 50 µm) after RIE plasma etching, optical image (yellow middle-inset) and
AFM image of the pattern defined in PPA resist by thermal tip nanolithography (NanoFrazor; top-left
inset shows the writing tip). (b) Optical profilometer cross section of a f = 100 µm lens; insets show
optical and AFM images. (c) Microscope images of PPA masks of the cubic phase structure (Airy
beam generator; a phase 0-2π phase map in the inset), 4-level fraxicon, binary grating with period of
5 µm. Optical profilometer cross section of the grating after RIE.

Etch depths were measured using optical profiler to estimate the etch depth into Si 449

and etch contrast with PPA. Figure A1 shows several examples of a grating, Fresnel lens, 450

and 3D phase patterns of cubic phase (Airy beam generator) and 4-level fraxicon. The etch 451

depth was approximately around 160-180 nm for 65-nm-deep pattern defined in PPA; PPA 452

thickness was 100 nm for 5% solution of PPA spin coated at 2000 rpm. This translated to an 453

etch selectivity (Si vs. PPA) of 2.6, which is less compared to best-performing examples with 454

NanoFrazor reaching 4-10 depending on pattern and etch conditions. The most probable 455

reason for the lower selectivity is introduction of O2 into the process, which might have 456

suppressed the selectivity of SF6 for Si. The RIE conditions were (Oxford, Plasmalab 100 457

ICP380): SF6 40 sccm, C4 F8 60 sccm, O2 10 sccm, at RF of 40 W, ICP at 1500 W, etch time 458

of 50 s and pressure of 10 mTorr. A RF strike power was increased to 50 W for the 5 s 459

plasma ignition step only. For binary patterns, a metal hard mask (e.g., Cr) can be used for 460

Si etching via Bosch process up to a large 150 µm depth as shown for THz optical vortex 461

generator elements based on a metasurface design [97]. 462

Typical pattern writing time for 60 nm structures of Fresnel lens in PPA (Fig. A1(a)) 463

was 5 min for 50 × 50 µm2 area of f = 50 µm lens and 4-level fraxicon. The residuals of 464

PPA are removed before with a gentle barrel ashing ∼5 s, 200 W, 400 sccm O2 . Also, a 465

developer Tetramethylammonium hydroxide (TMAH) can be used. 466

References 467

1. Maruo, S.; Nakamura, O.; Kawata, S. Three-dimensional microfabrication with two-photon-absorbed photopolymerization. Opt. 468

Lett. 1997, 22, 132–134. 469

2. Kawata, S.; Sun, H.B.; Tanaka, T.; Takada, K. Finer features for functional microdevices. Nature 2001, 412, 697–698. 470

3. Serbin, J.; Egbert, A.; Ostendorf, A.; Chichkov, B.N.; Houbertz, R.; Domann, G.; Schulz, J.; Cronauer, C.; Fröhlich, L.; Popall, M. 471

Femtosecond laser-induced two-photon polymerization of inorganic–organic hybrid materials for applications in photonics. Opt. 472

Lett. 2003, 28, 301–303. 473

4. Sun, H.B.; Kawakami, T.; Xu, Y.; Ye, J.Y.; Matuso, S.; Misawa, H.; Miwa, M.; Kaneko, R. Real three-dimensional microstructures 474

fabricated by photopolymerization of resins through two-photon absorption. Opt. Lett. 2000, 25, 1110–1112. 475

5. Takada, K.; Sun, H.B.; Kawata, S. Improved spatial resolution and surface roughness in photopolymerization-based laser 476

nanowriting. Applied Physics Letters 2005, 86, 071122. 477

6. Sun, H.B.; Tanaka, T.; Kawata, S. Three-dimensional focal spots related to two-photon excitation. Applied Physics Letters 2002, 478

80, 3673–3675. 479


Version February 21, 2024 submitted to Journal Not Specified 14 of 17

7. Farsari, M.; Chichkov, B. Two-photon fabrication. Nature Photonics 2009, 3, 450–452. 480

8. Joglekar, A.P.; hua Liu, H.; Meyhöfer, E.; Mourou, G.; Hunt, A.J. Optics at critical intensity: Applications to nanomorphing. 481

Proceedings of the National Academy of Sciences 2004, 101, 5856–5861. 482

9. Mao, B.; Siddaiah, A.; Liao, Y.; Menezes, P.L. Laser surface texturing and related techniques for enhancing tribological 483

performance of engineering materials: A review. Journal of Manufacturing Processes 2020, 53, 153–173. https://doi.org/https: 484

//doi.org/10.1016/j.jmapro.2020.02.009. 485

10. Korte, F.; Serbin, J.; Koch, J.; Egbert, A.; Fallnich, C.; Ostendorf, A.; Chichkov, B. Towards nanostructuring with femtosecond laser 486

pulses. Applied Physics A 2003, 77, 229–235. 487

11. Davis, K.M.; Miura, K.; Sugimoto, N.; Hirao, K. Writing waveguides in glass with a femtosecond laser. Opt. Lett. 1996, 488

21, 1729–1731. 489

12. Miura, K.; Qiu, J.; Inouye, H.; Mitsuyu, T.; Hirao, K. Photowritten optical waveguides in various glasses with ultrashort pulse 490

laser. Applied Physics Letters 1997, 71, 3329–3331. 491

13. Shimotsuma, Y.; Kazansky, P.G.; Qiu, J.; Hirao, K. Self-Organized Nanogratings in Glass Irradiated by Ultrashort Light Pulses. 492

Phys. Rev. Lett. 2003, 91, 247405. 493

14. Beresna, M.; Gecevičius, M.; Kazansky, P.G.; Gertus, T. Radially polarized optical vortex converter created by femtosecond laser 494

nanostructuring of glass. Applied Physics Letters 2011, 98, 201101. 495

15. Bellouard, Y.; Said, A.; Dugan, M.; Bado, P. Fabrication of high-aspect ratio, micro-fluidic channels and tunnels using femtosecond 496

laser pulses and chemical etching. Opt. Express, 12, 2120–2129. 497

16. Vitek, D.N.; Block, E.; Bellouard, Y.; Adams, D.E.; Backus, S.; Kleinfeld, D.; Durfee, C.G.; Squier, J.A. Spatio-temporally focused 498

femtosecond laser pulses for nonreciprocal writing in optically transparent materials. Opt. Express 2010, 18, 24673–24678. 499

17. Bellouard, Y.; Said, A.A.; Bado, P. Integrating optics and micro-mechanics in a single substrate: a step toward monolithic 500

integration in fused silica. Opt. Express 2005, 13, 6635–6644. 501

18. Liao, Y.; Ni, J.; Qiao, L.; Huang, M.; Bellouard, Y.; Sugioka, K.; Cheng, Y. High-fidelity visualization of formation of volume 502

nanogratings in porous glass by femtosecond laser irradiation. Optica 2015, 2, 329–334. 503

19. Bellouard, Y.; Champion, A.; Lenssen, B.; Matteucci, M.; Schaap, A.; Beresna, M.; Corbari, C.; Gecevičius, M.; Kazansky, P.; 504

Chappuis, O.; et al. Integrating optics and micro-mechanics in a single substrate: a step toward monolithic integration in fused 505

silica. J. Laser Micro/Nanoengineering 2012, 7, 1–10. 506

20. Osellame, R.; Hoekstra, H.; Cerullo, G.; Pollnau, M. Femtosecond laser microstructuring: an enabling tool for optofluidic 507

lab-on-chips. Laser Photon. Rev. 2011, 5, 442–463. 508

21. Sugioka, K.; Cheng, Y. Ultrafast lasers-reliable tools for advanced materials processing. Light Sci. Appl. 2014, 3, e149. 509

22. Öktem, B.; Pavlov, I.; Ilday, S.; Kalaycioglu, H.; Rybak, A.; Yavas, S.; Erdogan, M.; Ilday, F. Nonlinear laser lithography for 510

indefinitely large-area nanostructuring with femtosecond pulses. Nature Photonics 2013, 7, 897–901. 511

23. Liu, H.; Lin, W.; Hong, M. Hybrid laser precision engineering of transparent hard materials: challenges, solutions and applications. 512

Light Sci. Appl. 2021, 10, 162. 513

24. Wang, H.; Zhang, W.; Ladika, D.; Yu, H.; Gailevičius, D.; Wang, H.; Pan, C.F.; Suseela Nair, P.; Ke, Y.; Mori, T.; et al. Two-Photon 514

Polymerization Lithography for Optics and Photonics: Fundamentals, Materials, Technologies, and Applications. Adv. Func. 515

Mat. 2023, 33, 2214211. 516

25. Gamaly, E.G.; Rode, A.V.; Luther-Davies, B.; Tikhonchuk, V.T. Ablation of solids by femtosecond lasers: Ablation mechanism and 517

ablation thresholds for metals and dielectrics. Physics of Plasmas 2002, 9, 949–957. 518

26. Momma, C.; Chichkov, B.N.; Nolte, S.; von Alvensleben, F.; Tünnermann, A.; Welling, H.; Wellegehausen, B. Short-pulse laser 519

ablation of solid targets. Optics Communications 1996, 129, 134–142. 520

27. Moughames, J.; Porte, X.; Thiel, M.; Ulliac, G.; Larger, L.; Jacquot, M.; Kadic, M.; Brunner, D. Three-dimensional waveguide 521

interconnects for scalable integration of photonic neural networks. Optica 2020, 7, 640–646. 522

28. Malinauskas, M.; Zukauskas, A.; Belazaras, K.; Tikuisis, K.; Purlys, V.; Gadonas, R.; Piskarskas, A. Laser fabrication of various 523

polymer micro-optical components. Eur. Phys. J-Appl. Phys. 2012, 58, 20501. 524

29. Schumann, M.; Bückmann, T.; Gruhler, N.; Wegener, M.; Pernice, W. Hybrid 2D–3D optical devices for integrated optics by direct 525

laser writing. Light Sci. Appl. 2014, 3. 526

30. Bertoncini, A.; Liberale, C. Polarization Micro-Optics: Circular Polarization From a Fresnel Rhomb 3D Printed on an Optical 527

Fiber. Photon. Technol. Lett. 2018, 30, 1882 – 1885. 528

31. Thiele, S.; Pruss, C.; Herkommer, A.; Giessen, H. 3D printed stacked diffractive microlenses. Opt. Express 2019, 27, 35621–35630. 529

32. Jia, Y.; Wang, S.; Chen, F. Femtosecond laser direct writing of flexibly configured waveguide geometries in optical crystals: 530

fabrication and application. Opto-Electronic Advances 2020, 3, 190042. 531

33. Li, L.; Kong, W.; Chen, F. Femtosecond laser-inscribed optical waveguides in dielectric crystals: a concise review and recent 532

advances. Advanced Photonics 2022, 4, 024002. 533

34. Sakakura, M.; Lei, Y.; Wang, L.; Yu, Y.; Kazansky, P. Ultralow-loss geometric phase and polarization shaping by ultrafast laser 534

writing in silica glass. Light Sci. Appl. 2020, 9, 15. 535

35. Zukauskas, A.; Malinauskas, M.; Brasselet, E. Monolithic generators of pseudo-nondiffracting optical vortex beams at the 536

microscale. Appl. Phys. Lett. 2013, 103, 181122. 537


Version February 21, 2024 submitted to Journal Not Specified 15 of 17

36. Gissibl, T.; Thiele, S.; Herkommer, A.; Giessen, H. Sub-micrometre accurate free-form optics by three-dimensional printing on 538

single-mode fibres. Nat. Commun. 2016, 7, 11763. 539

37. Dietrich, P.I.; Blaicher, M.; Reuter, I.; Billah, M.; Hoose, T.; Hofmann, A.; Caer, C.; Dangel, R.; Offrein, B.; Troppenz, U.; et al. In 540

situ 3D nanoprinting of free-form coupling elements for hybrid photonic integration. Nat. Photon. 2018, 12, 241–247. 541

38. Melissinaki, V.; Farsari, M.; Pissadakis, S. A Fiber-Endface, Fabry–Perot Vapor Microsensor Fabricated by Multiphoton Polymer- 542

ization. IEEE J. Select. Top. Quant. Electron. 2015, 21, 5600110. 543

39. Gonzalez-Hernandez, D.; Varapnickas, S.; Bertoncini, A.; Liberale, C.; Malinauskas, M. Micro-Optics 3D Printed via Multi-Photon 544

Laser Lithography. Advanced Optical Materials 2023, 11, 2201701. 545

40. Mainik, P.; Spiegel, C.A.; Blasco, E. Recent Advances in Multi-Photon 3D Laser Printing: Active Materials and Applications. 546

Advanced Materials, n/a, 2310100. 547

41. Somers, P.; AM unchinger.; Maruo, S.; Moser, C.; Xu, X.; Wegener, M. The physics of 3D printing with light. Nature Reviews 548

Physics 2023. 549

42. Gissibl, T aand Thiele, S.; Herkommer, A.; Giessen, H. Two-photon direct laser writing of ultracompact multi-lens objectives. Nat. 550

Photonics 2016, 10, 554. 551

43. Bauer, J.; Crook, C.; Baldacchini, T. A sinterless, low-temperature route to 3D print nanoscale optical-grade glass. Science 2023, 552

380, 960–966. 553

44. Laakso, M.; Huang, P.H.; Edinger, P.; Hartwig, O.; Duesberg, G.S.; Errando-Herranz, C.; Stemme, G.; Gylfason, K.B.; Niklaus, F. 554

Three-dimensional printing of silica-glass structures with submicrometric features, 2020, [arXiv:physics.app-ph/2005.06805]. 555

45. Huang, P.H.; Laakso, M.; Edinger, P.; Hartwig, O.; Duesberg, G.; Lai, L.L.; Mayer, J.; Nyman, J.; Errando-Herranz, C.; Stemme, G.; 556

et al. Three-dimensional printing of silica glass with sub-micrometer resolution. Nature Communications 2023, 14, 3305. 557

46. Jin, F.; Liu, J.; Zhao, Y.Y.; Dong, X.Z.; Zheng, M.L.; Duan, X.M. λ/30 inorganic features achieved by multi-photon 3D lithography. 558

Nature Communications 2022, 13, 3305. 559

47. Liu, T.; Tao, P.; Wang, X.; Wang, H.; He, M.; Wang, Q.; Cui, H.; Wang, J.; Tang, Y.; Tang, J.; et al. Ultrahigh-printing-speed 560

photoresists for additive manufacturing. Nature Nanotechnology 2023. 561

48. Kato, J.i.; Takeyasu, N.; Adachi, Y.; Sun, H.B.; Kawata, S. Multiple-spot parallel processing for laser micronanofabrication. 562

Applied Physics Letters 2005, 86, 044102. 563

49. Kiefer, P.; Hahn, V.; Kalt, S.; Sun, Q.; Eggeler, Y.; Wegener, M. A multi-photon-focus 3D laser printer based on a 3D-printed 564

diffractive optical element and a 3D-printed multi-lens array. Light: Advanced Manufacturing 2024, p. online. 565

50. Malinauskas, M.; Žukauskas, A.; Bičkauskaitė, G.; Gadonas, R.; Juodkazis, S. Mechanisms of three-dimensional structuring of 566

photo-polymers by tightly focussed femtosecond laser pulses. Opt. Express 2010, 18, 10209–10221. 567

51. Katsura, T. Thermal diffusivity of silica glass at pressures up to 9 GPa. Physics and Chemistry of Minerals 1993, 20, 201–208. 568

52. Uemukai, M.; Matsumoto, N.; Suhara, T.; Nishihara, H.; Eriksson, N.; Larsson, A. Monolithically integrated InGaAs-AlGaAs 569

master oscillator power amplifier with grating outcoupler. IEEE Photonics Technology Letters 1998, 10, 1097–1099. https: 570

//doi.org/10.1109/68.701514. 571

53. Uenishi, K.; Uemukai, M.; Suhara, T. Rotation-Symmetric Multispot Focusing Phase-Shifted Grating Coupler for Integrated 572

Semiconductor Laser. Japanese Journal of Applied Physics 2012, 51, 058001. https://doi.org/10.1143/JJAP.51.058001. 573

54. Gourley, K.; Golub, I.; Chebbi, B. Demonstration of a Fresnel axicon. IEEE Photonics Technology Letters 2011, 50, 303–306. 574

55. Ovsianikov, A.; Viertl, J.; Chichkov, B.; Oubaha, M.; MacCraith, B.; Sakellari, I.; Giakoumaki, A.; Gray, D.; Vamvakaki, M.; Farsari, 575

M.; et al. Ultra-Low Shrinkage Hybrid Photosensitive Material for Two-Photon Polymerization Microfabrication. ACS Nano 576

2008, 2, 2257–2262. 577

56. Chatzinikolaidou, M.; Rekstyte, S.; Danilevicius, P.; Pontikoglou, C.; Papadaki, H.; Farsari, M.; Vamvakaki, M. Mater. Sci. Eng. C 578

2015, 48, 301–309. 579

57. Maciulaitis, J.; Rekstyte, S.; Bratchikov, M.; Gudas, R.; Malinauskas, M.; Pockevicius, A.; Usas, A.; Rimkunas, A.; Jankauskaite, 580

V.; Grigaliunas, V.; et al. Full length article Customization of direct laser lithography-based 3D scaffolds for optimized in vivo 581

outcome. Appl. Surf. Sci. 2019, 487, 692–702. 582

58. Amato, L.; Gu, Y.; Bellini, N.; Eaton, S.; Cerullo, G.; Osellame, R. Integrated three-dimensional filter separates nanoscale from 583

microscale elements in a microfluidic chip. Lab Chip 2012, 12, 1135–1142. 584

59. Zukauskas, A.; Matulaitiene, I.; Paipulas, D.; Niaura, G.; Malinauskas, M.; Gadonas, R. Tuning the refractive index in 3D direct 585

laser writing lithography: towards GRIN microoptics. Laser Photon. Rev. 2015, 9, 706–712. 586

60. Skarmoutsou, A.; Lolas, G.; Charitidis, C.; Chatzinikolaidou, M.; Vamvakaki, M.; Farsari, M. Nanomechanical properties of 587

hybrid coatings for bone tissue engineering. J. Mechan. Behav. Biomed. Mater. 2013, 25, 48–62. 588

61. Gailevicius, D.; Padolskyte, V.; Mikoliunaite, L.; Sakirzanovas, S.; Juodkazis, S.; Malinauskas, M. Additive-Manufacturing of 3D 589

Glass-Ceramics down to Nanoscale Resolution. Nanoscale Horiz. 2019, 4, 647–651. 590

62. Jonusauskas, L.; Gailevicius, D.; Mikoliunaite, L.; Sakalauskas, D.; Sakirzanovas, S.; Juodkazis, S.; Malinauskas, M. Optically 591

Clear and Resilient Free-Form µ-Optics 3D-Printed via Ultrafast Laser Lithography. Materials 2017, 10, 12. 592

63. Sakellari, I.; Kabouraki, E.; Karanikolopoulos, D.; Droulias, S.; Farsari, M.; Loukakos, P.; Vamvakaki, M.; Gray, D. Quantum dot 593

based 3D printed woodpile photonic crystals tuned for the visible. Nanoscale Adv. 2019, 1, 3413–3423. 594

64. Giakoumaki, A.; Kenanakis, G.; Klini, A.; Androulidaki, M.; Viskadourakis, Z.; Farsari, M.; Selimis, A. 3D micro-structured arrays 595

of Zn0 nanorods. Sci. Rep. 2017, 7, 2100. 596


Version February 21, 2024 submitted to Journal Not Specified 16 of 17

65. Seniutinas, G.; Weber, A.; Padeste, C.; Sakellari, I.; Farsari, M.; David, C. Beyond 100 nm resolution in 3D laser lithography — 597

Post processing solutions. Microelectron. Eng. 2018, 191, 25–31. 598

66. Sänger, J.C.; Pauw, B.R.; Sturm, H.; Günster, J. First time additively manufactured advanced ceramics by using two-photon 599

polymerization for powder processing. Open Ceramics 2020, 4, 100040. https://doi.org/10.1016/j.oceram.2020.100040. 600

67. Gonzalez-Hernandez, D.; Sanchez-Padilla, B.; Gailevičius, D.; Thodika, S.C.; Juodkazis, S.; Brasselet, E.; Malinauskas, M. Single- 601

Step 3D Printing of Micro-Optics with Adjustable Refractive Index by Ultrafast Laser Nanolithography. Advanced Optical 602

Materials 2023, 11, 2300258. 603

68. Luther-Davies, B.; Rode, A.; Madsen, N.; Gamaly, E. Picosecond high-repetition-rate pulsed laser ablation of dielectrics: The 604

effect of energy accumulation between pulses. Opt. Eng. 2005, 44, 051102. 605

69. Rekštytė, S.; Jonavicius, T.; Gailevičius, D.; Malinauskas, M.; Mizeikis, V.; Gamaly, E.G.; Juodkazis, S. Nanoscale precision of 3D 606

polymerisation via polarisation control. Adv. Opt. Mat. 2016, 4, 1209 – 1214. 607

70. Ayub, S.; Gentet, L.; Fiáth, R.; Schwaerzle, M.; Borel, M.; David, F.; Barthó, P.; Ulbert, I.; Paul, O.; Ruther, P. Hybrid intracerebral 608

probe with integrated bare LED chips for optogenetic studies. Biomed. Microdev. 2017, 19, 1–12. 609

71. Keppeler, D.; Schwaerzle, M.; Harczos, T.; Jablonski, L.; Dieter, A.; Wolf, B.; Ayub, S.; Vogl, C.; Wrobel, C.; Hoch, G.; et al. 610

Multichannel optogenetic stimulation of the auditory pathway using microfabricated LED cochlear implants in rodents. Sci. 611

Transl. Med. 2020, 12, eabb8086. 612

72. Anand, V.; Maksimovic, J.; Katkus, T.; Ng, S.H.; Ulčinas, O.; Mikutis, M.; Baltrukonis, J.; Urbas, A.; Šlekys, G.; Ogura, H.; et al. All 613

femtosecond optical pump and x-ray probe: holey-axicon for free electron lasers. Journal of Physics: Photonics 2021, 3, 024002. 614

https://doi.org/10.1088/2515-7647/abd4ef. 615

73. Torcal-Milla, F.; Lobera, J.; Lopez, A.; Palero, V.; Andres, N.; Arroyo, M. Mach-Zehnder-based measurement of light emitting 616

diodes temporal coherence. Optik 2022, 267, 169722. 617

74. Wei, S.; Cao, G.; Lin, H.; Mu, H.; Liu, W.; Yuan, X.; Somekh, M.; Jia, B. High tolerance detour-phase graphene-oxide flat lens. 618

Photonics Research 2021, 9, 2454–2463. 619

75. Dharmavarapu, R.; Bhattacharya, S.; Juodkazis, S. Diffractive optics for axial intensity shaping of Bessel beams. Journal of Optics 620

2018, 20, 085606. 621

76. Milster, T.D.; Kim, Y.S.; Wang, Z.; Purvin, K. Multiple-order diffractive engineered surface lenses. Appl. Opt. 2020, 59, 7900–7906. 622

https://doi.org/10.1364/AO.394124. 623

77. Apai, D.; Milster, T.D.; Kim, D.W.; Bixel, A.; Schneider, G.; Liang, R.; Arenberg, J. A Thousand Earths: A Very Large Aperture, 624

Ultralight Space Telescope Array for Atmospheric Biosignature Surveys. The Astronomical Journal 2019, 158, 83. https: 625

//doi.org/10.3847/1538-3881/ab2631. 626

78. Milster, T.D.; Wang, Z.; Kim, Y.S. Design aspects of large-aperture MODE lenses. OSA Continuum 2021, 4, 171–181. https: 627

//doi.org/10.1364/OSAC.410187. 628

79. Yoshida, M.; Katsuno, S.; Inoue, T.; Gelleta, J.; Izumi, K.; De Zoysa, M.; Ishizaki, K.; Noda, S. High-brightness scalable 629

continuous-wave single-mode photonic-crystal laser. Nature 2023, 618, 727–732. 630

80. Samsonas, D.; Skliutas, E.; Ciburys, A.; Kontenis, L.; Gailevičius, D.; Berzinš, J.; Narbutis, D.; Jukna, V.; Vengris, M.; Juodkazis, S.; 631

et al. 3D nanopolymerization and damage threshold dependence on laser wavelength and pulse duration. Nanophotonics 2023. 632

https://doi.org/10.1515/nanoph-2022-0629. 633

81. Li, Z.Z.; Fan, H.; Wang, L.; Zhang, X.; Zhao, X.J.; Yu, Y.H.; Xu, Y.S.; Wang, Y.; Wang, X.J.; Juodkazis, S.; et al. Super stealth dicing 634

of transparent solids with nanometric precision. arxiv 2023, p. arXiv:2308.02352v2. 635

82. Gailevičius, D.; Ryu, M.; Honda, R.; Lundgaard, S.; Suzuki, T.; Maksimovic, J.; Hu, J.; Linklater, D.P.; Ivanova, E.P.; Katkus, T.; 636

et al. Tilted black-Si: ∼0.45 form-birefringence from sub-wavelength needles. Opt. Express 2020, 28, 16012–16026. 637

83. Richardson, R.; Ibbotson, M.; Thompson, A.; Wise, A.; Fallon, J. Optical stimulation of neural tissue. Healthc. Technol. Lett. 2020, 638

7, 58–65. 639

84. Shannon, R. Multichannel electrical stimulation of the auditory nerve in man. II. Channel interaction. Hear. Res. 1983, 12, 1–16. 640

85. White MW, Merzenich MM, G.J. Multichannel cochlear implants. Channel interactions and processor design. Arch. Otolaryngol. 641

1984, 110, 493–501. 642

86. Friesen, L.; Shannon, R.; Baskent, D.; Wang, X. Speech recognition in noise as a function of the number of spectral channels: 643

comparison of acoustic hearing and cochlear implants. J. Acoust. Soc. Am. 2001, 110, 1150–63. 644

87. Wilson, B.; Finley, C.; Lawson, D.; Wolford, R.; Eddington, D.; Rabinowitz, W. Better speech recognition with cochlear implants. 645

Nature 1991, 352, 236–8. 646

88. Berenstein, C.; Mens, L.; Mulder, J.; Vanpoucke, F. Current steering and current focusing in cochlear implants: comparison of 647

monopolar, tripolar, and virtual channel electrode configurations. Ear Hear. 2008, 29, 250–60. 648

89. Bierer, J.; Litvak, L. Reducing channel interaction through cochlear implant programming may improve speech perception: 649

Current focusing and channel deactivation. Trends Hear. 2016, 20. 650

90. Dieter, A.; Duque-Afonso, C.; Rankovic, V.; Jeschke, M.; Moser, T. Near physiological spectral selectivity of cochlear optogenetics. 651

Nature Comm. 2019, 10, 1962. 652

91. Dieter, A.; Klein, E.; Keppeler, D.; Jablonski, L.; Harczos, T.; Hoch, G.; Rankovic, V.; Paul, O.; Jeschke, M.; Ruther, P.; et al. 653

µ-LED-based optical cochlear implants for spectrally selective activation of the auditory nerve. EMBO Mol. Med. 2020, p. e12387. 654
Version February 21, 2024 submitted to Journal Not Specified 17 of 17

92. Azees, A.; Thompson, A.; Thomas, R.; Zhou, J.; Ruther, P.; Wise, A.; Ajay, E.; Garrett, D.; Quigley, A.; Fallon, J.; et al. Spread of 655

activation and interaction between channels with multichannel optogenetic stimulation in the mouse cochlea. Hear. Res. 2023, 656

pp. 58–65. 657

93. Keppeler, D.; Kampshoff, C.; Thirumalai, A.; Duque-Afonso, C.; Schaeper, J.; Quilitz, T.; Topperwien, M.; Vogl, C.; Hessler, R.; 658

Meyer, A.; et al. Multiscale photonic imaging of the native and implanted cochlea. Proc. Natl. Acad. Sci. USA 2021, 118. 659

94. Jurgens, T.; Hohmann, V.; Buchner, A.; Nogueira, W. The effects of electrical field spatial spread and some cognitive factors on 660

speech-in-noise performance of individual cochlear implant users-A computer model study. PLoS One 2018, 13, e0193842. 661

95. Khurana, L.; Jablonski, D.K.L.; Moser, T. Model-based prediction of optogenetic sound encoding in the human cochlea by future 662

optical cochlear implants. Comput. Struct. Biotechnol. J. 2022, 20, 3621–9. 663

96. Dharmavarapu, R.; Ng, S.H.; Eftekhari, F.; Juodkazis, S.; Bhattacharya, S. MetaOptics: opensource software for designing 664

metasurface optical element GDSII layouts. Opt. Express 2020, 28, 3505–3516. 665

97. Dharmavarapu, R.; Izumi, K.; Katayama, I.; Ng, S.H.; Vongsvivut, J.; Tobin, M.J.; Kuchmizhak, A.; Nishijima, Y.; Bhattacharya, S.; 666

Juodkazis, S. Dielectric cross-shaped-resonator-based metasurface for vortex beam generation at mid-IR and THz wavelengths. 667

Nanophotonics 2019, 8, 1263–1270. 668

Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual 669

author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to 670

people or property resulting from any ideas, methods, instructions or products referred to in the content. 671

You might also like