We demonstrate direct printing from donor ink containing ferrite nanoparticles by employing laser induced forward transfer with an optical vortex possessing orbital angular momentum (OAM). We show, for the first time, that the as-printed dots are twisted and exhibit spinel Fe3O4 monocrystalline properties without the need for a sintering process. The helicity of the as-printed dots is shown to be selectively controlled merely by reversing the handedness of optical vortices. The diameter of the printed dots was typically measured to be less than 1/10th of the irradiated laser spot (diffraction limit). These results imply that the optical vortex twists and confines the sintered nanoparticles within its dark core to form chiral spinel monocrystalline dots. The observation of mono-crystallization with optical vortex induced forward transfer will offer new fundamental physics such as OAM light–matter interactions and could pave the way toward advanced printable magnetic devices, such as high-density magnetic data storage.

Multicomponent oxide systems have been intensely investigated because of their scientific and practical importance.1,2 Iron-based multicomponent oxide systems, referred to as ferrites, are particularly attractive as they possess excellent magnetic properties and exhibit multiple functional crystal forms, including spinel, garnet, orthoferrite, and hexaferrite.3,4

Nozzle-based ink-jet printing has been well established as a non-contact deposition technology for materials. The technique enables the on-demand development of electronics, photonics, and biomedical devices with cost and energy savings and reduced material waste.5–7 However, despite its broad applicability, it still struggles to directly print highly viscous or highly dense materials with a high spatial resolution owing to nozzle clogging effects. In addition, it frequently requires additional sintering steps to develop practical devices.

Laser induced forward transfer (LIFT) has been widely studied as an alternative approach to ink-jet printing. The technique enables the selective direct deposition of a myriad of materials (herein referred to as donors), such as solids, liquids, and even high viscosity pastes, onto a receiver substrate simply by irradiation with a laser pulse without nozzles.8–13 However, LIFT also struggles to directly print donors onto substrates with high spatial resolution at millimeter-scale working distances. Until now, there have been few reports on the direct printing of high viscous multicomponent oxide systems with non-contact deposition technologies (even including LIFT), and the direct printing of micron-scale monocrystalline multicomponent oxide systems has never been reported.

An optical vortex possesses unique features, such as a ring-shaped spatial form and an orbital angular momentum (OAM), characterized by a topological charge ℓ, associated with its helical wavefront with an on-axis phase-singularity.14–17 It has been intensely studied in a variety of research fields, including nano/micron-scale optical trapping and manipulation,18 optical/quantum space division multiplexing communications,19 and super-resolution microscopes with high spatial resolution beyond the diffraction limit.20 In addition, it has been discovered that the OAM of the optical vortex twists the irradiated materials, and it allows for the fabrication of chiral structures with the aid of a spin angular momentum (SAM, s = ±1) owing to the helical electric field of circularly polarized light.21–23 

In recent years, us and our co-workers have successfully proposed a new LIFT technology with an optical vortex pulse possessing a helical wavefront (instead of a conventional Gaussian beam with a planar wavefront), in which the OAM of the optical vortex pulse twists the irradiated donor material even if it has moderate-to-high viscosity (>1 Pa·s), thus resulting in the ejection of a spinning donor droplet with a perfectly straight flight path, somewhat akin to a rifling bullet.24 

This technique is herein referred to as optical vortex laser-induced forward transfer (OV-LIFT), and it enables the direct printing of microdots of a variety of donor materials with a high spatial resolution at an extremely long working distance.25–27 In addition, OV-LIFT is known to produce as-printed dots with excellent physical properties, such as high electric conductivity, without any additional sintering processes. Thus, the OV-LIFT technique has great potential to enable the development of next generation printed devices with time and cost savings.

In this paper, we demonstrate the direct printing of a ferrite nanoparticle ink with OV-LIFT. Surprisingly, we discover that the OV-LIFT produces twisted printed dots possessing spinel Fe3O4 monocrystalline properties without any sintering processes. In addition, we find that the helicity of printed dots can be selectively controlled merely by reversing the handedness of the irradiated optical vortex pulse. Furthermore, the diameter of printed dots is observed to be inversely proportional to the numerical aperture (NA) of a focusing lens, and its minimum value was measured to be 1–2 µm, corresponding to that of less than 1/10th of the irradiated laser spot (diffraction limit). These considerations suggest that the optical vortex twists and confines the sintered nanoparticles within its dark core to form chiral spinel monocrystalline printed dots.

The donor used in the experiment was formed of magnetic nanoparticles [∼1.0 wt. % (FeO, MnO) · Fe2O3, 100–300 nm] in suspension with a 57:43 mixture of purified water and glycerin (∼6 mPa s). Figure 1(a) shows a scanning electron microscope (SEM) image of magnetic nanoparticles. The donor solution was dropped onto a glass substrate to form a thin film with a thickness of ∼26 µm (an optical density of ∼0.3 at 532 nm) and a roughness of ∼6 µm.

FIG. 1.

(a) A SEM image of magnetic nanoparticles. (b) Experimental setup for OV-LIFT. A green laser with a wavelength of 532 nm and a pulse duration of 10 ns was converted into a circularly polarized optical vortex by a spatial light modulator and a quarter-wave plate, and it was focused and scanned onto a donor film by a focusing lens and a Galvano scanner.

FIG. 1.

(a) A SEM image of magnetic nanoparticles. (b) Experimental setup for OV-LIFT. A green laser with a wavelength of 532 nm and a pulse duration of 10 ns was converted into a circularly polarized optical vortex by a spatial light modulator and a quarter-wave plate, and it was focused and scanned onto a donor film by a focusing lens and a Galvano scanner.

Close modal

Figure 1(b) shows a schematic diagram of the experimental setup for OV-LIFT used in this study. A Q-switched nanosecond green laser with a wavelength of 532 nm and a pulse duration of 10 ns was used as a light source, and its output was converted into a circularly polarized optical vortex pulse [J = ±2 (ℓ = ±1, s = ±1)] by employing a spatial light moderator and a quarter-wave plate. The generated optical vortex was loosely focused to be an annular spot with a diameter of 75 µm onto the donor film by using a lens with a focal length of 300 mm. This was two-dimensionally scanned by using a galvanometer scanner. The fluence of the irradiated laser beam was then fixed to be ∼0.28 J/cm2. The distance between the donor film and the receiver substrate was controlled to be 0.6–1.5 mm.

With this system, a single droplet was launched and printed onto the receiver substrate. The temporal dynamics of the droplet ejection were observed from the side by employing an ultrahigh-speed camera with a frame rate of 2 × 106 frames/s. The printed dots were then annealed at ∼300 °C to remove the solvent.

A single optical vortex pulse created a blister near its central dark core (within ∼3 µs). Subsequently, a rotating jet was launched to eject a single droplet (a diameter of ∼11 µm) owing to the Plateau–Rayleigh instability at ∼30 µs after the pulse deposition [Fig. 2(a)]. Interestingly, a twisted gray core (herein referred to as a “magnetic core”) with a diameter of 7 µm (one-tenth of that of the irradiated laser spot) was deposited at the center of a printed dot of 20 µm in diameter (Fig. 3). The core diameter was then the same at any working distance (0.6–1.5 mm). Note that the twisted direction of the magnetic core was reversed by inverting the handedness of the optical vortex (OAM) and circular polarization (SAM). Conversely, an equivalent Gaussian pulse produced simultaneously multiple droplets, resulting in a non-uniform dot printed with more debris and without any core on the receiver [Figs. 2(b) and 3(b)]. The OV-LIFT includes many steps: laser-induced cavitation, OAM transfer effect, and cavitation pressure induced mass transport and ejection of spinning droplets through Plateau–Rayleigh instability (microsecond-scale hydro-fluid dynamics). Such microsecond-scale hydro-fluid dynamics should play an important role in the phenomenon, as shown in Fig. 2.

FIG. 2.

Temporal evolution of the ejection of the donor droplet. (a) OV-LIFT produces a spinning jet, thus resulting in the ejection of a single droplet via Plateau–Rayleigh instability. (b) Gaussian beam irradiation launches a non-spinning jet, and it induces the ejection of multiple droplets.

FIG. 2.

Temporal evolution of the ejection of the donor droplet. (a) OV-LIFT produces a spinning jet, thus resulting in the ejection of a single droplet via Plateau–Rayleigh instability. (b) Gaussian beam irradiation launches a non-spinning jet, and it induces the ejection of multiple droplets.

Close modal
FIG. 3.

(a) Printed dot with a magnetic core formed on a glass receiver substrate by employing OV-LIFT. (b) Printed dot by employing LIFT with a conventional Gaussian beam. (c) and (d) SEM images of the dot and its cross section printed by the irradiation of an optical vortex with J = 2 (ℓ = +1, s = +1). The white arrow shows the twisting direction of the magnetic core. (e) and (f) SEM images of the dot and its cross section printed by the irradiation of an optical vortex with J = −2 (ℓ = −1, s = −1).

FIG. 3.

(a) Printed dot with a magnetic core formed on a glass receiver substrate by employing OV-LIFT. (b) Printed dot by employing LIFT with a conventional Gaussian beam. (c) and (d) SEM images of the dot and its cross section printed by the irradiation of an optical vortex with J = 2 (ℓ = +1, s = +1). The white arrow shows the twisting direction of the magnetic core. (e) and (f) SEM images of the dot and its cross section printed by the irradiation of an optical vortex with J = −2 (ℓ = −1, s = −1).

Close modal

The diameters of magnetic cores and printed dots were approximately inversely proportional to the numerical aperture (NA) (Fig. 4). The minimum diameters of the magnetic core and printed dot were measured to be 2.6 and 10.7 µm at NA = 0.04. It is worth noting that the height of the magnetic core was typically in the range of 0.6–2 µm, and its minimum value was 0.6 µm at NA of 0.04. These results show that OV-LIFT can sinter and twist ferrite nanoparticle ink to form chiral magnetic cores owing to helicity transfer from the irradiated optical vortex pulse.

FIG. 4.

Diameter of the magnetic cores vs NA of the focusing lens. The broken line shows inverse linear proportionality.

FIG. 4.

Diameter of the magnetic cores vs NA of the focusing lens. The broken line shows inverse linear proportionality.

Close modal

The fabricated magnetic core was sliced into a 100 nm thick film by employing a focused ion beam milling method, and its physical properties were observed using a transmission electron microscope (TEM) (Fig. 5). Note that a carbon-coated layer was deposited on the magnetic core to prevent surface damage. Surprisingly, an ultrafine, hexagonal electron diffraction pattern, originating from the (122) face, was observed in more than 90% (surrounded by the red dashed line) of the entire core film, thus demonstrating that the magnetic core exhibits a well-formed octahedral single-crystal with a cubic spinel Fe3O4 structure (Fig. 5, Table I). The experimental spot intervals and lattice angles in the electron diffraction pattern of the film were fully identical to those of spinel-type crystals simulated by the open software ReciPro.28 The magnetic core (surrounded by the red dashed line) was formed of a single domain without any grain boundaries.

FIG. 5.

(a)–(c) Focused ion beam processing to fabricate a thin magnetic core flake. (a) Magnetic core, (b) its cross-section, and (c) its flake with a thickness of about 100 nm. Note that a carbon layer was coated on the sample to prevent flake damage. (d) Bright field image of the magnetic core observed by TEM. The core is formed of two domains surrounded by red (P1) and blue (P2) broken lines. (e) Electron diffraction pattern from the A (red) domain. This is identical to that of (f) the (122) cubic spinel Fe3O4 structure simulated by ReciPro software, thus demonstrating that the magnetic core is a well-formed octahedral single crystal. (g) Electron diffraction pattern from the B (blue) domain. The simulation is fully identical to that of the (h) (100) cubic spinel Fe3O4 structure simulated by ReciPro software. Note that the X- and Y-axes are tilted by 6.7° and 2.3°, respectively.

FIG. 5.

(a)–(c) Focused ion beam processing to fabricate a thin magnetic core flake. (a) Magnetic core, (b) its cross-section, and (c) its flake with a thickness of about 100 nm. Note that a carbon layer was coated on the sample to prevent flake damage. (d) Bright field image of the magnetic core observed by TEM. The core is formed of two domains surrounded by red (P1) and blue (P2) broken lines. (e) Electron diffraction pattern from the A (red) domain. This is identical to that of (f) the (122) cubic spinel Fe3O4 structure simulated by ReciPro software, thus demonstrating that the magnetic core is a well-formed octahedral single crystal. (g) Electron diffraction pattern from the B (blue) domain. The simulation is fully identical to that of the (h) (100) cubic spinel Fe3O4 structure simulated by ReciPro software. Note that the X- and Y-axes are tilted by 6.7° and 2.3°, respectively.

Close modal
TABLE I.

Experimental and simulated lattice constants in the red (P1) and blue (P2) domains. All lattice constants are normalized by the lattice constant ab. ∠abc (angle between crystal axes ab and bc) and ∠abd (angle between crystal axes ab and bd) indicate the inter-axis angles. These results show that there is good agreement between the experiments and the simulations.

P1P2
TEMSimulationTEMSimulation
DiffractionMiller index(122)DiffractionMiller index(122)
ac/ab 0.63 0.63 0.69 0.71 
ad/ab 0.60 0.63 1.00 1.00 
ae/ab 0.40 0.47 0.73 0.71 
abc 68.6° 69.7° 90.9° 90.8° 
abd 24.4° 25.0° 45.9° 45.4° 
P1P2
TEMSimulationTEMSimulation
DiffractionMiller index(122)DiffractionMiller index(122)
ac/ab 0.63 0.63 0.69 0.71 
ad/ab 0.60 0.63 1.00 1.00 
ae/ab 0.40 0.47 0.73 0.71 
abc 68.6° 69.7° 90.9° 90.8° 
abd 24.4° 25.0° 45.9° 45.4° 

These results suggest that optical vortex beams enable the close-packing and sintering effects of magnetic nanoparticles to form the monocrystalline magnetic core via the OV-LIFT process. Assuming that the latent heat for the phase transition of ferrites is negligible, the single optical vortex pulse allows for a temperature rise of the donor by ∼2400 K (the melting point of ferrites is ∼850 °C). The optical density and the specific heat of the donor film were then assumed to be 0.4 and 670 J/kg·K,29 respectively. In addition, the irradiated laser fluence was 0.28 J/cm2. Such a high temperature rise should be sufficient to sinter ferrite nanoparticles.

The temporal evolution of internal pressure in a cavitation bubble was analyzed by employing the Rayleigh–Plesset equation,
where PBt and Pt are the pressures within the bubble and at infinity outside the bubble, ρL is the density of the solvent, Rt is the bubble radius, νL is the kinematic viscosity of the solvent, and S is the surface tension. It is worth mentioning that Rayleigh–Plesset analysis assumes that PBt is uniform in the bubble. Under the assumption that the spatial form of the blister reflects the shape of the cavitation bubble, Rt was then estimated by fitting a blister with a hemisphere function [Fig. 6(a)]. The optical vortex enables the production of a negative (inner compression) cavitation pressure (it reached −0.5 MPa at ∼0.8 µs after the laser irradiation) in the donor blister through the rapid shrinkage of the cavitation bubble (the twisted jet formation then occurs at ∼1.8 µs after the laser irradiation),25,26,30 thus resulting in the aggregation and crystallization of ferrite nanoparticles (Fig. 6). In fact, the synthesis of high-density, single-layer ferrite submicron particles (spray pyrolysis method31) has been demonstrated under the nitrogen atmosphere with ∼0.3 MPa and a temperature above 900 °C. Thus, laser-induced sub-MPa level compression pressure and ∼2000 K level temperature rise in our experiment will allow the production of the single-crystalline magnetic core. In future work, the direct observation of the cavitation pressure and temperature rise in the bubble will be needed to fully understand the mechanism of this crystallization.
FIG. 6.

(a) Time evolution of the cavitation bubble after the laser irradiation. (b) Time evolution of cavitation pressure estimated using the Rayleigh–Plesset equation.

FIG. 6.

(a) Time evolution of the cavitation bubble after the laser irradiation. (b) Time evolution of cavitation pressure estimated using the Rayleigh–Plesset equation.

Close modal

Furthermore, a 2-dimensional array of magnetic cores was fabricated by employing OV-LIFT in combination with a galvanometer scanner. The optical vortex pulse was then focused to be a 34 µm spot on the film, and the working distance between the donor and receiver was also fixed to be 0.4 mm. Note that the as-printed dots were annealed at ∼300 °C below the melting point (∼850 °C) of ferrite to remove the solvent (Fig. 7). The magnetic cores were fabricated at a positional error of ∼5.6 µm (4.9 and 1.8 µm along the x and y axes, respectively). This value was limited by the nonuniform dispersion (nonuniform absorption of the film) of ferrite nanoparticles in the solvent. Further improvement of positional accuracy will be possible with the refinement of the fabrication technique for the donor film.

FIG. 7.

Two-dimensional magnetic core array with 100 µm intervals. The NA of the focusing lens was 0.04. The working distance between the donor and receiver substrates was 0.4 mm.

FIG. 7.

Two-dimensional magnetic core array with 100 µm intervals. The NA of the focusing lens was 0.04. The working distance between the donor and receiver substrates was 0.4 mm.

Close modal

We have discovered, for the first time, that OV-LIFT enables the direct printing of monocrystalline magnetic cores with a spinel structure. The effect occurs via cavitation dynamics, in which ferrite nanoparticles are aggregated and crystallized by optical vortex induced negative cavitation pressure. In addition, we have discovered that the magnetic cores are twisted by OAM transfer effects. The magnetic core diameter was measured to be typically one-tenth of that of the irradiated laser spot, even at larger and more practical working distances (>1 mm). The minimum diameter was then measured to be 2.6 µm at NA = 0.04.

Furthermore, a 2-dimensional array of magnetic cores can be fabricated simply by employing a galvanometer scanner. The positional accuracy of the printed cores was limited by the nonuniform dispersion (nonuniform absorption of the film) of ferrite nanoparticles in the solvent. An improvement in positional accuracy will be possible with the refinement of the fabrication technique for the donor film. The nanosecond OV-LIFT has been demonstrated in our current study. Femto/pico-second OV-LIFT via two photon absorption allows the spatially localized cavitation of donors, and it will further enable the super-resolution direct print of monocrystalline magnetic cores.

Such a crystallization phenomenon via OV-LIFT could be potentially extended to directly print and pattern a variety of crystals, including dielectrics, semiconductors, and even biomaterials, beyond the limits of conventional pulsed laser ablation32,33 and pulsed laser melting techniques in liquid.34–36 The resulting structures offer opportunities for new fundamental physics for the study of light–matter interactions based on the interaction of OAM with light.

The authors acknowledge the support in the form of Kakenhi Grants-in-Aid (Grant Nos. JP16H06507, JP17K19070, JP18H03884, JP18H05242, JP22H05131, JP22H05138, JP22K18981, and JP23H00270) from the Japan Society for the Promotion of Science (JSPS) and from the Core Research for Evolutional Science and Technology program (Grant No. JPMJCR1903) of the Japan Science and Technology Agency (JST).

The authors have no conflicts to disclose.

Akihiko Kaneko: Conceptualization (equal); Data curation (lead); Formal analysis (lead); Funding acquisition (equal); Investigation (equal); Methodology (equal); Project administration (equal); Software (supporting); Supervision (equal); Validation (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal). Muneaki Iwata: Conceptualization (equal); Data curation (equal); Formal analysis (equal); Funding acquisition (equal); Investigation (equal); Methodology (equal); Project administration (equal); Software (lead); Supervision (equal); Validation (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal). Rong Wei: Data curation (equal); Software (equal); Visualization (equal). Ken-ichi Yuyama: Conceptualization (equal); Data curation (lead); Formal analysis (equal); Funding acquisition (equal); Investigation (equal); Methodology (equal); Project administration (equal); Software (supporting); Supervision (equal); Validation (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal). Takashige Omatsu: Conceptualization (equal); Data curation (equal); Formal analysis (equal); Funding acquisition (equal); Investigation (equal); Methodology (equal); Project administration (equal); Software (lead); Supervision (equal); Validation (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal).

The data that support the findings of this study are available from the corresponding authors upon reasonable request.

1.
A. G.
Díez
,
M.
Rincón-Iglesias
,
S.
Lanceros-Méndez
,
J.
Reguera
, and
E.
Lizundia
, “
Multicomponent magnetic nanoparticle engineering: The role of structure-property relationship in advanced applications
,”
Mater. Today Chem.
26
,
101220
(
2022
).
2.
S.-G.
Cho
,
K.-W.
Jeon
,
J.
Kim
,
K. H.
Kim
, and
J.
Kim
, “
Synthesis and ferromagnetic properties of magnetic ink for direct printing
,”
IEEE Trans. Magn.
47
(
10
),
3157
3159
(
2011
).
3.
O.
Dehghani Dastjerdi
,
H.
Shokrollahi
, and
S.
Mirshekari
, “
A review of synthesis, characterization, and magnetic properties of soft spinel ferrites
,”
Inorg. Chem. Commun.
153
,
110797
(
2023
).
4.
M.
Sugimoto
, “
The past, present, and future of ferrites
,”
J. Am. Ceram. Soc.
82
(
2
),
269
280
(
1999
).
5.
J. R. C.
Pita
,
W. R. S.
Baxter
,
J.
Morgan
,
S.
Temple
,
G. D.
Martin
, and
I. M.
Hutchings
, “
Future, opportunities and challenges of inkjet technologies
,”
Atomization Spray
23
,
541
565
(
2013
).
6.
M. A.
Shah
,
D.-G.
Lee
,
B.-Y.
Lee
, and
S.
Hur
, “
Classifications and applications of inkjet printing technology: A review
,”
IEEE Access
9
,
140079
140102
(
2021
).
7.
H. S.
Kim
,
J. S.
Kang
,
J. S.
Park
,
H. T.
Hahn
,
H. C.
Jung
, and
J. W.
Joung
, “
Inkjet printed electronics for multifunctional composite structure
,”
Compos. Sci. Technol.
69
(
7–8
),
1256
1264
(
2009
).
8.
R. D.
Boehm
,
P. R.
Miller
,
J.
Daniels
,
S.
Stafslien
, and
R. J.
Narayan
, “
Inkjet printing for pharmaceutical applications
,”
Mater. Today
17
(
5
),
247
252
(
2014
).
9.
P.
Sopeña
,
J.
Arrese
,
S.
González-Torres
,
J. M.
Fernández-Pradas
,
A.
Cirera
, and
P.
Serra
, “
Low-cost fabrication of printed electronics devices through continuous wave laser-induced forward transfer
,”
ACS Appl. Mater. Interfaces
9
(
35
),
29412
29417
(
2017
).
10.
J. M.
Fernández-Pradas
,
P.
Sopeña
,
S.
González-Torres
,
J.
Arrese
,
A.
Cirera
, and
P.
Serra
, “
Laser-induced forward transfer for printed electronics applications
,”
Appl. Phys. A
124
(
2
),
214
(
2018
).
11.
E. C. P.
Smits
,
A.
Walter
,
D. M.
de Leeuw
, and
K.
Asadi
, “
Laser induced forward transfer of graphene
,”
Appl. Phys. Lett.
111
(
17
),
173101
(
2017
).
12.
M.
Makrygianni
,
I.
Kalpyris
,
C.
Boutopoulos
, and
I.
Zergioti
, “
Laser induced forward transfer of Ag nanoparticles ink deposition and characterization
,”
Appl. Surf. Sci.
297
,
40
44
(
2014
).
13.
A.
Karaiskou
,
I.
Zergioti
,
C.
Fotakis
,
M.
Kapsetaki
, and
D.
Kafetzopoulos
, “
Microfabrication of biomaterials by the sub-ps laser-induced forward transfer process
,”
Appl. Surf. Sci.
208–209
,
245
249
(
2003
).
14.
R. A.
Beth
, “
Mechanical detection and measurement of the angular momentum of light
,”
Phys. Rev.
50
(
2
),
115
125
(
1936
).
15.
J.
Leach
,
M. J.
Padgett
,
S. M.
Barnett
,
S.
Franke-Arnold
, and
J.
Courtial
, “
Measuring the orbital angular momentum of a single photon
,”
Phys. Rev. Lett.
88
(
25
),
257901
(
2002
).
16.
M.
Padgett
,
J.
Courtial
, and
L.
Allen
, “
Light’s orbital angular momentum
,”
Phys. Today
57
(
5
),
35
40
(
2004
).
17.
S. M.
Barnett
,
L.
Allen
,
R. P.
Cameron
,
C. R.
Gilson
,
M. J.
Padgett
,
F. C.
Speirits
, and
A. M.
Yao
, “
On the natures of the spin and orbital parts of optical angular momentum
,”
J. Opt.
18
(
6
),
064004
(
2016
).
18.
M.
Dienerowitz
,
M.
Mazilu
,
P. J.
Reece
,
T. F.
Krauss
, and
K.
Dholakia
, “
Optical vortex trap for resonant confinement of metal nanoparticles
,”
Opt. Express
16
(
7
),
4991
4999
(
2008
).
19.
W.
Shao
,
S.
Huang
,
X.
Liu
, and
M.
Chen
, “
Free-space optical communication with perfect optical vortex beams multiplexing
,”
Opt. Commun.
427
(
15
),
545
550
(
2018
).
20.
S.
Bretschneider
,
C.
Eggeling
, and
S. W.
Hell
, “
Breaking the diffraction barrier in fluorescence microscopy by optical shelving
,”
Phys. Rev. Lett.
98
(
21
),
218103
(
2007
).
21.
T.
Omatsu
,
K.
Chujo
,
K.
Miyamoto
,
M.
Okida
,
K.
Nakamura
,
N.
Aoki
, and
R.
Morita
, “
Metal microneedle fabrication using twisted light with spin
,”
Opt. Express
18
(
17
),
17967
17973
(
2010
).
22.
K.
Toyoda
,
K.
Miyamoto
,
N.
Aoki
,
R.
Morita
, and
T.
Omatsu
, “
Using optical vortex to control the chirality of twisted metal nanostructures
,”
Nano Lett.
12
(
7
),
3645
3649
(
2012
).
23.
K.
Toyoda
,
F.
Takahashi
,
S.
Takizawa
,
Y.
Tokizane
,
K.
Miyamoto
,
R.
Morita
, and
T.
Omatsu
, “
Transfer of light helicity to nanostructures
,”
Phys. Rev. Lett.
110
(
14
),
143603
(
2013
).
24.
R.
Nakamura
,
H.
Kawaguchi
,
M.
Iwata
,
A.
Kaneko
,
R.
Nagura
,
S.
Kawano
,
K.
Toyoda
,
K.
Miyamoto
, and
T.
Omatsu
, “
Optical vortex-induced forward mass transfer: Manifestation of helical trajectory of optical vortex
,”
Opt. Express
27
(
26
),
38019
38027
(
2019
).
25.
H.
Kawaguchi
,
K.
Umesato
,
K.
Takahashi
,
K.
Yamane
,
R.
Morita
,
K.
Yuyama
,
S.
Kawano
,
K.
Miyamoto
,
M.
Kohri
, and
T.
Omatsu
, “
Generation of hexagonal close-packed ring-shaped structures using an optical vortex
,”
Nanophotonics
11
(
4
),
855
864
(
2022
).
26.
T.
Omatsu
,
K.
Miyamoto
,
K.-I.
Yuyama
,
K.
Yamane
, and
R.
Morita
, “
Laser-induced forward-transfer with light possessing orbital angular momentum
,”
J. Photochem. Photobiol., C
52
,
100535
(
2022
).
27.
K.
Yuyama
,
H.
Kawaguchi
,
R.
Wei
, and
T.
Omatsu
, “
Fabrication of an array of hemispherical microlasers using optical vortex laser-induced forward transfer
,”
ACS Photonics
10
,
4045
4051
(
2023
).
28.
Y.
Seto
and
M.
Ohtsuka
, “
ReciPro: Free and open-source multipurpose crystallographic software integrating a crystal model database and viewer, diffraction and microscopy simulators, and diffraction data analysis tools
,”
J. Appl. Crystallogr.
55
(
2
),
397
410
(
2022
).
29.
F.
Kuwahara
, “
Nano fluids and convective heat transfer
,”
J. Heat Transfer Soc. Jpn.
58
(
242
),
2
8
(
2019
).
30.
D.
Zhang
,
B.
Gökce
, and
S.
Barcikowski
, “
Laser synthesis and processing of colloids: Fundamentals and applications
,”
Chem. Rev.
117
(
5
),
3990
4103
(
2017
).
31.
Q.
Li
,
C. M.
Sorensen
,
K. J.
Klabunde
, and
G. C.
Hadjipanayis
, “
Aerosol spray pyrolysis synthesis of magnetic manganese ferrite particles
,”
Aerosol Sci. Technol.
19
(
4
),
453
467
(
1993
).
32.
J.
Xiao
,
P.
Liu
,
C. X.
Wang
, and
G. W.
Yang
, “
External field-assisted laser ablation in liquid: An efficient strategy for nanocrystal synthesis and nanostructure assembly
,”
Prog. Mater. Sci.
87
,
140
220
(
2017
).
33.
S.
Ibrahimkutty
,
P.
Wagener
,
T. D. S.
Rolo
,
D.
Karpov
,
A.
Menzel
,
T.
Baumbach
,
S.
Barcikowski
, and
A.
Plech
, “
A hierarchical view on material formation during pulsed-laser synthesis of nanoparticles in liquid
,”
Sci. Rep.
5
(
1
),
16313
(
2015
).
34.
H.
Yoshihara
,
N.
Koshizaki
,
Y.
Yamauchi
, and
Y.
Ishikawa
, “
Size distribution evolution and viscosity effect on spherical submicrometer particle generation process by pulsed laser melting in liquid
,”
Powder Technol.
404
,
117445
(
2022
).
35.
Y.
Ishikawa
,
T.
Tsuji
,
S.
Sakaki
, and
N.
Koshizaki
, “
Pulsed laser melting in liquid for crystalline spherical submicrometer particle fabrication–mechanism, process control, and applications
,”
Prog. Mater. Sci.
131
,
101004
(
2023
).
36.
A.
Pyatenko
,
H.
Wang
, and
N.
Koshizaki
, “
Growth mechanism of monodisperse spherical particles under nanosecond pulsed laser irradiation
,”
J. Phys. Chem. C
118
(
8
),
4495
4500
(
2014
).