You are on page 1of 23

LASER & PHOTONICS

18638899, 2013, 1, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/lpor.201100046, Wiley Online Library on [10/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
REVIEWS Laser Photonics Rev. 7, No. 1, 22–44 (2013) / DOI 10.1002/lpor.201100046

Abstract Direct laser writing has become a versatile and rou-


tine tool for the mask-free fabrication of polymer structures with
lateral linewidths down to less than 100 nm. In contrast to its
planar counterpart, electron-beam lithography, direct laser writ-
ing also allows for the making of three-dimensional structures.
However, its spatial resolution has been restricted by diffraction.
Clearly, linewidths and resolutions on the scale of few tens of
nanometers and below are highly desirable for various applica-
tions in nanotechnology. In visible-light far-field fluorescence mi-
croscopy, the concept of stimulated emission depletion (STED)
introduced in 1994 has led to spectacular record resolutions
down to 5.6 nm in 2009. This review addresses approaches aim-
ing at translating this success in optical microscopy to optical
lithography. After explaining basic principles and limitations, pos-
sible depletion mechanisms and recent lithography experiments
by various groups are summarized. Today, Abbe’s diffraction
barrier as well as the generalized two-photon Sparrow criterion
have been broken in far-field optical lithography. For further fu-
ture progress in resolution, the development of novel tailored
ARTICLE
REVIEW

photoresists in combination with attractive laser sources is of


utmost importance.

Three-dimensional optical laser lithography beyond the


diffraction limit
Joachim Fischer * and Martin Wegener
1. Introduction toresist to be exposed. Moreover, these approaches are in-
trinsically limited to the surface, i. e., to planar lithography.
A dream of nanoscience is to shape matter from the atomic The beauty of far-field optics is that it also allows
scale to the macroscopic – in three dimensions. While for three-dimensional (3D) lithography, often referred to
bottom-up self-assembly approaches have long been ex- as 3D direct laser writing (DLW) [2–10]. In 3D DLW, a
pected to eventually take over top-down lithography, the (pulsed) laser is tightly focused to a diffraction-limited spot
latter remains dominant in many important areas such as within the volume of a thick-film photoresist. By exploit-
computer-chip or photonic-circuitry fabrication. Abbe’s fa- ing two-photon absorption and/or other nonlinearities (e. g.,
mous resolution formula states that the achievable resolution brought about by the photoresist itself), the effectively ex-
in far-field optical lithography as well as in microscopy is posed volume can be restricted to the focal region leading to
proportional to the wavelength of the (light) waves used. a volume element, the “voxel”, in analogy to the picture ele-
This has led to ultraviolet (UV), deep UV (DUV) or even ment commonly referred to as pixel. It is important to note
extreme UV (EUV) planar optical lithography that are used that the existence of nonlinearity (precisely, a super-linear
in today’s and tomorrow’s computer-chip fabrication lines, behavior) is crucial: Suppose that the photoresist reacted
respectively. The corresponding mask masters are fabricated purely linearly and we raster scanned a plane parallel to
using electron-beam lithography, which takes advantage of the substrate deep within the thick-film photoresist. Clearly,
the yet smaller de Broglie wavelength of accelerated elec- the photoresist absorption would have to be very small in
trons. In principle, optical near-field effects can overcome order to avoid unwanted intensity gradients into the depth.
Abbe’s barrier in lithography [1]. While interesting demon- Thus, every volume element in the bulk photoresist would
strations have been published indeed, such approaches do “see” the same number of photons passing through (inde-
have fundamental limitations in that the near-field-writing pendent of the z-position), hence a block of exposed pho-
head inherently needs to be in close proximity to the pho- toresist would result while only a plane was desired. The

Institute of Applied Physics, DFG-Center for Functional Nanostructures, Institute of Nanotechnology, Karlsruhe Institute of Technology (KIT), 76128
Karlsruhe, Germany
*
Corresponding author: e-mail: joachim.fischer@kit.edu
This is an open access article under the terms of the Creative Commons Attribution License, which permits use, distribution and reproduction
in any medium, provided the original work is properly cited.


C 2014 The Authors. Laser & Photonics Reviews published by Wiley-VCH Verlag GmbH & Co. KGaA Weinheim
REVIEW

18638899, 2013, 1, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/lpor.201100046, Wiley Online Library on [10/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
ARTICLE

Laser Photonics Rev. 7, No. 1 (2013) 23

nonlinearity concentrates the effective dose to the focal expose a photoresist film. Determining the exposure dose
volume, especially in the axial direction. Accordingly, non- is generally difficult. For the moment, we assume that the
linearity is not just a benefit, it is essential in 3D DLW [11]. local exposure dose is simply proportional to the density
Such regular 3D DLW is readily commercially available of molecules excited by light absorption. For linear (i. e.,
and lateral linewidths down to below 100 nm can routinely one-photon) absorption, the exposure dose is thus simply
be achieved in 3D. However, despite considerable efforts proportional to the light intensity and the exposure time.
over many years worldwide, no resolution on that scale Would nonlinearity help to improve the resolution de-
in the spirit of Abbe has been achieved. In the following fined along these lines? It would not. Suppose we ex-
section, we will recall Abbe’s considerations and point out ploit two-photon absorption. Our above reasoning does not
the crucial difference between linewidth and resolution for change at all, however, the exposure-dose pattern with grat-
parallel as well as serial lithography exposure schemes. ing period axy changes its shape (see Fig. 1). For linear ab-
sorption and a fixed exposure time, the cos 2 (πx/axy ) inten-
sity pattern results in a corresponding dose pattern. The two-
2. Abbe’s diffraction limit: photon absorption probability is proportional to the square
6 resolution
linewidth = of the intensity, thus it follows a cos 4 (πx/axy ) pattern. The
additional square reduces the width, the “linewidth”, of
In the derivation of his famous formula, Ernst Abbe consid- the grating lines by a factor of roughly 1.4. The maxima
ered a grating with period axy in the xy-plane inspected by of the two-photon exposure dose have a full width at half
a microscope lens with finite aperture. The optical axis is maximum (FWHM) given by
along z, a plane wave impinging along this direction serves √
for illumination. He argued that at least the zeroth and both 2 acos( 4 0.5) axy λ
FWHM = axy × ≈ = . (2)
first diffraction orders of the grating need to be collected by π 2.75 5.5 NA
the lens’ aperture in order to retain the information on the
For example, for NA = 1.4, the FWHM is nearly eight times
grating period. By using the usual Fraunhofer diffraction
smaller than the free-space wavelength. For yet higher-order
formula of a grating, it is straightforward to arrive at the
processes, the FWHM of the exposure profile would be
Abbe condition
yet smaller.
λ λ So far, we have not addressed the photoresist. The prop-
axy ≥ =: , (1) erties of the chemical product (e. g., the cross-linking den-
2n sin α 2 NA
sity) can depend super-linearly on the exposure dose, which,
where λ is the vacuum wavelength of light, n the refractive as discussed above, may also depend super-linearly on the
index of the material in which the grating and the lens are intensity of light. While the photochemical processes may
embedded, α is the half-opening angle of the microscope- be very complex in detail, one can often describe this chem-
lens aperture, and NA the numerical aperture. If we now ical nonlinearity in the context of a simple digital threshold
think of a parallel “single-shot” exposure of a photoresist model. Herein, the photoresist only withstands the devel-
(see Fig. 1), the resolution criterion stays the same: By reci- opment process provided the exposure dose has locally ex-
procity, i. e., by considering light waves impinging along ceeded a certain threshold dose. Reducing the exposure dose
the direction of the diffraction orders, an intensity grating towards the threshold dose by decreasing either exposure
of the same critical period is formed and can be used to intensity or exposure time leads to smaller linewidths. In

Figure 1 (online color at: www.lpr-


journal.org) Diffraction-limited intensity
grating (red) applicable to parallel expo-
sures of photoresist films. The squared
intensity (purple) is relevant for a parallel
two-photon exposure. For a fixed expo-
sure time, the threshold dose translates
into a threshold intensity, which deter-
mines the width of the exposed lines. The
incident vacuum wavelength is 800 nm
and the numerical aperture NA = 1.4.

www.lpr-journal.org 
C 2014 The Authors. Laser & Photonics Reviews published by Wiley-VCH Verlag GmbH & Co. KGaA Weinheim
LASER & PHOTONICS

18638899, 2013, 1, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/lpor.201100046, Wiley Online Library on [10/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
REVIEWS

24 J. Fischer and M. Wegener: 3D optical lithography off limits

this fashion, arbitrarily small linewidths could be achieved – (which is a conservative assumption). In this case, the re-
in principle. However, the grating period, i. e., the resolution, duced width of a single exposure (e. g., a line) can indeed
is still limited by the above Abbe condition. translate to enhanced resolution. The use of two-photon
This reasoning has shown that linewidth and resolution absorption in a serial scheme √therefore improves the resolu-
are generally not the same. In fact, for the above parallel ex- tion by roughly a factor of 2, as we will argue in Sect. 3.
posure, the resolution is fundamentally limited by the Abbe Due to the resist’s threshold behavior, one might further be
criterion, whereas the linewidth is not limited at all. Thus, tempted to assume that the photoresist “forgets” any below-
we have to carefully distinguish between the two notions of threshold contributions of a first line’s exposure, e. g., due to
linewidth and resolution. To make a claim of resolution, a diffusion exchange.1 In this case, a second line could indeed
grating with a certain period has to be demonstrated. Alter- be located within one linewidth (center-to-center) next to
natively, we could relax this a bit and ask that at least two the first line. The gap between the two lines could even
lines or features separated by that distance need to be shown be smaller. In fact, nothing would fundamentally limit this
(see next section). Claiming a certain linewidth on the other gap. Hence, resolution (i. e., minimum period) and linewidth
hand can be based on an isolated line or, more generally, on would essentially be the same.
an isolated object. The experiments to be discussed in Sect. 6 are indeed
When aiming at 3D structures, we also have to carefully consistent with the resolution improvement by two-photon
distinguish between lateral resolution (axy ) and axial reso- absorption. We will also see, however, that the reasoning
lution (az ). Abbe’s original formula makes no statement on regarding the threshold and the “forgetting photoresist” is
the axial resolution. We can, however, extend Abbe’s rea- highly oversimplified: First, a perfect threshold could al-
soning in a straightforward fashion by considering four (or low for arbitrarily small linewidths. In reality, the smallest
more) laser beams with different wave vectors k impinging accessible linewidths are roughly 40% of the two-photon ex-
through the microscope objective lens aperture, forming a posure’s FWHM. Second, an ideal “forgetting” photoresist
3D interference pattern [12, 13]. The minimum axial period would allow for arbitrarily small gaps (i. e., linewidth would
is determined by the accessible bandwidth of axial wave- equal resolution). However, photoresist‘s usually seem to
vector components, ∆kz . The maximum kz is determined by “remember” a lot. In other words, our above definition of the
a beam propagating along the optical axis, the minimum kz exposure dose, i. e., the assumption of a linear accumulation
by a beam impinging under the maximum angle compatible of sequential exposures, is often meaningful indeed.
with the NA. This leads to a minimum axial period given by

2π λ λ
az = ≥ = p . (3) 3. The generalized Sparrow criterion in 3D
∆kz n(1 − cos α) n − n2 − (NA)2
As in optical microscopy, one can arrive at similar defini-
The ultimate mathematical limit is obviously reached for
tions of “resolution” in optical lithography by reasonings
n = NA which is equivalent to α = 90◦ . This axial Abbe
different from those of Abbe. In fluorescence microscopy,
limit is only two times larger (i. e., worse) than the lateral
the Abbe criterion nearly coincides with the Sparrow crite-
Abbe limit. However, for realistic numerical apertures, the
rion. Sparrow originally investigated the diffraction-limited
axial resolution further deteriorates with respect to the lat-
resolving power of spectroscopes. He found that a spectral
eral one. For example, for n = 1.5 and NA = 1.4 (parameters
line pair (broadened by diffraction) is still resolvable as long
close to the experiments in the following sections), we ob-
as there is a local minimum in the middle of the signal [14].
tain az ≥ λ /0.96. This value is 2.92 times larger than the lat-
This condition can easily be translated to fluorescence mi-
eral Abbe limit under the same conditions, i. e., axy ≥ λ /2.8.
croscopy. Here, two slightly separated point emitters appear
Another example is n = 1.5 and NA = 1.0, for which we
broadened by the microscope’s point-spread function. In
obtain az ≥ λ /0.38. This value is 5.26 times larger than the
order to be resolvable, the sum of the shifted point-spread
lateral Abbe limit under the same conditions, i. e., axy ≥ λ /2.
functions must still have a local minimum. This criterion is
These examples emphasize that obtaining good axial resolu-
also relevant to serial lithography schemes like DLW, when
tion is even more of a challenge than obtaining good lateral
we assume that the entire doses of sequential exposures just
resolution in 3D optical lithography.
accumulate, i. e., the photoresist remembers everything. The
As argued above for the lateral Abbe limit, by using a
required local minimum in the sum of two shifted point ex-
threshold process, the axial linewidth or the axial feature
posures is then crucial for the separation of the points, even
sizes are not fundamentally limited by diffraction at all. This
for a perfectly sharp threshold. Translating this criterion to
means that a reliable measurement of the axial resolution
sequential two-photon exposures is achieved by replacing
must likewise study a structure that is periodic along the
axial direction or at least study two axially adjacent lines 1 The Schwarzschild effect found in photographic plates is a
or features (see next section). Measuring just linewidths or similar example for a “forgetting” photosensitive medium and,
axial feature sizes is not sufficient. hence, a deviation from our above simple exposure-dose definition
In contrast to this thought experiment of parallel interfer- with purely linear accumulation. The Schwarzschild effect corre-
ence exposure, DLW is a serial process. In a serial process, sponds to no darkening of the photographic plate if the incident
nonlinearities can be fully exploited. Let us first assume that light intensity is below a certain value – no matter how long the
doses of sequential exposures simply accumulate linearly plate is illuminated.


C 2014 The Authors. Laser & Photonics Reviews published by Wiley-VCH Verlag GmbH & Co. KGaA Weinheim www.lpr-journal.org
REVIEW

18638899, 2013, 1, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/lpor.201100046, Wiley Online Library on [10/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
ARTICLE

Laser Photonics Rev. 7, No. 1 (2013) 25

the intensity profiles by their squares. To evaluate the res- absorption, such a (single) exposure-dose profile is propor-
olution in this sense, one needs detailed knowledge of the tional to |E|2 · ∆t, where ∆t is a fixed exposure time. For
laser focus. Here, we complement the above simple analytic two-photon absorption, the exposure-dose profile is propor-
considerations by complete numerical calculations of the tional to |E|4 · ∆t. The results are depicted in Figure 3. For
laser focus using a vector Debye approach following [15]. one-photon exposure, we find the critical distances to be
We chose parameters corresponding to typical experiments, axy = 288 nm in the lateral direction and az = 726 nm in
in particular to an exposure wavelength of 800 nm. The the axial direction. The lateral value is very close to the
scalability of the Maxwell equations allows for translating analytic Abbe criterion (axy = 286 nm). For two-photon ex-
these results to other exposure wavelengths. The numerical posure, the critical distances are reduced to axy = 200 nm
aperture is NA = 1.4, the incident laser light is circularly in the lateral direction and az = 500 nm in the axial direc-
polarized. The objective lens is illuminated homogeneously, tion. Assuming Gaussian profiles in lateral as well as in
corresponding to a beam diameter that is significantly larger axial direction, one would expect that squaring the profiles
than the pupil diameter. would reduce their√extent, and hence also their critical dis-
The resulting distributions of the squared modulus of the tance, by a factor 2. For the lateral direction, this simple
(complex) electric-field vector, |E|2 , and the square of that, reasoning would √ lead to a two-photon-modified Abbe for-
|E|4 , are depicted in Fig. 2. Note that the former quantity is mula axy = λ /(2 2 NA) = 202 nm that nicely resembles
generally distinct from the intensity, i. e., the time-average the numerical results for the generalized two-photon Spar-
of the modulus of the Poynting vector (not depicted). For row criterion.
example, for linear polarization of the incident light and for As expected, the numerical calculations show that
the above tight-focusing conditions, the intensity distribu- the critical distances along the axial direction are signi-
tion in the focal plane is rotationally symmetric, whereas ficantly larger than the lateral ones, just like the FWHM
that of |E|2 is elongated along the direction of the incident of the exposure profiles along the two axes. The elon-
polarization. Typical photoinitiator molecules have electric- gation factors of FWHMs and critical distances for one-
dipole-allowed optical transitions and are hence driven by photon and two-photon exposure profiles range between
the electric-field component of the light alone. Thus, |E|2 2.45–2.52. Hence,√we suggest a further modified Abbe for-
is the relevant distribution for one-photon absorption and mula az = λ A/(2 2 NA) = 505 nm for the axial direction,
|E|4 that for two-photon absorption. (Strictly speaking, this where A = 2.5 is the aspect ratio of the exposure volume.
aspect should also be applied to the above considerations Again, this intuitive and handy formula nicely approximates
regarding the Abbe criterion.) As already expected from our the numerical results for the generalized two-photon Spar-
above simple analytic consideration, the axial extent of the row criterion. However, the aspect ratio depends on the nu-
profiles in Fig. 2 is significantly larger than the lateral one. merical aperture and the refractive index of the photoresist,
To obtain the resolution according to Sparrow, we add for which we have taken n = 1.52.
two numerically computed exposure profiles, that are shifted If the photoresist exhibits a threshold behavior, the width
with respect to each other in either the lateral or the axial of the photoresist voxels is not fundamentally limited by
direction and still yield a local minimum. For one-photon diffraction (in analogy to the discussion in the previous sec-
tion). However, as long as the resist “remembers” previous
below-threshold exposures, the resolution in the sense of
Sparrow is fundamentally limited by diffraction.

4. The basic idea of depletion


It is obviously desirable to effectively decrease the lateral
and/or the axial extent of the exposed volume. Diffraction
fundamentally inhibits doing that. It is Stefan Hell’s decep-
tively simple yet ingenious idea of introducing a depletion
laser that enables circumventing this limitation [16–23].
This idea and its derivatives have already revolutionized
fluorescence microscopy (see, e. g., the special issue [20]
for reviews and state of the art).
In regular DLW, light from a laser excites the photoresist.
Figure 2 (online color at: www.lpr-journal.org) Calculated focal This excitation eventually leads to an irreversible chemical
intensity distribution of a typical writing spot. (a) Iso-intensity reaction, e. g., polymerization. For a common negative-tone
surfaces. The profiles along the two black lines are depicted photoresists, these regions become insoluble whereas the
in b) and c). (b) Lateral profiles of |E|2 (red) and |E|4 (purple) unexposed regions can be removed in a development step.
correspond to one-photon exposure and two-photon exposure, The basic idea of depletion DLW [24–28] is to access the
respectively. (c) Axial profiles of |E|2 (red) and |E|4 (purple). The system at some intermediate state, aiming at inhibiting this
horizontal lines in b) and c) correspond to the iso-intensity values chemical reaction, i. e., effectively reducing the photoresist’s
of the surfaces in a). sensitivity. Only a part of the excited molecules and, hence,

www.lpr-journal.org 
C 2014 The Authors. Laser & Photonics Reviews published by Wiley-VCH Verlag GmbH & Co. KGaA Weinheim
LASER & PHOTONICS

18638899, 2013, 1, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/lpor.201100046, Wiley Online Library on [10/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
REVIEWS

26 J. Fischer and M. Wegener: 3D optical lithography off limits

Figure 3 (online color at: www.lpr-journal.org) Critical distances, axy and az , for two point exposures derived from numerical
calculations corresponding to the Sparrow criterion. All plots shown use slightly larger values as indicated (i. e., axy + 10 nm and
az + 20 nm) such that small local minima are still visible. The exposure-dose profiles along the black lines in the iso-intensity plots
(left-hand side of each panel) are plotted separately (right-hand side of each panel). The line plots show the contributions of the single
exposures (red) and the sum of both (blue). (a) Critical lateral distance, axy , for one-photon exposure. (b) Critical axial distance, az , for
one-photon exposure. (c,d) Same as (a,b), but for two-photon exposure. A typical value for the polymerization threshold in lithography
(75% of the peak dose) is used as dose value for the blue iso-surfaces, illustrating that the two point exposures are not separated any
more. Parameters are: 800 nm free-space wavelength, NA = 1.4, and circular polarization.

only a part of the original exposure dose contributes to the dose profile is not. The diffraction limit is broken by exploit-
chemical reaction. We will call this part the “effective” ex- ing photoresist switching in combination with tailored foci
posure dose. The inhibition or switching can be induced by of light.
a second laser, which will generally operate at a wavelength In Sect. 5, we will summarize possible microscopic de-
different from that of the excitation laser (to avoid excitation pletion mechanisms in photoresists suitable for depletion
by the depletion laser). Clearly, the inhibition is only pos- DLW. For now, we assume that the initial local excitation
sible within the lifetime of the intermediate state. It is also is multiplied with the factor 1/(1 + γIdepl ), where Idepl is
important to note that this inhibition needs to be reversible. the local depletion intensity and γ is a photoresist-specific
Otherwise, depleted regions would remain insensitive from constant containing, e. g., the cross-section of the depletion
thereon. This reversibility must not be confused with our mechanism and the intermediate-state lifetime. Section 5.1
above notion of a “forgetting” photoresist. will show that this scaling is expected for stimulated emis-
This switching alone together with two similarly shaped sion as well as for other depletion pathways.
laser foci does not yet lead to any resolution improvement.
However, if the spatial maximum of the excitation profile
coincides with a local zero of the depletion profile, only 4.1. Common depletion modes
the outer regions of the excitation volume are de-excited,
leading to an effectively reduced exposure volume. Upon Fortunately, the know-how for realizing suitable depletion
increasing the depletion power, this effective exposure vol- foci can be copied from fluorescence microscopy (see,
ume shrinks towards zero. While both the excitation and the e. g., [29]). A well-known and widely used focus is the so-
depletion profile are governed by diffraction, the exposure- called donut mode illustrated in Fig. 4a–c. It can be realized


C 2014 The Authors. Laser & Photonics Reviews published by Wiley-VCH Verlag GmbH & Co. KGaA Weinheim www.lpr-journal.org
REVIEW

18638899, 2013, 1, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/lpor.201100046, Wiley Online Library on [10/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
ARTICLE

Laser Photonics Rev. 7, No. 1 (2013) 27

Figure 4 (online color at: www.lpr-journal.org) Typical depletion profiles for depletion DLW. (a–d) Donut depletion focus. (a) Iso-intensity
surfaces of the donut depletion focus. The front xy-quadrant has been removed to reveal the interior. (b) Lateral intensity profile. (c) Axial
intensity profile. (d) Phase mask used for the generation of this focus. (f–h) Same as (a–d), but for a bottle depletion focus rather than a
donut. (e,j) The anticipated resolution scaling with increasing depletion power is illustrated with the extent of single voxels for each
depletion mode ((e) donut, (j) bottle-beam). Parameters are: 532 nm free-space wavelength, NA = 1.4, and circular polarization.

by introducing a helical phase mask as shown in Fig. 4d into the resulting bottle-beam focus, the intensity in the center
the beam. Here, the additionally accumulated optical phase of the focus is strictly zero. In contrast to the donut, how-
increases from zero to 2π in one turn along the azimuth. ever, the intensity not only rises radially but also along the
The phase mask plane may be imaged onto the back-focal optical axis. In fact, it rises more steeply along the optical
plane of the microscope objective lens. Upon using circu- axis (where the initial excitation volume is also elongated),
lar polarization of the correct handedness, a wavelength of resulting in a more pronounced depletion effect in the axial
532 nm and again a numerical aperture of NA = 1.4, the direction. This aspect will lead to a more pronounced axial
intensity distribution shown in Fig. 4a–c results. The opti- linewidth improvement compared to the lateral one. The
cal intensity on the entire optical axis is strictly zero. Thus, ultimate ratio of axial to lateral voxel extent is given by the
depletion will never occur here. While this focus nicely al- shape of the inner iso-intensity surfaces of the depletion
lows for reducing the lateral extent of the effective exposure focus. Equivalently, this ratio is determined by the ratio
profile, it leaves the axial distribution essentially unaffected. of the curvatures of the axial and lateral intensity profiles
This may be useful for planar optical lithography, but is (compare panels g and h in Fig. 4). For the parameters of
quite undesired for truly 3D optical lithography, where the Fig. 4, the ultimate aspect ratio is 0.7, i. e., the voxels can
worst spatial resolution matters for the making of arbitrary be smaller along the axial than along the lateral direction.
complex structures.
Thus, the bottle-beam focus shown in Fig. 4f–h is more The two different phase masks can also be effectively
attractive for 3D DLW. This focus can be achieved by replac- combined by (even coherently) superimposing two beams,
ing the helical phase mask by the one shown in Fig. 4i. Here, one with the helical phase mask and the other with the
a cylindrical disk in the center of the beam introduces a π cylindrical one. This combination allows for controlling
phase shift. For√a large beam, the diameter of the cylinder the relative steepness of the axial and the lateral intensity
needs to be 1/ 2 of the objective’s pupil diameter, such minimum via the relative phase and amplitude of the two
that the cylinder covers half of the relevant beam area. In beams. For many applications aiming at complex isotropic

www.lpr-journal.org 
C 2014 The Authors. Laser & Photonics Reviews published by Wiley-VCH Verlag GmbH & Co. KGaA Weinheim
LASER & PHOTONICS

18638899, 2013, 1, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/lpor.201100046, Wiley Online Library on [10/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
REVIEWS

28 J. Fischer and M. Wegener: 3D optical lithography off limits

three-dimensional nanostructures, the desired voxel aspect linewidth by a factor three, the depletion power needs to
ratio is 1. increase by about a factor ten, hence the two-photon coeffi-
cient has to be decreased by two orders of magnitude.
The Figs. 2–4 refer to a high numerical aperture of
4.2. Resolution scaling NA = 1.4. Using the depletion idea, one can also obtain
similarly fine features with lower numerical apertures. This
Let us get a first impression of what can be expected by may be desirable in some cases because a lower NA allows
using the simple threshold model. Usually, regions with for using larger working distances and also for not having
above-threshold excitation will remain in a negative-tone to use an immersion liquid (see Sect. 6 or [25]). However,
photoresist after the development step. However, if the de- it should be clear from the above that especially the axial
pletion factor 1/(1 + γIdepl ) reduces the effective local ex- linewidths will rapidly deteriorate when decreasing the NA
posure dose below the threshold, the regions will again (see Sect. 2). Yet, depletion DLW with air microscope objec-
wash out in the developer. As a result, the anticipated voxels tive lenses may still be interesting for applications in planar
shrink in size. Strictly speaking, not only the voxel (i. e., the optical lithography.
above-threshold region) but also the entire effective expo-
sure profile shrinks. This also includes the below-threshold
regions. Therefore, the resolution in the sense of Sparrow is 4.3. Noise sensitivity
expected to increase. For the donut depletion mode, only the
lateral width decreases, whereas lateral and axial extent de- Another beneficial aspect of depletion DLW is that it tends to
crease for the bottle-beam focus. Exemplary calculations for substantially reduce the sensitivity of the critical dimensions
two-photon excitation and depletion with either of the two (minimum feature size attainable) with respect to errors or
depletion profiles are depicted in Fig. 4e and j, respectively. fluctuations. For example, lithography industry demands for
Here, the local effective exposure dose is proportional to robustness against power fluctuations as large as 5%. The
2 × 1/(1 + γI effect of such fluctuations on DLW is studied in Fig. 5 using
Iexc depl ). The excitation profile Iexc and the de-
pletion profile Idepl are chosen according to Fig. 2 and Fig. 4, a hypothetical exposure power with 5% rms noise (Fig. 5a).
respectively. The threshold of the photoresist is chosen at When operated at rather high excitation powers (e. g., when
75% of the peak exposure dose. Panels e and j can directly the threshold is at the FWHM of the exposure profile), regu-
be compared, as the same depletion-beam power is used. lar DLW is rather insensitive to noise fluctuations (Fig. 5d).
The different depletion modes have different peak intensi- Clearly, the resulting feature sizes are rather large. However,
ties and therefore yield different scales on the horizontal when operated close to the threshold (e. g., when aiming for
axes. Upon continuously increasing the depletion power, the 100 nm linewidth), the noise translates into terrible linewidth
feature sizes converge to zero (see Fig. 4, right-hand side). fluctuations – the test line even gets disconnected (Fig. 5e).
There is no fundamental optics limit for the linewidth and Here, the 5% rms power fluctuations (or equivalently about
the resolution in optical lithography any more. 10% dose or threshold fluctuations) lead to linewidth fluctu-
It is interesting to study the asymptotic behavior. Sup- ations of 42% rms and peak fluctuations of 100%. In sharp
pose that the feature size is already strongly reduced and contrast, depletion DLW allows for having both at the same
entirely limited by the depletion power. The initial shape of time, small linewidths and robustness against fluctuations.
the excitation is not relevant anymore and we can assume This fact is illustrated in Fig. 5f. Here, we again consider
a constant exposure profile instead. In this case, the spatial 5% fluctuations of the excitation laser, but additionally also
minimum of the depletion profile can be approximated by 5% (uncorrelated) fluctuations of the depletion laser. Fairly
a parabola. The edge of an exposed feature is given by the small edge roughness results. This increased robustness
condition that the depletion intensity parabola exceeds a alone may be a major benefit of depletion-mode DLW for
certain value, such that the initial (spatially constant) excita- many applications. In many scientific applications, however,
tion is reduced below the polymerization threshold. Solving the excitation-laser fluctuations can typically be controlled
for the crossing point immediately leads to a feature size to nearly 0.1%, in which case this aspect is only of minor
that scales inversely with the square root of the depletion importance. We do emphasize, however, that small spatial
power. Note that this reasoning does not involve the above inhomogeneities of local photoresist properties (e. g., of the
1/(1 + γIdepl ) depletion factor and should therefore apply local concentration of photoinitiator and/or inhibiting oxy-
to all depletion mechanisms alike. Thus, for example, to re- gen), which lead to an effective local threshold variation,
duce the feature size by a factor of 10, one needs to increase would have the same effect and may not be avoidable.
the depletion power by a factor of 100. This means that
undesired but possibly finite one-photon absorption of the
depletion beam increases by factor 100, too. Two-photon 4.4. Time constants matter
absorption of the depletion beam increases even by factor
104 . This unfavorable scaling with depletion power suggests So far we have not addressed any time dependence of the
that parasitic processes may at some point overwhelm the depletion process. In reality, depletion will only be possible
desired depletion benefit. Suppose, for example, a situation for a certain time span after (pulsed) excitation of the pho-
in which the two-photon absorption of the depletion beam toresist. This is essential as it makes the process irreversible
is the limiting factor. If one aims at further reducing the after some time. Otherwise, two crossing lines could not be


C 2014 The Authors. Laser & Photonics Reviews published by Wiley-VCH Verlag GmbH & Co. KGaA Weinheim www.lpr-journal.org
REVIEW

18638899, 2013, 1, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/lpor.201100046, Wiley Online Library on [10/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
ARTICLE

Laser Photonics Rev. 7, No. 1 (2013) 29

Figure 5 (online color at: www.lpr-journal.org) Linewidth fluctuations caused by fluctuating laser power. (a) Fluctuating laser powers
for excitation and depletion (uncorrelated random numbers with 5% standard deviation). (b) Voxel width and relative voxel-width
fluctuations plotted vs. average excitation power for regular DLW. (c) Voxel width and relative voxel-width fluctuations plotted vs. average
depletion power in depletion DLW. The excitation power is held constant at 141% of the threshold power. The relative width fluctuations
caused by noise of the excitation laser (red) and noise of the depletion laser (green) are plotted separately. Selected working points are
marked with circles and illustrated in detail in (d–f). (d) Resulting line shape for regular DLW where the linewidth equals the FWHM of
the squared intensity. (e) Shape of a thinner line for regular DLW. The reduction in linewidth is realized by decreasing the excitation
power level to 107% of the threshold power. (f) Shape of a thinner line for depletion DLW. Excitation-power level like in d), but the
depletion power is increased. The depicted line shape accounts for both lasers’ fluctuations. The upper panels in (d–f) show calculated
squared intensity distributions of the excitation focus (red), intensity distributions of the depletion focus (green), the anticipated effective
exposure profile (purple) along the a lateral axis. The polymerization threshold is indicated in blue.

exposed as during the second exposure the first one would undesired, µs time constants are acceptable but may become
be depleted near the crossing point. The relevant linewidth borderline in the near future, and ns time constants are
together with this time constant τ of the depletion process compatible with very fast writing.
imposes a fundamental limitation on the maximum accessi- This discussion seems to favor extremely short depletion
ble writing speed (other limitations may come in addition). time-constants. However, there is a trade-off. The intensity
Suppose one wants to write lines with l = 20 nm feature required to deposit a certain energy necessary for the deple-
size with a certain linear scan velocity v using a common tion process increases like I ∝ 1/τ. Thus, larger intensities
donut mode. Clearly, the depletion process must be com- are required for small τ, which increases the relative weight
pleted when the laser foci are separated by one feature size of undesired two-photon absorption of the depletion beam
from a first point written on the line. One may even want that will likely occur in any photoresist at high intensities.
a certain safety margin to make sure that the process has Furthermore, the intensities required for relatively large
really become irreversible. Let us consider a conservative values of τ may still be accessible with (quasi-)continuous-
margin of a factor of 10. In this case, the writing speed is wave depletion lasers, whereas short depletion pulses are
limited by the inequality required for small τ.
l
v≤ . (4)
10 · τ 5. Different depletion mechanisms
For example, for τ = 1 ms we obtain v ≤ 2 µm/s, for τ = 1 µs
we obtain v ≤ 2 mm/s, and for τ = 1 ns we obtain v ≤ 2 m/s. The depletion necessary to enhance the resolution in 3D
Today, routine DLW writing speeds are in the range of DLW can be accomplished in different ways. Any photo-
100 µm/s. The largest speed employed in DLW so far is induced mechanism that is locally prohibiting the formation
190 mm/s using scanning galvanometer mirrors [30]. Future of an insoluble cross-linked polymer is suitable at first (see
commercial instruments will likely move towards this value. Fig. 6 and Fig. 7 for a list of possibilities). Usually, after
Thus, depletion time-constants in the ms range are clearly photoinitiator molecules are excited by light absorption,

www.lpr-journal.org 
C 2014 The Authors. Laser & Photonics Reviews published by Wiley-VCH Verlag GmbH & Co. KGaA Weinheim
LASER & PHOTONICS

18638899, 2013, 1, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/lpor.201100046, Wiley Online Library on [10/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
REVIEWS

30 J. Fischer and M. Wegener: 3D optical lithography off limits

this process chain may serve as a depletion mechanism in


3D DLW.
The following commonly used photoinitiators have been
shown to not work in this regard (at least for the experi-
mental conditions of the corresponding papers): Irgacure
369 [24, 26], Irgacure 1800 [26], Irgacure 819 [26], Irgacure
Darocur TPO [24, 26], and Irgacure 184 [24]. Less com-
monly used photoinitiating molecules have been shown to
not work either [24]: Brilliant Green, Rose Bengal, and
Tris(bipyridine)ruthenium(II) dichloride. Figure 8 summa-
rizes the currently known photoresist compositions discov-
ered by different groups that have the required depletion
capabilities. In the following section, we will review sev-
eral depletion mechanisms (illustrated in Figure 6) that can
explain the polymerization inhibition observed in the differ-
ent photoresists. Moreover, we will shortly discuss possi-
ble alternative mechanisms of an effective polymerization
inhibition (Figure 7). These alternative mechanisms may
also interfere with the above intended mechanisms without
being recognized and can either amplify or limit their per-
formance.

Figure 6 (online color at: www.lpr-journal.org) Schematic di-


agrams of the molecular states and transitions for the differ-
ent depletion mechanisms in STED-inspired DLW approaches.
(a) STED lithography: Photoinitiator (PI) molecules are excited via
two-photon absorption (TPA), relax to the S1 state and are brought
back to the ground state via stimulated emission (SE), before they
can proceed to the triplet (T1 ) via intersystem crossing (ISC),
generate radicals (R• ), initiate a propagation polymer chain (RM• )
and finally yield an irreversibly cross-linked polymer. (b) RAPID
lithography: PI molecules are excited via two-photon absorption
and generate an active species in a long-lived intermediate state.
Upon light excitation, the intermediate state is deactivated and
does not lead to a cross-linking polymerization. (c) Two-color
photo-initiation/inhibition (2PII) lithography: PI molecules are ex-
cited via one-photon absorption (OPA) and generate radicals
that can initiate the polymerization. Photoinhibitor molecules are Figure 7 (online color at: www.lpr-journal.org) Alternative mech-
activated via one-photon absorption at a different wavelength. anisms for photo-inhibited polymerization. (a) Excited-state ab-
The generated non-initiating radicals (Q• ) can scavenge initiating sorption: After PI excitation, several intermediate states can ab-
radicals and terminate propagating chains (RM• ). sorb the depletion light. From such highly excited states, non-
radiative decay to the ground state can occur. (b) Resist heat-
ing: Through repeated absorption from an excited-state and non-
radiative decay to the same state, the resist can be heated and
a major fraction of the molecules undergoes inter-system several resist properties can change with the increased tempera-
ture, disturbing the excitation, initiation or polymerization process.
crossing (ISC) to a long-lived and reactive triplet state. From
this triplet state, initiating radicals are generated with a cer-
tain yield (e. g., via cleavage or charge transfer to other
molecules). These radicals initiate a polymerization reac-
tion, which proceeds in acquiring monomer molecules until 5.1. Stimulated-emission-depletion lithography
the propagating radical chains are terminated or until no 5.1.1. Mechanism and experimental realization
monomer is left. As soon as the polymer is sufficiently
cross-linked it becomes insoluble for solvents and the ex- The original idea for diffraction-unlimited lithography is
posure has become irreversible. Any process that disturbs to force photoinitiator molecules to undergo stimulated


C 2014 The Authors. Laser & Photonics Reviews published by Wiley-VCH Verlag GmbH & Co. KGaA Weinheim www.lpr-journal.org
REVIEW

18638899, 2013, 1, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/lpor.201100046, Wiley Online Library on [10/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
ARTICLE

Laser Photonics Rev. 7, No. 1 (2013) 31

Figure 8 (online color at: www.lpr-journal.org) Ingredients of the known photoresist systems that provide a photo-deactivation
pathway. The resists contain cross-linkable multifunctional monomer(s), a photoinitiator and in one case a photoinhibitor. The relevant
(normalized) spectra of the corresponding photoinitiator (and the photoinhibitor) are plotted on the right hand side. The wavelengths
used for excitation and depletion are marked with arrows. The four rows correspond to the photoresists described in (a) [27] (spectra
for DETC dissolved in PETTA [36]), (b) [24] (spectrum for MGCB in ethanol), (c) [25] (spectra for photoinitiator and photoinhibitor in
CHCl3 [25]), and (d) [26] (spectra for ITX in ethanol [26]). The photoresist used in [28] exploits the same mechanism as the photoresist
in panel (c), together with another monomer. Therefore, this resist is not listed separately.

emission to the ground state after their excitation and be- crossing and the intersystem-crossing yield becomes
fore intersystem-crossing takes place (see Fig. 6a) [17]. This kISC ΦISC,0 ΦISC,0
stimulated-emission transition can be induced by the intense ΦISC = == kSTED
= , (6)
light of a depletion laser. The dependence of the depletion ktot 1 + ktot,0 1 + γIdepl
efficiency on the depletion power is easily derived when where γ is a photoinitiator-specific constant, Idepl is the de-
looking at the depopulation of the photoinitiator’s S1 state pletion laser intensity, and kSTED ∝ Idepl is used. According
(directly after its population by the excitation pulse). The to this simple model, the intersystem-crossing yield and,
depopulation rate constant in absence of a depletion beam hence, the effective exposure dose decreases continuously
is ktot,0 = kfl + kISC + knr , where kfl is the fluorescent rate, with increasing depletion intensity.
kISC is the intersystem-crossing rate, and knr is the combined Usually, the depletion-laser wavelength is chosen at
rate of all non-radiative decay channels. The undisturbed the red end of the fluorescence spectrum. This is a com-
intersystem-crossing quantum-yield is then given by promise between moving towards the wavelength of maxi-
mum fluorescence (and hence towards maximum stimulated-
kISC kISC emission cross-section) and avoiding the fundamental ab-
ΦISC,0 = = . (5) sorption band, i. e., limiting one-photon absorption of the
ktot,0 kfl + kISC + knr
depletion laser.
It has been found that common photoresists and pho-
When the depletion laser is switched on, stimulated emission toinitiator molecules are not suited for stimulated emission
kSTED is added as depopulation channel for the S1 , and the depletion (STED) [24,26]. For efficient STED, a large cross-
depopulation now follows ktot = kfl + kISC + knr + kSTED . section for stimulated emission is required, which is equiv-
Now stimulated emission competes with the intersystem- alent to a large oscillator strength between the S0 and the

www.lpr-journal.org 
C 2014 The Authors. Laser & Photonics Reviews published by Wiley-VCH Verlag GmbH & Co. KGaA Weinheim
LASER & PHOTONICS

18638899, 2013, 1, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/lpor.201100046, Wiley Online Library on [10/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
REVIEWS

32 J. Fischer and M. Wegener: 3D optical lithography off limits

S1 state of the molecule. Common photoinitiator molecules


have a n − π ∗ type transition with inherently low oscillator
strength (e. g., Irgacure 907 [31]), whereas the fluorescent
dyes used in STED microscopy have π − π ∗ type transitions
with large oscillator strengths. Moreover, the excited-state
lifetime of photoinitiators is often unfavorably short (around
100 ps) [24, 32]. To drive efficient stimulated emission, one
would need to use pulses shorter than 100 ps and addition-
ally increase the pulse energy considerably to compensate
for the small cross-sections. Clearly, this would quickly lead
to pronounced multi-photon absorption of the depletion
beam, limiting the depletion capability. Even more likely,
the small cross-section for stimulated emission is overcom-
pensated by a transient absorption from S1 to higher singlet Figure 9 (online color at: www.lpr-journal.org) Pump-probe data
states. Such states are often very reactive and tend to have of DETC in ethanol. Time-dependent change of the optical den-
enhanced intersystem-crossing rates, leading to enhanced sity (OD), ∆OD (dots), after excitation at 387.5 nm wavelength for
exposure instead of depletion [24]. Finally, if excited-state different probe wavelengths as indicated. ∆OD < 0 corresponds
absorption dominates over stimulated emission, repeated to increased probe transmittance upon optical pumping. The solid
curves result from a global fit of a simple rate-equation model
excited-state absorptions (followed by non-radiative relax-
to these data. Negative changes in transient optical density cor-
ation to S1 ) will also increase the temperature of the resist
respond to stimulated emission for wavelengths up to 630 nm.
and limit the applicable depletion power.
Reproduced from [35].
To overcome these obstacles, “dye-like photoinitia-
tors” [26, 27] with larger oscillator strengths and “pho- However, it is well known that the S1 lifetime of a mol-
toinitiating dyes” [24, 33] have been explored. However, ecule can depend on the solvent environment. Indeed, the
only recently one photoresist composition based on 7- S1 lifetime of DETC dissolved in PETTA increases to about
Diethylamino-3-thenoylcoumarin (DETC) has unambigu- 1 ns [36]. This reference speculates that this increase in the
ously been shown to support STED. DETC belongs to more viscous monomer is due to steric hindering of con-
the group of keto-coumarins, which are known as triplet- formational changes necessary for inter-system crossing. In
sensitizers and photoinitiators [34]. Being a coumarin PETTA, a simple pump-probe experiment is no longer pos-
derivative, DETC possesses a large oscillator strength sible, because the intense pump beam leads to irreversible
(40550 L/mol/cm [35]) that is comparable to state-of- polymerization of the photoresist inside the cuvette. Alter-
the-art green-emitting fluorescent dyes (e. g., Atto 425, natively, the S1 lifetime can be determined by a lithogra-
45000 L/mol/cm), which are usually coumarin derivatives as phy experiment with time-delayed excitation and depletion
well. The additional ketone group in DETC is very common pulses. Such experiments have been performed in [36] and
for photoinitiator molecules and promotes the photoiniti- are shown in Fig. 10. These experiments show a temporal
ating functionality. With fluorescence quantum yields of decay of the depletion effect composed of an initial rapid de-
roughly 2.5% in ethanol [36], 10% in benzene [34], and cay with about 1 ns time constant and a second slower decay
27% in the viscous monomer pentaerythritol tetraacrylate with a time constant between 12.5 ns and 1 µs (Fig. 10a). The
(PETTA) [36], DETC is a poor dye, but turns out to be rapid decay has been assigned to stimulated emission [36],
an efficient initiator for two-photon DLW [27]. DETC dis- the slower one is yet to be determined. The spectral de-
solved in PETTA possesses a one-photon absorption peak pendence of the fast component nicely follows the gain
at 420 nm wavelength, which is convenient for two-photon spectrum of DETC, supporting the interpretation that this
excitation with Ti:sapphire oscillator lasers or frequency- depletion channel is indeed stimulated emission (Fig. 10b).
doubled Er-doped fiber oscillators (Fig. 8a). The fluores- The slow component shows a different spectral dependence,
cence emission of DETC peaks at 480 nm wavelength and supporting the interpretation that two different depletion
decays to 10% at 550 nm. Thus, frequency-doubled Nd or mechanisms are at work in this photoresist.
Yb lasers at 532 nm can serve as attractive high-power de-
pletion laser sources. Moreover, keto-coumarins are widely
tunable in their absorption and emission wavelengths [34].
5.1.2. Properties and limitations
This makes other keto-coumarins promising candidates for
future optimized STED photoinitiators.
Compared to other mechanisms to be discussed below, stim-
Standard pump-probe experiments on DETC dissolved ulated emission offers the fastest time constant, as the life-
in ethanol [35] show that stimulated emission overwhelms time of the responsible S1 state is typically 0.1 ns - 4 ns.
excited-state absorption for relevant wavelengths within the Directly after stimulated emission (and subsequent vibronic
fluorescence band, including especially 532 nm. This can be relaxation) the photoinitiator molecules are ready for a
seen from the negative change of the optical density in Fig. 9. new exposure. This allows for very fast scan speeds in the
From fitting a simple rate-equation model to the measured range of m/s (see Sect. 4.4), but also requires high depletion-
data, the authors have derived a S1 lifetime of about 100 ps. laser intensities.


C 2014 The Authors. Laser & Photonics Reviews published by Wiley-VCH Verlag GmbH & Co. KGaA Weinheim www.lpr-journal.org
REVIEW

18638899, 2013, 1, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/lpor.201100046, Wiley Online Library on [10/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
ARTICLE

Laser Photonics Rev. 7, No. 1 (2013) 33

(a)

(b)

Figure 10 (online color at: www.lpr-


journal.org) Temporal (a) and spectral
(b) characteristics of the polymerization
deactivation with DETC as photoinitiator. The
temporal decay in (a) is assigned to the S1
lifetime and corresponds to a fast component of
the inhibition due to stimulated emission. The
offset is due to another depletion channel with
a considerably longer time constant. In (b) the
spectral characteristics of the slow and the fast
component are plotted vs. the anticipated gain of
DETC in PETTA. The fast effect resembles the
gain spectrum, consistent with STED being the
underlying mechanism. Reproduced from [36].

Stimulated emission is also unique among possible de- rhodamines [33]) are used as photoinitiators. The origi-
pletion mechanisms because it does not inject additional nal implementation used malachite green carbinol base
heat into the photoresist, but even ejects energy from the (MGCB). Initially, this molecule was expected to support
system. Other mechanisms have to be triggered by the ab- STED and was chosen due to its very high extinction co-
sorption of a depletion photon, the energy of which will efficient. MGCB has its fundamental absorption peak at
finally become thermal. In contrast, when using STED, a 620 nm, and another shoulder around 430 nm before the ab-
new photon which contains a major fraction of the initial ex- sorption rises strongly towards shorter wavelengths (see
citation energy is emitted and leaves the photoresist system. Fig. 8b). Two-photon absorption of femtosecond pulses
For efficient stimulated emission, high-peak-power de- around 800 nm wavelength is possible and will likely ex-
pletion pulses are required – especially if the photoinitiator cite the molecule to a higher singlet state Sn followed by
molecule is not perfectly optimized for the depletion wave- non-radiative relaxation to S1 . One would expect further red-
length. The use of stimulated emission comes along with the shifted fluorescence emission around 700 nm wavelength.
search for suitable, high-peak-power pulsed laser sources Hence, a potential STED depletion could also be performed
and their temporal alignment. Continuous-wave depletion with 800 nm light. Depletion pulse lengths >50 ps would
at relatively high power levels is possible in principle, but ensure that the depletion pulses do not cause efficient two-
might be limited by competing effects like excited-state ab- photon absorption [24]. However, during the preparation of
sorption. Fig. 8 we could not detect any luminescence from MGCB
in ethanol solution at the anticipated spectral position for
excitation wavelengths between 400 nm and 700 nm.
5.2. RAPID lithography Indeed, activating the depletion laser can induce an
effective polymerization inhibition in MGCB-based pho-
toresists [24]. However, the authors have argued that the
5.2.1. Mechanism and experimental realization observed effect can not be caused by stimulated emission,
because the depletion effect is insensitive to pulse delays
In Resolution Augmentation through Photo-Induced De- between excitation and depletion pulses up to 13 ns. They
activation (RAPID) lithography [24] special cationic dyes assigned the observed effect to the depletion of an interme-
(cationics diarylmethanes, cationic triarylmethanes, cationic diate state with longer lifetime (see Fig. 6b). This allows

www.lpr-journal.org 
C 2014 The Authors. Laser & Photonics Reviews published by Wiley-VCH Verlag GmbH & Co. KGaA Weinheim
LASER & PHOTONICS

18638899, 2013, 1, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/lpor.201100046, Wiley Online Library on [10/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
REVIEWS

34 J. Fischer and M. Wegener: 3D optical lithography off limits

for the use of a continuous-wave (cw) laser for depletion time units (corresponding to 0.05 and 0.02 velocity units)
and eliminates the necessity for a precise timing between were used to successfully model the observed proportional
excitation and depletion pulses. The exact nature of the linewidth scaling. Within this model, the exposure times
intermediate state, however, is yet to be determined [33]. used in the experiments lead to the conclusion that the deple-
Possible explanations are that after the excitation these mole- tion time-constant is between 15 ms and 350 ms. (We deduce
cules generate weakly reactive radical pairs [24] or solvated exposure times of 1–7 ms from Fig. 2j of [33]. These corre-
electrons [33] that can recombine to the parent molecules spond to the used writing speeds of 30 µm/s to 150 µm/s and
via photo-induced electron-backtransfer. a diffraction-limited excitation spot with a lateral intensity
FWHM of 310 nm.) This implies that RAPID photoinitia-
tors in general or at least the highly sensitive photoinitiator
5.2.2. Properties and limitations MGC-HCl in particular are not suitable for fast scanning
(see Sect. 4.4) with confinement along the direction of scan-
The photoinitiator MGCB absorbs from 400 nm to 700 nm ning (e. g., with a lateral donut depletion-mode). However,
wavelength, i. e., essentially throughout the entire visible axial confinement with a bottle beam together with lateral
spectrum. This fact has to be considered when handling this scanning is possible and has been demonstrated [24]. More-
photoresist to avoid accidental exposures by ambient light. over, a less commonly used depletion mode [18] could be
Moreover, remaining unconsumed photoinitiator molecules used to only improve the resolution in one lateral direction
in the final polymer structures might render them absorbing perpendicular to the scan direction.
for visible light.
Further investigations have shown that not only MGCB
but three broad classes of dyes are capable of being depleted 5.3. Photoinhibitor lithography
by the RAPID scheme [33]. For example, using malachite
green carbinol hydrochloride (MGC-HCl) the depletion ef- 5.3.1. Mechanism and experimental realization
fect is so sensitive that even the femtosecond excitation
pulses themselves lead to a depletion effect. This manifests A lithography approach based on another depletion mecha-
itself in an unexpected proportional dependence of the re- nism [25] has been named two-color photo-initiation / inhi-
sulting polymer linewidth on the writing speed. Usually, bition lithography (2PII) [37]. 2PII is not based on a direct
increased scan speeds yield decreased linewidths because deactivation of an intermediate excited photoinitiator state
the local exposure dose is decreased towards the threshold but on the activation of a counter-acting photoinhibitor (see
value due to shortened exposure times. In contrast, when Fig. 6c) [25]. While the photoinitiator system is allowed to
using special RAPID photoinitiators, increased scan speeds generate initiating radicals upon one-photon absorption, the
yield increased linewidths [33]. This is explained as fol- additional photoinhibitor is activated by one-photon absorp-
lows [33]: At faster scan speeds, more excited molecules, tion at a different wavelength (see Fig. 8c). The latter cleaves
which are in their vulnerable intermediate state, are left be- into two weakly reactive radicals which can efficiently scav-
hind the writing spot. These molecules can continue the enge the initiating radicals and also terminate propagating
chemical reaction undisturbed. At slower scan speeds, the radical polymer chains. Reference [25] has shown that this
same molecules in the intermediate state would have been mechanism can reduce the polymerization rate by up to a
partially depleted. This means that for faster scan speed the factor of 5. Thus, the mechanism can prevent the formation
effective dose increases, because the overall depletion is of a cross-linked insoluble polymer network.
reduced. This contribution overcompensates the expected The spectra of the compounds have to be chosen such
decrease in dose due to the smaller exposure time. Combin- that they can be excited separately with little cross-talk. In
ing such photoiniators (with proportional linewidth scaling) the original implementation of 2PII, the photoinitiating sys-
with common photoinitiators can yield a photoresist with tem camphorquinone + ethyl 4-(dimethylamino)benzoate
nearly scan-speed-independent linewidths [33]. is used and excited at 473 nm wavelength near its absorp-
This highly sensitive depletion mechanism allows for tion peak (see Fig. 8c). The activation of the photoinhibitor
low depletion powers and the use of low-cost cw laser (tetraethylthiuram disulfide) is performed with 364 nm radi-
diodes as depletion lasers. Low-power depletion is also ation, where the photoinitiator has an absorption minimum
beneficial because limiting effects (one-photon absorption, (see Fig. 8c). The resulting reduction in polymerization rate
two-photon absorption, and excited-state-absorption of the under these conditions is shown in Fig. 11 [25].
depletion laser) are less pronounced. Even wide-field paral-
lel exposures with tailored excitation and depletion patterns
to reduce linewidths seem to be in reach [33]. With multiple 5.3.2. Properties and limitations
exposures, the diffraction limit could again be circumvented.
Finally, the proportional linewidth scaling intuitively The use of one-photon absorption has distinct advantages
implies that the lifetime of the vulnerable intermediate state and disadvantages. In the original implementation of 2PII,
is at least on the scale of the effective exposure time used (or one-photon absorption is used for the excitation of both the
even longer). This assumption is in agreement with the nu- initiator and the inhibitor. This results in very low power
merical modeling in [33]. Therein, a decay time of around requirements in the 10 µW range that can easily be deliv-
1000 time units and exposure times between 20 and 50 ered by low-cost cw laser diodes. In contrast, two-photon


C 2014 The Authors. Laser & Photonics Reviews published by Wiley-VCH Verlag GmbH & Co. KGaA Weinheim www.lpr-journal.org
REVIEW

18638899, 2013, 1, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/lpor.201100046, Wiley Online Library on [10/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
ARTICLE

Laser Photonics Rev. 7, No. 1 (2013) 35

Figure 11 Effect of photo-inhibition on pho-


topolymerization rate in 2PII. Initial methacrylate
polymerization rate versus UV depletion inten-
sity during constant visible excitation irradiation
(15 mW/cm2 ). Reproduced from [25].

absorption usually demands for 100 fs pulses with < 10 mW low-viscosity photoresists with pronounced diffusion of the
average power that can be readily delivered by fs-fiber os- active species, this unique feature can be a major advantage.
cillators, yet at a cost more than ten times higher. The time constant for inhibition with 2PII is (co-) de-
One drawback of one-photon absorption is that it is termined by the lifetime of the activated inhibiting radi-
difficult if not impossible to fabricate complex 3D struc- cals. Once the depletion light is switched off, those can
tures – even if the photoresist offers sufficient nonlinearity recombine and again form the initial parent molecule. The
to provide 3D operation without two-photon absorption (see recombination reaction can be very fast, depending on the
Sect. 1). When focusing deeper into the photoresist volume, viscosity of the photoresist formulation (which affects the
both the excitation and depletion light is attenuated on its diffusion speed) and on the concentration of the activated
way through the absorbing material. Hence, the laser powers inhibitor (as the recombination is bimolecular). The concen-
to be applied are strongly depth-dependent. Furthermore, tration and, hence, the time constant, changes throughout
this can lead to a consumption of the initiator and/or in- the depletion profile. For tetraethylthiuram disulfide in low-
hibitor and to a continuous heating of the photoresist inside viscosity acetonitrile, Reference [38] determined a time con-
the complete cone of the focused laser beams. In contrast, stant around 2 µs. This value likely provides a lower bound.
using two-photon absorption for excitation, the laser light Reference [39] has estimated an upper bound of 200 ms for
is only absorbed within the focal volume. In addition, us- the same photoinhibitor in a higher-viscosity photoresist.
ing STED or RAPID, the depletion light is also exclusively
absorbed by molecules in their intermediate state (hence,
within the focal volume). Thus, the laser powers to be ap- 5.4. Alternative depletion pathways
plied are rather depth-independent, the anticipated photore-
sist heating is much less pronounced, and no consumption Apart from the mechanisms described so far, alternative
of one species is expected. depletion mechanisms based on excited-state absorption are
Applying two-photon absorption to both the photoacti- conceivable and might compete with the above mechanisms.
vation and the photoinhibition in 2PII would resolve these For example, the polymerization-inhibition effect observed
obstacles and provide full 3D capability. In this case, the for isopropyl thioxantone (ITX) [26] was originally ascribed
depletion laser would likely be centered around 800 nm to stimulated emission. Later it was shown that the transient
wavelength. Unfortunately, the shape of the minimum of a absorption spectra of ITX are dominated by excited-state
donut mode gets much broader (and hence less attractive for absorption [35]. These experiments are analogous to those
depletion DLW) when the intensity profile is squared (see, depicted for DETC in Fig. 9. In contrast to DETC, however,
e. g., Fig. 1 for comparison). the observed pump-induced change in optical density of
In contrast to the above approaches, 2PII does not only ITX in ethanol solution was found to be mainly positive
confine the effective excitation but also directly tackles the and subsequent analysis showed that both the S1 state and
blurring of the excitation pattern due to diffusion of initi- the subsequently occupied T1 state of ITX contribute to this
ating and propagating radicals. Once those species diffuse absorption [35].
(or propagate) out of the anticipated polymerization volume, Historically, some authors had likewise speculated that
the probability of their termination increases. Especially in the fluorescence depletion effect in STED microscopy might

www.lpr-journal.org 
C 2014 The Authors. Laser & Photonics Reviews published by Wiley-VCH Verlag GmbH & Co. KGaA Weinheim
LASER & PHOTONICS

18638899, 2013, 1, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/lpor.201100046, Wiley Online Library on [10/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
REVIEWS

36 J. Fischer and M. Wegener: 3D optical lithography off limits

Table 1 Summary of recent lithography experiments on depletion-DLW approaches. Gray shaded fields are experimental results.
Excitation wavelength λexc and numerical aperture (NA) of the publications are stated and used to determine diffraction-limited critical
distances in the lateral and the axial direction. For this purpose, the generalized Sparrow criterion described in Sect. 3 was numerically
evaluated; two-photon absorption was taken into account in cases where it was used. Experimental demonstrations for resolutions can
be directly compared to the value of this diffraction limit (diff. limit). Demonstrations of voxels (V) and lines (L) are listed separately.
Although feature sizes are not directly limited by diffraction, the reader may relate obtained feature sizes to each other by comparing
the corresponding diffraction limits. The aspect ratio is voxel height divided by voxel width.

actually not be caused by stimulated emission but rather range of many µs to ms, which would significantly limit the
be mediated by excited-state absorption [40]. Although it accessible DLW writing speeds (see Section 4.4).
turned out that STED is the far dominating mechanism in Broadly speaking, when aiming at better and better res-
STED microscopy, the proposed processes can still play an olutions, large laser intensities will be required. Any sort of
important role in depletion lithography. The corresponding absorptive process will eventually introduce heat into the
processes for lithography are illustrated in Fig. 7. Excited- system and may lead to (unwanted) temperature-dependent
state absorption might take place directly from the S1 state effects. In contrast, stimulated emission (followed by rapid
or, alternatively, from the triplet. Even the generated radicals relaxation from the S0∗ -state) carries energy out of the sys-
might absorb depletion light and contribute (not depicted tem (see Sect. 5.1.2).
in Fig. 7). Following light absorption, the highly excited
molecules might immediately relax non-radiatively to the
ground-state (Fig. 7a) in which case their contribution to the 6. Review of lithography experiments
exposure would be lost.
Another conceivable mechanism is that molecules are This section summarizes experimental results obtained by
excited several times from an excited state and quickly re- any form of depletion DLW in chronological order. So far,
lax to the same excited state non-radiatively (see Fig. 7b). only four groups worldwide (Fourkas [24], McLeod [25],
In this case, significant heat might be imposed onto the Wegener [26, 27], and Gu [28]) have published corre-
system and thermal effects could play the dominant role. sponding lithography results in journal papers. The papers
For example, the two-photon absorption cross-section of a show systematic improvements in lateral [25–28] and ax-
molecule can decrease considerably with increasing temper- ial [24, 27] size of single voxels [24–28] and lines [25–28].
ature (e. g., 50% reduction from 20 °C to 110 °C for Dis- So far, only one publication has addressed the resolution
perse Red 19 [41]). In this case, the depletion laser might in the sense of Abbe (see Sect. 2) and has broken the Abbe
quickly heat up the exposure volume, subsequent excitation criterion as well as the more appropriate generalized two-
pulses would find photoinitiator molecules with reduced photon Sparrow criterion (see Sect. 3) in the lateral and in
two-photon-absorption cross-sections, and the effective ex- the axial direction [27].
posure dose would drop below threshold. Alternatively, the Table 1 summarizes experimental results of these pub-
intersystem-crossing quantum yield might also decrease lications (first column). The exposure or excitation wave-
with increasing temperature due to temperature-dependent lengths and numerical apertures used are also stated (second
non-radiative decay from the S1 state. Further mechanisms and third column) to allow for a fair comparison. Columns
are conceivable as well. Such temperature-mediated mecha- four and six show the measured lateral and axial spatial reso-
nisms might lead to fairly low depletion-power requirements. lutions. For direct comparison, the fifth and seventh columns
However, the corresponding time constants are likely in the quote the corresponding generalized Sparrow criteria that


C 2014 The Authors. Laser & Photonics Reviews published by Wiley-VCH Verlag GmbH & Co. KGaA Weinheim www.lpr-journal.org
REVIEW

18638899, 2013, 1, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/lpor.201100046, Wiley Online Library on [10/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
ARTICLE

Laser Photonics Rev. 7, No. 1 (2013) 37

Figure 12 (online color at: www.lpr-journal.org) Results obtained with RAPID lithography: (a–f) SEM images of voxels created with
deactivation beam powers of 0 mW, 17 mW, 34 mW, 50 mW, 84 mW, and 100 mW, respectively. (g) Three-dimensional and contour AFM
images of the smallest voxel that have been created with RAPID lithography. (h) Corresponding images of the smallest voxel that could
be created with conventional DLW. The x and y dimensions of the voxels in (g) and (h) are exaggerated due to the width of the AFM tip,
whereas the z dimension (height) is accurate. (i) Dependence of the height and aspect ratio of voxels on the power of the deactivation
beam. The error bars represent ±1 SD based on measurements of four voxels. (j) Tower with rings created with conventional DLW.
(k) Tower with rings created with RAPID lithography. Reproduced from [24].

have been determined using Debye theory (see Sect. 3). For structures are aimed at, the axial resolution is the limit-
publications using two-photon absorption, the intensity dis- ing one. In [24] single voxels were fabricated close to the
tribution has been squared in these calculations. The refrac- substrate-photoresist interface and the fallen voxels were
tive index was chosen as n = 1.52 for all calculations. The characterized by scanning-electron microscopy (SEM) and
obtained improved feature sizes (columns eight and nine) atomic-force microscopy (AFM) (see Fig. 12). When in-
are not directly relatable to the calculated diffraction-limited creasing the depletion power, the aspect ratio of the fab-
critical distances as we have extensively argued in Sect. 2. ricated voxels decreased systematically from > 3 to 0.5
Finally, the measured reduced aspect ratios, i. e., the ratios (Fig. 12a–f and i). The axial extents of the smallest voxels
of axial to lateral feature size, are summarized in column ten. were measured to be 40 nm using AFM (Fig. 12g), equiva-
Where applicable, we have distinguished between results lent to λ /20. The corresponding voxels fabricated without
obtained for voxels (V) and lines (L). the depletion laser had an axial extent of 700 nm. Basic
three-dimensional operation was demonstrated as well (see
Fourkas group 2009 Fig. 12j–k). The axial resolution in the sense of Abbe was
In 2009, the Fourkas group and the McLeod group simul- not reported.
taneously published their approaches [24, 25]. In [24], the
Fourkas group introduced RAPID lithography (Section 5.2) McLeod group 2009
and addressed the axial resolution in two-photon DLW by In the second early publication [25], the McLeod group in-
using a bottle-beam for depletion (compare Fig. 4f–j). The troduced the 2PII scheme (see Sect. 5.3) utilizing a photoin-
aspect ratio in two-photon DLW is usually between 2.5 and hibitor for spatial confinement. A donut mode (see Fig. 4a–
5 [24, 42]. Consequently, once complex three-dimensional d) was used to reduce the lateral extent of single voxels

www.lpr-journal.org 
C 2014 The Authors. Laser & Photonics Reviews published by Wiley-VCH Verlag GmbH & Co. KGaA Weinheim
LASER & PHOTONICS

18638899, 2013, 1, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/lpor.201100046, Wiley Online Library on [10/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
REVIEWS

38 J. Fischer and M. Wegener: 3D optical lithography off limits

Figure 13 (online color at: www.lpr-journal.org)


Results obtained with photoinhibition lithography:
Scanning electron micrographs of polymerized fea-
tures. (a) Voxels polymerized on a microscope
slide using a 0.45-NA singlet lens and the coin-
cident Gaussian blue / donut-shaped UV irradia-
tion scheme, observed at 45◦ and normal to the
slide surface. The blue excitation power was held
constant at 10 mW while the UV depletion power
was progressively increased. The depletion power,
from left to right, was 0, 1, 2.5, 10, and 100 mW.
The exposure time was 8 s for each dot. Scale
bars, 10 µm. (b) Profile of a voxel similarly fabri-
cated but with 10 mW of blue power and 110 mW of
UV focused at NA = 1.3, then imaged via SEM at
normal incidence. The SEM intensity on the white
line (inset) is plotted as squares against the ex-
pected polymerization profile (green) obtained by
a double-parameter fit of the initiation (blue) and in-
hibition (violet) rate profiles, as shown. (c) Polymer
column fabricated by using the same conditions
as Fig. 13b. The focus was translated normal to
the glass slide at a velocity of 0.125 µm/s for 3 µm.
Scale bar, 200 nm. Reproduced from [25].

fabricated at the substrate-photoresist interface. Using a


singlet lens with NA = 0.45, the lateral extent of the fabri-
cated voxels could be reduced from 3.6 µm down to 200 nm
(Fig. 13a). This impressive improvement allows for the use
of inexpensive low-NA lenses with large working distances
while maintaining the resolution of higher-NA objective
lenses. Using a lens with NA = 1.3, even smaller features
were fabricated. The smallest voxels had a full width of
110 nm [25] (Fig. 13b). Scanning the beams at a scan veloc-
ity of 0.125 µm/s yielded a connected line with a width of
150–200 nm (Fig. 13c). The lateral resolution in the sense
of Abbe was not reported.

Wegener group 2010

The Wegener group [26] introduced another depletion-


DLW approach based on an ITX-containing photoresist
(see Sect. 5.4). The observed polymerization suppression
was found to be consistent with stimulated emission [26].
Later, it was found that stimulated emission is unlikely to be
the dominant mechanism [35]. Nevertheless, the reversible
polymerization inhibition could be used to reduce the lat-
eral width of fabricated polymer lines from 155 nm down to
65 nm, again using a donut mode (see Fig. 14). Increasing
the depletion power beyond the corresponding optimal value
led to increasing feature sizes, probably due to two-photon Figure 14 (online color at: www.lpr-journal.org) Results obtained
absorption of the depletion laser. A scan speed of 100 µm/s with ITX-based depletion DLW: Polymer linewidth versus power
was used. Moreover, it was shown that the inhibition effect of the continuous-wave 532 nm wavelength depletion beam that
is reversible for a time-delay of 200 ms [26]. This means is focused to a donut-shaped mode. The power of the two-photon
that the photoresist does completely recover after a depleted excitation beam centered around 810 nm wavelength is fixed to
exposure and the same regions can be re-exposed without 13.5 mW. The subpanels exhibit electron micrographs illustrating
any loss in sensitivity. The lateral resolution in the sense of the raw data underlying the data points. A scan speed of 100 µm/s
Abbe was not reported. was used. Reproduced from [26].


C 2014 The Authors. Laser & Photonics Reviews published by Wiley-VCH Verlag GmbH & Co. KGaA Weinheim www.lpr-journal.org
REVIEW

18638899, 2013, 1, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/lpor.201100046, Wiley Online Library on [10/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
ARTICLE

Laser Photonics Rev. 7, No. 1 (2013) 39

Figure 15 Results obtained with DETC-based depletion DLW: Electron micrographs of simple line gratings fabricated via regular
DLW ((a) and (c)) and depletion DLW ((b) and (d)). The center-to-center distance of the lines is axy = 200 nm and axy = 175 nm as
indicated within the panels. While regular DLW in (a) does not yield a well-defined grating at axy = 200 nm and suffers from polymer
clusters bridging the gaps, the quality of the grating fabricated via depletion DLW shown in (b) is excellent. For axy = 175 nm, regular
DLW does not allow for the fabrication of gratings with lines that are clearly separated from the substrate. Chosing the z-position of
the focus within the substrate, a periodic (yet very flat) height variation can be obtained (see (c)). In contrast, using depletion DLW,
elevated and simultaneously separated lines are still possible with reasonable quality as shown in (d). The depletion power of the donut
mode used is 50 mW in front of the microscope-objective-lens entrance pupil. Reproduced from [27].

Wegener group 2011 Fig. 16 have been used as test structures. Cross-sections of
fabricated structures are shown in Fig. 17a–f. Here, the axial
In [27], the Wegener group published a modified re- as well as the lateral extent decreased with increasing de-
sist formulation based on DETC (see Sect. 5.1). Later, pletion power (also see panel (g)). Furthermore, the aspect
they showed [35, 36] that this formulation supports true ratio of fabricated lines could be reduced from 2.5 to 1.4.
stimulated-emission depletion as well as another depletion Instead of using isolated voxels close to the substrate sur-
mechanism yet to be determined. Using a cw depletion laser face [24], polymerized lines inside complex 3D structures
(and hence, triggering both depletion channels simultane- were used for the evaluation [27]. This demonstrated the
ously), increased lateral and axial resolution in the sense full 3D capability of the approach. Furthermore, shrinkage-
of Abbe was demonstrated. A scan speed of 100 µm/s was based effects were ruled out since lines suffering from heavy
used throughout this publication. shrinkage would not hold complex structures like shown in
Using a donut mode for the lateral confinement, grat- Fig. 17a–f. To further ensure that the electron beam did not
ings with periods down to 175 nm were successfully alter the line shape during imaging, the developed polymer
demonstrated (see Fig. 15). The simple Abbe limit states structures were filled with ZnO via standard atomic-layer
axy = λ /(2 NA) = 289 nm but is not suited for the sequen- deposition (ALD) [27]. The resulting composite structures
tial two-photon exposure (see Sect. 2 and 3). The more suit- were opened via focused-ion-beam milling. During this step,
able generalized two-photon Sparrow criterion leads to a the initial polymer structures were partly calcined, such that
lower value of axy = 203 nm (see Table 1, compare Fig. 3c). essentially holes in ZnO blocks remained. In Fig. 17, the
To the best of our knowledge, both criteria were broken for
the first time in this publication [27]. All panels in Fig. 15
correspond to best results from a fine excitation-power
sweep in 1% steps and a z-position sweep (not shown).
Here, depletion DLW was better than the best regular DLW.
The gratings with axy = 200 nm and axy = 175 nm fab-
ricated using regular DLW (Fig. 15a,c) are still clearly mod-
ulated, yet not sufficiently separated. This means, that the
generalized two-photon Sparrow criterion – although most
suitable – is no sharp fundamental limitation. Otherwise,
we would expect to see no modulation at all, as the accu-
mulated exposure dose does not have local minima. Since
a “forgetting” photoresist can easily yield gratings below Figure 16 (online color at: www.lpr-journal.org) Scheme of a
the diffraction limit (see Sect. 2) this is not too surprising. three-dimensional woodpile photonic crystal [43]. The woodpile is
Obviously, the resist used in [27] is only “slightly forgetting” composed of a first layer (gray) of periodically arranged dielectric
and it is not possible to overcome the diffraction limit with rods with rod spacing a, a second orthogonal layer (blue), a third
regular DLW in a high-quality fashion. To the best of our layer (red) laterally displaced by half the rod spacing with respect
knowledge, no other “forgetting” photoresist formulation to the first layer, and a fourth layer (green) displaced with respect
has yet broken the diffraction-limit in regular DLW. to the second layer. This unit cell is repeated along the axial
In a second set of experiments, the axial resolution direction. For a face-centered-cubic√ (fcc) woodpile, the resulting
was addressed using a bottle-beam focus for depletion [27]. axial lattice constant is given by 2 · a. In a woodpile, the smallest
For this purpose, woodpile photonic crystals illustrated in axial distance az equals 3/4 of this axial lattice constant.

www.lpr-journal.org 
C 2014 The Authors. Laser & Photonics Reviews published by Wiley-VCH Verlag GmbH & Co. KGaA Weinheim
LASER & PHOTONICS

18638899, 2013, 1, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/lpor.201100046, Wiley Online Library on [10/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
REVIEWS

40 J. Fischer and M. Wegener: 3D optical lithography off limits

Figure 17 (online color at: www.lpr-journal.org) Improved aspect ratio via DETC-based depletion DLW: (a–f) Oblique-view electron
micrographs of ZnO-filled woodpile photonic crystals after focused-ion-beam milling. The viewing angle with respect to the normal is
54◦ . (g) Width, height and calculated aspect ratio of polymer rods inside the three-dimensional woodpiles ((a)-(f)). Height measurements
have been corrected for the viewing angle. The measurements are averaged over 10 rods. The error bars indicate ± one standard
deviation of the corresponding ensembles. The bars for height and width in (a–f) correspond to the averaged values shown in (g).
Reproduced from [27].

Figure 18 (online color


at: www.lpr-journal.org)
DETC-based depletion
DLW: (a) True-color
reflection-mode optical
micrographs of woodpile
photonic crystals fabri-
cated via regular DLW.
(b) Same, but using deple-
tion DLW. All woodpiles
have 24 layers and a
footprint of 20 µm × 20 µm.
The rod spacing is de-
creased from a = 450 nm
to a = 250 nm along the
vertical axis, the exposure
power is increased in
steps of 1% from left to
right. (c) and (d) Selected
(see asterisks in (a) and
(b)) transmittance (solid)
and reflectance (dashed)
spectra for DLW and de-
pletion DLW, respectively.
Reproduced from [27].

inner ZnO surfaces were used as a measure for the initial ing regular DLW (Fig. 18a,c) and STED-DLW (Fig. 18b,d).
polymer lineshape (see bars in panels (a)-(f)). Each square depicted in the reflection-mode optical micro-
The reduced aspect ratio and axial voxel extent were also graphs (panels a and b) is a woodpile with a footprint of
shown to directly translate to an increased axial resolution in 20 µm × 20 µm and 24 layers in the axial direction. The exci-
the sense of Abbe. Woodpile photonic crystals [5,11,43–46] tation power was increased along the horizontal direction of
with different lateral rod spacings, a, were fabricated us- the upper panels, to ensure a fair comparison with optimum


C 2014 The Authors. Laser & Photonics Reviews published by Wiley-VCH Verlag GmbH & Co. KGaA Weinheim www.lpr-journal.org
REVIEW

18638899, 2013, 1, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/lpor.201100046, Wiley Online Library on [10/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
ARTICLE

Laser Photonics Rev. 7, No. 1 (2013) 41

exposure doses for both approaches. The rod spacing was


varied along the vertical axis.
The smallest axial center-to-center distance between two
separated polymer rods in a woodpile photonic crystal √ is 3
times the layer separation [27], which was chosen as 2 a/4
(see Fig. 16). The finest woodpiles successfully fabricated
via STED-DLW had a = 275 nm, √and correspond to an an-
ticipated axial resolution of az = 2a · 3/4 = 291 nm . How-
ever, the actual writing trajectories were pre-compensated
for shrinkage along the z-direction and therefore separated
by a larger margin. To be conservative, the separation of
the voxels during√exposure was used as resolution measure,
leading to az = 2a · 3/4 · s = 373 nm, where s = 1.28 is
the shrinkage pre-compensation factor. Clearly, this az value
is below the generalized two-photon Sparrow criterion in
the axial direction of az = 506 nm (see Table 1, compare
Fig. 3d). This is the first time the axial diffraction-limit
was broken.
With the depletion laser switched off, the finest attain-

able woodpiles had a = 375 nm, and hence, az = 2a ·
3/4 · s = 510 nm. Again, the generalized two-photon Spar-
row criterion does not seem to be a sharp limit for regu-
lar DLW. However, to our knowledge the presented wood-
piles are the finest woodpiles in DLW literature (com-
pare [11, 45, 46]). This implies that high-quality structures
below the diffraction-limit are not (yet) attainable with reg-
ular DLW and a “forgetting” photoresist.
Moreover, one might be afraid that a photoresist for
depletion DLW comes along with unfavorable properties
limiting its usability or resolution (e. g., low reactivity, low
dynamic range, low mechanical stability of the resulting Figure 19 (online color at: www.lpr-journal.org) Results from a
structures). In this case, depletion DLW would not truly modified photoinhibition lithography approach: (a) The obtained
extend state-of-the-art regular DLW but would first have lateral voxel sizes are plotted as a function of the depletion power.
to compensate for the resist’s shortcomings before a true The voxels were exposed at 200 nW excitation power with ex-
improvement in resolution and quality could be expected. posure times stated in the figure legend. The red curves are
The fact, however, that this special DETC-based photoresist corresponding numerical predictions. (b) Voxel size for different
was used to demonstrate the finest published woodpiles exposure parameters (top) and SEM image of a voxel fabricated
using regular DLW allays these doubts. 200 nW excitation power and 6 µW depletion power at 0.4 s ex-
posure time (bottom). (c) Polymer lines fabricated with different
depletion-power levels (top-down): 0 µW, 1.0 µW, 1.5 µW, and
Gu group 2011
2.0 µW. The scanning speed was 3 µm/s, the excitation power
Recently, the Gu group [28] published systematic experi- was 200 nW. The scale bar is 200 nm. Reproduced from [28].
ments on a modified version of the photoresist from [25].
They used the same photoinitiator and photoinhibitor 7. Applications in 3D structures
but replaced the original monomer by the bi-functional
monomer NK Ester BPE-100 (2.2 Bis[4-(Methacryloxy Eventually, depletion 3D DLW optical lithography needs
Ethoxy)Phenyl]Propane). This leads to a higher photosensi- to demonstrate more than just interesting principles, small
tivity of the photoresist and higher mechanical stability of feature sizes, and test structures on spatial resolution – it
the fabricated lines [28]. The lateral extent of single voxels needs to successfully realize complex functional 3D nanos-
exposed directly at the substrate-photoresist interface could tructures that would be very difficult or even impossible to
be reduced from 500 nm to 40 nm by using appropriate inhi- make with other existing technologies. In particular, deple-
bition laser powers in the range of a few µW. As depicted tion DLW obviously needs to be better than the best existing
in Fig. 19a, the voxel sizes were also predicted numerically regular DLW (see, e. g., [9, 45]).
using a non-steady-state kinetic model. Lateral scanning Any purely periodic 3D nanostructure can likely be
at a velocity of 3 µm/s yielded polymer lines attached to made inexpensively and on a large scale by optical inter-
the substrate surface. The width of these lines was reduced ference lithography. Any simple layer-by-layer structure
from 400 nm down to roughly 130 nm. As in the McLeod can be made by standard electron-beam lithography of the
work [25], the attainable width of lines was larger than the individual layers and successive planarization and stacking
attainable width of single voxels. of layers. Challenges, however, do arise for these and other

www.lpr-journal.org 
C 2014 The Authors. Laser & Photonics Reviews published by Wiley-VCH Verlag GmbH & Co. KGaA Weinheim
LASER & PHOTONICS

18638899, 2013, 1, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/lpor.201100046, Wiley Online Library on [10/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
REVIEWS

42 J. Fischer and M. Wegener: 3D optical lithography off limits

technologies in case of seemingly simple problems like the top. Ideally, the structure should appear as a flat hence
achieving free-form surfaces in three-dimensional space or unsuspicious metal mirror. Most importantly in the context
for three-dimensional architectures which are intentionally of this review, the structure contains free-form surfaces (the
non-periodic. The so-called carpet invisibility cloak to be carpet) and intentional complex non-periodicities (the index
described in the remainder of this article comprises all of profile) at the same time. Using regular 3D DLW, opera-
these difficulties at the same time and, hence, represents an tion at 1.4 µm wavelength became possible in 2010, but
interesting and challenging test case with a specific function. visible operation wavelengths were out of reach for regular
Indeed, depletion 3D DLW has enabled the first (and so far 3D DLW.
also the only) three-dimensional polarization-independent Fig. 21a and b exhibit electron micrographs of a struc-
visible-frequency broadband carpet invisibility cloak. ture fabricated using depletion 3D DLW with the DETC-
Invisibility cloaking can be viewed as a demanding based photoresist (see Sects. 5.1 and 6) using a bottle-beam
benchmark example for the much broader and far-reaching for depletion [51]. (The fabricated structures are oriented up-
ideas of a design approach named transformation op- side down compared to Fig. 21.) Along these lines, a minia-
tics [47, 48]. Transformation optics connects geometry of turization from a woodpile rod spacing of a = 800 nm [50]
curved space and propagation of light in inhomogeneous op- (by regular 3D DLW) to a = 350 nm became possible [51].
tical media [47, 48]. Any imaginable distortion of space can At this rod spacing, the volume filling fraction can still be
be mapped onto spatially inhomogeneous optical materials, tuned over a considerable range (compare different Bragg-
which generally need to be anisotropic magneto-dielectrics. reflection colors for a = 350 nm in Fig. 18b) whereas regular
A simple yet impressive example is the carpet cloak illus- DLW can not realize a single filling fraction in the same
trated in Fig. 20. An object can be placed underneath the photoresist (Fig. 18a). To realize the refractive-index gra-
bump in the metallic floor. Of course, the bump leads to dients of the target distribution, the excitation laser power
an extreme distortion (or aberration) of the reflection an
observer sees from above, which would immediately raise
suspicions that something is hidden underneath. These dis-
tortions can be approximately reversed or compensated by
tailoring the locally isotropic refractive index of the sur-
rounding. The refractive index on top of the bump needs
to be higher than in the surrounding to compensate for the
shorter optical path length, the refractive index on the two
sides needs to be correspondingly lower. The false-color rep-
resentation in Fig. 20 illustrates the target refractive-index
profile, which can be mimicked by locally adjusting the vol-
ume filling fraction of a polymer woodpile photonic crystal
used in the long-wavelength limit [50]. The performance
of the structure can simply be tested by looking at it from

Figure 21 (online color at: www.lpr-journal.org) (a) Colored


oblique-view electron micrograph of the polymer reference (top)
and carpet cloak (bottom) structures (fabricated on glass sub-
strate and coated with 100 nm gold). The scale bar corresponds
to 10 µm. (b) Corresponding focused-ion-beam cuts of nominally
identical structures. The scale bar corresponds to 2 µm. (c) and
Figure 20 (online color at: www.lpr-journal.org) A three- (d) are true-color optical micrographs of the structures in (a) taken
dimensional carpet cloak. Looking through a spatially inhomo- with an optical microscope under circularly polarized illumination
geneous region of the woodpile photonic crystal shown here, a at 700 nm wavelength. Note the identical distortions due to the
bump in the gold film – the carpet – appears as a flat metal mirror. bump in both structures in (c) when inspected from the air side
The local refractive-index profile needed to create the illusion is (serving as a control experiment). When inspected from the glass-
calculated using the laws of transformation optics and can be real- substrate side in (d), the reference structure (top) still shows
ized by adjusting the local volume-filling fraction of the polymeric pronounced dark stripes. In sharp contrast, the stripes essen-
woodpile that leads to locally nearly isotropic optical properties. tially disappear for the cloaking structure (bottom). Reproduced
Reproduced from [49]. from [51].


C 2014 The Authors. Laser & Photonics Reviews published by Wiley-VCH Verlag GmbH & Co. KGaA Weinheim www.lpr-journal.org
REVIEW

18638899, 2013, 1, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/lpor.201100046, Wiley Online Library on [10/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
ARTICLE

Laser Photonics Rev. 7, No. 1 (2013) 43

was varied continuously while the depletion power was held what was believed for decades. In this sense, this type of
constant. This leads to a gradually changing effective re- optical laser lithography is “diffraction-unlimited”. Success-
fractive index. The resulting 3D structure is so complex ful photoresist development will likely require interdisci-
that usual characterization techniques such as electron mi- plinary efforts comprising physics, chemical physics, photo-
croscopy literally only scratch the surface. However, the chemistry, and polymer chemistry. Many researchers con-
invisibility-cloaking performance of the complete 3D de- sider photoresists an issue that has long reached industrial
vice can easily be tested by optical microscopy. We note status. However, corresponding basic research regarding
that the supplementary material of [50] explicitly shows that diffraction-unlimited optical lithography has just started.
the performance is very sensitive with respect to both, the
high-index and the low-index regions. Acknowledgements. We thank Tolga Ergin for fruitful discus-
Fig. 21c and d show true-color optical micrographs taken sions and help regarding the figure preparation, Andreas Frölich
at 700 nm wavelength of light. When inspected through the for the ZnO ALD, Georg von Freymann, Jonathan Müller, Andreas
glass substrate (see Fig. 21a), the bump without cloak (i. e., Naber, and Michael Thiel for helpful discussions. We acknowl-
the homogeneous woodpile) leads to two pronounced dark edge support by the Deutsche Forschungsgemeinschaft (DFG),
stripes in the image (Fig. 21d, top). These aberrations essen- the State of Baden-Württemberg, and the Karlsruhe Institute of
tially disappear in presence of the cloak (Fig. 21d, bottom). Technology (KIT) through the DFG Center for Functional Nanos-
The device is polarization insensitive [51], works for a large tructures (CFN) within subprojects A1.4 and A1.5. The project
range of angles in three-dimensional space [52], is spec- METAMAT is supported by the Bundesministerium für Bildung
trally broadband from about 600 nm wavelength to about und Forschung (BMBF). The PhD education of J. F. is embedded
3 µm, and not only works for the amplitude of the light wave in the Karlsruhe School of Optics & Photonics (KSOP).
but also for its phase [53]. Some years ago, such cloaking Received: 8 December 2011, Revised: 9 January 2012,
behavior was commonly believed impossible. Accepted: 31 January 2012
Published online: 20 March 2012

8. Conclusions and future challenges Key words: Direct laser writing, optical lithography, superresolu-
tion, stimulated emission, photoresist.
In optical microscopy, STED-based and STED-inspired ap-
proaches have revolutionized the field and have opened new
research possibilities, especially in life-science applications.
Joachim Fischer received his physics
Fluorescence optical microscopes on this basis have be-
diploma in 2008 from the Universität Karl-
come commercially available. The Abbe diffraction limit
sruhe (TH) in Germany. His diploma the-
is no longer the relevant limit and record spatial resolu- sis describes the fabrication and char-
tions of 5.6 nm have been achieved at visible frequencies. acterization of three-dimensional flexi-
If such spectacular resolutions could be translated from op- ble micro-scaffolds for cell-culture exper-
tical microscopy to optical lithography, many further new iments. Since then, he is working on
possibilities would arise. A revolution in nanotechnology his PhD project at the Karlsruhe Insti-
might result. tute of Technology (KIT). His research
In this review, we have described the underlying princi- interests include three-dimensional super-
ples and have presented a snapshot of the current experimen- resolution laser-lithography and its applications.
tal state of the art. Today, several groups have convincingly
demonstrated depletion effects in direct-laser-writing opti- Martin Wegener After completing his
cal lithography as well as reduced feature sizes in the lateral PhD in physics in 1987 at Johann Wolf-
and in the axial direction. However, much less experimen- gang Goethe-Universität Frankfurt (Ger-
tal work has actually shown spatial resolutions – in Ernst many), he spent two years as a postdoc at
Abbe’s sense of resolving a grating with a certain period AT&T Bell Laboratories in Holmdel (USA).
– beyond the diffraction limit. For applications aiming at From 1990–1995 he was professor at Uni-
complex 3D nanostructures, nanostructured effective ma- versität Dortmund, since 1995 he is pro-
terials, or 3D data storage [37, 54], it is spatial resolution fessor at Universität Karlsruhe (TH), now
that matters and not just linewidth or feature size. Today, Karlsruhe Institute of Technology (KIT).
the diffraction limit has been beaten both in the lateral and Since 2001 he has a joint appointment at Institut für Nan-
in the axial direction in optical lithography, too, but only by otechnologie. Since 2001 he is also the coordinator of the
a margin much smaller than in optical microscopy. DFG-Center for Functional Nanostructures (CFN) in Karlsruhe.
His research interests comprise ultrafast optics, (extreme) non-
To further progress towards lithographic resolutions of
linear optics, photonic crystals, photonic metamaterials, trans-
just a few tens of nanometers in three dimensions, new pho-
formation optics, and optical lithography. This research has
toresist systems featuring, e. g., reduced radical diffusion
led to various awards and honors, among which are the Al-
and reduced unwanted absorption at the depletion wave- fried Krupp von Bohlen und Halbach Research Award 1993,
length need to be developed [26, 55].While the optics part the Baden-Württemberg Teaching Award 1998, the DFG Got-
still deserves further improvements, it is clear that optics tfried Wilhelm Leibniz Award 2000, the European Union René
does not impose any fundamental limits – in contrast to

www.lpr-journal.org 
C 2014 The Authors. Laser & Photonics Reviews published by Wiley-VCH Verlag GmbH & Co. KGaA Weinheim
LASER & PHOTONICS

18638899, 2013, 1, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/lpor.201100046, Wiley Online Library on [10/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
REVIEWS

44 J. Fischer and M. Wegener: 3D optical lithography off limits

[27] J. Fischer and M. Wegener, Opt. Mater. Express 1, 614–624


Descartes Prize 2005, the Baden-Württemberg Research
(2011).
Award 2005, and the Carl Zeiss Research Award 2006. He is
[28] Y. Cao, Z. Gan, B. Jia, R. A. Evans, and M. Gu, Opt. Express
a member of Leopoldina, the German Academy of Sciences
19, 19486–19494 (2011).
(since 2006), Fellow of the Optical Society of America (since [29] J. Keller, A. Schönle, and S. W. Hell, Opt. Express 15, 3361–
2008), Fellow of the Hector Foundation (since 2008), and ad- 3371 (2007).
junct professor at the Optical Sciences Center, Tucson, USA [30] S. A. Pruzinsky and P. V. Braun, Adv. Funct. Mater. 15, 1995–
(since 2009). 2004 (2005).
[31] F. Morlet-Savary, X. Allonas, C. Dietlin, J. P. Malval, and
J. P. Fouassier, J. Photochem. Photobiol. A 197, 342–350
(2008).
References [32] C. S. Colley, D. C. Grills, N. A. Besley, S. Jockusch, P. Ma-
tousek, A. W. Parker, M. Towrie, N. J. Turro, P. M. W. Gill,
[1] T. C. Chong, M. H. Hong, and L. P. Shi, Laser Photonics Rev. and M. W. George, J. Am. Chem. Soc. 124, 14952-14958
4, 123–143 (2010). (2002).
[2] H.-B. Sun, S. Matsuo, and H. Misawa, Appl. Phys. Lett. 74, [33] M. P. Stocker, L. Li, R. R. Gattass, and J. T. Fourkas, Nature.
786–788 (1999). Chem 3, 223–227 (2011).
[3] S. Kawata, H.-B. Sun, T. Tanaka, and K. Takada, Nature 412, [34] D. P. Specht, P. A. Martic, and S. Farid, Tetrahedron 38, 1203–
697–698 (2001). 1211 (1982).
[4] M. Straub and M. Gu, Opt. Lett. 27, 1824–1826 (2002). [35] T. Wolf, J. Fischer, M. Wegener, and A.-N. Unterreiner, Opt.
[5] M. Deubel, G. von Freymann, M. Wegener, S. Pereira, Lett. 36, 3188–3190 (2010).
K. Busch, and C. M. Soukoulis, Nature Mater. 3, 444–447 [36] J. Fischer and M. Wegener, Adv. Mater. 24(10), OP65–OP69
(2004). (2012), doi:10.1002/adma.201103758.
[6] C. N. LaFratta, J. T. Fourkas, T. Baldacchini, and R. A. Farrer, [37] R. R. McLeod, B. A. Kowalski, and M. C. Cole, Proc. SPIE
Angew. Chem. Int. Ed. 46, 6238–6258 (2007). 7591, 759102 (2010).
[7] S. Maruo and J. T. Fourkas, Laser Photonics Rev. 2, 100–111 [38] V. F. Plyusnin, E. P. Kuznetzova, G. A. Bogdanchikov,
(2008). V. P. Grivin, V. N. Kirichenko, and S. V. Larionov, J. Pho-
[8] K.-S. Lee, R. H. Kim, D.-Y. Yang, S. H. Park, Prog. Polym. tochem. Photobiol. A, Chem. 68, 299–308 (1992).
Sci 33, 631–681 (2008). [39] B. A. Kowalski, T. F. Scott, C. N. Bowman, A. C. Sullivan,
[9] G. von Freymann, A. Ledermann, M. Thiel, I. Staude, S. Es- and R. R. McLeod, Proc. SPIE 7053, 70530E (2008).
sig, K. Busch, and M. Wegener, Adv. Funct. Mater. 20, 1038– [40] E. Rittweger, B. R. Rankin, V. Westphal, and S. W. Hell,
1052 (2010). Chem. Phys. Lett. 442, 483–487 (2007).
[10] Y.-L. Zhang, Q.-D. Chen, H. Xia, H.-B. Sun, Nano Today 5, [41] L. De Boni, E. Piovesan, L. Misoguti, S. C. Zilio, and
435–448 (2010). C. R. Mendonca, J. Phys. Chem. A 111, 6222–6224 (2007)
[11] M. Thiel, J. Fischer, G. von Freymann, and M. Wegener, Appl. [42] H. B. Sun, K. Takada, M. S. Kim, K. S. Lee, and S. Kawata,
Phys. Lett. 97, 221102 (2010). Appl. Phys. Lett. 83, 1104 (2003).
[12] M. Campbell, D. N. Sharp, M. T. Harrison, R. G. Denning, [43] K.-M. Ho, C. T. Chan, C. M. Soukoulis, R. Biswas, and
and A. J. Turberfield, Nature 404, 53–56 (2000). M. Sigalas, Solid State Comm. 89, 413–416 (1994).
[13] D. C. Meisel, M. Wegener, and K. Busch, Phys. Rev. B 70, [44] M. Deubel, M. Wegener, S. Linden, and G. von Freymann,
165104 (2004). Appl. Phys. Lett. 87, 221104 (2005).
[14] C. M. Sparrow, Astro J. 44, 76–86 (1916). [45] I. Staude, M. Thiel, S. Essig, C. Wolff, K. Busch,
[15] A. S. van de Nes, L. Billy, S. F. Pereira, and J. J. M. Braat, G. von Freymann, and M. Wegener, Opt. Lett. 35, 1094–1096
Opt. Express 12, 1281–1293 (2004). (2010).
[16] S. W. Hell and J. Wichmann, Opt. Lett. 19, 780–782 (1994). [46] W. Haske, V. W. Chen, J. M. Hales, W. Dong, S. Barlow,
[17] T. A. Klar, S. Jakobs, M. Dyba, A. Egner, and S. W. Hell, S. R. Marder, and J. W. Perry, Opt. Express 15, 3426–3436
Proc. Natl. Acad. Sci. 97, 8206–8210 (2000). (2007).
[18] V. Westphal and S. W. Hell, Phys. Rev. Lett. 94, 143903 [47] U. Leonhardt, Science 312, 1777–1780 (2006).
(2005). [48] J. B. Pendry, D. Schurig, and D. R. Smith, Science 312, 1780–
[19] S. W. Hell, Science 316, 1153–1158 (2007). 1782 (2006).
[20] S. W. Hell, Nature Methods 6, 24–32 (2009); also see the [49] M. Wegener and S. Linden, Physics Today 63, 32–36 (2010).
references cited in this special feature issue. [50] T. Ergin, N. Stenger, P. Brenner, J. B. Pendry, and M. We-
[21] S. W. Hell, R. Schmidt, and A. Egner, Nature Photon. 3, 381– gener, Science 328, 337–339 (2010).
387 (2009). [51] J. Fischer, T. Ergin, and M. Wegener, Opt. Lett. 36, 2059–
[22] K. I. Willig, S. O. Rizzoli, V. Westphal, R. Jahn, and 2061 (2011).
S. W. Hell, Nature 440, 935–939 (2006). [52] T. Ergin, J. Fischer, and M. Wegener, Phys. B,
[23] E. Rittweger, K. Y. Han, S. E. Irvine, C. Eggeling, and doi:10.1016/j.physb.2011.11.035 (2011).
S. W. Hell, Nature Photon. 3, 144–147 (2009). [53] T. Ergin, J. Fischer, and M. Wegener, Phys. Rev. Lett. 107,
[24] L. Li, R. R. Gattass, E. Gershgoren, H. Hwang, and 173901 (2011).
J. T. Fourkas, Science 324, 910–913 (2009). [54] T. Grotjohann, I. Testa, M. Leutenegger, H. Bock, N. T. Ur-
[25] T. F. Scott, B. A. Kowalski, A. C. Sullivan, C. N. Bowman, ban, F. Lavoie-Cardinal, K. I. Willig, C. Eggeling, S. Jakobs,
and R. R. McLeod, Science 324, 913–917 (2009). and S. W. Hell, Nature 478, 204–208 (2011).
[26] J. Fischer, G. von Freymann, and M. Wegener, Adv. Mater. [55] S. W. Hell, Phys. Lett. A 326, 140–145 (2004).
22, 3578–3582 (2010).


C 2014 The Authors. Laser & Photonics Reviews published by Wiley-VCH Verlag GmbH & Co. KGaA Weinheim www.lpr-journal.org

You might also like