ArticlePDF Available

Abstract and Figures

Coherent diffraction imaging (CDI) or lensless X-ray microscopy has become of great interest for high spatial resolution imaging of, e.g., nanostructures and biological specimens. There is no optics required in between an object and a detector, because the object can be fully recovered from its far-field diffraction pattern with an iterative phase retrieval algorithm. Hence, in principle, a sub-wavelength spatial resolution could be achieved in a high-numerical aperture configuration. With the advances of ultrafast laser technology, high photon flux tabletop Extreme Ultraviolet (EUV) sources based on the high-order harmonic generation (HHG) have become available to small-scale laboratories. In this study, we report on a newly established high photon flux and highly monochromatic 30 nm HHG beamline. Furthermore, we applied ptychography, a scanning CDI version, to probe a nearly periodic nanopattern with the tabletop EUV source. A wide-field view of about 15 × 15 μm was probed with a 2.5 μm−diameter illumination beam at 30 nm. From a set of hundreds of far-field diffraction patterns recorded for different adjacent positions of the object, both the object and the illumination beams were successfully reconstructed with the extended ptychographical iterative engine. By investigating the phase retrieval transfer function, a diffraction-limited resolution of reconstruction of about 32 nm is obtained.
Schematic view of the tabletop EUV source for high-numerical aperture ptychography. An 1 kHz Ti:sa amplifier delivers 35 fs–FWHM optical pulses with up to 8 mJ pulse energy and at an 800 nm central wavelength. The IR beam was focused with a 500 mm focal length lens into an 8 mm long gas cell fed with argon gas through a piezo-driven valve at a backing pressure of 1.5 bars. The resulting HHG beam passed from the source through an 1 mm–inner diameter differential pumping tube to the characterisation chamber, and then was characterised with a flat-field EUV spectrometer. To separate the IR beam, a 300 nm–thick Al foil was used as a spectral filter. At about 80 cm downstream from the HHG source, an 1 mm–diameter aperture was inserted as a spatial filter, resulting in a desired illumination profile for ptychography. A single harmonic at 30 nm was selected and focused with a pair of multilayer mirrors in a z-configuration to minimise astigmatism, including a flat bending mirror (M1) and an f2 = 110 mm spherical mirror (M2). A sample with lithographed nanopatterns was mounted on an xyz-translation stage and located at the focus of the 30 nm probe beam. An SEM image of the sample is provided in Fig. 2a. Diffraction patterns of the adjacent areas on the sample were recorded with an in-vacuum X-ray CCD camera at a distance z = 16.5 mm. A second 200 nm–thick Al filter was installed in front of the CCD camera to block the residual stray light. The lateral positioning and data recording were synchronised with a home-built LabVIEW program. Image reconstruction was performed with the extended ptychographical iterative engine (ePIE) on an NVIDIA Tesla K40 computing processor.
… 
This content is subject to copyright. Terms and conditions apply.
1
ScIeNTIfIc REPORTS | (2018) 8:16693 | DOI:10.1038/s41598-018-34257-2
www.nature.com/scientificreports
Coherent Tabletop EUV
Ptychography of Nanopatterns
Nguyen Xuan Truong
1,2, Reza Safaei3, Vincent Cardin3, Scott M. Lewis1, Xiang Li Zhong4,
François Légaré
3 & Melissa A. Denecke1,2
Coherent diraction imaging (CDI) or lensless X-ray microscopy has become of great interest for
high spatial resolution imaging of, e.g., nanostructures and biological specimens. There is no optics
required in between an object and a detector, because the object can be fully recovered from its far-eld
diraction pattern with an iterative phase retrieval algorithm. Hence, in principle, a sub-wavelength
spatial resolution could be achieved in a high-numerical aperture conguration. With the advances
of ultrafast laser technology, high photon ux tabletop Extreme Ultraviolet (EUV) sources based
on the high-order harmonic generation (HHG) have become available to small-scale laboratories. In
this study, we report on a newly established high photon ux and highly monochromatic 30 nm HHG
beamline. Furthermore, we applied ptychography, a scanning CDI version, to probe a nearly periodic
nanopattern with the tabletop EUV source. A wide-eld view of about 15 × 15 μm was probed with a
2.5 μmdiameter illumination beam at 30 nm. From a set of hundreds of far-eld diraction patterns
recorded for dierent adjacent positions of the object, both the object and the illumination beams were
successfully reconstructed with the extended ptychographical iterative engine. By investigating the
phase retrieval transfer function, a diraction-limited resolution of reconstruction of about 32 nm is
obtained.
Since the last three decades, coherent diraction imaging (CDI), also named as lensless X-ray microscopy, has
been of great interest as an alternative to the current state-of-the-art microscopy to achieve the atomic-level
resolution. Conventional X-ray microscopy oen requires multiple extremely precise and pricy optical condens-
ers and deectors, e.g., Fresnel zone plates or multilayer mirrors, which might introduce optical aberrations or
signicantly absorb X-rays. Nevertheless, the highest image resolution achieved with X-ray microscopy is about
10 nm1,2, which is well far away from the diraction-limited resolution. As a simple version of X-ray micros-
copy, CDI is the most ecient way of using photons with the spatial resolution essentially depending only on
the wavelength and the highest scattering angle (numerical aperture). In CDI, the exit-surface wave (ESW) dif-
fracted from an object can be fully recovered in both amplitude and phase from a single diraction pattern
(DP) measured in the far-eld. According to Sayre3, if an experimental DP is suciently oversampled, i.e., at
least twice the Nyquist frequency, the ESW can be reconstructed with the iterative phase retrieval (IPR) algo-
rithms, giving the physical image of the object. A number of IPR algorithms have been introduced to solve the
phase problem for a single experimental DP, such as the error reduction (ER)4, hybrid input-output (HIO)5, and
their modied versions. Basically, an IPR algorithm computes the ESW back and forth between the object- and
Fourier-domains, applying certain known constraints such as the Fourier modulus, support, and non-negativity
constraints5,6. Both ER and HIO algorithms have been widely used in numerous situations715, but they oen
suer from stagnation and trapping in local minima for noisy diraction patterns16,17. Intensive eorts have been
made to improve their performance, leading to the introduction of the noise-robust frameworks, e.g., the relaxed
averaged alternating reections (RAAR)18, noise-robust HIO19, dierence-map20, oversampling smoothness
(OSS)21, and optimisation-based IPR algorithms16,2224. Still, CDI works well only for isolated objects. A scanning
version of CDI, termed as ptychography, has been proposed for wide-eld imaging. In ptychography, multiple
well-overlapping areas of an object are sequentially probed with a probe beam and the corresponding (far-eld)
diraction patterns are measured. With the additional overlap constraint, ptychography is hence more robust
and reliable compared to the conventional CDI, resulting in a higher resolution of reconstruction. Since the rst
1School of Chemistry, The University of Manchester, M13 9PL, Manchester, UK. 2Dalton Nuclear Institute, The
University of Manchester, M13 9PL, Manchester, UK. 3INRS, Energie, Matériaux et Télécommunications, 1650 Bld.
Lionel Boulet, Varennes, Québec, J3X 1S2, Canada. 4School of Materials, The University of Manchester, M13 9PL,
Manchester, UK. Correspondence and requests for materials should be addressed to N.X.T. (email: xuantruong.
nguyen@manchester.ac.uk)
Received: 4 May 2018
Accepted: 17 September 2018
Published: xx xx xxxx
OPEN
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
2
ScIeNTIfIc REPORTS | (2018) 8:16693 | DOI:10.1038/s41598-018-34257-2
demonstrations in the 1990s2528, ptychography has been increasingly applied to study various kinds of samples
including nanostructures and biological cells2933. Recently, for instance, electron ptychographical microscopy has
achieved a sub-nm resolution for 3D-imaging of carbon nanotubes34. A number of iterative phase retrieval frame-
works have been introduced to recover both the object and illumination beam from a ptychographical data set,
including the basic and extended ptychographical iterative engines (PIE35 and ePIE36) and the dierence map30,37.
ere exist a few ptychographical solvers such as ptypy38, PyNX.Ptycho39, and SHARP40.
Due to the demand of a great number of coherent X-ray photons, high-resolution ptychography has been
oen demonstrated at large facilities such as synchrotrons and free electron lasers41,42. anks to the recent
advances in the ultrafast laser technology, coherent tabletop EUV to so X-ray sources based on the high-order
harmonic generation have become achievable in many small-scale laboratories. Future high photon ux and
long-term stable HHG sources might oer a unique tool for developing novel phase retrieval algorithms and
time-resolved CDI, among other things, in laboratories. So far, only a few groups have been able to demonstrate
Figure 1. Schematic view of the tabletop EUV source for high-numerical aperture ptychography. An 1 kHz
Ti:sa amplier delivers 35 fs–FWHM optical pulses with up to 8 mJ pulse energy and at an 800 nm central
wavelength. e IR beam was focused with a 500 mm focal length lens into an 8 mm long gas cell fed with argon
gas through a piezo-driven valve at a backing pressure of 1.5 bars. e resulting HHG beam passed from the
source through an 1 mm–inner diameter dierential pumping tube to the characterisation chamber, and then
was characterised with a at-eld EUV spectrometer. To separate the IR beam, a 300 nm–thick Al foil was used
as a spectral lter. At about 80 cm downstream from the HHG source, an 1 mm–diameter aperture was inserted
as a spatial lter, resulting in a desired illumination prole for ptychography. A single harmonic at 30 nm was
selected and focused with a pair of multilayer mirrors in a z-conguration to minimise astigmatism, including
a at bending mirror (M1) and an f2 = 110 mm spherical mirror (M2). A sample with lithographed nanopatterns
was mounted on an xyz-translation stage and located at the focus of the 30 nm probe beam. An SEM image of
the sample is provided in Fig.2a. Diraction patterns of the adjacent areas on the sample were recorded with
an in-vacuum X-ray CCD camera at a distance z = 16.5 mm. A second 200 nm–thick Al lter was installed
in front of the CCD camera to block the residual stray light. e lateral positioning and data recording were
synchronised with a home-built LabVIEW program. Image reconstruction was performed with the extended
ptychographical iterative engine (ePIE) on an NVIDIA Tesla K40 computing processor.
Figure 2. (a) An SEM image of the nanopatterned sample. (b) A representative diraction pattern of the
sample aer binning 2 × 2 pixels into 1 pixel and performing curvature correction. An almost six orders of
magnitude high dynamic range image was obtained without the use of a beamstop.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
3
ScIeNTIfIc REPORTS | (2018) 8:16693 | DOI:10.1038/s41598-018-34257-2
ptychographical imaging with tabletop EUV sources, mostly at around 30 nm4346 and 13 nm47,48. In an attempt
to achieve a 100 μm–wide eld of view within a single ePIE reconstruction, a polychromatic EUV beam around
29 nm was employed to provide a sucient photon ux44. A breakthrough has been recently made by Gardner et
al.47 in achieving a sub-wavelength spatial resolution for a periodic sample with a 13 nm HHG beam. However,
the experimental far-eld intensity of the probe beam was required for the image reconstruction47, termed as the
modulus enforced probe (MEP) approach.
In this study, we demonstrate the preparation of a nanostructured sample by means of the state-of-the-art
electron-beam lithography. We then introduce a high photon ux and highly monochromatic tabletop 30 nm
beamline, newly established at the Photon Science Institute of the University of Manchester. Finally, we present
the rst EUV ptychographical imaging of the nearly periodic nanopattern, reconstructed with a modied ePIE
without a prior knowledge of the probe.
Materials and Methods
Tabletop EUV source. e EUV beamline used for ptychography is shown in Fig.1, including an HHG, a
characterisation, and an imaging stage. Great details of the beamline have been given in our recent reports16,49.
In brief, a femtosecond infrared (IR) Ti:sa laser system provides laser pulses with pulse energy up to 8 mJ, a
full-width at half maximum (FWHM) duration of 35 fs and a central wavelength of 800 nm at 1 kHz repetition
rate. e IR beam was focused with an f1 = 500 mm lens into an 8 mm–long gas cell located in the HHG vacuum
chamber. Argon gas was fed into the gas cell with a piezo-driven jet (Attotech) at a backing pressure of 1.5 bars.
e gas jet operated at 1 kHz with a typical 250 μs opening time driven by a high-voltage controller. e temporal
delay between the IR laser pulses and the gas jet was varied to maximize the HHG yield. e position of the gas
cell could be precisely tuned with an xyz-translation and rotation stage. e HHG chamber was separated from
the dierential pumping chamber by a 100 mm long tube with an 1 mm inner diameter, reducing the gas load in
the successive vacuum chambers. e vacuum pressure was ~4 × 103 mbar in the HHG chamber, and ~106
mbar in the dierential pumping and the experimental chambers. Aer passing the dierential pumping stage,
the HHG beam entered the experimental chamber, where it was characterised with a home-built at-eld EUV
spectrometer49. e IR beam was ltered out with a 300 nm–thick aluminium foil. At about 80 cm downstream
from the HHG source, an 1 mm–diameter aperture was inserted as a spatial lter, resulting in a desired EUV
beam prole for ptychography.
For ptychographical imaging, a single harmonic at λ = 30 nm was selected using a pair of multilayer mirrors
(optiX fab) in a z-conguration, containing a at mirror and an f2 = 110 mm mirror. Each mirror has a reectivity
>33% and a FWHM bandwidth of ~1 nm. e fold angles were kept <5° to minimize the astigmatism of the
30 nm probe beam. e spectral bandwidth (λ/λ) is about 200, which is sucient to perform coherent imaging
in this work16,50,51. A sample with lithographed nanopatterns was mounted on a sub-nm xyz-translation stage
(SmarAct) at the focus of the concave multilayer mirror M2, perpendicular to the 30 nm probe beam. e light
diracted from the sample was detected with an in-vacuum X-ray CCD camera (Andor iKon-M 934, 1024 × 1024
pixels, and p0 = 13 µm pixel-size) placed at a distance z = 16.5 mm downstream. e CCD’s sensor was cooled
down to 95 °C to enhance the signal-to-noise ratio. e camera was rotated an angle θ 45° relative to the sam-
ple so that the strongest diracted signals go along the diagonal of the sensor, optimizing the use of the sensor. A
second 200 nm-thick aluminium foil was installed in between the sample and camera to block the residual stray
light. By measuring the 30 nm beam prole, a photon ux of 7 × 107 photons/s on the detector was obtained.
It corresponds to a photon ux of 1.1 × 108 photons/s impinging on the sample when taking the absorption
of the second Al foil into account. e FWHM-diameter of the focal area of the probe was about wp = 2.5 µm,
determined by xyz-scanning of a 2 µm–diameter iris and recording the probe beam with the X-ray CCD camera.
Sample Preparation. Nanopatterns were prepared on a 10 nm thick silicon nitride (Si3N4) membrane,
which acts as an EUV-transmitting window (~15 µm × 15 µm), on a silicon frame (5 mm × 5 mm × 200 µm) by
means of the electron beam lithography. To apply the resist to the Si3N4 window via a spinning process, poly(me-
thyl methacrylate) (PMMA) was used as a bonding agent. e spinning process requires the sample to be held on
by a vacuum, which might damage the Si3N4 membrane. To support the Si3N4 membrane sample, we rst applied
a mixture of 8% wt. of anisole and PMMA to the surface of a 10 mm × 10 mm silicon wafer and then placed the
sample onto the PMMA. e whole unit was baked on a hot plate at a temperature of 180 °C for 20 minutes, allow-
ing the anisole to evaporate and leaving the PMMA behind. Next, the resist/solvent material was applied to the
sample by spin-coating with a speed of 8000 rpm and duration of 40 s. e sample with resist/solvent was then so
baked at a temperature of 100 °C for 2 minutes, resulting in an 100 nm thick resist evenly coated on the sample.
Further details are given elsewhere52.
e nanopatterns were written using an FEI scanning electron microscope that was driven by a Raith Elphy
Quantum 6 MHz pattern generator. e pattern consisted of boxes that were 50 µm in length and their widths
varied from 1 µm to 100 nm, and the line space varied to match the width of the box. e pattern was exposed
onto the Si3N4 window using an acceleration voltage of 30 keV, a current of 50 pA, and a 4 nm step-size. Once the
exposure was completed, the sample was developed in hexane for 10 s and the result is shown in Fig.2a. e nal
step was to remove the silicon wafer from the sample by dissolving the PMMA bonding agent in an acetone bath
for a period of ten hours.
Image Reconstructions. Image reconstruction was performed with the extended ptychographical iterative
engine36, including the lateral translation correction53. In brief, at the mth iteration, a complex-valued probe beam
Pm(r) illuminates an object Om(r, Rj), coordinated by Rj as the lateral translation of the object relative to the probe
beam. e exit surface wave at the object-plane is given as,
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
4
ScIeNTIfIc REPORTS | (2018) 8:16693 | DOI:10.1038/s41598-018-34257-2
=⋅ .fPOrR rrR(, )()(,) (1)
m
jm mj
At the detector-plane in the far-eld, the waveeld is computed as the Fourier transform of the ESW,
=.FT fFk rR() {(,)}(2)
mmj
e modulus constraint replaces the calculated amplitude (|Fm(K)|) with the experimental one while main-
taining the phase,
=⋅ .
IFk kFkFk
() () ()/()
(3)
m,j
mod
jmm
By back-propagating to the object-plane, the modied ESW is then updated as
Figure 3. (Top) Typical normalised error as a function of the number of iteration (Eq. (7)). e ePIE algorithm
performed 500 iterations with probe updates aer 120 iterations and translation renement aer 250 iterations.
(Top, insets) e reconstructed images illustrate the visual quality of the object without probe updates (a),
with probe updates (b), and with translation correction (c), taken at the iteration marked by vertical arrows.
(Bottom) e nal reconstructed amplitude and phase of the object (d,e) and probe (f,g), respectively.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
5
ScIeNTIfIc REPORTS | (2018) 8:16693 | DOI:10.1038/s41598-018-34257-2
=.
{}
fFTrR Fk(, )()
(4)
m
mod
j
1
m,j
mod
Here, the support constraint might be applied5,54. Furthermore, probe and object updates are obtained by
applying the overlap constraint, namely,
α=+
||
+
OO
P
P
ffrR rR
r
r
rR rR(, )(,)
()
()
(, )(,)
(5)
m1 jmjm
mmax
2m
modjmj
β=+
||
+
PP
O
Offrr
rR
rR rR rR() ()
(, )
(, )(, )(,),
(6)
m1 mmj
mjmax
2m
modjmj
where the empirical parameters (α, β) are set to unity in this study. e above steps are sequentially repeated for
all available scan positions {Rj} to form a single ePIE-iteration. Aer each complete iteration, to monitor the ePIE
progress the normalised error is measured as36
σ=
∑∑
∑∑ .
I
I
kFk
k
() ()
()
(7)
m
k
k
jjm
2
jj
Oen, the initial probe guess is unity and only updated aer a few tens of iterations. Translation correction for
each scan position Rj was performed by calculating the relative shi δj(δjx, δjy) between the object’s estimates of the
successive iterations, i.e., Om+1(r, Rj) and Om(r, Rj), and modifying the current position to
γδ=+
RR
,
(8)
j
mod
jj
where the unitless magnication parameter
γγγ(, )
xy
is a function of the iteration number as suggested in the
original work53. The lateral shift δj is usually of the sub–0.01 pixel order and was calculated with the
cross-correlation technique53,55. Due to the invariance of the Fourier-transform to the object’s lateral transla-
tion36,56, we observed that the translation correction should be always applied (e.g., some tens of iterations) to
rene the object and probe functions.
Results and Discussion
In this study, the sample was raster-scanned with step-sizes (Δx, Δy) of a few hundreds of nm, i.e., less than wp/4
to ensure a sucient (>70%) overlap between the adjacent areas. A random oset of about 15% of the step-sizes
was added to each position to avoid the periodic artefacts (known as the raster grid pathology)37. At each scan
position, the diraction pattern was accumulated three times with 2 s exposure time and an 1 MHz readout speed.
Each diraction pattern was rotated with an angle θ to suit the sample’s coordinates, resulting in a resized image
(~1448 × 1448 pixels) with the strongest scattering signals along the x-axis. To improve the signal-to-noise ratio
and reduce the computing time, we numerically integrated the diraction intensity by binning 2 × 2 pixels into 1
pixel and cropped each diraction pattern to N × N pixels (with N = 600). All diraction patterns were then rem-
apped onto the Ewald sphere for curvature correction57,58, following by a 2D Gaussian smoothing kernel with a
standard deviation of 5. A representative diraction pattern of the sample is shown in Fig.2b, with the maximum
spatial frequency kmax 16 μm1. A high dynamic range close to six orders of magnitude was obtained, which is
crucial to achieve high-resolution imaging with the ePIE algorithm. e linear oversampling ratio is related to the
probe’s diameter as
λΘ=≈.zpw/(2)76
xy p
0
, which is entirely satised the oversampling requirement. Note
Figure 4. Phase retrieval transfer function (PRTF) computed from ve hundred independent ePIE
reconstructions. e resolution cuto of the ePIE, determined by the spatial frequency at which the PRTF
reaches a value of 1/e, is greater than the experimental cutokmax 16 μm1.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
6
ScIeNTIfIc REPORTS | (2018) 8:16693 | DOI:10.1038/s41598-018-34257-2
that the eective probe’s beam might be slightly larger than the used FWHM diameter wp, yielding a possible
smaller oversampling ratio. In principle, the smallest resolvable period on an object, i.e., the half-pitch distance
resolution, is given according to the Rayleigh criterion as,
λΔ= =.rzpN/(2)31 7
0
nm.
In the following, we rst explore in details the performance of the modied ePIE framework with a small data
set of 169 DPs, probing an area of about 5 × 5 μm of the object. e spatial resolution of reconstruction can be
drawn from multiple independent ePIE reconstructions. Second, we present a full eld of view of the sample with
a data set of 900 DPs, which covers the whole Si3N4 window.
e rst ptychographical scan includes 13 × 13 positions with the step-sizes Δx = Δy = 300 nm. e ePIE
started with a probe guess of unity and a random object guess. e ePIE performed 500 iterations with probe
updates aer 120 iterations and translation renement aer 250 iterations. e ePIE code was written in Matlab
(R2016a) and ran on a Tesla K40 GPU accelerator. Figure3 (top) shows the normalised error as a function of
the number of iteration, monitoring the progress of image reconstruction. Snap-shots of the on-the-y recon-
structed object are also depicted to illustrate the improvement (Fig.3a–c) when dierent numerical techniques
were applied. Essentially, by applying probe update or translation correction, the error metric quickly drops by
Figure 5. Amplitude and phase of the object (a,b) and probe (c,d), respectively, obtained aer a thousand ePIE
iterations from a collection of 900 far-eld diraction patterns of the sample. (c) An intensity line prole (blue-
lled) is extracted along the green line in (a) compared to its corresponding part of the SEM image (yellow-
lled) shown in Fig.2a. Note the dierent imaging methods between the ePIE image (transmission) and the
SEM (reection).
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
7
ScIeNTIfIc REPORTS | (2018) 8:16693 | DOI:10.1038/s41598-018-34257-2
about an order of magnitude within a few tens of iterations. It then reduces exponentially slow with iteration,
representing the characteristics of the well-known ER algorithm built within the ePIE. e nal reconstructed
amplitude and phase of the object and probe are shown in Fig.3d–g. e retrieved lines are quite homogenous
in the phase’s picture, while their amplitudes are signicantly modulated (~40%) from line to line along the
x-axis. is eect has been also observed in many other independent reconstructions. We strongly believe that the
cross-talk between the object and the probe mainly accounts for the eect. We have performed additional ePIE
reconstructions applying the modulus enforced probe approach47 and observed much less modulation. However,
the corresponding normalised errors are about an order of magnitude higher. e far-eld probe’s modulus was
measured by moving the sample out of the 30 nm beam.
e spatial resolution can be directly determined if a well-characterised knife-edge sample is available. In CDI,
a phase retrieval transfer function (PRTF) has been oen used as a powerful mean to gauge the spatial resolution
of reconstruction and is given as14,16,59,60
=
|〈 〉|FT f
I
k
rR
k
PRTF ()
{(,)}
() ,
(9)
j
2
j
where
〈〉frR(, )
j
is the mean ESW at a xed position Rj, calculated from several independent reconstructions. e
resolution cuto of the ePIE reconstruction can be dened as the spatial frequency at which the PRTF reaches a
value of 1/e 0.37. We note, however, that the PRTF might strongly depend on the applied numerical meth-
ods14,16. Further, care must be taken into account to remove the outliers (failed solutions) before computing the
PRTF61.
Figure4 shows the PRTF obtained from five hundred independent reconstructions with the same set-
ting parameters. Here, the resolution cuto exceeds the experimental cuto kmax, which corresponds to the
diraction-limited resolution of 31.7 nm. e resolution of our setup is currently limited by the highest scattering
angle recorded with the CCD camera. In addition to using a larger CCD’s sensor, dierent modications might
be applied to increase the dynamic range of the measured DPs, which is directly related to the highest scatter-
ing angle. First, a mechanical beamstop is used to block the brightest undiracted light, allowing to measure
the high-angle diracted signals with longer exposure time. However, the use of a beamstop oen complicates
the experimental design and is very time-consuming. Second approach is to stitch dierent diraction patterns
recorded for various exposure durations, while removing the oversaturated signal44. Care must be considered in
reading the sensor output, because artefacts (e.g., saturation trail) might occur for oversaturated CCD sensors.
For a full-field scan, we consider a ptychographical data set of 30 × 30 positions with the step-sizes
Δx = Δy = 400 nm raster-scanning over the whole window (~15 × 15 μm). e ePIE reconstruction performed
one thousand iterations with a random object estimate, following with probe updates after 120 iterations,
and translation correction aer 300 iterations. Figure5 shows the reconstructed amplitude and phase of the
object5(a,b) and the probe5(c,d), respectively. A comparison between the ePIE and SEM image is given in
Fig.5e, showing excellent agreement between the two approaches. Note that the ePIE image is obtained in a
transmission conguration while the SEM image is obtained in a reection mode. Compared to the SEM image,
the ePIE object shows blurred edges, indicating possible sample’s defects from the e-beam lithography’s prepa-
ration. We observe slightly modulated transmission (~15% in average) of the object (5a) along the y-axis with a
mean period of about 310 nm, which is probably due to the nearly periodic features of the sample leading to the
strong cross-talking between object and probe. In a recent report47, this artefact is greatly reduced when the MEP
method was applied.
roughout this work we use for each position a loose support calculated from the inverse FT of the exper-
imental diraction pattern6. We however strongly believe that the visual quality of the object and the progress
of image reconstruction might be greatly improved with a dynamic support scheme (e.g., shrink-wrap54). e
acquisition time of the nine hundred diraction patterns was about two and half hours. Consequently, we limited
ourselves to the thermal dri and mechanical vibrations present in the laboratory. Future developed photon-rich
ux EUV sources with high repetition rates (100 kHz)62 might help to reduce exposure time and enhance the
quality of the experimental data.
References
1. Sadinawat, A. & Attwood, D. Nanoscale X-ray imaging. Nat. Photonics 4, 840–848, https://doi.org/10.1038/nphoton.2010.267
(2010).
2. Mimura, H. et al. Breaing the 10 nm barrier in hard-X-ray focusing. Nat. Phys. 6, 122–125, https://doi.org/10.1038/Nphys1457
(2010).
3. Sayre, D. Some Implications of a Theorem Due to Shannon. Acta Crystallogr. 5, 843–843, https://doi.org/10.1107/
S0365110x52002276 (1952).
4. Gerchberg, . W. & Saxton, W. O. Practical Algorithm for Determination of Phase from Image and Diraction Plane Pictures. Opti
35, 237–246 (1972).
5. Fienup, J. . Phase etrieval Algorithms - a Comparison. Appl. Opt. 21, 2758–2769, https://doi.org/10.1364/AO.21.002758 (1982).
6. Fienup, J. ., Crimmins, T. . & Holsztynsi, W. econstruction of the Support of an Object from the Support of Its Auto-
Correlation. J. Opt. Soc. Am. 72, 610–624, https://doi.org/10.1364/Josa.72.000610 (1982).
7. Shechtman, Y. et al. Phase etrieval with Application to Optical Imaging. IEEE Signal Proc. Mag. 32, 87–109, https://doi.org/10.1109/
Msp.2014.2352673 (2015).
8. Miao, J. W. et al. Quantitative image reconstruction of GaN quantum dots from oversampled diraction intensities alone. Phys. ev.
Lett. 95, 085503, https://doi.org/10.1103/PhysevLett.95.085503 (2005).
9. Taajo, H., Taahashi, T., Itoh, . & Fujisai, T. econstruction of an object from its Fourier modulus: development of the
combination algorithm composed of the hybrid input-output algorithm and its converging part. Appl. Opt. 41, 6143–6153, https://
doi.org/10.1364/Ao.41.006143 (2002).
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
8
ScIeNTIfIc REPORTS | (2018) 8:16693 | DOI:10.1038/s41598-018-34257-2
10. B ausche, H. H., Combettes, P. L. & Lue, D. . Phase retrieval, error reduction algorithm, and Fienup variants: a view from convex
optimization. J. Opt. Soc. Am. A 19, 1334–1345, https://doi.org/10.1364/Josaa.19.001334 (2002).
11. Spence, J. C. H., Weierstall, U. & Howells, M. Phase recovery and lensless imaging by iterative methods in optical, X-ray and electron
diraction. Phil. Trans. . Soc. A 360, 875–895, https://doi.org/10.1098/rsta.2001.0972 (2002).
12. Weierstall, U. et al. Image reconstruction from electron and X-ray diraction patterns using iterative algorithms: experiment and
simulation. Ultramicroscopy 90, 171–195, https://doi.org/10.1016/S0304-3991(01)00134-6 (2002).
13. Lei, N. Direct Bragg-pea phase retrieva l by a hybrid-input-output algorithm with proper intensity normalization. Acta Crystallogr.
A 63, 66–76, https://doi.org/10.1107/S0108767306049579 (2007).
14. Chapman, H. N. et al. High-resolution ab initio three-dimensional x-ray diraction microscopy. J. Opt. Soc. Am. A 23, 1179–1200,
https://doi.org/10.1364/Josaa.23.001179 (2006).
15. obinson, I. & Harder, . Coherent X-ray diraction imaging of strain at the nanoscale. Nat. Mater. 8, 291–298, https://doi.
org/10.1038/nmat2400 (2009).
16. Truong, N. X., Whittaer, E. & Denece, M. A. Phase retrieval of coherent diractive images with global optimization algorithms. J.
Appl. Crystallogr. 50, 1637–1645, https://doi.org/10.1107/S1600576717013012 (2017).
17. Fienup, J. . & Wacerman, C. C. Phase-etrieval Stagnation Problems and Solutions. J. Opt. Soc. Am. A 3, 1897–1907, https://doi.
org/10.1364/Josaa.3.001897 (1986).
18. Lue, D. . elaxed averaged alternating reections for diraction imaging. Inverse Probl. 21, 37–50, https://doi.org/10.1088/0266-
5611/21/1/004 (2005).
19. Martin, A. V. et al. Noise-robust coherent diractive imaging with a single diraction pattern. Opt. Express 20, 16650–16661, https://
doi.org/10.1364/Oe.20.016650 (2012).
20. Elser, V. Phase retrieval by iterated projections. J. Opt. Soc. Am. A 20, 40–55, https://doi.org/10.1364/Josaa.20.000040 (2003).
21. o driguez, J. A., Xu, ., Chen, C. C., Zou, Y. F. & Miao, J. W. Oversampling smoothness: an eective algorithm for phase retrieval of
noisy diraction intensities. J. Appl. Crystallogr. 46, 312–318, https://doi.org/10.1107/S0021889813002471 (2013).
22. Marchesini, S. Phase retrieval and saddle-point optimization. J. Opt. Soc. Am. A 24, 3289–3296, https://doi.org/10.1364/
Josaa.24.003289 (2007).
23. Chen, C. C., Miao, J., Wang, C. W. & Lee, T. . Application of optimization technique to noncrystalline x-ray diraction microscopy:
Guided hybrid input-output method. Phys. ev. B 76, 064113, https://doi.org/10.1103/Physrevb.76.064113 (2007).
24. Colombo, A., Galli, D. E., De Caro, L., Scattarella, F. & Carlino, E. Facing the phase problem in Coherent Diractive Imaging via
Memetic Algorithms. Sci. ep. 7, 42236, https://doi.org/10.1038/Srep42236 (2017).
25. Nellist, P. D., Mccallum, B. C. & odenburg, J. M. esolution Beyond the Information Limit in Transmission Electron-Microscopy.
Nature 374, 630–632, https://doi.org/10.1038/374630a0 (1995).
26. Chapman, H. N. Phase-retrieval X-ray microscopy by Wigner-distribution deconvolution. Ultramicroscopy 66, 153–172, https://doi.
org/10.1016/S0304-3991(96)00084-8 (1996).
27. Plamann, T. & odenburg, J. M. Electron ptychography. II eory of three-dimensional propagation eects. Acta Crystallogr A 54,
61–73, https://doi.org/10.1107/S0108767397010507 (1998).
28. Nellist, P. D. & odenburg, J. M. Electron ptychography. I Experimental demonstration beyond the conventional resolution limits.
Acta Crystallogr A 54, 49–60, https://doi.org/10.1107/S0108767397010490 (1998).
29. odenburg, J. M. et al. Hard-x-ray lensless imaging of extended objects. Phys. ev. Lett. 98, 034801, https://doi.org/10.1103/
PhysevLett.98.034801 (2007).
30. Thibault, P. et al. High-resolution scanning x-ray diffraction microscopy. Science 321, 379–382, https://doi.org/10.1126/
science.1158573 (2008).
31. Gieweemeyer, . et al. Quantitative biological imaging by ptychographic x-ray diraction microscopy. P. Natl. Acad. Sci. USA 107,
529–534, https://doi.org/10.1073/pnas.0905846107 (2010).
32. Dierolf, M. et al. Ptychographic X-ray computed tomography at the nanoscale. Nature 467, 436–439, https://doi.org/10.1038/
nature09419 (2010).
33. Pfeier, F. X-ray ptychography. Nat. Photonics 12, 9–17, https://doi.org/10.1038/s41566-017-0072-5 (2018).
34. Gao, S. et al. Electron ptychographic microscopy for three-dimensional imaging. Nat. Commun. 8, 163, https://doi.org/10.1038/
s41467-017-00150-1 (2017).
35. odenburg, J. M. & Faulner, H. M. L. A phase retrieval algorithm for shiing illumination. Appl Phys Lett 85, 4795–4797, https://
doi.org/10.1063/1.1823034 (2004).
36. Maiden, A. M. & odenburg, J. M. An improved ptychographical phase retrieval algorithm for diractive imaging. Ultramicroscopy
109, 1256–1262, https://doi.org/10.1016/j.ultramic.2009.05.012 (2009).
37. Thibault, P., Dierolf, M., Bun, O., Menzel, A. & Pfeiffer, F. Probe retrieval in ptychographic coherent diffractive imaging.
Ultramicroscopy 109, 338–343, https://doi.org/10.1016/j.ultramic.2008.12.011 (2009).
38. Enders, B. & ibault, P. A computational framewor for ptychographic reconstructions. Proc. oyal Soc. A 472, 20160640, https://
doi.org/10.1098/spa.2016.0640 (2016).
39. Mandula, O., Aizarna, M. E., Eymery, J., Burghammer, M. & Favre-Nicolin, V. PyNX.Ptycho: a computing library for X-ray coherent
diraction imaging of nanostructures. J. Appl. Crystallogr. 49, 1842–1848, https://doi.org/10.1107/S1600576716012279 (2016).
40. Marchesini, S. et al. SHAP: a distributed GPU-based ptychographic solver. J. Appl. Crystallogr. 49, 1245–1252, https://doi.
org/10.1107/S1600576716008074 (2016).
41. Schropp, A. et al. Full spatial characterization of a nanofocused x-ray free-electron laser beam by ptychographic imaging. Sci. ep.
3, 1633, https://doi.org/10.1038/Srep01633 (2013).
42. Seiboth, F. et al. Perfect X-ray focusing via tting corrective glasses to aberrated optics. Nat. Commun. 8, https://doi.org/10.1038/
Ncomms14623 (2017).
43. Shanblatt, E. . et al. Quantitative Chemically Specic Coherent Diractive Imaging of eactions at Buried Interfaces with Few
Nanometer Precision. Nano Lett. 16, 5444–5450, https://doi.org/10.1021/acs.nanolett.6b01864 (2016).
44. Bash, P. D. et al. Wide-eld broadband extreme ultraviolet transmission ptychography using a high-harmonic source. Opt. Lett. 41,
1317–1320, https://doi.org/10.1364/Ol.41.001317 (2016).
45. Zhang, B. S. et al. High contrast 3D imaging of surfaces near the wavelength limit using tabletop EUV ptychography. Ultramicroscopy
158, 98–104, https://doi.org/10.1016/j.ultramic.2015.07.006 (2015).
46. Seaberg, M. D. et al. Tabletop nanometer extreme ultraviolet imaging in an extended reection mode using coherent Fresnel
ptychography. Optica 1, 39–44, https://doi.org/10.1364/Optica.1.000039 (2014).
47. Gardner, D. F. et al. Subwavelength coherent imaging of periodic samples using a 13.5 nm tabletop high-harmonic light source. Nat.
Photonics 11, 259, https://doi.org/10.1038/Nphoton.2017.33 (2017).
48. Porter, C. L. et al. General-purpose, wide eld-of-view reection imaging with a tabletop 13 nm light source. Optica 4, 1552–1557,
https://doi.org/10.1364/Optica.4.001552 (2017).
49. Truong, N. X., Strashnov, I., Whittaer, E., Zhong, X. L. & Denece, M. A. Coherent diractive imaging of graphite nanoparticles
using a tabletop EUV source. Phys. Chem. Chem. Phys. 19, 29660–29668, https://doi.org/10.1039/c7cp03145a (2017).
50. Sandberg, . L. et al. Lensless diractive imaging using tabletop coherent high-harmonic so-x-ray beams. Phys. ev. Lett. 99,
098103, https://doi.org/10.1103/Physrevlett.99.098103 (2007).
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
9
ScIeNTIfIc REPORTS | (2018) 8:16693 | DOI:10.1038/s41598-018-34257-2
51. Dinh, . B., Le, H. V., Hannaford, P. & Van Dao, L. Coherent diractive imaging microscope with a high-order harmonic source.
Appl. Opt. 54, 5303–5308, https://doi.org/10.1364/Ao.54.005303 (2015).
52. Lewis, S. M. et al. Use of Supramolecular Assemblies as Lithographic esists. Angew. Chem. Int. Ed 56, 6749–6752, https://doi.
org/10.1002/anie.201700224 (2017).
53. Zhang, F. C. et al. Translation position determination in ptychographic coherent diraction imaging. Opt. Express 21, 13592–13606,
https://doi.org/10.1364/Oe.21.013592 (2013).
54. Marchesini, S. et al. X-ray image reconstruction from a diraction pattern alone. Phys. ev. B 68, 140101, https://doi.org/10.1103/
Physrevb.68.140101 (2003).
55. Guizar-Sicairos, M., urman, S. T. & Fienup, J. . Ecient subpixel image registration algorithms. Opt. Lett. 33, 156–158, https://
doi.org/10.1364/Ol.33.000156 (2008).
56. Fienup, J. . Invariant error metrics for image reconstruction. Appl. Opt. 36, 8352–8357, https://doi.org/10.1364/Ao.36.008352
(1997).
57. Sandberg, . L. et al. High numerical aperture tabletop so x-ray diraction microscopy with 70-nm resolution. Proc. Natl. Acad.
Sci. USA 105, 24–27, https://doi.org/10.1073/pnas.0710761105 (2008).
58. Zurch, M. et al. eal-time and Sub-wavelength Ultrafast Coherent Diraction Imaging in the Extreme Ultraviolet. Sci. ep. 4, 7356,
https://doi.org/10.1038/Srep07356 (2014).
59. Shapiro, D. et al. Biological imaging by so x-ray diraction microscopy. Proc. Natl. Acad. Sci. USA 102, 15343–15346, https://doi.
org/10.1073/pnas.0503305102 (2005).
60. imura, T. et al. Imaging live cell in micro-liquid enclosure by X-ray laser diffraction. Nat. Commun. 5, 3052, https://doi.
org/10.1038/Ncomms4052 (2014).
61. van der Schot, G. et al. Imaging single cells in a beam of live cyanobacteria with an X-ray laser. Nat. Commun. 6, 5704, https://doi.
org/10.1038/Ncomms6704 (2015).
62. Hadrich, S. et al. High photon ux table-top coherent extreme-ultraviolet source. Nat. Photonics 8, 779–783, https://doi.org/10.1038/
Nphoton.2014.214 (2014).
Acknowledgements
N.X.T. acknowledges the Dalton Nuclear Institute and an endowment from BNFL for nancial support. We thank
the Photon Science Institute for providing us with the ultrafast laser facility.
Author Contributions
N.X.T. designed, built and optimised the EUV beamline for ptychography. S.M.L. designed and fabricated the
sample, and X.L.Z. characterised the sample. N.X.T., R.S. and V.C. conducted the experiments. N.X.T. analysed
the experimental data and prepared the manuscript. M.A.D. and F.L. planed and initiated the project. All authors
contributed to the manuscript.
Additional Information
Competing Interests: e authors declare no competing interests.
Publisher’s note: Springer Nature remains neutral with regard to jurisdictional claims in published maps and
institutional aliations.
Open Access This article is licensed under a Creative Commons Attribution 4.0 International
License, which permits use, sharing, adaptation, distribution and reproduction in any medium or
format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Cre-
ative Commons license, and indicate if changes were made. e images or other third party material in this
article are included in the article’s Creative Commons license, unless indicated otherwise in a credit line to the
material. If material is not included in the article’s Creative Commons license and your intended use is not per-
mitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the
copyright holder. To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/.
© e Author(s) 2018
Content courtesy of Springer Nature, terms of use apply. Rights reserved
1.
2.
3.
4.
5.
6.
Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers and authorised users (“Users”), for small-
scale personal, non-commercial use provided that all copyright, trade and service marks and other proprietary notices are maintained. By
accessing, sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of use (“Terms”). For these
purposes, Springer Nature considers academic use (by researchers and students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and conditions, a relevant site licence or a personal
subscription. These Terms will prevail over any conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription
(to the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of the Creative Commons license used will
apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may also use these personal data internally within
ResearchGate and Springer Nature and as agreed share it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not
otherwise disclose your personal data outside the ResearchGate or the Springer Nature group of companies unless we have your permission as
detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial use, it is important to note that Users may
not:
use such content for the purpose of providing other users with access on a regular or large scale basis or as a means to circumvent access
control;
use such content where to do so would be considered a criminal or statutory offence in any jurisdiction, or gives rise to civil liability, or is
otherwise unlawful;
falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
use bots or other automated methods to access the content or redirect messages
override any security feature or exclusionary protocol; or
share the content in order to create substitute for Springer Nature products or services or a systematic database of Springer Nature journal
content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a product or service that creates revenue,
royalties, rent or income from our content or its inclusion as part of a paid for service or for other commercial gain. Springer Nature journal
content cannot be used for inter-library loans and librarians may not upload Springer Nature journal content on a large scale into their, or any
other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not obligated to publish any information or
content on this website and may remove it or features or functionality at our sole discretion, at any time with or without notice. Springer Nature
may revoke this licence to you at any time and remove access to any copies of the Springer Nature journal content which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or guarantees to Users, either express or implied
with respect to the Springer nature journal content and all parties disclaim and waive any implied warranties or warranties imposed by law,
including merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published by Springer Nature that may be licensed
from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a regular basis or in any other manner not
expressly permitted by these Terms, please contact Springer Nature at
onlineservice@springernature.com
... Therefore, high-brightness HHG light sources [9][10][11] potentially serve as a unique laboratory-scale tool for the frontier of spintronic applications. Owing to its high spatial coherence, HHG radiation has been used for imaging of nanostructures with a large-scale view [12][13][14], magnetic imaging with diffraction-limited spatial resolution, and enhanced magnetic contrasts using holographic methods [15]. Furthermore, the extended degree of the control of spin [4][5][6] and orbital angular momentum [7,8] of HHG pulses via the manipulation of driving lasers provides the unique ability for resolving and controlling the magnetic and topological excitations of magnetic textures. ...
... The agreement is not only because of the small wavelength of the x ray but also because of the longitudinal coherence (monochromaticity) of the beam. For phase-resolved applications, such as large-scale view nanoimaging [12][13][14], and enhanced magnetic contrasts using holographic methods [15], spatial coherence is also important. Note that since HHG is an "electron recollision" process in the microscopic picture [64], particular x-ray photon energies can be generated through different electron trajectories in the optical cycles of the driver laser field, which may lead to a reduction in spatial coherence. ...
... Therefore, high-brightness HHG light sources [9][10][11] potentially serve as a unique laboratory-scale tool for the frontier of spintronic applications. Owing to its high spatial coherence, HHG radiation has been used for imaging of nanostructures with a large-scale view [12][13][14], magnetic imaging with diffraction-limited spatial resolution, and enhanced magnetic contrasts using holographic methods [15]. Furthermore, the extended degree of the control of spin [4][5][6] and orbital angular momentum [7,8] of HHG pulses via the manipulation of driving lasers provides the unique ability for resolving and controlling the magnetic and topological excitations of magnetic textures. ...
... The agreement is not only because of the small wavelength of the x ray but also because of the longitudinal coherence (monochromaticity) of the beam. For phase-resolved applications, such as large-scale view nanoimaging [12][13][14], and enhanced magnetic contrasts using holographic methods [15], spatial coherence is also important. Note that since HHG is an "electron recollision" process in the microscopic picture [64], particular x-ray photon energies can be generated through different electron trajectories in the optical cycles of the driver laser field, which may lead to a reduction in spatial coherence. ...
Article
Full-text available
Development of ultrafast table-top x-ray sources that can map various spin, orbital, and electronic configurations and reordering processes on their natural time and length scales is an essential topic for modern condensed matter physics as well as ultrafast science. In this work, we demonstrate spatiotemporally resolved resonant magnetic scattering (XRMS) to probe the inner-shell 4d electrons of a rare-earth (RE) composite ferrimagnetic system using a bright > 200 e V soft x-ray high harmonic generation (HHG) source, which is relevant for future energy-efficient, high-speed spintronic applications. The XRMS is enabled by direct driving of the HHG process with power-scalable, high-energy Yb laser technology. The optimally phase-matched broadband plateau of the HHG offers a record photon flux ( > 2 × 1 0 9 p h o t o n s / s / 1 % bandwidth) with excellent spatial coherence and covers the entire resonant energy range of RE’s N 4 , 5 edges. We verify the underlying physics of our x-ray generation strategy through the analysis of microscopic and macroscopic processes. Using a CoTb alloy as a prototypical ferrimagnetic system, we retrieve the spin dynamics, and resolve a fast demagnetization time of 500 ± 126 f s , concomitant with an expansion of the domain periodicity, corresponding to a domain wall velocity of ∼ 750 m / s . The results confirm that, far from cross-contamination of low-energy absorption edges in multi-element systems, the highly localized states of 4 d electrons associated with the N 4 , 5 edges can provide high-quality core-level magnetic information on par with what can be obtained at the M edges, which is currently accessible only at large-scale x-ray facilities. The analysis also indicates the rich material-, composition-, and probing-energy-dependent driving mechanism of RE-associated multicomponent systems. Considering the rapid emergence of high-power Yb lasers combined with novel nonlinear compression technology, this work indicates potential for next-generation high-performance soft x-ray HHG-based sources in future extremely photon-hungry applications on the table-top scale, such as probing electronic motion in biologically relevant molecules in their physiological environment (liquid phase), and advanced coherent imaging of nano-engineered devices with 5 ∼ 8 n m resolution.
... Ptychographic imaging has been performed with various illumination sources, such as hard x-ray, electron, LED, or laser, [11][12][13][14] for different applications ranging from materials science to biology. [15][16][17] Its many advantages notwithstanding, a ptychography-based tool also brings many challenges, such as its computational cost, coherent light source, and the necessity of an efficient diffraction recording system. Since the technique does not project object images in a direct manner, the optimization can be more complex. ...
Article
Full-text available
Background: With the adoption of extreme ultraviolet (EUV) lithography in the semiconductor manufacturing, actinic EUV mask metrology has become a crucial technology to ensure the required defect sensitivity and throughput for high-volume manufacturing. Reflection-mode EUV scanning microscope (RESCAN) is a lensless actinic microscope dedicated to EUV mask metrology based on coherent diffraction imaging (CDI) as an alternative approach for EUV mask metrology and inspection. Aim and Approach: In CDI, the complex-amplitude image of the sample is obtained through its measured diffraction. Though this approach can overcome the disadvantages and limitations of conventional imaging systems, the quality of the recorded diffraction data is crucial for the reli- able reconstruction of a high-resolution image. Ultimately, the signal-to-noise ratio of the recorded diffraction data depends on several parameters, such as the sample’s reflectance, the quantum efficiency of the detector, its full well capacity, and the intensity of the illumination. Result: We investigate the optimal photon flux for RESCAN and provide a systematic study on the relation between the image resolution and the illumination intensity for CDI-based imaging at EUV wavelength. Conclusions: The insights provided will be helpful for the optimization of CDI for EUV imaging, in particular for increasing the throughput of EUV mask inspection with low power sources.
... 16,17) Through recent progress in high harmonic generation (HHG) using a tabletop laser, EUV and soft X-rays have been successfully employed in small-scale facilities. 18,19) In this study, we demonstrated an actinic mask imaging method using an EUV ptychography microscope that employs coherent EUV light produced by HHG and ptychographic algorithms. The phase and amplitude information of the periodic patterns are reconstructed by a regularized ptychographical iterative engine (rPIE) 20) with a modulus enforced probe (MEP) constraint 21) to enforce the illumination probe. ...
Article
Full-text available
Extreme ultraviolet (EUV) lithography is expected to be used for 3 nm technology nodes and beyond, yet the need for actinic mask metrology and inspection remains a critical challenge. In this study, we demonstrate an EUV ptychography microscope as a high-harmonic generation (HHG)-based actinic mask imaging tool. A series of diffraction patterns on an EUV mask is used to reconstruct both the amplitude and phase information of the periodic patterns using ptychographic algorithms. The results show that the EUV ptychography microscope has the potential for determining the actinic metrology of EUV masks and providing phase information for EUV mask development.
... To counteract the loss in sensitivity to the surface, this should be accompanied by the illumination by longer wavelengths of the photon beam. Moreover, with the advent of lab EUV sources, this method present a real alternative to scanning methods 21,22 . We report, on the dimensional reconstruction of a lamellar grating using EUV scattering and its comparison to a GISAXS reconstruction. ...
Preprint
Increasing miniaturization and complexity of nanostructures require innovative metrology solutions with high throughput that can assess complex 3D structures in a non-destructive manner. EUV scatterometry is investigated for the characterization of nanostructured surfaces. The reconstruction is based on a rigorous simulation using a Maxwell solver based on finite-elements and is statistically validated with a Markov-Chain Monte Carlo sampling method. Here it is shown that this method is suitable for the dimensional characterization of the nanostructures and the investigation of oxide or contamination layers. In comparison to grazing-incidence small-angle X-rayscattering (GISAXS) EUV allows to probe smaller areas. The influence of the divergence on the diffracted intensities in EUV is much lower than in GISAXS, which also reduces the computational effort of the reconstruction.
Article
Extreme ultraviolet (EUV) lithography uses reflective optics and a thick mask absorber, leading to mask 3D (M3D) effects. These M3D effects cause disparities in the amplitudes and phases of EUV mask diffractions, impacting mask imaging performance and reducing process yields. Our findings demonstrate that wrinkles in the EUV pellicle can exacerbate M3D effects. This imbalance results in critical dimension variation, image contrast loss, and pattern shift in mask images. Therefore, the use of a pellicle material with thermodynamic characteristics that minimize wrinkles when exposed to EUV rays is imperative.
Article
Ptychography is a lensless imaging technique that is aberration-free and capable of imaging both the amplitude and the phase of radiation reflected or transmitted from an object using iterative algorithms. Working with extreme ultraviolet (EUV) light, ptychography can provide better resolution than conventional optical microscopy and deeper penetration than scanning electron microscope. As a compact lab-scale EUV light sources, high harmonic generation meets the high coherence requirement of ptychography and gives more flexibilities in both budget and experimental time compared to synchrotrons. The ability to measure phase makes reflection-mode ptychography a good choice for characterising both the surface topography and the internal structural changes in EUV multilayer mirrors. This paper describes the use of reflection-mode ptychography with a lab-scale high harmonic generation based EUV light source to perform quantitative measurement of the amplitude and phase reflection from EUV multilayer mirrors with engineered substrate defects. Using EUV light at 29.6nm from a tabletop high harmonic generation light source, a lateral resolution down to ∼88nm and a phase resolution of 0.08rad (equivalent to topographic height variation of 0.27nm) are achieved. The effect of surface distortion and roughness on EUV reflectivity is compared to topographic properties of the mirror defects measured using both atomic force microscopy and scanning transmission electron microscopy. Modelling of reflection properties from multilayer mirrors is used to predict the potential of a combination of on-resonance, actinic ptychographic imaging at 13.5nm and atomic force microscopy for characterising the changes in multilayered structures.
Conference Paper
We generate multi-wavelength extreme-ultraviolet vortex beams via high-harmonic generation. We characterize the wavefronts of these high orbital angular mo-mentum beams using ptychography.
Chapter
Interferometry-based quantitative phase imaging methods support precise phase measurements and are widely used for optical metrology and biological imaging. However, these methods rely on the high degree of coherence of light beams to achieve interference, and their accuracy heavily depends on the quality of optical elements adopted and the stability of working environment. Thus, it is difficult to apply interferometric methods for imaging with short wavelength including X-ray and high-energy electron beam, where high quality optical elements are not available, and the coherence of radiation beam is much lower than that of common laser. Furthermore, the high requirement of interferometry on the mechanical stability of instrument strictly limits their applications in lots of circumstances. Non-interferometric methods, including coherent diffraction imaging, phase diversity and transport of intensity equation (TIE), etc., can provide a powerful alternative solution to quantitative phase imaging problems without using complex optical alignment and highly coherent illumination. This chapter outlines the principle of such non-interferometric quantitative phase imaging techniques systematically followed by several numerical simulation in MATLAB®.
Article
Increasing miniaturization and complexity of nanostructures require innovative metrology solutions with high throughput that can assess complex 3D structures in a non-destructive manner. EUV scatterometry is investigated for the characterization of nanostructured surfaces and compared to grazing-incidence small-angle X-ray scattering (GISAXS). The reconstruction is based on a rigorous simulation using a Maxwell solver based on finite-elements and is statistically validated with a Markov-Chain-Monte-Carlo sampling method. It is shown that in comparison to GISAXS, EUV allows to probe smaller areas and to reduce the computation times obtaining comparable uncertainties.
Article
Full-text available
Coherent diffractive imaging (CDI) or lensless microscopy has recently been of great interest as a promising alternative to electron microscopy in achieving atomic spatial resolution. Reconstruction of images in real space from a single experimental diffraction pattern in CDI is based on applying iterative phase-retrieval (IPR) algorithms, such as the hybrid input–output and the error reduction algorithms. For noisy data, these algorithms might suffer from stagnation or trapping in local minima. Generally, the different local minima have many common as well as complementary features and might provide useful information for an improved estimate of the object. Therefore, a linear combination of a number of chosen minima, termed a basis set, gives an educated initial estimate, which might accelerate the search for the global solution. In this study, a genetic algorithm (GA) is combined with an IPR algorithm to tackle the stagnation and trapping in phase-retrieval problems. The combined GA–IPR has been employed to reconstruct an irregularly shaped hole and has proven to be reliable and robust. With the concept of basis set, it is strongly believed that many effective local and global optimization frameworks can be combined in a similar manner to solve the phase problem.
Article
Full-text available
Knowing the three-dimensional structural information of materials at the nanometer scale is essential to understanding complex material properties. Electron tomography retrieves three-dimensional structural information using a tilt series of two-dimensional images. In this paper, we report an alternative combination of electron ptychography with the inverse multislice method. We demonstrate depth sectioning of a nanostructured material into slices with 0.34 nm lateral resolution and with a corresponding depth resolution of about 24–30 nm. This three-dimensional imaging method has potential applications for the three-dimensional structure determination of a range of objects, ranging from inorganic nanostructures to biological macromolecules.
Article
Full-text available
Due to their short wavelength, X-rays can in principle be focused down to a few nanometres and below. At the same time, it is this short wavelength that puts stringent requirements on X-ray optics and their metrology. Both are limited by today’s technology. In this work, we present accurate at wavelength measurements of residual aberrations of a refractive X-ray lens using ptychography to manufacture a corrective phase plate. Together with the fitted phase plate the optics shows diffraction-limited performance, generating a nearly Gaussian beam profile with a Strehl ratio above 0.8. This scheme can be applied to any other focusing optics, thus solving the X-ray optical problem at synchrotron radiation sources and X-ray free-electron lasers.
Article
Full-text available
Coherent Diffractive Imaging is a lensless technique that allows imaging of matter at a spatial resolution not limited by lens aberrations. This technique exploits the measured diffraction pattern of a coherent beam scattered by periodic and non-periodic objects to retrieve spatial information. The diffracted intensity, for weak-scattering objects, is proportional to the modulus of the Fourier Transform of the object scattering function. Any phase information, needed to retrieve its scattering function, has to be retrieved by means of suitable algorithms. Here we present a new approach, based on a memetic algorithm, i.e. a hybrid genetic algorithm, to face the phase problem, which exploits the synergy of deterministic and stochastic optimization methods. The new approach has been tested on simulated data and applied to the phasing of transmission electron microscopy coherent electron diffraction data of a $\text{SrTiO}_\text{3}$ sample. We have been able to quantitatively retrieve the projected atomic potential, and also image the oxygen columns, which are not directly visible in the relevant high-resolution transmission electron microscopy images. Our approach proves to be a new powerful tool for the study of matter at atomic resolution and opens new perspectives in those applications in which effective phase retrieval is necessary.
Article
X-ray ptychographic microscopy combines the advantages of raster scanning X-ray microscopy with the more recently developed techniques of coherent diffraction imaging. It is limited neither by the fabricational challenges associated with X-ray optics nor by the requirements of isolated specimen preparation, and offers in principle wavelength-limited resolution, as well as stable access and solution to the phase problem. In this Review, we discuss the basic principles of X-ray ptychography and summarize the main milestones in the evolution of X-ray ptychographic microscopy and tomography over the past ten years, since its first demonstration with X-rays. We also highlight the potential for applications in the life and materials sciences, and discuss the latest advanced concepts and probable future developments.
Article
Lensless imaging with short-wavelength light is a promising method for achieving high-resolution, chemically sensitive images of a wide variety of samples. The use of 13 nm illumination is of particular interest for materials science and the imaging of next-generation nanofabricated devices. Prior to this work, there was an unmet need for a microscope that can image general samples with extreme ultraviolet light, which requires a reflection geometry. Here, we fulfill this need by performing lensless imaging using a 13 nm high-harmonic beam at grazing incidence, where most materials are reflective. Furthermore, we demonstrate to our knowledge the first 13 nm reflection-mode lensless microscope on a tabletop by using a compact high-harmonic generation source. Additionally, we present an analytic formalism that predicts when general lensless imaging geometries will yield Nyquist sampled data. Our grazingincidence ptychographic approach, which we call GLIDER, provides the first route for achieving wide field-of-view, high-resolution, lensless images of general samples with extreme ultraviolet and soft x-ray light.
Article
Structural information of nanostructures plays a key role in the synthesis of novel nano-sized materials for promising applications such as high-performance nanoelectronics and nanophotonics. In this study, we apply for the first time the state-of-the-art coherent diffractive imaging method to characterize the structure of graphite nanoparticles. A sample with nanographites on a Si3N4 support was exposed to 30 nm radiation from a tabletop laser-driven high-order harmonic generation extreme ultraviolet (EUV) source. From the measured far-field diffraction pattern, we were able to reconstruct the distribution of the graphite nanoparticles with a spatial resolution of ~330 nm using the standard iterative phase retrieval algorithms. A closer look at the reconstructed images reveals possible absorption effects of graphite nanoparticles. This experiment demonstrates the first step towards wide-field and high-resolution imaging of nuclear materials using the newly established lab-scale EUV source. Having such a source opens the door to performing investigations of nuclear graphite and other radioactive material in the lab, thus avoiding the need to transport samples to external facilities.
Article
A new resist material for electron beam lithography has been created that is based on a supramolecular assembly. Initial studies revealed that with this supramolecular approach, high-resolution structures can be written that show unprecedented selectivity when exposed to etching conditions involving plasmas.
Article
Coherent diffractive imaging is unique, being the only route for achieving high spatial resolution in the extreme ultraviolet and X-ray regions, limited only by the wavelength of the light. Recently, advances in coherent short-wavelength light sources, coupled with progress in algorithm development, have significantly enhanced the power of X-ray imaging. However, so far, high-fidelity diffraction imaging of periodic objects has been a challenge because the scattered light is concentrated in isolated peaks. Here, we use tabletop 13.5 nm high-harmonic beams to make two significant advances. First, we demonstrate high-quality imaging of an extended, nearly periodic sample for the first time. Second, we achieve subwavelength spatial resolution (12.6 nm) imaging at short wavelengths, also for the first time. The key to both advances is a novel technique called ‘modulus enforced probe’, which enables robust and quantitative reconstructions of periodic objects. This work is important for imaging next-generation nano-engineered devices.