You are on page 1of 7

Articles

https://doi.org/10.1038/s41566-019-0536-x

Single-shot quantitative phase gradient microscopy


using a system of multifunctional metasurfaces
Hyounghan Kwon1,2, Ehsan Arbabi   1,2, Seyedeh Mahsa Kamali1,2, MohammadSadegh Faraji-Dana1,2
and Andrei Faraon   1,2*

Quantitative phase imaging (QPI) of transparent samples plays an essential role in multiple biomedical applications, and
miniaturizing these systems will enable their adoption into point-of-care and in vivo applications. Here, we propose a compact
quantitative phase gradient microscope (QGPM) based on two dielectric metasurface layers, inspired by a classical differential
interference contrast (DIC) microscope. Owing to the multifunctionality and compactness of the dielectric metasurfaces, the
QPGM simultaneously captures three DIC images to generate a quantitative phase gradient image in a single shot. The volume
of the metasurface optical system is on the order of 1 mm3. Imaging experiments with various phase resolution samples verify
the capability to capture quantitative phase gradient data, with phase gradient sensitivity better than 92.3 mrad μm−1 and
single-cell resolution. The results showcase the potential of metasurfaces for developing miniaturized QPI systems for
label-free cellular imaging and point-of-care devices.

O
ptical phase microscopy techniques have been widely on-chip spectrometers28,29 and endoscopes30. In addition, verti-
investigated for imaging transparent specimens like cells1–4. cal integration of multiple metasurfaces has been introduced to
For these weakly scattering samples, phase information achieve enhanced functionalities31–34. Despite these great advances,
represents the optical path difference of light passing through applications of metasurfaces for QPI have not previously been
the cell, which is usually directly related to its morphological and explored. Although different types of spatial field differentiator,
chemical properties3. Moreover, phase imaging techniques do not which may be regarded as qualitative phase imaging devices, have
require contrast agents and avoid several issues faced in fluores- been proposed, their investigation has been limited to optical com-
cence microscopy such as photobleaching and phototoxicity5. puting and optical signal processing35,36.
Although conventional phase imaging methods such as phase con- Here, we propose a miniaturized quantitative phase gradient
trast1 and differential interference contrast (DIC) microscopy2 only microscope (QPGM) inspired by the classical DIC microscope and
capture qualitative phase information, quantitative phase imaging based on an integrated system of multifunctional dielectric meta-
(QPI) has been rapidly growing in the past two decades3,4,6. For surfaces. As we fully exploit the two unique properties of metasur-
example, techniques like digital holographic microscopy7, tomo- faces, compactness and multifunctionality, via both polarization
graphic QPI8,9, Fourier ptychography10 and lensless imaging11 over- and spatial multiplexing methods, two metasurface layers that are
come limitations of qualitative phase imaging methods to acquire cascaded vertically operate as a miniaturized QPGM. We experi-
quantitative phase data. mentally demonstrate that the millimetre-scale optical device can
Miniaturized microscopes have garnered great interest in recent capture quantitative phase gradient images (PGIs) from phase
decades12–14 because they enable and facilitate in  vivo biological resolution targets and biological samples.
imaging in freely moving objects14 and in portable applications.
Miniaturized systems have only been demonstrated as different Results
forms of amplitude imaging modules such as single-photon12 or Concept of the metasurface-based QPGM. Figure 1a illustrates
two-photon13 fluorescence microscopes. This is mainly because the concept of a miniaturized QPGM consisting of two cascaded
QPI systems usually require an interference ‘set-up’ to retrieve the metasurface layers. The roles of each layer are explained visually in
phase information, and such set-ups need complicated and bulky Fig. 1b. Metasurface layer 1 captures two images for transverse elec-
optical systems. This left the miniaturized QPI microscopes that tric (TE) and transverse magnetic (TM) polarizations with focal
are of interest in various fields such as biomedicine4 out of reach points that are separated along the y axis. In addition, it splits the
until now. captured light equally into three separate directions towards the
Dielectric metasurfaces are a category of diffractive optical three metasurface lenses in layer 2. To implement this multifunc-
elements consisting of nanoscatterers15,16 that enable the con- tionality, polarization20 and spatial multiplexing techniques23,24 are
trol of light in subwavelength scales17–22. In addition, metasur- employed in the design of metasurface layer 1 (see Supplementary
faces can simultaneously provide multiple distinct functionalities Note 1 for details). Metasurface layer 2, which is composed of
through various schemes such as spatial multiplexing23,24 or more three birefringent off-axis lenses, forms three DIC images with
sophisticated designs of the nanoscatterers25,26. These capabilities, three different phase offsets between the TE and TM polarizations.
compactness, low weight and compatibility with conventional Effectively, each metasurface of the second layer constitutes a sepa-
nanofabrication processes have made them suitable candidates for rate DIC microscope system with the metasurface layer 1. With two
miniaturized optical devices such as miniaturized microscopes27, linear polarizers at the input and output aligned to 45° and −45°,

T. J. Watson Laboratory of Applied Physics and Kavli Nanoscience Institute, California Institute of Technology, Pasadena, CA, USA. 2Department of Electrical
1

Engineering, California Institute of Technology, Pasadena, CA, USA. *e-mail: faraon@caltech.edu

Nature Photonics | VOL 14 | February 2020 | 109–114 | www.nature.com/naturephotonics 109


Articles Nature Photonics

a b Metasurface Metasurface
layer 1 layer 2 QPGM
I1 I2 I3
I1 I2 I3

Camera
with a
polarizer

Metasurface z TE TM
layer 2
y
x
c d e ∇y ϕ
Metasurface ϕ(x,y) (rad) I1 I2 I3 (a.u.) (rad µm–1)
z π 1 1.5
layer 1
y
Phase y
objects
x –π 0 –1.5
x

Fig. 1 | Schematic of a metasurface-based QPGM and its operation principle. a, Schematic of the QPGM employing two metasurface layers, where the
second layer is composed of three separate metasurface lenses. The first metasurface, together with each of the lenses in the second metasurface layer,
forms a different image of the object. A polarizer and the polarization-sensitive metasurfaces then result in three interference patterns. b, Illustration of the
roles of the two metasurface layers. Metasurface 1 makes two sheared focuses for TE and TM polarizations and splits the field in three different directions
towards the three lenses in the second layer. With the polarizer, the three metasurface lenses in layer 2 form three DIC images (I1, I2 and I3) having different
phase offsets between the TE and TM polarizations. The combination of the two layers forms the QPGM shown on the right. c, A binary phase sample
with unity amplitude used as an example target. d, Set of three DIC images of the phase sample shown in c. e, PGI formed from combining I1, I2 and I3
from d, showing the phase gradient along the y axis. The y axis lies along the shear direction of the system.

respectively, the two metasurface layers capture three separate DIC Second, the vertical integration of the two multifunctional meta-
images with different phase offsets. As an example, a binary phase surface layers uniquely enables capturing the PGI in a single shot
target is shown in Fig. 1c, which has optical fields with unity ampli- with compact implementation of the system.
tude, U(x, y) = eiϕ(x,y). The QPGM simultaneously captures three We should mention that while single polarization-sensitive
DIC images (I1, I2 and I3), as shown in Fig. 1d. Specifically, I1, I2 and bifocal metasurface lenses have been demonstrated before for
I3 are written as polarization splitting and imaging20,38, their potential application
for phase imaging has not been explored. Moreover, the working
Ij ¼ jUðx; yÞ � eiϕj Uðx; y � ΔyÞj ð1Þ principle used here is conceptually similar to gradient light inter-
ference microscopy, as in both methods several DIC images are
where ϕj ¼ ϕ0 þ 2π3 ðj � 1Þ and Δy is the sheared distance between utilized to calculate the quantitative phase gradient image along
TE and ITM polarizations at the object plane. I1, I2 and I3 in Fig. 1d one axis39. Various devices also have been proposed to develop
show a strong contrast at the top and bottom edges of the sample compact QPI techniques40–44. However, they are designed as add-
because each DIC image results from the interference of the two on devices to a conventional microscope system and their imaging
sheared optical fields along the y axis by equation (1). Using I1, I2 performances mainly rely on the microscope, which is fundamen-
and I3, one can calculate the unidirectional gradient of the phase tally hard to miniaturize. The system demonstrated here differs
sample with respect to y, ∇y ϕðx; yÞ, through the three-step phase from previous approaches in that the highly miniaturized system
shifting method37: I replaces the bulky phase or phase gradient microscope systems
and captures the phase gradient information in a single shot with-
1 pffiffiffi I2 � I3 out additional phase shifting elements like variable liquid-crystal
∇y ϕ ¼ arctanð 3 Þ � ∇y ϕcali ð2Þ
Δy 2I1 � I2 � I3 retarders or spatial light modulators.
In the following we discuss two implementations of the QPGM,
Here, ∇y ϕcali is the PGI calculated in the absence of the sample that one based on metasurfaces on two separate substrates and one com-
is usedIfor calibration (see Supplementary Note 3 for details about posed of metasurfaces on both sides of a single substrate. Although
the three-step phase shifting method). Figure 1e shows the PGI the first implementation provides large aperture size and field
calculated from the three DIC images in Fig. 1d. of view, the second implementation provides compactness and
The QPGM consists of the two multifunctional metasur- mechanical robustness.
face layers for two main reasons (see Supplementary Note 1
and Supplementary Fig. 1 for detailed discussion based on wave Implementation of the QPGM by two separate dielectric meta-
propagation simulations). First, at least two polarization-sen- surface layers. To implement the two metasurface layers, we utilized
sitive bifocal lenses are needed to capture clear DIC images. In a high-contrast transmitarray platform consisting of rectangular
other words, a single birefringent metasurface lens with a regular amorphous silicon nanoposts on a fused-silica substrate, as shown
refractive lens is not capable of capturing the DIC images clearly in Fig. 2a (see Methods and Supplementary Fig. 4 for details)20.
(see Supplementary Note 2 and Supplementary Fig. 2 for theo- We designed the metasurfaces for an operation wavelength of
retical and experimental results with the single metasurface lens). 850 nm. In Fig. 2b,c, schematics of the metasurface-based QPGM

110 Nature Photonics | VOL 14 | February 2020 | 109–114 | www.nature.com/naturephotonics


Nature Photonics Articles
a b
|tTE|ETEeiϕTE |tTM|ETMeiϕTM
Gold

z z
ETE ETM

y y
Unit cell
c
Amorphous
silicon SU-8
Dy 660 µm
z Fused y
y silica y
Dx 600 µm 600 µm
Nanoposts
x x
x

d e

Fig. 2 | Design and fabrication of the metasurfaces. a, Schematics of a uniform array of rectangular nanoposts (top) and a single unit cell (bottom),
showing the parameter definitions. The rectangular amorphous silicon nanoposts are located on a fused-silica substrate, cladded by an 8-μm-thick SU-8
layer for protection. The transmission phase of the two orthogonal polarizations can be independently controlled using the nanoposts. The amorphous
silicon layer is 664 nm thick, and the lattice constant is 380 nm. b, Schematic side views of metasurface layers 1 (left) and 2 (right), showing the gold
apertures used to block unwanted diffraction and the external noise. c, Schematic top views of metasurface layers 1 (left) and 2 (right). A magnified
view of the array of nanoposts is shown at the centre. d, Optical images of the fabricated metasurfaces. Nine copies of the fabricated metasurface-based
QPGM system are shown. e, SEM images of a portion of the fabricated metasurfaces. Scale bars, 2 μm (left) and 1 μm (right).

are presented. The left and right images correspond to metasur- resolution target (Quantitative Phase Microscopy Target, Benchmark
face layers 1 and 2 in Fig. 1a,b, respectively. To minimize the effects Technologies). As shown in Fig. 3a, the QPGM captured three
of geometric aberrations, the phase profiles of the metasurfaces DIC images of the 314-nm-thick resolution target in a single shot
were further optimized using the ray-tracing method (Zemax (see Methods and Supplementary Fig. 6 for details of the optical
OpticStudio) over a field of view (FOV) with diameter of 140 μm set-up). The resulting PGI, shown in Fig. 3b, is calculated from
(see Methods, Supplementary Note 1, Supplementary Table 1 and the DIC images in Fig. 3a using equation (2). As expected in the
Supplementary Fig. 5 for details of the optimized phase profiles and ideal cases in Fig. 1d,e, the unidirectional phase gradient imaging
corresponding point spread functions). in the y direction causes a strong contrast only at the top and
Conventional nanofabrication techniques were used to fabricate bottom edges, and very weak contrast at the left and right edges.
the metasurfaces (see Methods for the process details). All four This is also seen in the measurement results shown in Fig. 3a,b. To
metasurface lenses have identical diameters of 600 μm. To block the verify the QPGM capability, we used seven parts of the resolution
stray light caused by imperfect operation of the metasurfaces and target with thicknesses ranging from 54 to 371 nm. Figure 3c shows
limit the device aperture, circular gold apertures were patterned the resulting PGIs for 105-nm-, 207-nm- and 314-nm-thick resolu-
using photolithography. Figure 2d,e presents optical and scanning tion targets (see Supplementary Fig. 7 for the four remaining PGIs).
electron microscope (SEM) images of the two layers of the fabri- The results show a clear increase in the phase gradient as the thick-
cated metasurfaces. In this design, the whole QPGM system would ness of the structures increases. For a more rigorous analysis, Fig. 3d
fit within a cube measuring 1.92 × 1.26 × 2.70 mm3, including the shows the target thicknesses estimated from the PGIs, in addition
space between the metasurfaces. The magnification and objective to the values measured using atomic force microscopy (AFM). The
numerical aperture (NA) of the QPGM are ×1.98 and 0.4, respec- agreement between these measurements shows the ability of the
tively. Although the phase map is optimized for the central area system to retrieve quantitative phase data. To estimate the target
(140 μm diameter), the total FOV of the system is 336 μm in diam- thickness, the phase gradient is integrated at the edges of the tar-
eter. In addition, the separation between the optical axes for TE and gets along the y axis to calculate the phase. The thickness is then
TM polarizations is 1.5 μm, which results in a derivation step of estimated from the phase, the refractive index of the polymer con-
Δy = 2.25 μm (see Supplementary Fig. 3 for details about Δy). Δy is stituting the target and the wavelength. The QPGM can clearly cap-
larger than 1.06 μm, the theoretical diffraction limit of the imaging ture phase gradient information as small as 92.3 mrad μm−1, which
system, and it imposes a constraint on the resolution along the corresponds to a phase of 207 mrad (see Supplementary Fig. 7a for
y axis. However, we should point out that Δy can easily be adjusted details). In addition, the measured spatial and temporal noise levels
to below the diffraction-limited resolution by reducing the optical are 36.9 ± 0.7 and 11.4 mrad μm−1, respectively (see Supplementary
axis separation accordingly. Fig. 8 for details). Furthermore, the lateral resolutions achieved in
the experiment along the x and y axes are 2.76 μm and 3.48 μm,
Imaging with the QPGM based on two separate metasurface respectively (see Supplementary Fig. 9 for details). Compared with
layers. We characterized the QPGM, consisting of the fabricated the 1.06 μm theoretical diffraction limit, the reduced resolutions
metasurfaces, with a commercially available 1951 USAF phase result from the geometric aberration of the device, misalignment in

Nature Photonics | VOL 14 | February 2020 | 109–114 | www.nature.com/naturephotonics 111


Articles Nature Photonics

a b
I1 (a.u.) ∇y ϕ
1

(rad µm–1)
0.7

0
I2 I3
c 0

–0.7

d 400 e
(rad µm–1)
300 1.54
Thickness (nm)

200
0
100 AFM
QPGM
0 –1.54
1 2 3 4 5 6 7
Sample no.

f g h
5.09 mm (rad µm–1)
(a.u.)
1 0.7

0
Object
y
Metasurface Metasurface CMOS
layer 1 layer 2 image
z sensor 0 –0.7

Fig. 3 | Imaging experiment with the QPGM based on two separate metasurface layers. a, Three DIC images of a 314-nm-thick QPI 1951-USAF
resolution target captured by the QPGM. Scale bars, 25 μm. b, The PGI calculated from the DIC images in a. Scale bar, 25 μm. c, The PGIs
captured for three parts of the phase targets with thicknesses of 105 nm, 207 nm and 314 nm (left to right), respectively. Scale bars, 15 μm.
d, Thicknesses of seven different phase targets calculated by the QPGM, and those measured by AFM. The plotted thicknesses estimated
with the QPGM are averaged over 100 arbitrarily chosen points on the sample edges. Error bars represent standard deviations of the estimated
values. e, Schematic of a sea urchin cell and its corresponding PGIs. Scale bars, 40 μm. f, Schematic of the miniaturized optical set-up.
The microscope consists of the two metasurfaces shown in Fig. 2d, a CMOS image sensor and a linear polarizer attached to the sensor.
g, Three DIC images captured by the set-up in f. The 314-nm-thick resolution target used in a is imaged. Scale bars, 50 μm. h, The PGI
calculated from the DIC images in g. Scale bar, 50 μm.

the optical set-up and imperfect fabrication. To compare our results of the PGI in Fig. 3h are comparable to the results shown in Fig. 3b
with one of the state-of-the-art QPI techniques, the resolution tar- (see Supplementary Fig. 11c,d for additional measurement results).
gets were also characterized by Fourier ptychography techniques10
(see Supplementary Fig. 10 for details). Finally, to demonstrate the QPGM based on monolithically integrated double-sided meta-
capability of the QPGM to image biological samples, we imaged surfaces. To further miniaturize the device, we designed and
several sea urchin samples. As seen in a few sample PGIs, plotted in fabricated a monolithically integrated double-sided metasurface
Fig. 3e, the QPGM can capture the edges of the sea urchin as well as QPGM. Figure 4a schematically illustrates the double-sided QPGM.
the detailed morphology inside the cells. Thus, the QPGM is able to Although conceptually similar to the system discussed in the previ-
measure the phase gradient information of the biological samples, ous section, the double-sided QPGM is more compact, mechanically
which is directly related to cellular mass transport39. robust and does not need further alignment after fabrication. The
To demonstrate further miniaturization of the system, we per- two metasurface layers are based on the same platform discussed
formed additional measurements with an off-the-shelf CMOS in the previous section, but they are fabricated on the two sides of
image sensor. Figure 3f presents a schematic of the optical system. a 1-mm-thick fused-silica substrate (see Methods, Supplementary
The distance from the object plane to the CMOS image sensor is just Note 1 and Supplementary Table 2 for details about the design and
5.09 mm (see Supplementary Fig. 11a,b for details of the compact fabrication). The optical images of the device are shown in Fig. 4b.
optical set-up). Figure 3g shows three raw DIC images captured The total volume of the QPGM is 0.62 × 0.41 × 1.00 mm3, it has a
by the set-up in Fig. 3f when using the imaging target shown in Fig. 3a. magnification of ×1.60 and a FOV with diameter of 140 μm. To
Using equation (2), the three DIC images in Fig. 3g result in the verify the capability of the double-sided QPGM, we used the same
PGI plotted in Fig. 3h. The edges and the measured phase gradient resolution targets as in Fig. 3c,d. The results, shown in Fig. 4c,d,

112 Nature Photonics | VOL 14 | February 2020 | 109–114 | www.nature.com/naturephotonics


Nature Photonics Articles
a Side Amorphous
b
Gold view silicon

SU-8

Fused
1 mm
silica
z

Bottom Nanoposts Top


view view

210 µm

y 200 µm 200 µm

x
d
c 400
(rad µm–1)

Thickness (nm)
0.7 300

200
0
100 AFM
QPGM
0
–0.7 1 2 3 4 5 6 7
Sample no.

Fig. 4 | Imaging with the doublet QPGM formed from monolithic integration of two metasurface layers on the same substrate. a, Schematic of the
miniaturized QPGM using the double-sided metasurface from different viewing angles. The two metasurface layers are patterned on the two sides of a
1-mm-thick fused-silica substrate. b, Optical images of bottom (left) and top (right) views of the 8 × 8 array of the doublet QPGM, in addition to enlarged
images for portions of the devices. Scale bars, 200 μm c, PGIs captured using the doublet QPGM from the same parts of the resolution target as used
in Fig. 3c. Scale bars, 15 μm. d, Thicknesses of seven different parts of the phase target calculated from the PGIs captured by the metasurface QPGM,
and those measured by AFM. The plotted thicknesses estimated with the QPGM are averaged over 100 arbitrarily chosen points on the sample edges.
Error bars represent standard deviations of the estimated values.

verify that the double-sided metasurface QPGM generates PGIs Outlook and summary
that are comparable to those of the QPGM based on the two sepa- It is worth noting that the multifunctionality via both polariza-
rate metasurface layers. In particular, the estimated thicknesses of tion and spatial multiplexing schemes, which is the key property
the phase samples plotted in Fig. 4d are in good agreement with for the design of the miniaturized QPGM, is very hard to achieve
the values measured by AFM. Phase images of sea urchin samples, in any other platform, if at all. We envision that these and other
captured by the double-sided metasurface QPGM, are shown in versatile properties of metasurfaces enable various types of quan-
Supplementary Fig. 12. titative phase imaging devices. For example, three different images
at different axial positions can be captured to calculate the phase
Discussion information rather than the phase gradient information through
One limitation of the proposed system, especially for the double- the transport-of-intensity equation11,45. In addition, different focus
sided metasurface device, is its FOV. The small FOV mostly results scanning schemes can be integrated with the QPGM to achieve a
from the fact that our system is close to a 4 − f configuration that tomographic quantitative phase imaging device with fast axial scan-
requires the sum of the focal lengths of the two lenses to be compa- ning46–48. Moreover, we expect that around ten to a hundred of the
rable to their separation. To increase the FOV, a possible solution is miniaturized microscopes could be integrated on a single CMOS
the recently reported folded metasurface platform29, which might sensor for highly parallelized microscopy. Finally, a synergistic com-
be able to achieve a large FOV and a small footprint simultaneously. bination of the metasurfaces and computational optics is an emerg-
In this platform, the propagation space between the two lenses is ing area for enhancing the potentials of metasurface optical systems.
folded inside the substrate through multiple reflections, and there- Considering the computational aspects of quantitative phase imag-
fore the effective distance between the lenses can be significantly ing, we believe that new kinds of miniaturized quantitative phase
larger than the substrate thickness. Furthermore, adding another imaging device that will benefit from the enhanced optical control
metasurface layer can be used to mitigate the geometric aberrations of the metasurfaces can be proposed.
further and increase the FOV31. In conclusion, we utilized two multifunctional metasurface layers
The QPGM system is sensitive to the equality of optical axis to realize miniaturized quantitative phase gradient microscopes.
separation for TE and TM polarizations in the two metasurface The design and working principle were proposed and investigated
lenses. Nevertheless, this separation can be controlled very pre- both through simulation and experiment. Utilizing vertically inte-
cisely because metasurfaces allow for implementation of almost grated multifunctional metasurfaces, we experimentally captured
arbitrary phase profiles in subwavelength scales. If the separations phase gradient images of several transparent samples and verified
between the optical axes are identical for the two metasurfaces, the the quantitative phase imaging capability of the systems. This work
structure becomes robust against lateral and axial misalignment clearly demonstrates potentials of dielectric metasurface platforms
(Supplementary Fig. 13). in quantitative phase imaging systems to make miniaturized imaging

Nature Photonics | VOL 14 | February 2020 | 109–114 | www.nature.com/naturephotonics 113


Articles Nature Photonics

devices such as miniaturized microscopes or endoscopes30. With the 27. Arbabi, E. et al. Two-photon microscopy with a double-wavelength
current great interest in quantitative phase imaging and miniaturized metasurface objective lens. Nano Lett. 18, 4943–4948 (2018).
28. Zhu, A. Y. et al. Compact aberration-corrected spectrometers in the visible
microscopes, we envision that the metasurfaces will play a significant using dispersion-tailored metasurfaces. Adv. Opt. Mater. 7, 1801144 (2018).
role in the development of these technologies. 29. Faraji-Dana., M. et al. Compact folded metasurface spectrometer.
Nat. Commun. 9, 4196 (2018).
Online content 30. Pahlevaninezhad, H. et al. Nano-optic endoscope for high-resolution optical
coherence tomography in vivo. Nat. Photon. 12, 540–547 (2018).
Any methods, additional references, Nature Research reporting
31. Arbabi, A. et al. Miniature optical planar camera based on a wide-angle
summaries, source data, extended data, supplementary informa- metasurface doublet corrected for monochromatic aberrations. Nat. Commun. 7,
tion, acknowledgements, peer review information; details of author 13682 (2016).
contributions and competing interests; and statements of data and 32. Arbabi, A., Arbabi, E., Horie, Y., Kamali, S. M. & Faraon, A. Planar
code availability are available at https://doi.org/10.1038/s41566- metasurface retroreflector. Nat. Photon. 11, 415–420 (2017).
33. Avayu, O., Almeida, E., Prior, Y. & Ellenbogen, T. Composite functional
019-0536-x. metasurfaces for multispectral achromatic optics. Nat. Commun. 8,
14992 (2017).
Received: 22 March 2019; Accepted: 13 September 2019; 34. Zhou, Y. et al. Multilayer noninteracting dielectric metasurfaces for
Published online: 28 October 2019 multiwavelength metaoptics. Nano Lett. 18, 7529–7537 (2018).
35. Guo, C., Xiao, M., Minkov, M., Shi, Y. & Fan, S. Photonic crystal slab Laplace
References operator for image differentiation. Optica 5, 251–256 (2018).
1. Zernike, F. How I discovered phase contrast. Science 121, 345–349 (1955). 36. Kwon, H., Sounas, D., Cordaro, A., Polman, A. & Alù, A. Nonlocal
2. Lang, W. Nomarski Differential Interference-Contrast Microscopy metasurfaces for optical signal processing. Phys. Rev. Lett. 121, 173004 (2018).
(Carl Zeiss, 1982). 37. Huang, P. S. & Zhang, S. Fast three-step phase-shifting algorithm. Appl. Opt. 45,
3. Popescu, G. Quantitative Phase Imaging of Cells and Tissues 5086–5091 (2006).
(McGraw Hill, 2011). 38. Arbabi, E., Kamali, S. M., Arbabi, A. & Faraon, A. Full Stokes imaging
4. Park, Y., Depeursinge, C. & Popescu, G. Quantitative phase imaging in polarimetry using dielectric metasurfaces. ACS Photon. 5, 3132–3140 (2018).
biomedicine. Nat. Photon. 12, 578–589 (2018). 39. Nguyen, T. H., Kandel, M. E., Rubessa, M., Wheeler, M. B. & Popescu, G.
5. Alford, R. et al. Toxicity of organic fluorophores used in molecular imaging: Gradient light interference microscopy for 3D imaging of unlabeled
literature review. Mol. Imaging 8, 341–354 (2009). specimens. Nat. Commun. 8, 210 (2017).
6. Lee, K. et al. Quantitative phase imaging techniques for the study of cell 40. Wang, Z. et al. Spatial light interference microscopy (SLIM). Opt. Express 19,
pathophysiology: from principles to applications. Sensors 13, 4170–4191 (2013). 1016–1026 (2011).
7. Marquet, P. et al. Digital holographic microscopy: a noninvasive contrast 41. Shaked, N. T. Quantitative phase microscopy of biological samples using a
imaging technique allowing quantitative visualization of living cells with portable interferometer. Opt. Lett. 37, 2016–2018 (2012).
subwavelength axial accuracy. Opt. Lett. 30, 468–470 (2005). 42. Bon, P., Maucort, G., Wattellier, B. & Monneret, S. Quadriwave lateral
8. Choi, W. et al. Tomographic phase microscopy. Nat. Methods 4, 717–719 (2007). shearing interferometry for quantitative phase microscopy of living cells.
9. Kim, T. et al. White-light diffraction tomography of unlabelled live cells. Opt. Express 17, 13080–13094 (2009).
Nat. Photon. 8, 256–263 (2014). 43. Baek, Y., Lee, K., Yoon, J., Kim, K. & Park, Y. White-light quantitative phase
10. Zheng, G., Horstmeyer, R. & Yang, C. Wide-field, high-resolution Fourier imaging unit. Opt. Express 24, 9308–9315 (2016).
ptychographic microscopy. Nat. Photon. 7, 739–745 (2013). 44. Bouchal, P. et al. Geometric-phase microscopy for quantitative phase imaging
11. Greenbaum, A. et al. Wide-field computational imaging of pathology slides of isotropic, birefringent and space-variant polarization samples. Sci. Rep. 9,
using lens-free on-chip microscopy. Sci. Transl. Med. 6, 267ra175 (2014). 3608 (2019).
12. Ghosh, K. K. et al. Miniaturized integration of a fluorescence microscope. 45. Paganin, D. & Nugent, K. A. Noninterferometric phase imaging with partially
Nat. Methods 8, 871–878 (2011). coherent light. Phys. Rev. Lett. 80, 2586–2589 (1998).
13. Helmchen, F., Fee, M. S., Tank, D. W. & Denk, W. A miniature head-mounted 46. Arbabi, E. et al. MEMS-tunable dielectric metasurface lens. Nat. Commun. 9,
two-photon microscope: high-resolution brain imaging in freely moving 812 (2018).
animals. Neuron 31, 903–912 (2001). 47. Kamali, S. M., Arbabi, E., Arbabi, A., Horie, Y. & Faraon, A. Highly tunable
14. Ziv, Y. et al. Long-term dynamics of CA1 hippocampal place codes. elastic dielectric metasurface lenses. Laser Photon. Rev. 10, 1002–1008 (2016).
Nat. Neurosci. 16, 264–266 (2013). 48. She, A., Zhang, S., Shian, S., Clarke, D. & Capasso, F. Adaptive metalenses
15. Jahani, S. & Jacob, Z. All-dielectric metamaterials. Nat. Nanotechnol. 11, with simultaneous electrical control of focal length, astigmatism and shift.
23–36 (2016). Sci. Adv. 4, eaap9957 (2018).
16. Genevet, P., Capasso, F., Aieta, F., Khorasaninejad, M. & Devlin, R. Recent
advances in planar optics: from plasmonic to dielectric metasurfaces. Optica 4, Acknowledgements
139–152 (2017). This work was supported by the Caltech Innovation Initiative programe. The device
17. Lalanne, P., Astilean, S., Chavel, P., Cambril, E. & Launois, H. Blazed binary nanofabrication was performed at the Kavli Nanoscience Institute at Caltech. We thank
subwavelength gratings with efficiencies larger than those of conventional C. Choi and C. Yang for Fourier ptychography microscope measurements and helpful
échelette gratings. Opt. Lett. 23, 1081–1083 (1998). discussions. H.K. acknowledges a fellowship from Ilju organization.
18. Lin, D., Fan, P., Hasman, E. & Brongersma, M. L. Dielectric gradient
metasurface optical elements. Science 345, 298–302 (2014).
19. Zhan, A. et al. Low-contrast dielectric metasurface optics. ACS Photon. 3,
Author contributions
H.K. and A.F. conceived the project. H.K., E.A., S.M.K. and M.F.-D. designed and
209–214 (2016).
fabricated the samples. H.K. performed the simulations and measurements. H.K., E.A.,
20. Arbabi, A., Horie, Y., Bagheri, M. & Faraon, A. Dielectric metasurfaces for
S.M.K. and M.F. analysed the data. H.K., E.A. and A.F. co-wrote the manuscript.
complete control of phase and polarization with subwavelength spatial
All authors discussed the results and commented on the manuscript.
resolution and high transmission. Nat. Nanotechnol. 10, 937–943 (2015).
21. Arbabi, E., Arbabi, A., Kamali, S. M., Horie, Y. & Faraon, A. Controlling the
sign of chromatic dispersion in diffractive optics with dielectric metasurfaces. Competing interests
Optica 4, 625–632 (2017). H.K., E.A. and A.F. have submitted a patent application based on the idea presented
22. Chen, W. T. et al. A broadband achromatic metalens for focusing and in this work.
imaging in the visible. Nat. Nanotechnol. 13, 220–226 (2018).
23. Maguid, E. et al. Photonic spin-controlled multifunctional shared-aperture
antenna array. Science 352, 1202–1206 (2016). Additional information
24. Arbabi, E., Arbabi, A., Kamali, S. M., Horie, Y. & Faraon, A. Multiwavelength Supplementary information is available for this paper at https://doi.org/10.1038/
polarization-insensitive lenses based on dielectric metasurfaces with s41566-019-0536-x.
meta-molecules. Optica 3, 628–633 (2016). Correspondence and requests for materials should be addressed to A.F.
25. Kamali, S. M. et al. Angle-multiplexed metasurfaces: encoding independent Reprints and permissions information is available at www.nature.com/reprints.
wavefronts in a single metasurface under different illumination angles.
Phys. Rev. X 7, 041056 (2017). Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims
26. Shi, Z. et al. Single-layer metasurface with controllable multiwavelength in published maps and institutional affiliations.
functions. Nano Lett. 7, 041056 (2017). © The Author(s), under exclusive licence to Springer Nature Limited 2019

114 Nature Photonics | VOL 14 | February 2020 | 109–114 | www.nature.com/naturephotonics


Nature Photonics Articles
Methods chemical vapour deposition. Metasurface layer 2 was fabricated using the same
Simulation and design. The simulation results presented in Supplementary Fig. 4 process employed for the single-sided metasurfaces. Metasurface 2 was provided
were obtained by finding the transmission coefficients of the corresponding with an SU-8 polymer cladding layer (SU-8 2002, MicroChem), which served to
periodic metasurfaces using the rigorous coupled wave analysis technique49. protect this metasurface during fabrication of metasurface 1. A 4-μm-thick layer
The amorphous silicon nanoposts were assumed to be 664 nm tall and the square of SU-8 was spin-coated on the sample, baked at 90 °C for 4 min, and reflowed at
lattice constant was 380 nm. The nanoposts were capped with an 8-μm-thick SU-8 200 °C for 5 min to achieve a completely planarized surface. The SU-8 polymer
polymer layer. The refractive indices of amorphous silicon, fused silica and SU-8 was then exposed with ultraviolet light and cured by baking at 200 °C for another
for the operation wavelength of 850 nm in the simulations were 3.56, 1.44 and 90 min. To align metasurface layers 1 and 2, a second set of alignment marks
1.58, respectively. The side lengths of the nanoposts, Dx and Dy, were varied in were patterned on the side of metasurface 1 and aligned to the alignment marks
simulations to achieve full and independent phase control over the 2π range for TE on the side of metasurface layer 2 using optical lithography. Next, metasurface 1
and TM polarizations (see Supplementary Fig. 4a–d for the simulation results)20. was fabricated by the same method as the single-sided metasurfaces. However,
We optimized Dx and Dy as functions of ϕTE and ϕTM to provide high transmission the Al2O3 mask was not removed from the top of metasurface layer 1 because a
and the desired phase shifts. The optimized maps of the side lengths as functions mixture of ammonia and hydrogen peroxide at 80 °C would damage the SU-8
of the phase delays for TE- and TM-polarized light are plotted in Supplementary cladding layer of metasurface layer 2. It should be noted that the 60-nm-thick
Fig. 4e,f (see Supplementary Fig. 4g,h for the simulated transmittance corresponding Al2O3 layer was considered an integral part of metasurface layer 1 in the design
to the TE and TM polarizations). For the double-sided metasurface device, process. Like metasurface layer 2, metasurface layer 1 was also provided with a
metasurface layer 2 was designed using the same look-up table as in Supplementary 4-μm-thick SU-8 layer. Next, the circular apertures on both sides were patterned
Fig. 4e,f, while metasurface layer 1 was composed of nanoposts with a 60-nm-thick by photolithography, the deposition of chrome and gold (10 nm/100 nm) layers and
Al2O3 layer on top of the amorphous silicon layer. To analyse the presence of the liftoff. Finally, additional 4-μm-thick SU-8 polymer layers were spin-coated on the
Al2O3 layer, we performed additional simulations. In particular, the refractive index both sides to protect the apertures.
of Al2O3 for the operation wavelength in the simulation was 1.76. The optimized
results are plotted in Supplementary Fig. 4i–l. Measurement procedure. The imaging performance of the QPGM was characterized
We used the wave propagation method for the numerical studies of using the set-ups shown schematically in Supplementary Fig. 6. An 850 nm
Supplementary Fig. 1 to calculate the optical fields for TE- and TM-polarized light-emitting diode (LED; Thorlabs LED851L) was used as the light source.
light. The interference intensity patterns at the image plane can then be directly A linear polarizer (Thorlabs, LPVIS100-MP2) was placed in front of the LED
calculated from the TE- and TM-polarized fields. Metasurfaces are treated and set at 45° to confirm the polarization state of the input light. The sample
as phase plates in simulations, and their phase profiles used in the numerical image was captured by the two metasurface layers, which were mounted on
simulations are provided in Supplementary Note 1. It is worth noting explicitly three-axis translation stages to enable alignment. The field at the image plane of
that the optimal phase profiles of the two metasurface layers in Supplementary the QPGM was captured by a custom-built optical microscope. The microscope
Fig. 1e were obtained through optimization with the ray-tracing technique using consisted of an objective lens (Olympus, LMPlanFL 10X) and a tube lens (Thorlabs
commercial optical design software (Zemax OpticStudio, Zemax) to minimize AC254-150-B-ML, focal length of 15 cm). The second linear polarizer, set at −45°,
geometric aberrations. was inserted between the objective lens and the tube lens to form interference
between TE- and TM-polarized light. An optical bandpass filter (Thorlabs,
FL850-10) in front of the camera was used to limit the bandwidth of the LED
Device fabrication. Metasurface layers 1 and 2, shown in Fig. 2d, were and remove the background. For the double-sided metasurface devices, the optical
fabricated on two different 1-mm-thick fused-silica substrates. A 664-nm-thick system was almost the same as the one used for the two separated metasurface
layer of amorphous silicon was deposited using plasma-enhanced chemical layers, with the double-sided metasurface device mounted on one three-axis
vapour deposition. For nanopatterning, an ~300-nm-thick positive electron translation stage instead of two.
resist (ZEP-520A) was used. In addition, an ~60-nm-thick water-soluble For the set-up shown in Supplementary Fig. 11a, a CMOS image sensor
conductive polymer (aquaSAVE, Mitsubishi Rayon) was spin-coated for charge (MT9J001, Arducam) replaced the custom-built microscope. A 260-μm-thick
dissipation. The patterns were generated using electron-beam lithography linear polarizer (LPVIS100-MP2, Thorlabs) was attached on top of a glass
(EBPG-5000+, Raith). The conductive polymer was then dissolved in water substrate protecting the CMOS image sensor by double-sided tape. As shown
and the resist was developed in resist developer solution (ZED-N50, Zeon in Supplementary Fig. 11a, the bandpass filter was placed between the LED and
Chemicals). A 60-nm-thick Al2O3 layer was deposited by electron-beam lens (LB1761-C, Thorlabs).
evaporation. The pattern was transferred to the Al2O3 layer by a liftoff process in
solvent (Remover PG, MicroChem). The patterned Al2O3 layer worked as a hard
mask to etch the amorphous silicon layer. The dry etching step was performed Data availability
in a mixture of SF6 and C4F8 plasmas using an inductively coupled plasma The data that support the findings of this study are available from the
reactive ion etching process. The Al2O3 mask was dissolved in a 1:1 mixture of corresponding author upon request.
ammonium hydroxide and hydrogen peroxide heated to 80 °C.
The double-sided metasurfaces shown in Fig. 4b were patterned on both sides References
of a 1-mm-thick fused-silica substrate. Two 664-nm-thick layers of amorphous 49. Liu, V. & Fan, S. S4: a free electromagnetic solver for layered periodic
silicon were deposited on both sides of the substrate using plasma-enhanced structures. Comput. Phys. Commun. 183, 2233–2244 (2012).

Nature Photonics | www.nature.com/naturephotonics

You might also like