ArticlePDF Available

Super-Resolution Imaging of a Dielectric Microsphere Is Governed by the Waist of Its Photonic Nanojet

Authors:
  • Shenzhen Institutes of Advanced Technology, Chinese Academy of Sciences

Abstract and Figures

Dielectric microspheres with appropriate refractive index can image objects with super-resolution, i.e. with a precision well beyond the classical diffraction limit. A microsphere is also known to generate, upon illumination, a photonic nanojet, which is a scattered beam of light with a high-intensity main lobe and very narrow waist. Here we report a systematic study of the imaging of water-immersed nanostructures by barium titanate glass microspheres of different size. A numerical study of the light propagation through a microsphere points out the light focusing capability of microspheres of different size and the waist of their photonic nanojet. The former correlates to the magnification factor of the virtual images obtained from linear test nanostructures, the biggest magnification being obtained with microspheres of ~ 6-7 µm in size. Analyzing the light intensity distribution of microscopy images allows determining analytically the point spread function of the optical system and thereby quantifies its resolution. We find that the super-resolution imaging of a microsphere is dependent on the waist of its photonic nanojet, the best resolution being obtained with a 6 μm Ø microsphere, which generates the nanojet with the minimum waist. This comparison allows elucidating the super-resolution imaging mechanism.
Content may be subject to copyright.
Super-Resolution Imaging of a Dielectric Microsphere Is Governed by
the Waist of Its Photonic Nanojet
Hui Yang,
Raphaël Trouillon, Gergely Huszka, and Martin A. M. Gijs*
Laboratory of Microsystems, École Polytechnique Fé
dé
rale de Lausanne, 1015 Lausanne, Switzerland
*
SSupporting Information
ABSTRACT: Dielectric microspheres with appropriate refractive index can image
objects with super-resolution, that is, with a precision well beyond the classical
diraction limit. A microsphere is also known to generate upon illumination a
photonic nanojet, which is a scattered beam of light with a high-intensity main lobe
and very narrow waist. Here, we report a systematic study of the imaging of water-
immersed nanostructures by barium titanate glass microspheres of dierent size. A
numerical study of the light propagation through a microsphere points out the light
focusing capability of microspheres of dierent size and the waist of their photonic
nanojet. The former correlates to the magnication factor of the virtual images
obtained from linear test nanostructures, the biggest magnication being obtained
with microspheres of 67μm in size. Analyzing the light intensity distribution of
microscopy images allows determining analytically the point spread function of the
optical system and thereby quanties its resolution. We nd that the super-resolution
imaging of a microsphere is dependent on the waist of its photonic nanojet, the best
resolution being obtained with a 6 μm Ø microsphere, which generates the nanojet with the minimum waist. This comparison
allows elucidating the super-resolution imaging mechanism.
KEYWORDS: Microsphere, super-resolution imaging, photonic nanojet, optical microscopy, point spread function
Conventional optical microscopes are limited by the so-
called diraction limit and can resolve features of around
half of the wavelength of illumination λ, as they are only
capable of transmitting the propagating wave components
emanating from the illuminated object.
1,2
The evanescent
components that carry the ne information about the object
decay exponentially in a medium with positive permittivity and
permeability and are lost before reaching the image plane.
Several eorts to this date have been pursued to overcome the
diraction limit using various approaches including near-eld
scanning probes,
36
super/hyper-lenses driven by surface-
plasmon excitation,
2,712
and uorescence microscopy with
molecular excitation.
1317
However, the applications of these
methods have been limited in part due to their sophisticated
engineering designs or pretreatment steps. Alternatively, it has
been demonstrated that the use of nanoscale lenses,
1820
polymer microdroplets,
21
and especially dielectric micro-
spheres
2230
on top of the objects can achieve near-eld
focusing and magnication, which in turn results in the
capability to resolve features beyond the diraction limit. Wang
et al. rst proposed the use of fused silica beads with refractive
index n1.46 and diameter from 2 to 9 μm in combination
with a conventional optical microscope to achieve a resolution
of 50 nm in air with a white light source.
22
Later on, it was
reported that large polystyrene microspheres (above 30 μm)
can also achieve super-resolution imaging with large eld-of-
view in air without immersion liquid.
24
Moreover, it was
demonstrated that high-index microspheres (n1.92.1), fully
immersed in liquid, actually allow enhanced imaging with
minimum resolved feature sizes of λ/7 with white-light
illumination for imaging of nanofeatures
23
and adenoviruses,
25
as well as with uorescent microscopic setup for resolving the
structures of subcellular organelles.
26
More recently, micro-
spheres were used in combination with confocal microscopy for
achieving a 25 nm lateral resolution under 408 nm wavelength
illumination.
27
In these works, the microspheres are placed on
top of the sample object, where they collect the underlying
samples near-eld nanofeatures and subsequently transform
the near-eld evanescent waves into far-eld propagating waves,
creating a magnied image in the far-eld, which is collected by
a conventional optical microscope.
22,26,30
The super-resolution
capability of microspheres is linked to two factors: (i) the
microsphere performs as a solid-immersion lens and provides
local enhancement of the refractive index and a reduction of the
illumination wavelength; (ii) the development of the photonic
nanojet.
26,30
A substantial literature has developed regarding
the existence, properties, and potential applications of the
photonic nanojet.
3156
In principle, a photonic nanojet is a
narrow light beam with high optical intensity that can be
generated by a transparent dielectric symmetric body, like a
microcylinders or microsphere, upon illumination. The nanojet
that emerges from a microsphere locates in the immediate
Received: March 24, 2016
Revised: July 5, 2016
Published: July 11, 2016
Letter
pubs.acs.org/NanoLett
© 2016 American Chemical Society 4862 DOI: 10.1021/acs.nanolett.6b01255
Nano Lett. 2016, 16, 48624870
Figure 1. Mechanism of nanojet generation and imaging of a dielectric microsphere. (a) Focusing of a plane wave light beam into a nanojet at the
rear surface of a free-standing microsphere. At the front surface of the microsphere, the light is refracted at low incidence angle, while at higher
incidence angle it gets mostly reected. (b) FEM simulation of the light propagation through a 6 μm Ø barium titanate microsphere in water. The
linear region where substantial refracted light enters the microsphere at its front surface is referred to as L, while the width of the exiting beam at the
rear surface is denoted as l; the waist of the nanojet is referred to as w. (c,d) FEM simulation of the light propagation through 2 and 16 μm Ø barium
titanate microspheres in water medium, respectively. (e) When a microsphere is positioned on a grating structure with line width dand illuminated
from the front, the light reected by the grating allows detecting a virtual image with magnication factor M. When the distance hbetween the
microsphere and the grating is small enough (of order of the illumination wavelength λ), the near-eld evanescent wave carrying the ne details of
Nano Letters Letter
DOI: 10.1021/acs.nanolett.6b01255
Nano Lett. 2016, 16, 48624870
4863
vicinity of the rear-surface of the sphere. It is a nonresonant
phenomenon that appears for spheres with a diameter of the
order of 10100 λand with a ratio of their refractive index nms
to the refractive index of the background medium (water in this
study) nwthat is smaller than about 2. The nanojet can
maintain a subwavelength full width at half-maximum (fwhm)
transverse beam width along a path that can extend more than
2λbeyond the sphere, and the minimum fwhm beam width,
referred to as waistin this paper, can be smaller than the
classical diraction limit, in fact as small as λ/3.
The imaging resolution of a classical microscope depends on
the size of the spot that is generated by focusing the incident
propagating wave in the far-eld and is therefore limited by
diraction. For microsphere-assisted optical microscopy,
evanescent waves close to the surface of the microsphere play
a signicant role. Even though it was suggested that the
development of the photonic nanojet is essential to the super-
resolution imaging capability of a microsphere,
26,30
the exact
link remained unclear. Here we study, in a quantitative way the
role of the photonic nanojet for super-resolution imaging. First
a systematic numerical study of the light propagation through
microspheres of dierent size using the nite element method
(FEM) is performed. This allows characterizing the photonic
nanojet at the rear-surface of a microsphere and relating the
microspheres theoretical magnication factor to the light
focusing capability of the photonic nanojet. Second, we perform
an experimental study in which a systematic series of barium
titanate glass microspheres with diameter from 3 to 21 μm are
used to image linear test nanostructures that are immersed in
water. The experimental magnication factor and the point
spread function that is analytically determined from the images
allow evaluating the resolution of the optical system, which is
shown to directly correlate with the calculated properties of a
microspheres photonic nanojet.
As schematically illustrated in Figure 1a, when a microsphere
is illuminated by a propagating light beam from the far-eld, the
light is mostly refracted on its frontal surface at small incident
angle and reected at higher incident angle, the limiting angle
between the two regions given approximately by the Brewster
angle of the optical interface. When the size of the microsphere
is much bigger than the illumination wavelength, it is a good
approximation to use ray optics to explain how the light that is
incident on the top-surface of the microsphere propagates. The
refracted light propagates through the microsphere and
generates a photonic nanojet near the rear-surface of the
microsphere (details described in the Supporting Information).
An FEM study of the light wave propagation through a barium
titanate glass microsphere (nms = 1.92), with diameter Ø
ranging from 2 to 20 μm in surrounding water medium (nw=
1.33) is performed (details described in Methods). An
electromagnetic wave with wavelength λ= 600 nm is applied
to a boundary with a length that is the same as the size of the
microsphere and that is far away from its top-surface. As the
size of the boundary is much longer than the wavelength, a
plane-wave incident light beam is a good approximation for
studying the photonic nanojet. Figure 1bd shows the light
intensity distributions near a 6, 2, and 16 μm microsphere,
respectively. When the microsphere is 6 μm in size (Figure 1b),
the photonic nanojet directly emerges from the microsphere
surface. When decreasing the microspheres size, the focal point
of the nanojet moves into the microsphere, as seen for the
simulation result of a 2 μm Ø microsphere (Figure 1c). For a
microsphere size bigger than 6 μm, the external focal point of
the nanojet moves away from the rear-surface of the
microsphere and Figure 1d shows the simulation result
obtained for a 16 μm Ø microsphere. To quantitatively study
how light is focused by the microsphere into the photonic
nanojet, we determined from the model calculation the linear
region where substantial refracted light enters the microsphere
at its front surface, referred to as Lin Figure 1bd, while the
width of the exiting beam at the rear surface is denoted as l. The
waist of the nanojet is referred to as w(see Figure 1b), which is
the fwhm of the nanojet along the x-axis at the peak intensity of
the y-axis. It should be noted that the light source in the FEM
study is actually coherent, while the light source in conventional
optical microscopy is incoherent. The comparison on the light
intensity distribution in the photonic nanojet under coherent
and incoherent illumination is further discussed in the
Supporting Information. The study shows that the intensity
prole along the nanojet waist does not signicantly change,
when using either an incoherent or a coherent light source,
indicating that a coherent light-based simulation can provide
sucient information to study the resolution of the imaging
system.
The imaging mechanism of a dielectric microsphere is
schematically illustrated in Figure 1e. When the focused light
that exits the microsphere illuminates the sample object (in this
case a linear grating structure with feature size d), no nanojet is
generated but instead the reected light follows a reection-
symmetric optical path, while evanescent waves that contain the
high spatial-frequency information on the object are converted
into propagating waves within the microsphere (see Supporting
Information) when the distance hbetween the microsphere and
the grating is small enough (of order of the illumination
wavelength λ). In the meanwhile, a magnied virtual image is
generated in the far-eld with magnication factor M. The
imaging capability of the microsphere directly correlates with
the formation of the focused photonic nanojet as obtained from
the numerical study, as the same optical paths are involved.
Figure 1f,g shows the calculated light focusing capability of a
microsphere, expressed by the ratio L/l, and the waist of the
photonic nanojet normalized by the illumination wavelength w/
λas a function of the microsphere diameter, respectively. A
bigger L/lindicates a better light focusing by the microsphere
and corresponds to a smaller waist of the nanojet. According to
our simulations, the microsphere with diameter of 6 μm shows
the best light focusing capability and smallest waist of the
photonic nanojet.
To experimentally study the super-resolution imaging eect,
grating structures consisting of 120 nm wide lines with 100 nm
Figure 1. continued
the grating can become propagating in the high refractive index sphere, and later in the medium where it is to be collected by the microscope
objective. (f) FEM simulation results of the light focusing capability of a microsphere, expressed by the ratio L/l, as a function of the microsphere
diameter. The dots are obtained from the simulation, while the red dotted line is a guide to the eye. (g) FEM simulation results of the normalized
waist of the photonic nanojet w/λ, as a function of the microsphere diameter. The dots are obtained from the simulation, while the red dotted line is
a guide to the eye.
Nano Letters Letter
DOI: 10.1021/acs.nanolett.6b01255
Nano Lett. 2016, 16, 48624870
4864
interspacing (see Figure 2a) immersed in water are used as test
samples to be imaged with transparent barium titanate glass
microspheres that are loosely positioned on top of them. A
conventional reection microscope (Zeiss Axioplan micro-
scope) equipped with a CCD camera (AxioCam MRm camera)
and a 40×water immersion objective with numerical aperture
(NA) of 0.8 is used to take images. A halogen lamp is used as
the white-light illumination source with a wide-band spectrum,
which is from 400 to 700 nm and the peak appears at λ=
600 nm (more details are shown in the Supporting
Information). Figure 2b is a microcopy image of the gratings
taken by the 40×objective along, the inset is a 5×magnied
image, clearly showing that the conventional optical microscope
cannot resolve the nanopatterns with a feature size of 100 nm.
However, when a microsphere is placed on top of the nano-
objects, the subdiraction features become observable. Figure
2c,e,g shows microspheres with diameters of 4.2 μm, 7.1, and
11.8 μm positioned on the sample, respectively. The focal plane
of the microscope coincides with the plane of maximum sphere
diameter, hence, on these pictures, the grating nanostructures
are out-of-focus. When the focal plane of the microscope is set
to the image plane of the microsphere, images obtained with
the microspheres of Figure 2c,e,g are shown in Figure 2d,f,h,
respectively. Compared to the size of the objects, the feature
sizes of the line structures in the images are clearly magnied.
As already schematically illustrated in Figure 1e, comparing the
size of the object with that of the image allows determining the
magnication factor M. The experimental Mvalues versus the
microsphere diameter are plotted in Figure 2i. We can see that
the biggest magnication is obtained with microspheres of 6
7μm in size. When we compare the experimental magnication
factor Mwith the theoretical light focusing capability L/l,as
obtained from the simulations, a positive correlation with a
Pearsons correlation coecient of 0.91 is obtained, indicating
that a better light focusing capability of a microsphere logically
results in a higher magnication factor.
A remaining issue is whether the light focusing capability of a
microsphere has any impact on its super resolution imaging.
For solving this question, linear test nanostructures with 300
nm wide lines and 900 nm interspacing are imaged through the
microspheres with dierent size by using the same optical
microscope setup. As illustrated in Figure 3a, the microsphere is
Figure 2. Imaging of a grating nanostructure using dierent size microspheres and a water-immersion objective. (a) Scanning electron microscopy
image of the silicon grating test structure containing 120 nm wide lines with an interspacing of 100 nm. (b) Optical microscopy image of the
nanostructure taken by a 40×water-immersion objective with NA of 0.8. The insert is a 5×magnied image, clearly showing that conventional
microscopy cannot resolve the nanostructures with this feature size. (ch) Optical microscopy images obtained by positioning on the grating
microspheres with sizes of (c,d) 4.2 μm, (e,f) 7.1 μm, and (g,h) 11.8 μm, respectively. The images of (c,e,g) are focused on the microspherescenter
plane, while the corresponding images (d,f,h) are focused on the virtual image plane, showing that the grating nanostructure is imaged with a
dierent magnication factor Mfor microspheres of dierent sizes. (i) The dots represent the experimental magnication factor as a function of the
microsphere diameter, while the solid line is a guide to the eye. (j) The experimental magnication factor Mas a function of the light focusing
capability L/lobtained from the simulations. The solid line represents a linear tting curve with a Pearsons correlation coecient of 0.91.
Nano Letters Letter
DOI: 10.1021/acs.nanolett.6b01255
Nano Lett. 2016, 16, 48624870
4865
positioned in the middle of two lines and the magnied image
can be obtained when the distance hbetween the microsphere
and the grating is of order of the illumination wavelength λ(see
Figure 1e), therefore only the interspacing at the center of the
microsphere (darker zone in the image) and its neighboring
two lines (clearer zones in the image) are visible in the far-eld.
Figure 3b,c shows the microscopic images obtained with the
40×water-immersion objective (NA = 0.8) focusing on a 6.4
μm microsphere and on the image plane, respectively. The light
intensity prole along the dashed line in Figure 3c is shown in
Figure 3f. Moreover, similarly taken images obtained from a 9.9
μm microsphere are shown in Figure 3d,e, while the light
Figure 3. Quantication of the experimental resolution using the analytical point spread function model and correlation with the waist of the
photonic nanojet. (a) A microsphere is positioned specically in the middle of two lines of a dedicated test grating (line width of 300 nm and
interspacing of 900 nm) to characterize the sharpness of line/interspacing boundary in the virtual image. (be) Optical microscopy images obtained
by positioning on the grating microspheres with sizes of (b,c) 6.4 and (d,e) 9.9 μm, respectively. The images of (b,d) are focused on the
microspherescenter plane, while the corresponding images (c,e) are focused on the virtual image plane. Each dashed line indicates where the
intensity prole will be taken that is to be tted with the analytical point spread function model. (f) Intensity distribution along the dashed line in
panel c with x1and x2the positions of the descending and ascending steps obtained from the t using eq 5. (g) Intensity distribution along the
dashed line in panel e. (h) The actual image standard deviation σthat is obtained from the t, related to the true resolution of the system, as a
function of microsphere size. The dots are obtained from the ts with the analytical model, while the solid curve is a guide to the eye. (i) The
correlation between σand the normalized waist of the photonic nanojet w/λ. The solid line represents a linear tting curve with a Pearsons
correlation coecient of 0.88.
Nano Letters Letter
DOI: 10.1021/acs.nanolett.6b01255
Nano Lett. 2016, 16, 48624870
4866
intensity prole along the dashed line in Figure 3e is shown in
Figure 3g. Comparing Figure 3fwithFigure 3g, clear
dierences on the peak intensities and on the intensity proles
are observed.
This type of image is further analyzed using an analytical
point spread function (PSF) model to study the resolution in a
quantitative way. The PSF is an important factor characterizing
the mechanisms underlying the formation of an image in an
optical system, dened as the response of the imaging system to
a point object.
5759
This three-dimensional function is
characteristic of the imaging system and can have a rather
complicated analytical expression, hence sometimes requiring
simplications or approximations to facilitate its use. However,
the shape of the PSF is directly related to the resolution of the
system, as the narrower the PSF, the better the resolution. The
imaging process described here is based on the collimation of
light that is incident on top of the microsphere into the nanojet.
For detection of an object, the light is reected back from the
object into the detector, essentially following the same mirror-
symmetric trajectory as during incidence. The width of the
nanojet land more precisely the quantity L/ldetailed in Figure
1b is a measure of the focusing capability of a microsphere.
Indeed, the smaller land the larger L/lis, the easier it will be
for an initially evanescent wave, when it is reected from the
object, to be converted to a wave with higher spatial frequency
that becomes propagating inside the microsphere. If higher
spatial frequencies can be detected, this will result in a sharper
image, hence in a better resolution and a narrower PSF. That is
why the theoretical simulations of the nanojet proles,
provided, for instance, in Figure 1, are indicative of the
diraction processes in the optical system and can be useful to
compare with the experimental PSF. Additionally, simulations
hint that the illumination prole partially describes the blurring
introduced by the system and is at least indicative of the PSF in
the absence of severe mismatches in the refractive indices.
60
As
a consequence, these illumination proles are used below to
establish assumptions on the overall PSF, and are justied a
posteriori.
Typically, the PSF can be approximated through models and
numerical simulations. However, the purpose of this section is
to nd an experimental approach to evaluate the lateral
resolution of the system through the PSF and to compare it to
some of the results predicted by the simulations. In the case
considered in this work, that is, where a plane is imaged, only
the xyprole of the PSF at the waist of the nanojet is relevant
to the resulting image. Moreover, because of the cylindrical
geometry of the problem, as shown in Figure 1bd, only the
expression of the radial component at the waist (that is, along
the x-axis in Figure S2c) is required to describe the PSF.
Furthermore, only sections of the images located at the center
of the microsphere are analyzed. This guarantees that
distortions due to the spherical shape of the microlens are
limited, and that small translations along the x- and y-axes at
the vicinity of the central axis of the system do not dramatically
alter the image. This is supported experimentally by the images
shown in Figure 2d,f,h where the gratings are clearly resolved at
the center of the microsphere and are parallel, hence ensuring
that no dramatic radial distortion occurs. As a consequence, at
the center of the microsphere the imaging device can be
assumed to be largely shift independent in the xyplane.
Combined with the linearity of the system, this fact suggests
that the image can be expected to be solely determined by the
intensity prole of the PSF.
61
Furthermore, as shown by the
shape of the illumination proles discussed in the Section S6 of
the Supporting Information, the shape of the nanojet is
independent of phase shifts. As the nal image is formed by the
illumination light reected by the object, one can assume that
the whole system and therefore the PSF are not dependent on
phase shifts. Briey, it allows to relate the recorded two-
dimensional (2D) image Im to the input object Ob through a
convolution operation (denoted as *)
62,63
≡*xy xyIm( , ) (PSF Ob)( , ) (1)
The PSF can be dened as the impulse response function of
the system, that is, the image obtained from the imaging of a
point. Mathematically, this impulse, or point, can be expressed
as the Dirac function δ, the function returning 0 for x0 and y
0, and whose integral is 1 over
2. More intuitively, δcan be
approximated as an innitely high, innitely sharp peak
centered over (0, 0), the function being equal to 0 anywhere
else. Furthermore, δis the unit element for convolution, hence
δ=*xy xy
P
SF( , ) (PSF )( , ) (2)
Direct measurement of the PSF can be challenging, as
obtaining a pure point as the object to image is impossible. It
can be approximated by imaging a very small disk for instance,
but the result can be distorted if the object is below the imaging
capabilities of the device. A numerical simulation, as shown in
Figure 1, or an exact analytical solution can be used, but this is
not always available. Moreover, the purpose of this analysis is to
conrm the results of the numerical simulation (Figure 1f,g)
with an experimental approach. To experimentally characterize
the PSF, it has been suggested to image a step, corresponding
to a Heavyside function Halong one of the two dimensions of
the image, here, for instance, along the x-axis. This is a more
rigorous and elegant way to evaluate the PSF and also the
reason why the grating nanostructure is used as sample in this
work. Indeed, δis the derivative of Halong the x-axis,
64
and the
convolution operation is stable through dierentiation, leading
to
=
*xxHx
P
SF( ) (PSF )( )
(3)
It is commonly assumed that the PSF is accurately described
with a 2D Gaussian,
6571
which can be reduced to one-
dimension (1D) in the case of imaging a step along the x-axis.
Even though an Airy function can be considered, as it describes
the diraction pattern generated by a small circular aperture on
the xy-plane, numerical investigations show that a Gaussian is
also a very common t in the xy-plane for the PSF of a confocal
microscope.
72
To conrm the validity of using a Gaussian
approximation in contrast to an Airy approximation, the
intensity prolesatthewaistalongthex-axis for the
simulations of the nanojet proles shown in Figure 1bd
were t with a Gaussian and an Airy function. In both cases,
these ttings were found to be imperfect (Figure S5) but
nevertheless resulted in comparably good ts (R2> 0.966). As
the use of a Gaussian facilitates considerably the calculations
and the numerical analysis, this function was chosen to
approximate the overall PSF of the system in the rest of this
study.
Thanks to this approximation, the image can now be
obtained by convoluting the step prole (the object) with a 1D
Gaussian (the PSF) characterized by a standard deviation σ
(see eq S4). By integrating eq 3 along the x-axis, the image
Nano Letters Letter
DOI: 10.1021/acs.nanolett.6b01255
Nano Lett. 2016, 16, 48624870
4867
actually results in the integral of a Gaussian (the PSF) and is by
denition described by the error function erf
π
xe
t
erf( ) 2d
xt
0
2
(4)
By tting the prole of the image with erf, we can extract the
σassociated with the PSF, which is directly associated with the
lateral resolution of the imaging system.
73
Experimentally, the microsphere sits in the middle of two
lines. As shown in Figure 3c,e, two stripes are observed for each
microsphere, resulting in two opposite steps. In agreement with
eq 4, the obtained image prole should be described by the
function f
α
σσ
=
+
+
f
xxx xx
() erfc ()
2erf ()
2constant
12
(5)
where erfc is the complementary error function, αis a constant
and x1and x2are the positions of the descending and ascending
steps, respectively (shown in Figure 3f), which together dene
the magnication of the image. The image standard deviation σ
is obtained from the tting, and has to be divided by the
amplication factor associated with the microsphere size (as
shown in Figure 2i) to obtained the actual σ. A description of a
typical experimental image, along with the tted f, is shown on
Figure 3f. From this tting, the characteristic σfor each
microsphere size can be computed and is presented on Figure
3h. The other tting parameters are discussed in the Supporting
Information and support the analysis presented here with x2x1
900 nm, which is in good agreement with the geometry of
the sample, and α100 for all the beads considered. In
particular, the fact that the distance x2x1was correctly
evaluated further validates the shift invariance assumption. If
this was not the case, as this parameter is measured over a large
part of the eld of view appearing on the microsphere, a
signicant distortion would have occurred thus preventing and
accurate measurement of the intergrating distance. The best σis
obtained with 6 μm size microspheres and is below 100 nm.
The experimental σis also in good agreement with the nanojet
waist values derived from the simulations. The actual σtted
from the experimental data as a function of w/λobtained from
the simulations is plotted in Figure 3i: a positive correlation
with the Pearsonseciency of 0.88 is obtained, indicating that
the imaging resolution of the microsphere is clearly dependent
on the waist of the photonic nanojet. Additionally, wis dened
at the fwhm of the nanojet along x, which in a Gaussian
approximation is σ
σ
2
2 ln(2) 2.355 . As the simulated wis
240 nm (for λ= 600 nm) and the experimental σis 100 nm,
the correspondence between the simulation and the exper-
imental analysis is excellent. This good agreement validates a
posteriori the assumptions on the overall PSF shape drawn
from considering the simulated nanojet proles. It also supports
the validity of the simulations, and that the resolution of the
system is largely controlled by the illumination pattern, that is,
the width of the waist of the nanojet.
Considering that the illumination source has a spectral range
from 400 to 700 nm (see Supporting Information) and a 100
nm structure is resolved in the experiments, the resolution res is
therefore in between λ/4 and λ/7 with λthe illumination
wavelength in vacuum. According to the Rayleigh criterion, the
minimum feature size that can be resolved by a water-
immersion objective with NA = 0.8 is 188 nm under an
illumination wavelength of 400 nm. The 100 nm resolution is
therefore obtained thanks to the use of microspheres, which
convert the evanescent waves with the high spatial-frequency
information on the object into propagating waves within the
microsphere (see Supporting Information).
Moreover, the minimum waist of the photonic nanojet is
0.4 λ(Figure 3i). The relationship between the resolution and
the nanojet waist can hence be written as res 0.36w0.63w.
The diraction limit of an objective or lens is dened as λ/
(2NA). Our method can resolve an object with feature size of
100 nm under a wide-band illumination with spectrum from
400 to 700 nm, so that the NA of the imaging system is
obtained as 23.5 when considering the minimum and
maximum illumination wavelength, respectively. This means
that in our study the use of the 40×microscope objective
together with a 6 μm Ø microsphere would permit a resolution
that would be provided by a hypothetical (as nonexisting)
microscope objective with NA = 23.5.
In conclusion, we reported imaging of a samples nano-
features beyond the classical diraction limit by using a
conventional optical microscope in combination with a series of
barium titanate glass dielectric microspheres of dierent sizes. A
FEM study on light propagation revealed the light focusing
capability of a microsphere of a given size and the generation of
the photonic nanojet. By comparing the experimental imaging
results with the numerical study, we found that the
magnication factor obtained from the virtual images is highly
correlated to the calculated light focusing capability of a
microsphere. Moreover, we quantitatively studied the reso-
lution of the microspheres of dierent sizes by analyzing the
images and tting the results with a mathematical model based
on the PSF. Our work demonstrated the intimate link of the
super-resolution imaging mechanism of a dielectric micro-
sphere with its light focusing capability and the development of
the photonic nanojet. Indeed, the combination of refractive and
interferometric eects of the incident light produce the narrow-
waist photonic nanojet that exits the microsphere. In the
imaging mode of the microsphere, identical optical paths are
used for generating the magnied image and, therefore, the
degree of focusing of the incident light into a nanojet is closely
related to the possibility of a microsphere to transform a high
spatial frequency evanescent wave generated by the object into
a propagating wave that becomes detectable in the far eld. We
believe that due to these physical insights, dielectric micro-
spheres will be increasingly used in the future, providing a
straightforward and robust tool to be integrated with a
conventional microscope for super-resolution optical micros-
copy. This will allow aordable super-resolution imaging of a
whole range of samples and biological objects, such as virus
particles, labeled nucleic acids and molecules.
Methods. Numerical Simulation. The numerical study on
light propagation through the microspheres and surrounding
water medium is carried out by FEM in COMSOL Multi-
physics software. A scalar equation is used to study transverse
electric waves in a 2D model. A light source with wavelength of
600 nm and the same width of the microsphere size is set away
from the front-surface of the microspheres, 600 nm
corresponding to the peak of the halogen lamp that is used
as the white-light illumination source in the experiments. After
meshing of the model, the element size is 22 nm, that is, 1/30
of the wavelength, which is suciently small to obtain a precise
solution. Microspheres with a size ranging from 2 to 20 μm are
analyzed in individual models. After each model is solved, Lis
Nano Letters Letter
DOI: 10.1021/acs.nanolett.6b01255
Nano Lett. 2016, 16, 48624870
4868
obtained by measuring the distance between the two points
with maximum light intensity on the front-surface of the
microsphere incident with the illumination light, lis obtained
by measuring the distance between the two points with
maximum light intensity on the rear-surface where light exits
the microsphere, and wis the fwhm of the nanojet along the x-
axis at the peak intensity of the y-axis.
Experimental Section. The silicon grating structures
consisting of 120 nm wide lines with 100 nm interspacing (in
Figure 2) are on a MetroChip microscope calibration target,
which is obtained from Pelco (Redding, CA, U.S.A.). The
grating structure with 300 nm wide lines and 900 nm
interspacing (in Figure 3) is made by chromium on a glass
substrate by photolithography. The optical microscopic images
are obtained by using a Zeiss Axioplan microscope mounted
with a AxioCam MRm camera (Carl Zeiss GmbH,
Oberkochen, Germany) and a 40×water immersion objective
with NA of 0.8. A halogen lamp is used as the white-light
illumination source with a peak at λ= 600 nm.
ASSOCIATED CONTENT
*
SSupporting Information
The Supporting Information is available free of charge on the
ACS Publications website at DOI: 10.1021/acs.nano-
lett.6b01255.
Additional information, gures, and table(PDF)
AUTHOR INFORMATION
Corresponding Author
*E-mail: martin.gijs@ep.ch.
Present Address
(H.Y.) IMEC, Leuven, Belgium
Notes
The authors declare no competing nancial interest.
ACKNOWLEDGMENTS
The authors would like to thank the European Research
Council (ERC-2012-AdG-320404) and the Swiss National
Science Foundation (200020-152948) for providing funding of
this work.
REFERENCES
(1) Pendry, J. B. Phys. Rev. Lett. 2000,85, 39663969.
(2) Fang, N.; Lee, H.; Sun, C.; Zhang, X. Science 2005,308, 534
537.
(3) Hecht, B.; Sick, B.; Wild, U. P.; Deckert, V.; Zenobi, R.; Martin,
O. J. F.; Pohl, D. W. J. Chem. Phys. 2000,112, 77617774.
(4)Lewis,A.;Taha,H.;Strinkovski,A.;Manevitch,A.;
Khatchatouriants, A.; Dekhter, R.; Ammann, E. Nat. Biotechnol.
2003,21, 13781386.
(5) Sun, S.; Leggett, G. J. Nano Lett. 2004,4, 13811384.
(6) Dunn, R. C. Chem. Rev. 1999,99, 28912927.
(7) Smolyaninov, I. I.; Hung, Y.-J.; Davis, C. C. Science 2007,315,
16991701.
(8) Liu, Z.; Lee, H.; Xiong, Y.; Sun, C.; Zhang, X. Science 2007,315,
1686.
(9) Liu, Z.; Durant, S.; Lee, H.; Pikus, Y.; Fang, N.; Xiong, Y.; Sun,
C.; Zhang, X. Nano Lett. 2007,7, 403408.
(10) Zhang, X.; Liu, Z. Nat. Mater. 2008,7, 435441.
(11) Rho, J.; Ye, Z.; Xiong, Y.; Yin, X.; Liu, Z.; Choi, H.; Bartal, G.;
Zhang, X. Nat. Commun. 2010,1, 143.
(12) Xiong, Y.; Liu, Z.; Sun, C.; Zhang, X. Nano Lett. 2007,7, 3360
3365.
(13) Hell, S. W. Science 2007,316, 11531158.
(14) Huang, B.; Babcock, H.; Zhuang, X. Cell 2010,143, 10471058.
(15) Simonson, P. D.; Rothenberg, E.; Selvin, P. R. Nano Lett. 2011,
11, 50905096.
(16) Jungmann, R.; Steinhauer, C.; Scheible, M.; Kuzyk, A.;
Tinnefeld, P.; Simmel, F. C. Nano Lett. 2010,10, 47564761.
(17) Engelhardt, J.; Keller, J.; Hoyer, P.; Reuss, M.; Staudt, T.; Hell,
S. W. Nano Lett. 2011,11, 209213.
(18) Lee, J. Y.; Hong, B. H.; Kim, W. Y.; Min, S. K.; Kim, Y.;
Jouravlev, M. V.; Bose, R.; Kim, K. S.; Hwang, I.-C.; Kaufman, L. J.;
Wong, C. W.; Kim, P.; Kim, K. S. Nature 2009,460, 498501.
(19) Kim, K. S. Proc. SPIE 2010,DOI: 10.1117/2.1201004.002883.
(20) McLeod, E.; Nguyen, C.; Huang, P.; Luo, W.; Veli, M.; Ozcan,
A. ACS Nano 2014,8, 73407349.
(21) Kang, D.; Pang, C.; Kim, S. M.; Cho, H. S.; Um, H. S.; Choi, Y.
W.; Suh, K. Y. Adv. Mater. 2012,24, 17091715.
(22) Wang, Z.; Guo, W.; Li, L.; Lukyanchuk, B.; Khan, A.; Liu, Z.;
Chen, Z.; Hong, M. Nat. Commun. 2011,2, 218.
(23) Darafsheh, A.; Walsh, G. F.; Negro, L. D.; Astratov, V. N. Appl.
Phys. Lett. 2012,101, 141128.
(24) Lee, S.; Li, L.; Ben-Aryeh, Y.; Wang, Z.; Guo, W. J. Opt. 2013,
15, 125710.
(25) Li, L.; Guo, W.; Yan, Y.; Lee, S.; Wang, T. Light: Sci. Appl. 2013,
2, e104.
(26) Yang, H.; Moullan, N.; Auwerx, J.; Gijs, M. A. M. Small 2014,
10, 17121718.
(27) Yan, Y.; Li, L.; Feng, C.; Guo, W.; Lee, S.; Hong, M. ACS Nano
2014,8, 18091816.
(28) Darafsheh, A.; Limberopoulos, N. I.; Derov, J. S.; Walker, D. E.,
Jr.; Astratov, V. N. Appl. Phys. Lett. 2014,104, 061117.
(29) Darafsheh, A.; Guardiola, C.; Nihalani, D.; Lee, D.; Finlay, J. C.;
Cá
rabe, A. Proc. SPIE 2015,9337, 933705.
(30) Yang, H.; Gijs, M. A. M. Microelectron. Eng. 2015,143,8690.
(31) Chen, Z.; Taflove, A.; Backman, V. Opt. Express 2004,12,
12141220.
(32) Li, X.; Chen, Z.; Taflove, A.; Backman, V. Opt. Express 2005,13,
526533.
(33) Lecler, S.; Takakura, Y.; Meyrueis, P. Opt. Lett. 2005,30, 2641
2643.
(34) Itagi, A. V.; Challener, W. A. J. Opt. Soc. Am. A 2005,22, 2847
2858.
(35) Chen, Z.; Taflove, A.; Li, X.; Backman, V. Opt. Lett. 2006,31,
196198.
(36) Heifetz, A.; Huang, K.; Sahakian, A. V.; Li, X.; Taflove, A.;
Backman, V. Appl. Phys. Lett. 2006,89, 221118.
(37) Kapitonov, A. M.; Astratov, V. N. Opt. Lett. 2007,32, 409411.
(38) Yi, K. J.; Wang, H.; Lu, Y. F.; Yang, Z. Y. J. Appl. Phys. 2007,
101, 063528.
(39) Lecler, S.; Haacke, S.; Lecong, N.; Cré
gut, O.; Rehspringer, J.-L.;
Hirlimann, C. Opt. Express 2007,15, 49354942.
(40)Wu,W.;Katsnelson,A.;Memis,O.G.;Mohseni,H.
Nanotechnology 2007,18, 485302.
(41) Heifetz, A.; Simpson, J. J.; Kong, S.-C.; Taflove, A.; Backman, V.
Opt. Express 2007,15, 1733417342.
(42) Gerlach, M.; Rakovich, Y. P.; Donegan, J. F. Opt. Express 2007,
15, 1734317350.
(43) Ferrand, P.; Wenger, J.; Devilez, A.; Pianta, M.; Stout, B.;
Bonod, N.; Popov, E.; Rigneault, H. Opt. Express 2008,16, 6930
6940.
(44) Kong, S.-C.; Sahakian, A. V.; Heifetz, A.; Taflove, A.; Backman,
V. Appl. Phys. Lett. 2008,92, 211102.
(45) Kong, S.-C.; Sahakian, A.; Taflove, A.; Backman, V. Opt. Express
2008,16, 1371313719.
(46) Yang, S.; Astratov, V. N. Appl. Phys. Lett. 2008,92, 261111.
(47) McLeod, E.; Arnold, C. B. Nat. Nanotechnol. 2008,3, 413417.
(48) Cui, X.; Erni, D.; Hafner, C. Opt. Express 2008,16, 13560
13568.
(49) Devilez, A.; Stout, B.; Bonod, N.; Popov, E. Opt. Express 2008,
16, 1420014212.
Nano Letters Letter
DOI: 10.1021/acs.nanolett.6b01255
Nano Lett. 2016, 16, 48624870
4869
(50) Gé
rard, D.; Wenger, J.; Devilez, A.; Gachet, D.; Stout, B.;
Bonod, N.; Popov, E.; Rigneault, H. Opt. Express 2008,16, 15297
15303.
(51) Heifetz, A.; Kong, S.-C.; Sahakian, A. V.; Taflove, A.; Backman,
V. J. Comput. Theor. Nanosci. 2009,6, 19791992.
(52) Geints, Y. E.; Panina, E. K.; Zemlyanov, A. A. Opt. Commun.
2010,283, 47754781.
(53) Geints, Y. E.; Zemlyanov, A. A.; Panina, E. K. Quantum Electron.
2011,41, 520525.
(54) Liu, Y.; Kuang, C. F.; Ding, Z. H. Opt. Commun. 2011,284,
48244827.
(55) Kuang, C.; Liu, Y.; Hao, X.; Luo, D.; Liu, X. Opt. Commun.
2012,285, 402406.
(56) Kim, M.-K.; Scharf, T.; Mühlig, S.; Rockstuhl, C.; Herzig, H. P.
Opt. Express 2011,19, 1020610220.
(57) Shaw, P. J.; Rawlins, D. J. J. Microsc. 1991,163, 151165.
(58) Nasse, M. J.; Woehl, J. C. J. Opt. Soc. Am. A 2010,27, 295302.
(59) Novotny, L.; Hecht, B. Principles of Nano-Optics, 2nd ed.;
Cambridge University Press: New York, 2012.
(60) Haeberlé
, O.; Ammar, M.; Furukawa, H.; Tenjimbayashi, K.;
Török, P. Opt. Express 2003,11, 29642969.
(61) Yang, I. T. Methods Cell Biol. 1989,30,145.
(62) Knutsson, H.; Westin, C.-F. Proc. CVPR 1993, 515523.
(63) Zhang, X.; Kashti, T.; Kella, D.; Frank, T.; Shaked, D.; Ulichney,
R.; Fischer, M.; Allebach, J. P. Proc. SPIE 2012,8293, 829307.
(64) Venton, B. J.; Troyer, K. P.; Wightman, R. M. Anal. Chem. 2002,
74, 539546.
(65) Passarelli, M. K.; Wang, J.; Mohammadi, A. S.; Trouillon, R.;
Gilmore, I.; Ewing, A. G. Anal. Chem. 2014,86, 94739480.
(66) Kirshner, H.; Aguet, F.; Sage, D.; Unser, M. J. Microsc. 2013,
249,1325.
(67) Betzig, E.; Patterson, G. H.; Sougrat, R.; Lindwasser, O. W.;
Olenych, S.; Bonifacino, J. S.; Davidson, M. W.; Lippincott-Schwartz,
J.; Hess, H. F. Science 2006,313, 16421645.
(68) Rust, M. J.; Bates, M.; Zhuang, X. Nat. Methods 2006,3, 793
796.
(69) Hess, S. T.; Girirajan, T. P. K.; Mason, M. D. Biophys. J. 2006,
91, 42584272.
(70) Lacoste, T. D.; Michalet, X.; Pinaud, F.; Chemla, D. S.;
Alivisatos, A. P.; Weiss, S. Proc. Natl. Acad. Sci. U. S. A. 2000,97,
94619466.
(71) Small, A.; Stahlheber, S. Nat. Methods 2014,11, 267279.
(72) Zhang, B.; Zerubia, J.; Olivo-Marin, J. C. Appl. Opt. 2007,46,
181929.
(73) Price, R. L.; Jerome, W. G. Basic Confocal Microcopy; Springer:
New York, 2011.
Nano Letters Letter
DOI: 10.1021/acs.nanolett.6b01255
Nano Lett. 2016, 16, 48624870
4870
... In [1] it was shown that by examining objects with the help of dielectric particles, it is possible to resolve structures beyond the diffraction limit, i.e. whose size is significantly less than half the wavelength. Various theories have been proposed to explain this phenomenon [2][3][4][5][6][7]. However, no clear explanation of the nature of this phenomenon has been presented. ...
... The first step requires constructing a source field. We used data on the spectrum of this source from work [2], see Fig. 2, since in the original work only the peak wavelength was indicated. Note that further results do not change if we consider a uniform spectrum from 425 nm to 675 nm. ...
... Probability density distribution for source frequencies used in our the calculations according[2]. ...
Preprint
Full-text available
The approach that makes it possible to explain the phenomenon of super resolution in a virtual image using dielectric microparticles is presented. A resolution of 100 nm in the visible range is demonstrated, resolutions of the order of 50 nm are being discussed. The presented two-dimensional model uses the FTDT code to analyze the generation of radiation and propagation through a system of slits in a metal screen and a dielectric microparticle located above it. The image is constructed using the back propagation method.
... 5 The dielectric particle with a circular cross-section focuses light nearby to create PNJs or confine light inside the cavity to create whispering gallery modes [11][12][13] (WGMs). PNJs show several possible potential applications: single nanoscale-particle detection 14,15 and manipulation, [16][17][18] microsphere assisted-nanoscopy, 19 superresolution imaging [20][21][22] and applications in bio-sensing. 23,24 PNJs have been utilized to enhance the optical force in optical tweezers, 25,26 enhance Raman microscopy signals, 27,28 super-resolution nanopatterning 29 with tuned optical, 8,30 mechanical and electrical properties. ...
... However, further improvements are necessary to achieve a smaller spot size of under 1 μm [27][28][29] Recently, microsphere-assisted super-resolution imaging has emerged as a technique that enhances the magnification of optical system while overcoming the diffraction limit [30][31][32][33][34] . The photonic nanojet effect is one of the most representative theories for super-resolution enhancement by the microspheres, although its principles have not been fully elucidated [35][36][37][38] . The photonic nanojet refers to an energy concentration point on the shadow side of the microsphere, which is known to convert evanescent waves into propagating waves, resulting in superresolution images within the virtual imaging plane. ...
Article
Full-text available
As semiconductor devices shrink and their manufacturing processes advance, accurately measuring in-cell critical dimensions (CD) becomes increasingly crucial. Traditional test element group (TEG) measurements are becoming inadequate for representing the fine, repetitive patterns in cell blocks. Conventional non-destructive metrology technologies like optical critical dimension (OCD) are limited due to their large spot diameter of approximately 25 μm, which impedes their efficacy for detailed in-cell structural analysis. Consequently, there is a pressing need for small-spot and non-destructive metrology methods. To address this limitation, we demonstrate a microsphere-assisted hyperspectral imaging (MAHSI) system, specifically designed for small spot optical metrology with super-resolution. Utilizing microsphere-assisted super-resolution imaging, this system achieves an optical resolution of 66 nm within a field of view of 5.6 μm × 5.6 μm. This approach effectively breaks the diffraction limit, significantly enhancing the magnification of the system. The MAHSI system incorporating hyperspectral imaging with a wavelength range of 400–790 nm, enables the capture of the reflection spectrum at each camera pixel. The achieved pixel resolution, which is equivalent to the measuring spot size, is 14.4 nm/pixel and the magnification is 450X. The MAHSI system enables measurement of local uniformity in critical areas like corners and edges of DRAM cell blocks, areas previously challenging to inspect with conventional OCD methods. To our knowledge, this approach represents the first global implementation of microsphere-assisted hyperspectral imaging to address the metrology challenges in complex 3D structures of semiconductor devices.
... A higher-resolution image from a conventional fluorescence microscope can be achieved by using an objective lens with higher magnification (typically also with a higher numerical aperture), but it is typically at the cost of a reduced field of view. One effective approach for increasing the resolution while maintaining a large field of view involves the use of scanning dielectric microspheres [4][5][6][7][8][9][10][11][12][13] . When the dielectric microsphere has a refractive index higher than that of the outer medium, the propagated light is focused from the inside of the microsphere, and a highly localized electromagnetic beam is generated near its surface, which is known as a photonic nanojet, and it provides superresolution imaging below the diffraction limit. ...
Article
Full-text available
Large-field nanoscale fluorescence imaging is invaluable for many applications, such as imaging subcellular structures, visualizing protein interactions, and high-resolution tissue imaging. Unfortunately, conventional fluorescence microscopy requires a trade-off between resolution and field of view due to the nature of the optics used to form the image. To overcome this barrier, we developed an acoustofluidic scanning fluorescence nanoscope that simultaneously achieves superior resolution, a large field of view, and strong fluorescent signals. The acoustofluidic scanning fluorescence nanoscope utilizes the superresolution capabilities of microspheres that are controlled by a programmable acoustofluidic device for rapid fluorescence enhancement and imaging. The acoustofluidic scanning fluorescence nanoscope resolves structures that cannot be resolved with conventional fluorescence microscopes with the same objective lens and enhances the fluorescent signal by a factor of ~5 without altering the field of view of the image. The improved resolution realized with enhanced fluorescent signals and the large field of view achieved via acoustofluidic scanning fluorescence nanoscopy provides a powerful tool for versatile nanoscale fluorescence imaging for researchers in the fields of medicine, biology, biophysics, and biomedical engineering.
Article
In this paper, we numerically investigate a photonic hook (PH) generated by a dielectric micro-cylinder illuminated under an electric dipole using the finite element method. At first, we show the properties of the photonic nanojet (PNJ) produced by the micro-cylinder under an electric dipole. Then, we find that a PH can be obtained when the electric dipole is moved at a certain distance along the direction perpendicular to the center line of the cylinder. Considering the rotation symmetry of the model, we firstly present that a PH can be generated by changing the orientation of the electric dipole. And the power flow of the PHs produced by different orientated dipoles are studied. Finally, the PHs properties affected by the gap distance between the micro-cylinder and the electric dipole and the refractive index of the micro-cylinder are investigated. Our results provide a method to generate a PH under a dipole illumination. And it may have potential applications in the quantum dot detection and microsphere super-resolution.
Article
Wide-field mid-infrared (MIR) hyperspectral imaging offers a promising approach for studying heterogeneous chemical systems due to its ability to independently characterize the molecular properties of different regions of a sample. However, applications of wide-field MIR microscopy are limited to spatial resolutions no better than ∼1 μm. While methods exist to overcome the classical diffraction limit of ∼λ/2, chromatic aberration from transmissive imaging reduces the achievable resolution. Here we describe the design and implementation of a simple MIR achromatic lens combination that we believe will aid in the development of resolution-enhanced wide-field MIR hyperspectral optical and chemical absorption imaging. We also examine the use of this doublet lens to image through polystyrene microspheres, an emerging and simple means for enhancing spatial resolution.
Conference Paper
The integration of microspheres within the instruments of optical metrology and mask-less lithography could already show a significant enhancement of their lateral resolution. Exposing complex large structures exploiting this high resolution requires the lateral movement of the microsphere over the substrate. Challenging remains the accurate lateral and axial positioning of the microsphere ensuring the constant exposure conditions at every point. Preserving the advantage of optical instruments to not actually contact the specimen, the microsphere must be kept at a nanometer-close, yet constant distance from the surface. Here, we introduce the, to our best knowledge, novel approach to combine the principle of the differential confocal microscope with a scanning microsphere. This produces a differential signal towards the surface allowing a nanometer-sensitive and fast control of the axial position of the microsphere above the substrate. In preliminary experiments we show the repeatable pick-up of microspheres and their precise lateral scanning using a nanopositioning and nanomeasuring machine as well as axial depth responses and differential signals from the realized microsphere assisted differential confocal probe.
Article
Full-text available
Optical super-resolution imaging is a desirable technique in many fields, including medical and material sciences and nanophotonics. We demonstrated feasibility of optical microscopy, with resolution improvement by a factor of 2-3, by using microsphere-embedded coverslips composed of barium titanate glass microspheres (refractive index n~1.9-2.1) fixed in a transparent layer of polydimethylsiloxane. Imaging was performed by a conventional microscope and a fluorescent microscope with the microsphere-embedded layer placed in a contact position with various biological specimens and semiconductor nanostructures. We investigated a biomedical application of microsphere-assisted imaging technique by immunostaining of kidney sections where cellular distribution of a motor protein Myo1c and a podocyte specific protein ZO-1 was analyzed. A significant visual enhancement in the distribution pattern of the proteins was noted in the stained glomeruli by using microsphere-assisted imaging technique. Our results suggest that microsphere-assisted imaging technique is a promising candidate for applications in medical and cancer research, as well as in microfluidics and nanophotonics.
Article
Full-text available
We comprehensively study the least-squares Gaussian approximations of the diffraction-limited 2D-3D paraxial-nonparaxial point-spread functions (PSFs) of the wide field fluorescence microscope (WFFM), the laser scanning confocal microscope (LSCM), and the disk scanning confocal microscope (DSCM). The PSFs are expressed using the Debye integral. Under an L(infinity) constraint imposing peak matching, optimal and near-optimal Gaussian parameters are derived for the PSFs. With an L1 constraint imposing energy conservation, an optimal Gaussian parameter is derived for the 2D paraxial WFFM PSF. We found that (1) the 2D approximations are all very accurate; (2) no accurate Gaussian approximation exists for 3D WFFM PSFs; and (3) with typical pinhole sizes, the 3D approximations are accurate for the DSCM and nearly perfect for the LSCM. All the Gaussian parameters derived in this study are in explicit analytical form, allowing their direct use in practical applications.
Article
Full-text available
We demonstrate a series of advantages of microsphere-assisted imaging over confocal and solid immersion lens microscopies including intrinsic flexibility, better resolution, higher magnification, and longer working distances. We discerned minimal feature sizes of ̃50-60 nm in nanoplasmonic arrays at the illumination wavelength λ = 405 nm. It is demonstrated that liquid-immersed, high-index (n ̃ 1.9-2.1) spheres provide a superior image quality compared to that obtained by spheres with the same index contrast in an air environment. We estimate that using transparent microspheres at deep UV wavelengths of ̃200 nm might make possible imaging of various nanostructures with extraordinary high ̃30 nm resolution.
Book
First published in 2006, this book has become the standard reference on nano-optics. Now in its second edition, the text has been thoroughly updated to take into account new developments and research directions. While the overall structure and pedagogical style of the book remain unchanged, all existing chapters have been expanded and a new chapter has been added. Adopting a broad perspective, the authors provide a detailed overview of the theoretical and experimental concepts that are needed to understand and work in nano-optics, across subfields ranging from quantum optics to biophysics. New topics of discussion include: optical antennas; new imaging techniques; Fano interference and strong coupling; reciprocity; metamaterials; and cavity optomechanics. With numerous end-of-chapter problem sets and illustrative material to expand on ideas discussed in the main text, this is an ideal textbook for graduate students entering the field. It is also a valuable reference for researchers and course teachers.
Book
Basic Confocal Microscopy, Second Edition builds on the successful first edition by keeping the same format and reflecting relevant changes and recent developments in this still-burgeoning field. This format is based on the Confocal Microscopy Workshop that has been taught by several of the authors for nearly 20 years and remains a popular workshop for gaining basic skills in confocal microscopy. While much of the information concerning fluorescence and confocal microscopy that made the first edition a success has not changed in the six years since the book was first published, confocal imaging is an evolving field and recent advances in detector technology, operating software, tissue preparation and clearing, image analysis, and more have been updated to reflect this. Several of these advances are now considered routine in many laboratories, and others such as super resolution techniques built on confocal technology are becoming widely available.
Book
Most researchers agree that biological confocal microscopy was jump-started by the confocal design first published by White and Amos in 1985 in the Journal of Cell Biology. As a result, this remains a relatively young field. Yet the use of the technique has grown phenomenally since those early efforts, with new users joining the ranks daily. The publication of Basic Confocal Microscopy reflects the burgeoning need to train new students, technologists, and faculty wishing to use confocal microscopy in their research. A direct outgrowth of the authors’ five-day intensive course in the subject begun in 2005, this book covers the basics and includes all the information required to design, implement, and interpret the results of, biological experiments based on confocal microscopy. Concise yet comprehensive, the volume begins by covering the core issues of fluorescence, specimen preparation and labeling, before moving on to address the analog-to-digital conversion of specimen data gathered using confocal microscopy. Subsequent chapters detail the practicalities of operating confocal microscopes, providing all the information necessary to begin practicing confocal microscopy as well as optimizing the material obtained. The final block of chapters examine 3-dimensional analysis and the reconstruction of data sets, outline some of the ethical considerations in confocal imaging, and then supply a number of resources that the authors have found useful in their own work. Once readers have mastered the information this book presents, the resources found in its pages will be an excellent guide to continued learning about the more advanced forms of confocal microscopy.
Article
An organic lateral resolution test device has been developed to measure the performance of imaging mass spectrometry (IMS) systems. The device contains periodic gratings of polyethylene glycol (PEG) and lipid bars covering a wide range of spatial frequencies. Microfabrication technologies were employed to produce well-defined chemical interfaces, which allow lateral resolution to be assessed using the edge-spread function (ESF). In addition, the design of the device allows for the direct measurement of the modulation transfer function (MTF) to assess image quality. Scanning electron microscopy (SEM) and time-of-flight secondary ion mass spectrometry (ToF-SIMS) were used to characterize the device. ToF-SIMS imaging was used to measure the chemical displacement of biomolecules in MALDI matrix crystals. In a proof-of-concept experiment, the platform was also used to evaluate MALDI matrix application methods, specifically aerosol spray and sublimation methods.
Article
Nano-structured optical components, such as nanolenses, direct light at subwavelength scales to enable, among others, high-resolution lithography, miniaturization of photonic circuits, and nanoscopic imaging of biostructures. A major challenge in fabricating nanolenses is the appropriate positioning of the lens with respect to the sample while simultaneously ensuring it adopts the optimal size and shape for the intended use. One application of particular interest is the enhancement of contrast and signal-to-noise ratio (SNR) in the imaging of nanoscale objects, especially over wide fields-of-view (FOVs), which typically come with limited resolution and sensitivity for imaging nano-objects. Here we present a self-assembly method for fabricating time- and temperature-tunable nanolenses based on the condensation of a polymeric liquid around a nanoparticle, which we apply to the high-throughput on-chip detection of spheroids smaller than 40 nm, rod-shaped particles with diameter smaller than 20 nm, and bio-functionalized nanoparticles, all across an ultra-large FOV of > 20 mm^2. Previous nanoparticle imaging efforts across similar FOVs have detected spheroids no smaller than 100 nm, and therefore our results demonstrate the detection of particles >15 fold smaller in volume, which in free space have >240 times weaker Rayleigh scattering compared to the particle sizes detected in earlier wide-field imaging work. This entire platform, with its tunable nanolens condensation and wide-field imaging functions, is also miniaturized into a cost-effective and portable device, which might be especially important for field use, mobile sensing, and diagnostics applications, including e.g., the measurement of viral load in bodily fluids.
Article
The microsphere optical nanoscopy (MONS) technique recently demonstrated the capability to break the optical diffraction limit with a microsphere size of 2-9 μm fused silica. We report that larger polystyrene microspheres of 30, 50 and 100 μm diameters can overcome the diffraction limit in optical imaging. The sub-diffraction features of a Blu-ray Disc and gold nano-patterned quartz were experimentally observed in air by coupling the microspheres with a standard optical microscope in the reflected light illumination mode. About six to eight times magnification was achieved using the MONS. The mechanism of the MONS was theoretically explained by considering the transformation of near-field evanescent waves into far-field propagating waves. The super-resolution imaging was demonstrated by experiments and theoretical simulations.