ArticlePDF Available

Abstract and Figures

We analyze the effects of partial coherence in the image formation of a digital in-line holographic microscope (DIHM). The impulse response is described as a function of cross-spectral density of the light used in the space-frequency domain. Numerical simulation based on the applied model shows that a reduction in coherence of light leads to broadening of the impulse response. This is also validated by results from experiments wherein a DIHM is used to image latex beads using light with different spatial and temporal coherence.
Content may be subject to copyright.
Coherence effects in digital in-line holographic
microscopy
Unnikrishnan Gopinathan,*Giancarlo Pedrini, and Wolfgang Osten
Institut für Technische Optik, Universitäat Stuttgart, Pfaffenwaldering 9, 70569 Stuttgart, Germany
*
Corresponding author: unni.gopinathan@gmail.com
Received June 12, 2008; accepted July 21, 2008;
posted August 12, 2008 (Doc. ID 97336); published September 17, 2008
We analyze the effects of partial coherence in the image formation of a digital in-line holographic microscope
(DIHM). The impulse response is described as a function of cross-spectral density of the light used in the space-
frequency domain. Numerical simulation based on the applied model shows that a reduction in coherence of
light leads to broadening of the impulse response. This is also validated by results from experiments wherein
a DIHM is used to image latex beads using light with different spatial and temporal coherence. © 2008 Op-
tical Society of America
OCIS codes: 030.1640, 030.1670, 030.4070, 090.1995, 170.0180.
1. INTRODUCTION
In-line holography with spherical waves using a lensless
configuration was originally proposed by Gabor [1] while
attempting to address the problem of lens aberrations in
electron microscopy. To address the practical issues asso-
ciated with the hologram reconstruction in the original
method, one of the approaches was to record the hologram
digitally and use numerical techniques for reconstruction.
These methods have been generically referred to as digi-
tal in-line holography (DIH) [2]. Many novel recording
and reconstruction methods have been proposed to ex-
tract information regarding three-dimensional complex
amplitude distribution [35] uncorrupted by twin image
and three-dimensional complex spatial coherence distri-
bution [6].
Over the years, DIH has been successfully applied to
imaging microscopic objects with electrons [7,8] and pho-
tons [914]. Most of these works use configurations with-
out a lens. In addition to eliminating aberrations, the
lensless configuration is well suited for a wide range of
sources including x rays and deep ultraviolet for improved
resolution [1517].
In many applications using DIH, it is common to treat
the source as completely coherent, both spatially and tem-
porally. This assumption, though valid in many cases, will
not hold for all. For instance, x-ray sources are partially
coherent. Partially coherent light is also used for many
applications in optical microscopy [1820]. These applica-
tions would benefit from a generalized treatment that
takes into account the state of coherence of light. Effects
of spatial coherence in lensless Fresnel holography
[21,22] and in-line Gabor holography [23] have been stud-
ied earlier. However these works have mainly dealt with
the spatial coherence effects for the case of quasi-
monochromatic light.
In this paper we provide a theoretical framework for
analyzing the partial coherence effects in the image for-
mation of a digital in-line holographic microscope
(DIHM). The analysis presented is valid for primary or
secondary sources of any state of coherence. The state of
coherence of the source is characterized by the cross-
spectral density (CSD) [2426] in the space-frequency do-
main. The analysis in the space-frequency domain allows
us to account for any dispersive effect of the object. We
provide an expression for the impulse response in terms
of CSD of the source. The impulse response for the case of
a Gaussian Schell source is simulated. It is seen that a re-
duction in coherence of the light leads to broadening of
the impulse response. This is also validated by results
from experiments wherein a DIHM is used to image latex
beads using light with different spatial and temporal co-
herence.
2. IMPULSE RESPONSE FOR PARTIALLY
COHERENT LIGHT
A. Hologram Formation
The geometry of the system discussed is shown in Fig. 1.
Let
denote the illumination pupil of a planar secondary
source illuminated by a primary source directly or indi-
rectly (through an optical system) with a coherence state
characterized by the CSD function Wp1,p2,
of the
fluctuating field Up1,tand Up2,tat two points P1and
P2in the secondary source illumination pupil
;p1and p2
are the position vectors of two points P1and P2.Ifthe
fluctuations Up1,tand Up2,tare members of the sta-
tistical ensemble Up,t=Up,
ej
tthat is stationary
at least in the wide sense, Wolf [26] has shown that
Wp1,p2,
may be expressed as a correlation function in
the space-frequency domain as
Wp1,p2,
=U*p1,
Up2,
兲典
.1
To find out the impulse response of the system, we con-
sider the interference between the light emanating from
the source and the light scattered from a point object lo-
cated at the point Qwith a position vector ri. The scat-
Gopinathan et al. Vol. 25, No. 10/ October 2008/J. Opt. Soc. Am. A 2459
1084-7529/08/102459-8/$15.00 © 2008 Optical Society of America
tered wave from the object may be considered to be a
spherical wave emanating from the point Qwith an am-
plitude ri,
. The scattered and unscattered compo-
nent of the light from the point P1reaches the point Pfol-
lowing two paths P1QP and P1Pgiven by
V1r,
=ri,
Up1,
ejkrip1
rip1
ejkrri
rri+Up1,
ejkrp1
rp1.
2
If the distance of the scatterer rifrom the source is quite
large compared to the source dimension, the term ri
p1may be approximated as risi·p1in the numerator
and as riin the denominator; siis the unit vector along
the vector ri. In the numerator, rip1appears in the ex-
ponential term multiplied by the factor k. Hence, the ap-
proximation to rip1used is different in the numerator
and the denominator. Similarly, if the point Plies suffi-
ciently far away from the source, the terms rriand r
p1may be approximated as rs·riand rs·p1in the
numerator and as rin the denominator, respectively; sis
the unit vector along the vector r. Using the far field ap-
proximations given above we may write Eq. (2) as
V1rs,
=ri,
Up1,
ejkri
ri
ejkr
rejksi·p1+s·ri
+Up1,
ejkr
rejks·p1.3
A similar expression may be written for the light from the
point P2reaching the point Pfollowing two paths P2QP
and P2Pgiven by
V2rs,
=ri,
Up2,
ejkri
ri
ejkr
rejksi·p2+s·ri
+Up2,
ejkr
rejks·p2.4
The spectral density at the point Pis given by
Srs,
=
V1
*rs,
V2rs,
兲典d2p1d2p2.5
The integration in Eq. (5) is over the source domain
.All
integrals in the following discussion unless otherwise
stated run from to +. Substituting Eqs. (3)(5) we ob-
tain
Ss,
=
Wp1,p2,
ri,
兲兩2
ri
2ejksi·p2p1
+ejks·p2p1+ejkri
ri
ri,
ejksi·p2+s·ris·p1
+ejkri
ri
*ri,
ejksi·p1+s·ris·p2
d2p1d2p2.6
If we interchange p1and p2in the fourth term and use
the relation Wp2,p1,
=W*p1,p2,
(valid for any CSD
function [24,25]), it may be noted that the fourth term in
Eq. (6) is the complex conjugate of the third term. It may
be seen that S.is a function of only sand not rsince Pis
a point in the far field. The first term in Eq. (6) represents
the light reaching Pafter scattering at Q, whereas the
second term represents light reaching the point Pdirectly
from the source without any scattering. The last two
terms represent the interference of the scattered light
with the unscattered light.
The CSD of a statistically stationary source of any state
of coherence may be expanded as a linear combination of
orthonormal functions
np,
[26]:
Wp1,p2,
=
n
n
n
*p1,
np2,
,7
where the functions
np,
are the eigenfunctions and
n
are the eigenvalues of the homogeneous Fredholm
integral equation
Wp1,p2,
np1,
d2p1=
n
np2,
.8
In other words, Eq. (7) states that CSD may be expressed
as the sum of contributions from spatially completely co-
herent elementary sources if it is a continuous function
and bounded throughout the source volume [26]. Substi-
tuting Eq. (7) into Eq. (6) we obtain an expression for the
spectral density at a point Pin the far field,
Ss,
=ri,
兲兩2
ri
2
n
n
兲兩
˜
nksi,
兲兩2
+
n
n
兲兩
˜
nks,
兲兩2
+ejkri
ri
ri,
ejks·ri
n
n
˜
nksi,
˜
n
*ks,
+ejkri
ri
*ri,
ejks·ri
n
n
*
˜
n
*ksi,
˜
nks,
,
9
where sand siare the projections of the vectors sand
Fig. 1. Schematic illustrating the notations used in the text. SF
denotes spatial filter (see Fig. 4).
2460 J. Opt. Soc. Am. A/ Vol. 25, No. 10/ October 2008 Gopinathan et al.
sionto the z=0 plane and
˜
n.is the spatial Fourier
transform of
n.defined as
˜
n.,
=
np,
ejp·.d2p.10
Taking a temporal Fourier transform of the spectral
density gives us the intensity at the point Precorded as a
hologram denoted by the function H.and given by
Hs,.=
Ss,
ej
.d
.11
The first term in the expression for the hologram, ob-
tained by substituting Eq. (9) into Eq. (11), is proportional
to the intensity of the scattered light from the particle,
and the second term is proportional to the reference beam
without contribution from the object. Often the reference
beam is recorded separately and subtracted from the ho-
logram [7,9] to minimize the effects of the zero-order
term. The third and the fourth terms result in the object
image and a twin image of the object during reconstruc-
tion.
B. Numerical Reconstruction
The image of the object is formed by reconstructing the
hologram numerically using the Helmholtz–Kirchhoff in-
tegral [7,9]. The reconstructed intensity at the point rcin
the reconstructed volume is given by
Urc,
=
Hs,
ejks·rcd2s.12
The integral in the above equation runs over the area of
the detector plane. Hereafter, we focus our attention on
the contribution of the third term in the expression for
the hologram given by Eq. (11) [in conjunction with Eq.
(9)]. This term results in the formation of an image of a
point object located at ridenoted as Uiand obtained by
substituting the expression for the hologram in Eq. (12):
Uirc,
=
ri,
ejkri
ri
n
n
˜
nksi,
˜
n*ks,
ejks·rcrid2sej
d
.13
Equation (13) describes the image formation in a DIHM
with light of any state of coherence. It may be seen that
the image formation for partially coherent light may be
described in terms of the sum of contributions from com-
pletely spatially coherent modes,
Uirc,
=
n
Ui
nrc,
,14
where
Ui
nrc,
=
ri,
ejkri
ri
n
˜
nksi,
˜
n*ks,
ejks·rcrid2sej
d
15
is the contribution of the nth spatially coherent mode to
the image formed. If the object is considered to be consti-
tuted of a collection of point objects in a volume the im-
age of such an object may be obtained by integrating Eq.
(13) over all the object points in the volume :
Urc,
=
ri,
ejkri
ri
n
n
˜
nksi,
˜
n*ks,
ejks·rcrid2sej
d
d3ri.
16
We may write Eq. (16) as
Urc,
=
ri,
ejkri
ri
hircri,
ej
d
d3ri,
17
where hi.,
is the impulse response of the system given
by
hi.,
=
n
n
˜
nksi,
˜
n*ks,
ejks·.d2s.
18
It may be seen that the impulse response is a function of
the position of the object through the term
n
˜
ksi,
,
which makes the system shift variant for partially coher-
ent light. We now proceed to discuss a few special cases.
C. Spatially Coherent Light
When the light is completely spatially coherent then the
CSD may be represented by the lowest order mode n
=0. For this case, the impulse response is given by
hi
coh.,
=
0
˜
0ksi,
˜
0*ks,
ejks·.d2s.
19
D. Narrowband Light
For narrowband light the bandwidth
is very small
compared to center frequency
0. If the assumption that
the modulus and the phase of the CSD function W.is
constant over
is valid, then the CSD function may be
approximated as T
Wp1,p2,
0,T
being the spec-
tral density of light. For this case the impulse response
may be written as
hi
nb.,
=T
n
n
0
˜
nk0si,
0
˜
n*k0s,
0ejk0s·.d2s.20
Gopinathan et al. Vol. 25, No. 10/ October 2008/J. Opt. Soc. Am. A 2461
E. Gaussian Schell Sources
A Schell model source is characterized by CSD of the
form [25]
Wp1,p2,
=Sp1,
Sp2,
p1p2,
,21
where Sp,
is the spectral intensity at pand
p1
p2,
is the complex degree of spatial coherence. For the
Gaussian Schell model sources the spectral intensity dis-
tribution and degree of coherence are Gaussian functions
given by
Sp,
=A
ep2/2
s
2
,22
p1p2,
=ep1p22/2
2
.23
For the Gaussian Schell model sources, the eigenfunc-
tions and eigenvalues,
np,
and n
, that solve the
Fredholm integral equation [Eq. (8)] are given by [27]
nx=
2c
1/4 1
2nn!Hnx2cecx2,24
n=A
a+b+c
冊冉
b
a+b+c
n
,25
where Hn.is a hermite polynomial of order n,c
=a2+2ab,a=1/4
s
2
, and b=1/2
2
. In the above
equations the dependence of a,b, and con
is dropped
for convenience. The parameter
[27], defined as the ra-
tio of
to
s, is the measure of degree of global coherence
[27] of the source. When
1, the source may be said to
be globally coherent and is well approximated by the low-
est order mode. When
1 the source may be considered
to be globally incoherent, and a larger number of modes of
the order of 1/
are necessary to represent the source
properly. The Fourier transform of
nxin Eq. (24) is
given by
˜
n
=
2c
1/4 1
2nn!H
˜
n
2c
e
2/4c.26
Substituting Eq. (26) into Eq. (20) we get the impulse re-
sponse when a Gaussian Schell source with narrow tem-
poral bandwidth centered at frequency
0is used,
hi
Schell.,
=T
n
2c
1/2
n
0
2nn!H
˜
n*
k0
2csi
H
˜
n
k0
2cs
ek0
2共兩s2+s02/4cejk0s·.d2s.
27
One could express the impulse response in the time
Fig. 2. Impulse response evaluated at rx=ry=0 for rzranging from −0.1 to 0.1 mm for a Gaussian Schell source with (a)
=5,
=0.2 and
temporal FWHM bandwidth equal to 7 nm, (b)
=5,
=0.2 and temporal FWHM bandwidth equal to 14 nm, (c) temporal FWHM band-
width equal to 7 and 14 nm with
=5, and (d) temporal FWHM bandwidth equal to 7 and 14 nm with
=0.2. The Xaxis in all four plots
indicates distance rz in millimeters. The Yaxis shows normalized amplitude.
2462 J. Opt. Soc. Am. A/ Vol. 25, No. 10/ October 2008 Gopinathan et al.
domain by taking the temporal Fourier transform of
Eq. (27).
hi
Schell.,
=
n
2c
1/2
n
0
2nn!H
˜
n*
k0
2csi
H
˜
n
k0
2cs
ek0
2共兩s2+s02/4cejk0s·.d2s,
28
where
=
T
ej
d
.29
We simulate the impulse response of a DIHM as given in
Eq. (28) (for si=0) when a narrowband Gaussian Schell
model source is used. The distance Dbetween the source
and CCD was assumed to be 15 mm. The CCD was as-
sumed to have a square pixel of width 6.7
m. The tem-
poral frequency distribution of the light T
was as-
sumed to have a Gaussian distribution with a center
wavelength of 670 nm. The temporal coherence of the
light is varied by varying the temporal bandwidth mea-
sured as full bandwidth at half wavelength (FBHW) of
the Gaussian distribution. Figure 2shows the plots of the
absolute value of the impulse response hrwith the x
and ythe components of r,rx, and ryequal to zero and rz
ranging from −0.1 to 0.1 mm in steps of 5
m. In the
plots shown in Figs. 2(a) and 2(b) the value of
is
changed keeping the temporal coherence constant. As the
value of
decreases from 5 to 0.2 (decreasing spatial co-
herence), the impulse response broadens in the Zdirec-
tion. For
=5, the lowest order mode, n=0, was used to
calculate the impulse response. For the other values of
,
the first 21 modes (n=0 to 20) were used. In the plots
shown in Figs. 2(c) and 2(d) the value of
is kept con-
stant, and the temporal bandwidth of the light is
changed. It may be observed that an increase in temporal
bandwidth of the light (decreasing temporal coherence)
results in the broadening of the impulse response. Fig-
ures 3(a)3(d) show the corresponding plots of the abso-
lute value of impulse response hrwhen rz=0. Again one
observes a broadening of the impulse response in the XY
plane as the spatial and temporal coherence of light de-
creases.
3. EXPERIMENTS
A. System Description
The schematic of the setup used to study the effects of
temporal and spatial coherence in a lensless digital in-
line holographic is shown in Fig. 4. Light from a primary
source illuminates a spatial filter (SF) that serves as a
secondary source. The objects used in this paper are
spherical latex beads of 6
m in diameter mounted onto a
Fig. 3. (Color online) Impulse response evaluated in the plane rz= 0 for a Gaussian Schell source with (a)
=5,
=0.2 and temporal
FWHM bandwidth equal to 7 nm, (b)
=5,
=0.2 and temporal FWHM bandwidth equal to 14 nm, (c) temporal FWHM bandwidth equal
to 7 and 14 nm with
=5, and (d) temporal FWHM bandwidth equal to 7 and 14 nm with
=0.2. The Xand Yaxes in all four plots
indicate rxand ryin pixels. Each pixel translates to a physical distance 3
m. The Zaxis shows normalized amplitude.
Gopinathan et al. Vol. 25, No. 10/ October 2008/J. Opt. Soc. Am. A 2463
microscopic slide. The light emanating from the pinhole is
scattered by the object. The object beam interferes with
the unscattered reference beam, and the hologram is re-
corded by a CCD (1000 pixels 1000 pixels, pixel size
6.7
m6.7
m) placed at a distance 15 mm from the
source. From each object hologram the zero-order term is
removed by subtracting the reference wave intensity (re-
corded separately). The holograms are then used for the
numerical reconstruction performed using Eq. (12) over
volume x⫻⌬y⫻⌬zcentered on the position of the object.
The reconstruction space coordinates are given by r
=rcr0and has an origin at rc=r0. The resolution in the
reconstructed volume is D/Nx
dx in the Xdirection and
D/Ny
dy in the Ydirection, where Dis the distance be-
tween the source and the CCD,
dx and
dy are the pixel
pitches of the CCD in the Xand the Ydirections, and Nx
and Nyare the number of hologram pixels in the Xand
the Ydirections.
B. Results and Discussion
Two light sources with different temporal coherence were
used in the experiment. (1) LD1, a laser diode with center
wavelength 670 nm and temporal full width at half-
maximum (FWHM) bandwidth of 2 nm. (2) LD2, a laser
diode with center wavelength 635 nm and FWHM band-
width of 12 nm. The spatial coherence was varied by us-
ing spatial filters of two different diameters, 1 and 5
m.
As the diameter of the pinhole increases, the spatial co-
herence of light decreases. The light emanating from the
pinhole can be considered to be spatially coherent if the
absolute value of degree of coherence of light stays close
to unity within the pinhole. If the degree of coherence of
light drops considerably from the unity value within the
pinhole, then the light is spatially incoherent. In light of
the discussion in Section 2, partially spatially coherent
Fig. 4. Schematic of a lensless DIH setup. An SF, which acts as
a secondary source, is illuminated via an imaging system by a
primary source (S). The light scattered by the micro-object and
the unscattered light forms an in-line hologram at the CCD
plane.
Fig. 5. Experimental result showing the reconstructed amplitude at rx =ry = 0 for rz ranging from −0.1 to 0.1 mm for (a) source LD1
FWHM= 2 nmwith spatial filters of 1 and 5
m in diameter, (b) source LD2 FWHM= 12 nmwith spatial filters of 1 and 5
min
diameter, (c) source LD1 and LD2 with spatial filter of 1
m in diameter, and (d) source LD1 and LD2 with spatial filter of 5
min
diameter. The Xaxis in all the four plots indicates distance rz in millimeters. The Yaxis shows normalized amplitude.
2464 J. Opt. Soc. Am. A/ Vol. 25, No. 10/ October 2008 Gopinathan et al.
light can be described as the summation of a large num-
ber of spatially coherent modes. Shown in Figs. 5(a)5(d)
are plots of the reconstructed amplitude at rx =ry =0 for
rz varying from −0.1 to 0.1 mm at intervals of 10
min
the reconstruction space. Figure 5(a) shows plots when
two different spatial filters of 1 and 5
m in diameter are
used with light source LD1. Figure 5(b) shows the corre-
sponding plots for light source LD2. Figure 5(c) shows the
plots when two light sources LD1 and LD2 are used with
the spatial filter of 1
m in diameter. Figure 5(d) shows
the corresponding plots for the spatial filter of 5
m in di-
ameter. For a given temporal coherence of the light, as the
spatial coherence decreases, the image of the bead is
broader along the Zdirection. Shown in Figs. 6(a)6(d)
are three-dimensional plots of the reconstructed ampli-
tude in the plane rz =0 of the reconstruction space. Given
in Figs. 6(a) and 6(b) are the plots when the light source
LD1 is used with spatial filters of 1 and 5
m in diameter,
respectively. Figures 6(c) and 6(d) are the corresponding
plots for the light source LD2. It may be observed that a
decrease in the spatial and temporal coherence of light
leads to a broadening of the reconstructed image of beads.
4. CONCLUSION
We have theoretically analyzed the effects of partial co-
herence in a lensless DIH system. Our analysis is valid
for primary or secondary sources of any state of coher-
ence, though in this paper we have considered a planar
secondary source. It was found that the impulse response
of the system for light of any state of coherence is a func-
tion of the cross-spectral density of the light. The impulse
response was simulated for the case of a Gaussian Schell
model source. We have provided results from experiments
used to image a spherical latex bead using a lensless DIH
microscope using light from two different sources with
varying temporal bandwidth in conjunction with spatial
filters of two different sizes. The experimental results
were found to be in general agreement with the predic-
tions of theory and simulation.
ACKNOWLEDGMENTS
U. Gopinathan gratefully acknowledges the financial sup-
port of the Alexander Von Humboldt Foundation. This
work was also supported by the German Science Founda-
tion (DFG) grant OS. 111/19-2.
REFERENCES
1. D. Gabor, “A new microscopic principle,” Nature 161,
777–778 (1948).
2. U. Schnars, H.-J. Hartman, and W. Jüptner, Digital
Holography (Springer-Verlag, 2004).
3. J. J. Barton, “Removing multiple scattering and twin
Fig. 6. (Color online) Experimental result showing the reconstructed amplitude in the plane rz= 0 for (a) source LD1 with spatial filter
of 1
m in diameter, (b) source LD1 with spatial filter of 5
m in diameter, (c) source LD2 with spatial filter of 1
m in diameter, and (d)
source LD2 with spatial filter of 5
m in diameter. The Xand Yaxes in all four plots indicate rx and ry in pixels. Each pixel corresponds
toadistanceof3
m in the reconstruction space. The Zaxis shows normalized amplitude.
Gopinathan et al. Vol. 25, No. 10/ October 2008/J. Opt. Soc. Am. A 2465
images from holographic images,” Phys. Rev. Lett. 67,
3106–3109 (1991).
4. T. Latychevskaia and H.-W. Fink, “Solution to the twin
image problem in holography,” Phys. Rev. Lett. 98, 233901
(2007).
5. G. Situ, J. P. Ryle, U. Gopinathan, and J. T. Sheridan,
“Generalised in-line digital holographic technique based on
intensity measurements at two different planes,” Appl.
Opt. 47, 711–717 (2008).
6. M. Takeda, W. Wang, Z. Duan, and Y. Miyamoto,
“Coherence holography,” Opt. Express 13, 9629–9635
(2005).
7. J. J. Barton, “Photoelectron holography,” Phys. Rev. Lett.
61, 1356–1359 (1988).
8. H.-W. Fink, W. Stocker, and H. Schmid, “Holography with
low-energy electrons,” Phys. Rev. Lett. 65, 1204–1206
(1990).
9. J. Garcia-Sucerquia, W. Xu, S. K. Jericho, P. Klages, M. H.
Jericho, and H. J. Kreuzer, “Digital in-line holographic
microscopy,” Appl. Opt. 45, 836–850 (2006).
10. W. Xu, M. J. Jericho, I. A. Meinertzhagen, and H. J.
Kreuzer, “Digital in-line holography of microspheres,”
Appl. Opt. 41, 5367–5375 (2002).
11. W. Xu, M. J. Jericho, H. J. Kreuzer, and I. A.
Meinertzhagen, “Tracking particles in four dimensions
with in-line holographic microscopy,” Opt. Lett. 28,
164–166 (2003).
12. L. Repetto, E. Piano, and C. Pontiggia, “Lensless digital
holographic microscope with light-emitting diode
illumination,” Opt. Commun. 29, 1132–1134 (2004).
13. L. Repetto, R. Chittofrati, E. Piano, and C. Pontiggia,
“Infrared lensless holographic microscope with a vidicon
camera for inspection of metallic evaporations on silicon
wafers,” Opt. Commun. 251, 44–50 (2005).
14. B. Javidi, I. Moon, S. Yeom, and E. Carapezza, “Three-
dimensional imaging and recognition of microorganism
using single-exposure on-line (SEOL) digital holography,”
Opt. Express 13, 4492–4506 (2005).
15. S. Mayo, T. Davis, T. Gureyev, P. Miller, D. Paganin, A.
Pogany, A. Stevenson, and S. Wilkins, “X-ray phase
contrast microscopy and microtomography,” Opt. Express
11, 2289–2302 (2003).
16. D. Gao, S. W. Wilkins, D. J. Parry, T. E. Gureyev, P. R.
Miller, and E. Hansen, “X-ray ultramicroscopy using
integrated sample cells,” Opt. Express 14, 7889–7894
(2006).
17. G. Pedrini, F. Zhang, and W. Osten, “Digital holographic
microscopy in the deep 193 nmultraviolet,” Appl. Opt. 46,
7829–7835 (2007).
18. J. Pomarico, U. Schnars, H.-J. Hartman, and W. Jüptner,
“Digital recording and numerical reconstruction of
holograms: a new method for displaying light in flight,”
Appl. Opt. 34, 8095–8099 (1995).
19. G. Pedrini and H. J. Tiziani, “Short-coherence digital
microscopy by use of a lensless holographic imaging
system,” Appl. Opt. 41, 4489–4496 (2002).
20. L. Martínez-León, G. Pedrini, and W. Osten, “Applications
of short-coherence digital holography in microscopy,” Appl.
Opt. 44, 3977–3984 (2005).
21. T. Kozacki and R. Józ
´wiki, “Near field hologram
registration with partially coherent illumination,” Opt.
Commun. 237, 235–242 (2004).
22. T. Kozacki and R. Józ
´wiki, “Digital reconstruction of a
hologram recorded using partially coherent illumination,”
Opt. Commun. 252, 188–201 (2005).
23. J. Cheng and S. Han, “On x-ray in-line Gabor holography
with a partially coherent source,” Opt. Commun. 172,
17–24 (1999).
24. L. Mandel and E. Wolf, Optical Coherence and Quantum
Optics (Cambridge U. Press, 1995).
25. E. Wolf, Introduction to the Theory of Coherence and
Polarization of Light (Cambridge U. Press, 2007).
26. E. Wolf, “New theory of partial coherence in the space-
frequency domain. Part I: spectra and cross spectra of
steady-state sources,” J. Opt. Soc. Am. 72, 343–351 (1982).
27. A. Starikov and E. Wolf, “Coherent-mode representation of
Gaussian Schell-model sources and their radiation fields,”
J. Opt. Soc. Am. 72, 923–928 (1982).
2466 J. Opt. Soc. Am. A/ Vol. 25, No. 10/ October 2008 Gopinathan et al.
... Learning-based approaches have also been used for phase aberration compensation in digital holographic microscopy 58,152,205 . Again, phase aberration compensation can be formulated as a classification 58 or a regression 152,205 problem. ...
... Learning-based approaches have also been used for phase aberration compensation in digital holographic microscopy 58,152,205 . Again, phase aberration compensation can be formulated as a classification 58 or a regression 152,205 problem. In the work by Nguyen et al. 58 , the role DNN plays is to segment the reconstructed and unwrapped phase. ...
... In contrast, the regression approach proposed by Xiao et al. 205 endeavors to optimize the coefficients for constructing the phase aberration map that act as responses corresponding to the input aberrated phase image. Embedded physics DNN can be used for this problem as well. ...
Article
Full-text available
With the explosive growth of mathematical optimization and computing hardware, deep neural networks (DNN) have become tremendously powerful tools to solve many challenging problems in various fields, ranging from decision making to computational imaging and holography. In this manuscript, I focus on the prosperous interactions between DNN and holography. On the one hand, DNN has been demonstrated to be in particular proficient for holographic reconstruction and computer-generated holography almost in every aspect. On the other hand, holography is an enabling tool for the optical implementation of DNN the other way around owing to the capability of interconnection and light speed processing in parallel. The purpose of this article is to give a comprehensive literature review on the recent progress of deep holography, an emerging interdisciplinary research field that is mutually inspired by holography and DNN. I first give a brief overview of the basic theory and architectures of DNN, and then discuss some of the most important progresses of deep holography. I hope that the present unified exposition will stimulate further development in this promising and exciting field of research.
... In this way, for a fixed pinhole-sample distance the spatial coherence of the light source can be changed by the simple variation of the pinhole diameter d p . A complementary discussion on the effect of the spatial partial coherence in DLHM can be read for instance in Gopinathan et al. (2008). ...
... This latter image was obtained by means of a violet laser λ=405 nm illuminating a pinhole 0.5 µm in diameter. For panels A and B, the low spatial coherence of the light source leads to reconstructed images where the overall shape of the simple is visible, but few details of the internal structure are seen, namely, the spatial resolution of the microscope has been reduced due to the widening of the point spread function of it (Gopinathan et al., 2008). As the spatial coherence is increased over the plane of the simple, it can be seen a larger number of the internal details of the specimen and the edges of the simple are sharper, see panels C, D y E. Panels D, E and F show a comparable spatial resolution of the spatial partial coherent DLHM with that of fully coherent DLHM, however the former shows a better signal-to-noise ratio (SNR) than the fully spatial coherence DLHM. Figure 4. Effect of spatial coherence in DLHM. ...
Article
Full-text available
The principles and some applications of the digital lensless holographic microscopy (DLHM) to study the microworld are shown in this paper. The recording and reconstruction processes of the DLHM are analyzed to study its lateral resolution power. The effect of the spatial coherence in the study of a section of the head of a Drosophila melanogaster fly are presented. DLHM is applied to study dwynamic and static colloidal systems as a proof of the capability of achieving micrometer spatial lateral resolution without the use of lenses.
... In this way, for a fixed pinhole-sample distance the spatial coherence of the light source can be changed by the simple variation of the pinhole diameter d p . A complementary discussion on the effect of the spatial partial coherence in DLHM can be read for instance in Gopinathan et al. (2008). ...
... This latter image was obtained by means of a violet laser λ=405 nm illuminating a pinhole 0.5 µm in diameter. For panels A and B, the low spatial coherence of the light source leads to reconstructed images where the overall shape of the simple is visible, but few details of the internal structure are seen, namely, the spatial resolution of the microscope has been reduced due to the widening of the point spread function of it (Gopinathan et al., 2008). As the spatial coherence is increased over the plane of the simple, it can be seen a larger number of the internal details of the specimen and the edges of the simple are sharper, see panels C, D y E. Panels D, E and F show a comparable spatial resolution of the spatial partial coherent DLHM with that of fully coherent DLHM, however the former shows a better signal-to-noise ratio (SNR) than the fully spatial coherence DLHM. Figure 4. Effect of spatial coherence in DLHM. ...
Article
En este artículo se discuten los principios y algunas aplicaciones de la microscopia holográfica digital sin lentes (MHDSL) como una herramienta para el estudio del micromundo. Se analizan los procesos deregistro y reconstrucción de MHDSL para estudiar su poder de resolución lateral. Se presenta el efecto de la coherencia espacial en el estudio de una sección de la cabeza de una mosca Drosophila Melanogasterpor medio de la MHDSL. Se aplica la MHDSL al estudio de sistemas coloidales dinámicos y estáticos como una prueba de la capacidad de resolución espacial micrométrica sin la necesidad del uso de lentes.
... Despite the numerous advantages of DH, its susceptibility to noise remains a substantial challenge, hindering its application across various research fields. To mitigate this issue, researchers have explored the use of partially coherent light sources like LEDs [10][11][12][13]. These sources, characterized by a broad full-width at half maximum (FWHM), effectively suppress speckle noise. ...
... For many applications, the ultrashort pulse duration and high peak brightness of these beams requires single-shot characterizations of the radiation properties. First, the knowledge of the coherence is crucial to understand the physics behind the generation process [1,4] and can improve data analysis [5][6][7][8][9][10][11]. Furthermore, diffraction is based on the spatial coherence [12]. ...
Article
Full-text available
Spatial coherence is an impactful source parameter in many applications ranging from atomic and molecular physics to metrology or imaging. In lensless imaging, for example, it can strongly affect the image formation, especially when the source exhibits shot-to-shot variations. Single-shot characterization of the spatial coherence length of a source is thus crucial. However, current techniques require either parallel intensity measurements or the use of several masks. Based on the method proposed by González et al. [J. Opt. Soc. Am. A 28, 1107 (2011)JOAOD60740-323210.1364/JOSAA.28.001107], we designed a specific arrangement of a two-dimensional non-redundant array of apertures, which allows, through its far field interference pattern, for a single-shot measurement of the spatial coherence, while being robust against beam-pointing instabilities. The strategic configuration of the pinholes allows us to disentangle the degree of spatial coherence from the intensity distribution, thus removing the need for parallel measurement of the beam intensity. An experimental validation is performed using a high-harmonic source. A statistical study in different regimes shows the robustness of the method.
... [202][203][204] LED light, given appropriate source dimensions and optical bandwidth, can be utilized as a spatially coherent light source for coherent imaging techniques. 205,206 Furthermore, the possibility of utilizing multicolored RGB LEDs as a multiwavelength source does not only add the ability to obtain color images, 207 but also offers the possibility to reconstruct the phase information in dense samples. 192 Among other microscopy methods, digital inline-holography provides a promising way of using micro-LED arrays as a spatially coherent light source in combination with an image sensor to build a compact digital microscope [ Fig. 12(a)]. ...
Article
Full-text available
Gallium nitride (GaN) light-emitting-diode (LED) technology has been the revolution in modern lighting. In the last decade, a huge global market of efficient, long-lasting, and ubiquitous white light sources has developed around the inception of the Nobel-prize-winning blue GaN LEDs. Today, GaN optoelectronics is developing beyond solid-state lighting, leading to new and innovative devices, e.g., for microdisplays, being the core technology for future augmented reality and visualization, as well as point light sources for optical excitation in communications, imaging, and sensing. This explosion of applications is driven by two main directions: the ability to produce very small GaN LEDs (micro-LEDs and nano-LEDs) with high efficiency and across large areas, in combination with the possibility to merge optoelectronic-grade GaN micro-LEDs with silicon microelectronics in a hybrid approach. GaN LED technology is now even spreading into the realm of display technology, which has been occupied by organic LEDs and liquid crystal displays for decades. In this review, the technological transition toward GaN micro- and nanodevices beyond lighting is discussed including an up-to-date overview on the state of the art.
... Apart from the signal-to-noise ratio and limited sampling rate of the sensor, the coherence of light has a direct influence on the resolution of a reconstructed object. Limited spatial or temporal coherence imparts a limit on the maximum observable fringe frequency, which, in turn, limits the maximum resolution [12][13][14]. On the other hand, high spatial and temporal coherence leads to speckle noise and fringe contamination generated due to the multiple reflections at different layers between the sensor and source [15,16]. ...
Article
Full-text available
Lensless microscopy is a simple, portable, and cost-effective method of microscopy. It has been extensively investigated with coherent light sources such as laser setups or partially coherent light sources such as an LED filtered with a pinhole aperture. The coherence of light has a direct influence on the resolution of the reconstructed object. This paper presents lensless microscopy with a spatially extended light source (white LED flashlight without any subsequent engineering). A reconstruction method based on constrained and regularized optimization is presented. The resolution reduced due to decreased coherence is gained by the presented method of object estimation.
... While conventional DH captures interferograms using high-coherence illumination such as that provided by lasers, the resolution and quality of the resulting image can be degraded by a large amount of speckle noise. Recently, alternative DH systems utilizing low-coherence light sources such as light-emitting diodes (LEDs) have been proposed in several studies [7][8][9][10][11][12][13]. In comparison to conventional laser-based configurations, such LED-based DH systems can suppress the overall speckle noise to achieve a better image quality. ...
Article
Full-text available
We propose a new low-coherence interferometry system for dual-wavelength off-axis digital holography. By utilizing diffraction gratings, two beams with narrower bandwidths and different center wavelengths could be filtered in a single light-emitting diode. The characteristics of the system are analytically determined to extend the coherence length and field-of-view enough for off-axis configuration. The proposed system enables the fast and accurate measurement of the surface profile with more than a micrometer step height and less noise. The performance of the system is verified by the experimental results of a standard height sample.
... Illuminating the SLM surface with low coherence LEDs results in smoother images in comparison to high coherence laser sources. LEDs as sources in digital holography, have previously been investigated in holographic microscopy [13] and effect of partial spatial coherence in digital holographic microscopy was also studied [14,15]. ...
Conference Paper
Full-text available
Holographic Displays (HDs) provide 3D images with all natural depth cues via computer generated holograms (CGHs) implemented on spatial light modulators (SLMs). HDs are coherent light processing systems based on interference and diffraction, thus they generally use laser light. However, laser sources are relatively expensive, available only at some particular wavelengths and difficult to miniaturize. In addition, highly coherent nature of laser light makes some undesired visual effects quite evident, such as speckle noise, interference due to stray light or defects of optical components. On the other hand, LED sources are available in variety of wavelengths, has small die size, and no speckle artifact. However, their finite spatial size introduce some degree of spatial incoherence in an HD system and degrade image resolution, which is the subject of the study in this paper. Our theoretical analysis indicates that the amount of resolution loss depends on the distance between hologram and SLM image planes. For some special configurations, the source size has no effect at all. We also performed experiments with different configurations using lasers and LEDs with different emission areas that vary from 50 μm to 200 μm, and determined Contrast Transfer Function (CTF) curves which agree well with our theoretical model. The results show that it is possible to find configurations where LEDs combined with pinholes almost preserve natural resolution limit of human eye while keeping the loss in light efficiency within tolerable limits.
Article
Full-text available
The X-ray ultramicroscope (XuM), based on using a scanning electron microscope as host, provides a new approach to X-ray projection microscopy. The right-angle-type integrated sample cells described here expand the capabilities of the XuM technique. The integrated sample cell combines a target, a spacer, a sample chamber, and an exit window in one physical unit, thereby simplifying the instrumentation and providing increased mechanical stability. The XuM imaging results presented here, obtained using such right-angle integrated sample cells, clearly demonstrate the ability to characterize very small features in objects, down to of order 100nm, including their use for dry, wet and even liquid samples.
Article
Platinum atoms near a (111) single-crystal face have been imaged using photoelectron holography. Electron angular intensity patterns were collected at equally spaced wavenumbers from 6 to 12 A(exp -1). Images of atoms near expected atomic positions are obtained from single-wavenumber analyses over the range of the data set. Positions are detected further from the emitter than we have seen previously, and symmetry assumptions are not required. We have also adopted a three dimensional means of representing the data in order to help understand the results. Twin image suppression and artifact reduction in the holographically reconstructed data are set are obtained when images at different wavenumbers are correctly phase-summed. We are assessing the capability of the technique for rendering true three-dimensional structural information for unknown systems.
Article
The effect of using quasi-monochromatic, partially spatially coherent illumination in recording a hologram on the object information available via digital holographic reconstruction is studied theoretically. The analysis concerns digital lensless Fresnel holography. Three object illumination models: a stationary illumination field, an illumination field generated by an incoherent source in the near field of the object, and a Gaussian Schell-model source (GSM), are explored in detail. We investigate the imaging properties of a Fresnel digital holographic system employing these illumination fields, through study of the impulse and frequency responses.
Article
It is shown that, under very general conditions, the cross-spectral density of a steady-state source of any state of coherence may be expressed in terms of certain new modes of oscillations, each of which represents a completely spatially coherent elementary excitation. Making use of this result, a statistical ensemble of strictly monochromatic oscillations, all of the same temporal frequency, is then introduced that yields the cross-spectral density as a correlation function in the space-frequency domain. From these results two new expressions for the Wiener-Khintchine spectrum of the source and also a new mode representation of the cross-correlation function of the source follow at once.
Article
Preface; 1. Elementary coherence phenomena; 2. Mathematical preliminaries; 3. Second-order coherence phenomena in the space-time domain; 4. Second-order coherence phenomenon in the space-frequency domain; 5. Radiation from sources of different states of coherence; 6. Coherence effects in scattering; 7. Higher-order coherence effects; 8. Elementary theory of polarization of stochastic electromagnetic beams; 9. Unified theory of polarization and coherence; Appendices; Index.
Article
A recently formulated theory of partial coherence in the space-frequency domain is used to determine the mode structure of an important class of partially coherent sources and of the radiation fields generated by them. The effective number of modes is found to depend in a fundamental way on the ratio of the coherence length to the effective size of the source. The contribution of the effective modes to the far-field intensity is also analyzed.
Article
Basing on the coherence theory the effect of quasi-monochromatic partially coherent illumination in lensless Fresnel holography is studied theoretically. A simple formula describing holographic fringe formation process with partially coherent illumination is found. The problem solution is restricted to the stationary and some types of nonstationary object illumination mutual intensity distributions. The detailed discussion related to the object information content in holographic fringes under various types of partially coherent object illumination is presented.
Article
The lensless digital in-line microscope is proposed as a simple tool for post-production inspection of metallic film patterns on silicon wafers. Transmission holographic imaging is achieved by using a superluminescent diode emitting in the near infrared coupled into a single-mode high numerical aperture optical fiber and an infrared vidicon camera. Images of 300 nm thick vapor-deposited aluminum bolometers are presented.
Article
Partially coherent X-ray in-line Gabor holography is studied based on the theory of partial coherence. Using asymptotic expansion methods, the relations between the coherent degree and the resolution (contrast) are obtained. Both transverse and longitudinal resolution and image contrast decrease when the coherent degree decreases and the improvement of X-ray coherence has little effect on the imaging quality if the coherence is high enough. The numerical examples are in good agreement with the analytic results. A criterion on the coherent length is proposed based on the calculating results to get high quality holography.
Article
This study has demonstrated the possibility of displaying light in flight via numerical reconstruction of holograms recorded directly on a high-resolution CCD sensor. This way, no chemical or physical development steps are required, and optical recording and numerical reconstruction can be integrated into one system. Using digitally available information, the coherence length of the laser is measured. Also demonstrated is an easy way to increase artificially the lateral dimensions of the CCD sensor by introducing delays in different parts of the reference wave, hence allowing the observation of the evolution of a wave front.