ArticlePDF Available

Full-field optical coherence microscopy with optimized ultrahigh spatial resolution

Optica Publishing Group
Optics Letters
Authors:

Abstract and Figures

Full-field optical coherence microscopy (FF-OCM) with isotropic spatial resolution of 0.5 μm (in water), at 700 nm center wavelength, is reported. A theoretical study of the FF-OCM axial response is carried out for maximizing the axial resolution of the system, considering the effect of optical dispersion. The lateral resolution is optimized by using water-immersion microscope objectives with a numerical aperture of 1.2. This ultrahigh-resolution FF-OCM system is applied to animal and human skin tissue imaging, revealing ultra-fine in-depth structures at the sub-cellular level.
This content is subject to copyright. Terms and conditions apply.
Full-field optical coherence microscopy with
optimized ultrahigh spatial resolution
ANTOINE FEDERICI AND ARNAUD DUBOIS*
Laboratoire Charles Fabry, CNRS UMR 8501, Institut dOptique Graduate School, Univ. Paris-Sud, 2 avenue Augustin Fresnel,
91127 Palaiseau Cedex, France
*Corresponding author: arnaud.dubois@institutoptique.fr
Received 1 October 2015; revised 21 October 2015; accepted 21 October 2015; posted 21 October 2015 (Doc. ID 251019);
published 10 November 2015
Full-field optical coherence microscopy (FF-OCM) with
isotropic spatial resolution of 0.5 μm (in water), at 700 nm
center wavelength, is reported. A theoretical study of the
FF-OCM axial response is carried out for maximizing the
axial resolution of the system, considering the effect of op-
tical dispersion. The lateral resolution is optimized by using
water-immersion microscope objectives with a numerical
aperture of 1.2. This ultrahigh-resolution FF-OCM system
is applied to animal and human skin tissue imaging,
revealing ultra-fine in-depth structures at the sub-cellular
level. © 2015 Optical Society of America
OCIS codes: (180.1655) Coherence tomography; (170.4500) Optical
coherence tomography; (110.3175) Interferometric imaging; (180.6900)
Three-dimensional microscopy; (170.3660) Light propagation in
tissues.
http://dx.doi.org/10.1364/OL.40.005347
Optical coherence tomography (OCT) is a well-established tech-
nique for two- and three-dimensional (3D) imaging of semi-
transparent samples with micrometer-scale spatial resolution
[13]. A feature of OCT is that the axial resolution is usually
governed by the spectral properties of the signal, whereas the
transverse (lateral) resolution depends on the numerical aperture
(NA) of the lenses used for imaging [2]. By using a broadband
light source, the best reported axial resolution in OCT is
0.56 μm (in water) at 700 nm center wavelength, but with
important side lobes in the axial point spread function (PSF)
[4]. Besides, the transverse resolution in conventional OCT, es-
pecially in the frequency-domain version of OCT, is limited be-
cause relatively low-NA lenses are used to preserve a sufficient
depth of field [5]. Optical coherence microscopy (OCM) is a
particular approach of OCT that acquires en face tomographic
images. Since there is no depth of field constraint in OCM,
lenses of higher NA can be used to achieve better transverse res-
olution. Two general approaches for OCM have been developed
to date. The first approach is based on the combination of scan-
ning confocal microscopy with low-coherence interferometry
[6]. This approach requires transverse scanning using a spatially
coherent light source such as a superluminescent diode or a
short-pulse laser [6,7]. Transverse imaging resolution of 2 μm
has been achieved with OCM using waterimmersion micro-
scope objectives with NA of 0.3 [8]. Another approach of
OCM is full-field optical coherence microscopy (FF-OCM), also
often referred to as full-field optical coherence tomography (FF-
OCT) [9]. FF-OCM is based on an interference microscope for
acquisition of en face tomographic images using an area camera
[10]. In FF-OCM, a low spatially coherent light source is em-
ployed for whole-field illumination without requiring transverse
scanning. Low coherence light sources such as a halogen lamp or
a light emitting diode (LED) are used in FF-OCM, ensuring
continuous broadband emission without spikes or modulations,
leading to higher quality of the axial PSF compared to that ob-
tained with broadband spatially coherent light sources [4,11].
Thus, FF-OCM is particularly appropriate for high-axial resolu-
tion imaging, while being adapted to high-transverse resolution
owing to the possibility of using high-NA microscope objectives
[12,13]. Three-dimensional high-resolution imaging of various
ex-vivo biological samples has been extensively performed with
FF-OCM [14,15], providing images close to those obtained
in conventional histology. Water-immersion microscope objec-
tives of NA up to 1.0 have been employed in FF-OCM for im-
aging of highly transparent samples [16,17]. Oil-immersion
objectives of NA 1.25 were used with quasi-monochromatic
light to achieve spatial rather than temporal coherence gating
[18]. Imaging with axial resolution of 1.0 μm was then demon-
strated, but with relatively weak penetration depth.
In this Letter, we present a FF-OCM system using broad-
band illumination and detection with 1.2 NA microscope
objectives to perform unprecedented ultrahigh-resolution 3D
imaging of highly scattering samples. The imaging axial reso-
lution has been optimized using a theoretical analysis of the
effect of optical dispersion in FF-OCM.
As mentioned earlier, assuming sufficiently broadband
illumination, the axial resolution in OCT is governed by the
temporal coherence of the light source [12]. According to the
WienerKhinchin theorem, the temporal coherence function is
related to the power spectral density of the source by a Fourier
transform. If supposing a Gaussian shape of the power spectral
density with respect to wave number σ1λ, the imaging
axial resolution can be expressed as
Letter Vol. 40, No. 22 / November 15 2015 / Optics Letters 5347
0146-9592/15/225347-04$15/0$15.00 © 2015 Optical Society of America
Δzlc
2ns
2ln2
πns
1
Δσ2ln2
πns
λmaxλmin
Δλ;(1)
where lcis the temporal coherence length; Δσand Δλλmax
λmin are the full width at half-maximum (FWHM) of the spec-
trum expressed with respect to wavenumber and wavelength,
respectively; and nsis the refractive index of the sample. By
introducing λ0as the geometric mean of λmin and λmax , Eq. (1)
can be rewritten as
Δz2ln2
πns
λ2
0
Δλ;(2)
which is the usual expression of the axial resolution in OCT
[2]. If the spectrum is not too broad, λ0can be approximately
considered as the center wavelength or the peak wavelength of
the spectrum.
Equation (1) shows that for a given spectral width Δλ,a
lower λmin results in a better axial resolution. To estimate
the best axial resolution that can be achieved in OCT, we as-
sume that λmin is fixed, and we look for an optimized value of
λmax. Equation (1) can then be expressed with respect to only
one variable, λmax. Since Δzstrictly decreases for λmax greater
than λmin, we can conclude that for a Gaussian frequency spec-
trum with a fixed value of λmin, the larger the spectral width Δλ,
the better the axial resolution. We point out that if λ0is con-
sidered as the arithmetic mean of λmin and λmax , instead of the
geometric mean, an optimum for the axial resolution is found
for λmax 3λmin, which is not true.
The above analysis was directly derived from Eq. (1) and did
not take into account optical dispersion or absorption within
the sample, which are major issues when imaging in depth.
Absorption, which is mostly due to water, the main component
of biological tissues, affects the shape of the spectrum especially
for wavelengths beyond 1.2 μm. Dispersion mismatch in the
interferometer can lead to axial resolution deterioration when
the group velocity dispersion of the immersion medium differs
from that of the sample. It is then tightly linked to the spectral
width. Other than developing a FF-OCM system dedicated
to a particular sample, such a mismatch cannot be totally com-
pensated for and must be considered. On the other hand,
dispersion mismatch resulting from imperfections of the beam
splitter, or from differences between the two objectives, can easily
be corrected by setting a glass plate in each arm of the interfer-
ometer [19]. Hence, only dispersion mismatch induced by the
sample is considered in the analysis presented below.
In [20], an analytical expression has been established dem-
onstrating the effect of dispersion on the effective coherence
length with a Gaussian spectrum. However, this study did not
consider the change of the immersion medium thickness result-
ing from the dynamic focusing process specific to FF-OCM
[21]. The formula given in [20] needs to be adapted, taking
into account the variation of thickness of the immersion
medium, called δLim, to express the effective axial resolution
Δzd, at depth Ls, in the sample,
Δz2
dΔz2F2ϕ2;sLsϕ2;i mδLim
nsΔz2
2
:(3)
In the above expression, Δzrepresents the axial resolution at
the surface of the sample (i.e., without dispersion), ϕ2;s and
ϕ2;im are the group velocity dispersion [20] of the sample
and of the immersion medium, respectively. Fis a constant
factor whose value depends on the definition of the axial res-
olution. With the usual FWHM criterion, F4c2ln 2.To
better reveal the spread of the interferogram when dispersion
mismatch occurs, it is more appropriate to define the axial res-
olution as the full width of the fringe envelop at 1/e of the
maximum. In that case, the effective axial resolution is denoted
by Δzd;1eand F4c2. Since δLim is equal to Lsnimns[19]
where nsand nim are the mean refractive indexes of the sample
and the immersion medium, respectively, Δzd;1eonly depends
on the distance Ls. Then, it is worth noticing that even though
the immersion medium has the same mean refractive index
as the sample, the GVDs have to be identical to avoid axial
resolution deterioration.
Let us consider human skin epidermis as a sample to be
imaged. Knowing its refractive index properties [22], we can
simulate the evolution of the axial resolution Δzd;1eas a func-
tion of the spectral width Δλat several distinct imaging depths,
by varying λmax and keeping λmin constant. We choose to fix
λmin to 500 nm, considering that for lower wavelengths light
absorption in skin is too important. Results of the numerical
simulations are presented in Fig. 1(a). At a given imaging depth
in the epidermis, a unique value of Δλ(i.e., a value of λmax )is
found to optimize the axial resolution. As expected, the larger
the imaging depth, the narrower the spectrum must be to re-
duce the effect of dispersion mismatch. Figure 1(b) shows the
evolution of the optimal axial resolution as a function of the
imaging depth in epidermis, considering water as an immersion
medium. In spite of the optimization of the spectrum width
according to the depth, a degradation of the optimal axial res-
olution with depth cannot be avoided. If the spectrum is not
adjusted during the acquisition of a z-stack of images, a com-
promise has to be done between superficial and deep axial
resolution, considering the fact that a stretching of the axial
response also reduces the signal contrast and, therefore, the
imaging penetration depth.
A way to avoid optical dispersion in OCT is to use a narrow
spectrum combined with high-NA, objectives as done in [18].
However, although potentially high axial resolution can be
achieved, based on spatial coherence gating due to narrow
depth of field, this induces important side lobes in the axial
PSF [12]. This leads to a degradation of the sectioning ability
and image quality. We measured the axial PSF of a FF-OCM
system equipped with high-NA microscope objectives, consid-
ering either a broad spectrum (KG1 filter from Schott) or a
narrow spectrum (10 nm wide FB710-10 filter). By represent-
ing the amplitude of the axial PSF in logarithmic scale
[Fig. 2(b)], which is commonly done as post data treatment
Fig. 1. (a) Axial resolution Δzd;1e(at 1/e) with respect to the
FWHM of the Gaussian spectrum, at different depths Lsin human
epidermis, taking into account dispersion mismatch. The cross mark-
ers point out the maximal reachable axial resolution for a given epi-
dermis layer thickness. (b) Maximal reachable axial resolution versus
depth in the epidermis. The cross markers correspond to those of (a).
5348 Vol. 40, No. 22 / November 15 2015 / Optics Letters Letter
for displaying OCT images, a significant degradation is high-
lighted when sectioning is achieved with spatial coherence
gating rather than temporal coherence gating.
The sectioning ability provided by high-NA objectives asso-
ciated to broadband illumination leads to reduced side lobes ow-
ing to the combination of the two coherence gating effects.
However, with a very broad spectrum, the effect of spatial coher-
ence gating is in practice negligible compared to the effect of tem-
poral coherence gating and has no impact on the axial resolution.
In this Letter, we report on a FF-OCM system whose im-
aging axial resolution has been optimized using the theoretical
study presented previously. The experimental setup is depicted
in Fig. 3.
The FF-OCM system uses a halogen lamp as the light
source in a Köhler illumination arrangement. The interferom-
eter is based on the Linnik configuration [12,13], employing
two 1.2 NA plan apochromatic water-immersion microscope
objectives (CFI Plan Apo VC 60XWI from Nikon) with a
300 μm-working distance. Since these high-NA objectives
are aberration corrected for imaging through a 145 μm thick
cover glass, we use the bottom surface of such a cover glass (see
Fig. 3) as a reference surface offering and, at the same time, a
low reflectivity (4%) as required for optimizing the detection
sensitivity in FF-OCM [13]. An identical cover glass is
placed above the sample, as indicated in Fig. 3. These high-
performance objectives are equipped with a correction collar
system permitting fine aberration correction depending on the
cover glass thickness. Two glass windows, referred to as DCP in
Fig. 3, are placed in the interferometer to compensate for
residual dispersion mismatch. Precise angular alignment of the
two windows was carried out to maximize the axial resolution.
A piezoelectric transducer (PZT) is used to oscillate the refer-
ence surface to perform phase-shifting interferometry [13]. En
face tomographic images are acquired by translating the refer-
ence arm step by step (0.3 μm step) to move the temporal co-
herence gate into the tissue, while the microscope objective of
the sample arm is moved simultaneously to adjust the focus.
Because of the small working distance of the objectives, the
imaging depth is limited to 150 μm.Theimagesareprojected
via an achromatic 400 mm focal length tube lens onto a 2D
CMOS-detector (OWL-CL from Raptor Photonics) with 320 ×
256 pixels, a frame rate up to 345 Hz, and a full well capacity of
170,000 electrons. Based on an InGaAs photodiode array with
extended sensitivity in the visible, this camera has a broad spec-
tral response expanding from 400 to 1800 nm (see Fig. 4).
A large choice of spectral bands is made possible by inserting
appropriate optical filters in the illumination path to match the
optimal configuration predicted by theory. To optimize the
spectrum regarding the axial resolution for imaging at depths
in skin between 10 and 50 μm(see Fig. 1), the KG1 band
pass filter is placed in the illumination path (resultant detected
power spectral density shown in Fig. 4). The theoretical and
experimental interferograms obtained with the KG1 filter are
shown in Fig. 5.
One can notice the high similarity between the simulation
[Fig. 5(a)] and the measurement [Fig. 5(b)], both indicating a
superficial axial resolution of Δz0.50 μmin water at a
center wavelength of 700 nm. Owing to the use of high-
NA microscope objectives, the system is capable of 0.5 μm spa-
tial resolution imaging in all spatial directions. To the best of
our knowledge, such an isotropic spatial resolution has never
been reached in OCT.
To extract the OCT signal from the acquired interferometric
images, a phase-shifting method based on a sinusoidal oscillation
of the reference surface is implemented [12]. Accumulation of
images is done to increase the detection sensitivity as usual in
Fig. 2. Axial PSF amplitude measured with an effective NA of 0.7
with broadband (blue curve) and narrow-band (red curve) illumina-
tion, represented in (a) linear and (b) logarithmic scale.
Fig. 3. FF-OCM experimental setup. AD, aperture diaphragm; F,
KG1 filter; FD, field diaphragm; BS, beam splitter; DCP, dispersion
compensation plate; MO, 1.2 NA microscope objective; PZT, piezo-
electric transducer; L1, lens; L2, tube lens (doublet).
Fig. 4. Detected power spectral density considering the absorption
by the immersion medium (water) of the microscope objectives, with-
out (dotted black curve) and with (blue curve) a KG1 filter inserted
into the illumination path. Without a KG1 filter, the power spectral
density is mainly limited by the quantum efficiency of the camera.
Fig. 5. (a) Theoretical and (b) experimental interferograms, using a
KG1 filter. The FWHMs of the theoretical and experimental interfero-
grams are identical, equal to 0.50 μm (in water).
Letter Vol. 40, No. 22 / November 15 2015 / Optics Letters 5349
FF-OCM [13,19]. By accumulating 50 en face images, leading
to an effective frame rate of 1Hz, a detection sensitivity of
71 dB is achieved. This relatively low detection sensitivity com-
pared to other FF-OCM systems is due to the high electronical
noise of the InGaAs array detector [19,23].
En face and reconstructed B-scan images of Xenopus Laevis
frog skin are presented in Fig. 6. The high spatial resolution
enables to see fine sub-cellular structures, such as cell mem-
branes and nuclei. One can observe that the density of the cells
increases with depth. These types of structures are also visible in
human skin as shown in Fig. 7. In this sample, the nuclei ap-
pear dark aside the light cytoplasm [24,25]. While a high axial
resolution is achieved, the imaging penetration depth is a bit
lower than in conventional FF-OCM [1315], as a conse-
quence of the larger numerical aperture of the objectives. The
larger numerical aperture results in higher sensitivity to optical
aberrations, which reduces the signal amplitude as the imaging
depth increases. Nevertheless, one can note, especially in the
B-scan images, that the axial resolution does not perceptibly
degrade with depth and enables the distinction of thin layers
in the whole imaged volume.
We have developed a FF-OCM system with the highest spa-
tial resolution (0.5 μm in water, isotropic) ever achieved in
OCT, to the best of our knowledge. Besides, this was achieved
at a center wavelength of 700 nm, which is well adapted to
the imaging of biological tissues regarding optical absorption
and scattering. A theoretical study of the effect of optical
dispersion mismatch in the interferometer has been carried
out, providing the optimal spectral width depending on the
imaging depth. An experimental validation of this study has
been performed by 3D imaging of skin samples with a penetra-
tion depth close to 100 μm. As the 1.2 NA objectives used in
this FF-OCM system are very sensitive to optical aberrations,
implementation of adaptive optics could significantly improve
the imaging penetration depth [26].
Funding. TecSan (ANR-08-TECS-012-03).
Acknowledgment. The authors are grateful to Odile
Bronchain (Institut des Neurosciences Paris-Saclay).
REFERENCES
1. D. Huang, E. A. Swanson, C. P. Lin, J. S. Schuman, W. G. Stinson, W.
Chang, M. R. Hee, T. Flotte, K. Gregory, C. A. Puliafito, and J. G.
Fujimoto, Science 254, 1178 (1991).
2. A. F. Fercher, J. Biomed. Opt. 1, 157 (1996).
3. A. M. Zysk, F. T. Nguyen, A. L. Oldenburg, D. L. Marks, and S. A.
Boppart, J. Biomed. Opt. 12, 051403 (2007).
4. B. Povazay, K. Bizheva, A. Unterhuber, B. Hermann, H. Sattmann,
A. F. Fercher, W. Drexler, A. Apolonski, W. J. Wadsworth, J. C.
Knight, P. S. J. Russell, M. Vetterlein, and E. Scherzer, Opt. Lett.
27, 1800 (2002).
5. M. Wojtkowski, R. Leitgeb, A. Kowalczyk, T. Bajraszewski, and A. F.
Fercher, J. Biomed. Opt. 7, 457 (2002).
6. J. A. Izatt, E. A. Swanson, J. G. Fujimoto, M. R. Hee, and G. M. Owen,
Opt. Lett. 19, 590 (1994).
7. A. Aguirre, P. Hsiung, T. Ko, I. Hartl, and J. Fujimoto, Opt. Lett. 28,
2064 (2003).
8. Y. Chen, S.-W. Huang, A. D. Aguirre, and J. G. Fujimoto, Opt. Lett. 32,
1971 (2007).
9. E. Beaurepaire, A. C. Boccara, M. Lebec, L. Blanchot, and H. Saint-
Jalmes, Opt. Lett. 23, 244 (1998).
10. L. Vabre, A. Dubois, and A. C. Boccara, Opt. Lett. 27, 530 (2002).
11. W. Drexler, U. Morgner, F. Kärtner, C. Pitris, S. Boppart, X. Li, E.
Ippen, and J. Fujimoto, Opt. Lett. 24, 1221 (1999).
12. A. Dubois, L. Vabre, A. C. Boccara, and E. Beaurepair e, Appl. Opt. 41,
805 (2002).
13. A. Dubois, K. Grieve, G. Moneron, R. Lecaque, L. Vabre, and C.
Boccara, Appl. Opt. 43, 2874 (2004).
14. A. Dubois, G. Moneron, K. Grieve, and A. C. Boccara, Phys. Med.
Biol. 49, 1227 (2004).
15. K. Grieve, M. Paques, A. Dubois, J. Sahel, C. Boccara, and J.-F. Le
Gargasson, Invest. Ophthalmol. Visual Sci. 45, 4126 (2004).
16. W. J. Choi, K. S. Park, T. J. Eom, M.-K. Oh, and B. H. Lee, Opt. Lett.
37, 2784 (2012).
17. W. J. Choi, D. I. Jeon, S.-G. Ahn, J.-H. Yoon, S. Kim, and B. H. Lee,
Opt. Express 18, 23285 (2010).
18. A. Safrani and I. Abdulhalim, Opt. Lett. 37, 458 (2012).
19. A. Dubois, G. Moneron, and C. Boccara, Opt. Commun. 266, 738
(2006).
20. C. K. Hitzenberger, A. Baumgartner, W. Drexler, and A. F. Fercher, J.
Biomed. Opt. 4, 144 (1999).
21. S. Labiau, G. David, S. Gigan, and A. C. Boccara, Opt. Lett. 34, 1576
(2009).
22. H. F. Ding, J. Q. Lu, W. A. Wooden, P. J. Kragel, and X. H. Hu, Phys.
Med. Biol. 51, 1479 (2006).
23. W. Oh, B. Bouma, N. Iftimia, S. Yun, R. Yelin, and G. Tearney, Opt.
Express 14, 726 (2006).
24. M. Rajadhyaksha, M. Grossman, D. Esterowitz, and R. H. Webb, J.
Invest. Dermatol. 104, 946 (1995).
25. C. C. Tsai, C. K. Chang, K. Y. Hsu, T. S. Ho, M. Y. Lin, J. W. Tjiu, and
S. L. Huang, Biomed. Opt. Express 5, 3001 (2014).
26. A. Kumar, W. Drexler, and R. A. Leitgeb, Opt. Express 21, 10850
(2013).
Fig. 6. En face images of Xenopus Laevis frog skin (a)(d) at 10, 14,
18, and 30 μm depth, respectively, and reconstructed B-scan image (e).
Scale bars: 10 μm.
Fig. 7. En face FF-OCM images of human skin acquired at depths
of (a) 8 and (b) 30 μm. B-scan image (c) obtained from a z-stack. Scale
bars: 10 μm.
5350 Vol. 40, No. 22 / November 15 2015 / Optics Letters Letter
... In contrast, in interference microscopy resolution enhancement techniques are rarely applied till now. Nevertheless, techniques such as microsphere-assistance or immersion systems known from conventional microscopic imaging are being used in interference microscopy too [4][5][6][7][8][9]. ...
... Otherwise, if M min = 0, M max = 1 the object comprises two irregularities in a distance d in x-direction. Using equation (8) and introducing the surface height factor s 0 one can easily transform any amplitude object into a phase object, where: ...
Article
Full-text available
The lateral resolution in microscopic imaging generally depends on both, the wavelength of light and the numerical aperture of the microscope objective lens. To quantify the lateral resolution Ernst Abbe considered an optical grating illuminated by plane waves. In contrast, the Rayleigh criterion holds for two point sources or point scatterers separated by a lateral distance, which are supposed to emit spherical waves. A portion of each spherical wave is collected by the objective lens and results in an Airy disc corresponding to a diffraction limited intensity point spread function (PSF). If incoherent illumination is employed the intensity PSFs related to different scatterers on an object are added resulting in the well-known Rayleigh resolution criterion. In interference microscopy instead of the intensity the electric field scattered or diffracted by an object will be affected by the transfer function of the optical imaging system. For a reflective object the lateral resolution of an interference microscope can be again characterized by the Abbe limit if the object under investigation is a grating. However, if two irregularities on a flat surface are being imaged the resolution no longer obeys the Rayleigh criterion. Instead, it corresponds to an optical system with an annular aperture and thus surpasses the prediction given by the Rayleigh criterion. This holds true for both, amplitude as well as phase objects, as it will be elucidated in this study by theoretical considerations, simulation results and an experimental proof of principle.
... [28] source spatialement incohérente [30], donc une source spatialement large et non ponctuelle afin d'éviter naturellement l'effet de speckle (ou cross-talk) qui détériore la qualité des images finales [31], [32], même si l'élimination de telles interférences parasites a été étudiée [33]- [35]. Les lampes halogènes étaient privilégiées en FF-OCM car peu onéreuses et émettant des spectres larges de corps noirs [36], [37], pour une très bonne résolution axiale [38]. Aujourd'hui, on utilise des LED qui arrivent maintenant à émettre des spectres assez larges pour l'OCT. ...
... Faire de l'imagerie en face demande tout de même une variation de la phase du signal, à démoduler par la suite pour faire apparaître l'image. En OCM, cette modulation était assurée par l'oscillation du miroir de référence, tout comme en FF-OCM [37], [38], et aussi en LC-OCT [67]. Cette solution est cependant imparfaite puisque, pour travailler en temps réel, il faut que l'oscillation soit extrêmement rapide, de l'ordre de 10 kHz, ce qui rend son mouvement sinusoïdal et favorise l'apparition d'artefacts. ...
Thesis
Les cancers de la peau sont le type de cancer le plus commun, dont les conséquences peuvent cependant être facilement stoppées en cas de détection précoce. Seulement, les diagnostics sont aujourd'hui réalisés après un simple examen visuel, éventuellement suivi d'un prélèvement de peau suspecte. Des coupes histologiques sont alors extraites de cet échantillon en le découpant en fines tranches puis en le colorant. Malgré la complexité de la procédure, l'observation de ces coupes au microscope permet de confirmer ou non la présence d'une tumeur. La tomographie par cohérence optique confocale linéaire (LC-OCT) est une méthode optique récente particulièrement adaptée à l'imagerie de la peau : elle en forme des images jusqu'à ~500 µm de profondeur, ce qui est suffisant pour observer la jonction épiderme-derme où la prolifération des cellules cancéreuses devient dangereuse. La LC-OCT offre une résolution quasi-isotrope de 1 µm qui permet l'observation des tissus cutanés à l'échelle cellulaire. Pour cela, elle met à profit les interférences d'ondes lumineuses qui rendent possible l'imagerie in vivo de manière non invasive, sans avoir recours aux biopsies. Cette facilité d'imager la peau entraine ainsi un meilleur suivi des patients. Les systèmes de LC-OCT sont habituellement basés sur un interféromètre de Linnik utilisant deux objectifs de microscope. Cette thèse présente le développement d'un système de LC-OCT basé cette fois sur un interféromètre de Mirau qui n'a besoin que d'un seul objectif dans le but d'en faciliter l'intégration dans une sonde portative et d'augmenter la vitesse d'acquisition des images. Ce système d'imagerie parvient ainsi à acquérir des coupes verticales de taille 940 µm×350 µm (latéral x × axial z) à 17 Hz, en plus d'images 3D de taille 940 µm×600 µm×350 µm (x×y×z) avec une résolution de 1,9 µm×1,3 µm×1,1 µm (x×y×z) de qualité similaire à celles des systèmes de LC-OCT conventionnels.
... Commercial diagnostic OCT systems have a typical axial resolution of 5 to 15 µm. Much higher resolution, up to ∼1 µm, has been demonstrated in the research community by using a broader optical bandwidth spanning nearly half an octave [1][2][3][4][5][6][7][8][9][10]. This ultrahigh resolution OCT (UHR-OCT), although instrumentally more complex, is needed to examine cellular-level features that are not easily visible in conventional-resolution OCT. ...
Article
Full-text available
Ultrahigh resolution optical coherence tomography (UHR-OCT) can image microscopic features that are not visible with the standard OCT resolution of 5-15 µm. In previous studies, high-speed UHR-OCT has been accomplished within the visible (VIS) and near-infrared (NIR-I) spectral ranges, specifically within 550-950 nm. Here, we present a spectral domain UHR-OCT system operating in a short-wavelength infrared (SWIR) range from 1000 to 1600 nm using a supercontinuum light source and an InGaAs-based spectrometer. We obtained an axial resolution of 2.6 µm in air, the highest ever recorded in the SWIR window to our knowledge, with deeper penetration into tissues than VIS or NIR-I light. We demonstrate imaging of conduction fibers of the left bundle branch in freshly excised porcine hearts. These results suggest a potential for deep-penetration, ultrahigh resolution OCT in intraoperative applications.
... OCM is based on a combination of optical coherence tomography (OCT) and confocal microscopy [1]. OCM allows three-dimensional mapping of the refractive index of the sample and has a submicron resolution [2]. In addition, one of the methods, phase-sensitive OCM, has subnanometer axial sensitivity to changes in optical path-length [3]. ...
Article
Full-text available
We report on the development of a new approach for studying the internal structure of polymer integrated nanophotonics devices using phase-sensitive optical coherence microscopy. Visualization and flaw detection of devices and their internal structure was carried out using the example of coupling gratings and prisms for a miniature Otto configuration with a characteristic gap height of 50-300 nm.
... In practice, pathologists utilize knowledge regarding both cytoplasmic morphology and cellular nucleus properties when diagnosing diseases. Typically, cell nuclei in biological tissues, such as epithelial cells, exhibit poor reflection characteristics under OCM owing to the homogeneous distribution of DNAs and RNAs [18,19]. Therefore, other techniques are required to examine nuclei effectively. ...
Article
Full-text available
A technique using Linnik-based optical coherence microscopy (OCM), with built-in fluorescence microscopy (FM), is demonstrated here to describe cellular-level morphology for fresh porcine and biobank tissue specimens. The proposed method utilizes color-coding to generate digital pseudo-H&E (p-H&E) images. Using the same camera, colocalized FM images are merged with corresponding morphological OCM images using a 24-bit RGB composition process to generate position-matched p-H&E images. From receipt of dissected fresh tissue piece to generation of stitched images, the total processing time is <15 min for a 1-cm² specimen, which is on average two times faster than frozen-section H&E process for fatty or water-rich fresh tissue specimens. This technique was successfully used to scan human and animal fresh tissue pieces, demonstrating its applicability for both biobank and veterinary purposes. We provide an in-depth comparison between p-H&E and human frozen-section H&E images acquired from the same metastatic sentinel lymph node slice (∼10 µm thick), and show the differences, like elastic fibers of a tiny blood vessel and cytoplasm of tumor cells. This optical sectioning technique provides histopathologists with a convenient assessment method that outputs large-field H&E-like images of fresh tissue pieces without requiring any physical embedment.
Article
Full-text available
Line-field confocal optical coherence tomography (LC-OCT) is an optical technique based on low-coherence interference microscopy with line illumination, designed for tomographic imaging of semi-transparent samples with micrometer-scale spatial resolution. A theoretical model of the signal acquired in LC-OCT is presented. The model shows that a refractive index mismatch between the sample and the immersion medium causes a dissociation of the coherence plane and the focal plane, leading to a decrease in the signal amplitude and a degradation of the image’s lateral resolution. Measurements are performed to validate and illustrate the theoretical predictions. A mathematical condition linking various experimental parameters is established to ensure that the degradation of image quality is negligible. This condition is tested experimentally by imaging a phantom. It is verified theoretically in the case of skin imaging, using experimental parameters corresponding to those of the commercially available LC-OCT device.
Article
Full-text available
Significance: Full-field optical coherence tomography (FF-OCT) enables en face views of scattering samples at a given depth with subcellular resolution, similar to biopsy without the need of sample slicing or other complex preparation. This noninvasive, high-resolution, three-dimensional (3D) imaging method has the potential to become a powerful tool in biomedical research, clinical applications, and other microscopic detection. Aim: Our review provides an overview of the disruptive innovations and key technologies to further improve FF-OCT performance, promoting FF-OCT technology in biomedical and other application scenarios. Approach: A comprehensive review of state-of-the-art accomplishments in OCT has been performed. Methods to improve performance of FF-OCT systems are reviewed, including advanced phase-shift approaches for imaging speed improvement, methods of denoising, artifact reduction, and aberration correction for imaging quality optimization, innovations for imaging flux expansion (field-of-view enlargement and imaging-depth-limit extension), new implementations for multimodality systems, and deep learning enhanced FF-OCT for information mining, etc. Finally, we summarize the application status and prospects of FF-OCT in the fields of biomedicine, materials science, security, and identification. Results: The most worth-expecting FF-OCT innovations include combining the technique of spatial modulation of optical field and computational optical imaging technology to obtain greater penetration depth, as well as exploiting endogenous contrast for functional imaging, e.g., dynamic FF-OCT, which enables noninvasive visualization of tissue dynamic properties or intracellular motility. Different dynamic imaging algorithms are compared using the same OCT data of the colorectal cancer organoid, which helps to understand the disadvantages and advantages of each. In addition, deep learning enhanced FF-OCT provides more valuable characteristic information, which is of great significance for auxiliary diagnosis and organoid detection. Conclusions: FF-OCT has not been completely exploited and has substantial growth potential. By elaborating the key technologies, performance optimization methods, and application status of FF-OCT, we expect to accelerate the development of FF-OCT in both academic and industry fields. This renewed perspective on FF-OCT may also serve as a road map for future development of invasive 3D super-resolution imaging techniques to solve the problems of microscopic visualization detection.
Article
Full-text available
Photodynamic therapy (PDT) is a promising tool for least-invasive alternative methods for the treatment of brain tumors. The newly discovered PDT-induced opening of the blood–brain barrier (BBB) permeability open novel strategies for drug-brain delivery during post-surgical treatment of glioblastoma GBM. Here we discuss mechanisms of PDT-mediated opening of the BBB and age differences in PDT-related increase in BBB permeability, including with formation of brain edema. The meningeal lymphatic system plays a crucial role in the mechanism of brain drainage and clearance from metabolites and toxic molecules. We discuss that noninvasive photonic stimulation of fluid clearance via meningeal lymphatic vessels, and application of optical coherence tomography (OCT) for bed-side monitoring of meningeal lymphatic drainage has the promising perspective to be widely applied in both experimental and clinical studies of PDT and improving guidelines of PDT of brain tumors.
Article
Full-text available
With a Gaussian-like broadband light source from high brightness Ce³⁺:YAG single-clad crystal fiber, a full-field optical coherence tomography using a home-designed Mirau objective realized high quality images of in vivo and excised skin tissues. With a 40 × silicone-oil-immersion Mirau objective, the achieved spatial resolutions in axial and lateral directions were 0.9 and 0.51 μm, respectively. Such a high spatial resolution enables the separation of lamellar structure of the full epidermis in both the cross-sectional and en face planes. The number of layers of stratum corneum and its thickness were quantitatively measured. This label free and non-invasive optical probe could be useful for evaluating the water barrier of skin tissue in clinics. As a preliminary in vivo experiment, the blood vessel in dermis was also observed, and the flowing of the red blood cells and location of the melanocyte were traced.
Article
Full-text available
We describe a scanning white-light interferometer for high-precision surface structure analysis. Interferograms for each of the image points in the field of view of the instrument are generated simultaneously by scanning the object in a direction perpendicular to the object surface, while recording detector data in digital memory. These interferograms are then transformed into the spatial frequency domain and the surface height for each point is obtained by examination of the complex phase as a function of frequency. The final step is the creation of a complete three-dimensional image constructed from the height data and corresponding image plane coordinates. The measurement repeatability is better than 0·5 nm r.m.s. for a surface height range of 100 μm.
Article
Full-text available
Full-field optical coherence microscopy is an established optical technology based on low-coherence interference microscopy for high-resolution imaging of semitransparent samples. In this Letter, we demonstrate an extension of the technique using a visible to short-wavelength infrared camera and a halogen lamp to image in three distinct bands centered at 635, 870, and 1170 nm. Reflective microscope objectives are employed to minimize chromatic aberrations of the imaging system operating over a spectral range extending from 530 to 1700 nm. Constant 1.9-μm axial resolution (measured in air) is achieved in each of the three bands. A dynamic dispersion compensation system is set up to preserve the axial resolution when the imaging depth is varied. The images can be analyzed in the conventional RGB color channels representation to generate three-dimensional images with enhanced contrast. The capability of the system is illustrated by imaging different samples.
Article
Full-text available
We introduce a microscopic method that determines quantitative optical properties beyond the optical diffraction limit and allows direct imaging of unstained living biological specimens. In established holographic microscopy, complex fields are measured using interferometric detection, allowing diffraction-limited phase measurements. Here, we show that non-invasive optical nanoscopy can achieve a lateral resolution of 90 nm by using a quasi-2π-holographic detection scheme and complex deconvolution. We record holograms from different illumination directions on the sample plane and observe subwavelength tomographic variations of the specimen. Nanoscale apertures serve to calibrate the tomographic reconstruction and to characterize the imaging system by means of the coherent transfer function. This gives rise to realistic inverse filtering and guarantees true complex field reconstruction. The observations are shown for nanoscopic porous cell frustule (diatoms), for the direct study of bacteria (Escherichia coli), and for a time-lapse approach to explore the dynamics of living dendritic spines (neurones).
Article
Full-text available
This chapter describes the phase-measurement interferometry techniques. For all techniques, a temporal phase modulation is introduced to perform the measurement. By measuring the interferogram intensity as the phase is shifted, the phase of the wavefront can be determined with the aid of electronics or a computer. Phase modulation in an interferometer can be induced by moving a mirror, tilting a glass plate, moving a grating, rotating a half-wave plate or analyzer, using an acousto-optic or electro-optic modulator, or using a Zeeman laser. Phase-measurement techniques using analytical means to determine phase all have some common denominators. There are different equations for calculating the phase of a wavefront from interference fringe intensity measurements. The precision of a phase-measuring interferometer system can be determined by taking two measurements, subtracting them, and looking at the root-meansquare of the difference wavefront. The chapter discusses the simulation results. The elimination of the errors that reduce the measurement accuracy depends on the type of measurement being performed. Phase-measurement interferometry (PMI) can be applied to any two-beam interferometer, including holographic interferometers. Applications can be divided into: surface figure, surface roughness, and metrology.
Article
Full-text available
This paper proposes a sub-aperture correlation based numerical phase correction method for interferometric full field imaging systems provided the complex object field information can be extracted. This method corrects for the wavefront aberration at the pupil/ Fourier transform plane without the need of any adaptive optics, spatial light modulators (SLM) and additional cameras. We show that this method does not require the knowledge of any system parameters. In the simulation study, we consider a full field swept source OCT (FF SSOCT) system to show the working principle of the algorithm. Experimental results are presented for a technical and biological sample to demonstrate the proof of the principle.
Article
A compact and efficient algorithm for digital envelope detection in white light interferograms is derived from a well-known phase-shifting algorithm. The performance of the new algorithm is compared with that of other schemes currently used. Principal criteria considered are computational efficiency and accuracy in the presence of miscalibration. The new algorithm is shown to be near optimal in terms of computational efficiency and can be represented as a second-order nonlinear filter. In combination with a carefully designed peak detection method the algorithm exhibits exceptionally good performance on simulated interferograms. Key words: interferometry, white light interferometry, interference microscopy, phase-shifting algorithm, low coherence, nonlinear filter, envelope detection, demodulation.
Article
In this paper, we discuss the possibility of making tomographic reconstruction of the refractive index of a microscopic sample using a quadriwave lateral shearing interferometer, under incoherent illumination. A Z-stack is performed and the acquired incoherent elecromagnetic fields are deconvoluted before to retrieve in a quantitative manner the refractive index. The results are presented on polystyrene beads and can easily be expanded to biological samples. This technique is suitable to any white-light microscope equipped with nanometric Z-stack module.