You are on page 1of 6

Compact and flexible raster scanning multiphoton

endoscope capable of imaging unstained tissue


David R. Rivera1,2, Christopher M. Brown1, Dimitre G. Ouzounov, Ina Pavlova, Demirhan Kobat,
Watt W. Webb2, and Chris Xu
School of Applied and Engineering Physics, Cornell University, 271 Clark Hall, Ithaca, NY 14853

Contributed by Watt W. Webb, September 7, 2011 (sent for review April 28, 2011)

We present a compact and flexible endoscope (3-mm outer dia- their miniaturized dimensions (e.g., o:d: ≈ 1 mm) and fast image
meter, 4-cm rigid length) that utilizes a miniaturized resonant/non- acquisition speeds [e.g., 8 frames∕s with 512 × 512 pixels per
resonant fiber raster scanner and a multielement gradient-index frame, approximately 200-μm diameter field-of-view (FOVxy )]
lens assembly for two-photon excited intrinsic fluorescence and (14); however, these resonant devices are fundamentally limited
second-harmonic generation imaging of biological tissues. The by nonuniform spatial coverage and sampling time in comparison
miniaturized raster scanner is fabricated by mounting a commercial to current miniaturized raster scanners. Current miniaturized
double-clad optical fiber (DCF) onto two piezo bimorphs that are raster scanners are, however, limited in terms of their physical
aligned such that their bending axes are perpendicular to each dimensions and/or scan speeds. Le Harzic et al. previously de-
other. Fast lateral scanning of the laser illumination at 4.1 frames∕s monstrated a piezo-driven X-Y scanner (length ¼ 34 mm,
(512 lines per frame) is achieved by simultaneously driving the DCF width ¼ 1.9 mm) capable of a uniformly sampled FOVxy up to
cantilever at its resonant frequency in one dimension and nonre- 420 μm by 420 μm, but this device is limited by its frame rate
sonantly in the orthogonal axis. The implementation of a DCF into (i.e., 0.1 frames∕s with 512 × 512 pixels per frame) (19). Slow
the scanner enables simultaneous delivery of the femtosecond image acquisition speeds are not ideal because of the motion
pulsed 800-nm excitation source and epi-collection of the signal. artifacts typically faced in real-time in vivo clinical imaging envir-
Our device is able to achieve a field-of-view (FOVxy ) of 110 μm by onments. Additionally, although 2D MEMS scanning mirrors
110 μm with a highly uniform pixel dwell time. The lateral and axial with miniature dimensions (e.g., 750 μm × 750 μm mirror size)
resolutions for two-photon imaging are 0.8 and 10 μm, respec- have recently demonstrated fast line acquisition rates on the
tively. The endoscope’s imaging capabilities were demonstrated order of 1–3 kHz, the overall miniaturization of these MEMS
by imaging ex vivo mouse tissue through the collection of intrinsic scanners (i.e., their probe o.d.s) is limited by the die size of the
fluorescence and second-harmonic signal without the need for actuator, which is typically 3 mm × 3 mm (24, 25).
Downloaded from https://www.pnas.org by 111.249.51.128 on October 29, 2023 from IP address 111.249.51.128.

staining. The results presented here indicate that our device can be Here, we present an endoscope that utilizes a miniaturized
applied in the future to perform minimally invasive in vivo optical resonant/nonresonant cantilever fiber raster scanner design that
biopsies for medical diagnostics. is compact (width∕thickness ≤ 1 mm, total length ≈ 2.6 cm),
achieves ≥650 μm fiber-tip deflection for both the resonant
nonlinear optical endoscopy ∣ real-time optical diagnosis ∣ scanning fiber and nonresonant axes, and allows for imaging at approximately
endoscopy ∣ microendoscopy ∣ endogenous fluorescence 4.1 frames∕s (512 × 512 pixels). The fiber scanner’s small dimen-
sions enable it to be packaged along with a miniaturized high
N.A. gradient-index (GRIN) assembly to form a compact and
M ultiphoton imaging techniques such as two-photon fluores-
cence (TPF) and second-harmonic generation (SHG)
microscopy hold great promise for the future of medical diagnosis
flexible TPF/SHG endoscope, with an outside diameter of 3 mm
and a rigid length of 4 cm. The temporal pulse widths at different
because of their potential to replace surgical biopsies with mini- endoscope output powers were characterized as well as the ima-
mally invasive optical diagnosis of tissue health (1–7). In a clinical ging resolution of the device. Finally, we demonstrate that our
setting, these diagnostic techniques will be capable of acquiring multiphoton endoscope prototype is able to obtain ex vivo tissue
real-time, high-resolution, in vivo images without the need for images solely based on intrinsic fluorescence and SHG signal.
contrast agents. However, a challenge in translating these bene- These results show the potential for our endoscope prototype
ficial imaging technologies into the clinic lies in successfully to be used as a real-time in vivo diagnostic tool for medical diag-
miniaturizing bulky tabletop microscope components into a com- nostics.
pact probe without degrading the overall imaging performance of
the system. These microendoscopes would not only have the Results and Discussion
potential to be used as diagnostic tools capable of early cancer Miniaturized Raster Scanner. The miniaturized fiber raster scanner
detection, but could also be used for such applications as photo- is driven by high-performance two-layer piezoelectric actuators
dynamic therapy and microsurgery (8, 9). or bimorphs (T215-A4CL, Piezo Systems, Inc.). These bimorphs
A number of groups have demonstrated miniaturized instru- are suitable for the fabrication of a miniaturized scanning
ments capable of confocal, optical coherence tomography mechanism because of their small dimensions (e.g., width ¼
(OCT), TPF, and SHG imaging (10–25). The primary constitu- 0.5–1 mm, thickness ¼ 380 μm) and ability to achieve deflection
ents of these devices are typically a miniaturized scanning me-
chanism and lens assembly that is encapsulated in a protective Author contributions: D.R.R., C.M.B., W.W.W., and C.X. designed research; D.R.R., C.M.B.,
housing with dimensions suitable for minimally invasive proce- D.G.O, and I.P. performed research; D.R.R., C.M.B., D.G.O., D.K., W.W.W., and C.X. analyzed
dures (i.e., a probe outer diameter on the order of a few milli- data; and D.R.R., C.M.B., W.W.W., and C.X. wrote the paper.
meters with a rigid length of several centimeters). Within these The authors declare no conflict of interest.
microendoscopes, various distal miniaturized scanners have been Freely available online through the PNAS open access option.
demonstrated, including resonant-based [e.g., Lissajous (10, 11) 1
D.R.R. and C.M.B. contributed equally to this work.
or spiral (12–16) scan pattern] and non-resonant-based cantilever 2
To whom correspondence may be addressed. E-mail: drr66@cornell.edu or www2@
fiber scanners (16–19) as well as microelectromechanical systems cornell.edu.
(MEMS) scanning mirrors (20–25). Of these scanners, the reso- This article contains supporting information online at www.pnas.org/lookup/suppl/
nant-based spiral scanners are the most successful in terms of doi:10.1073/pnas.1114746108/-/DCSupplemental.

17598–17603 ∣ PNAS ∣ October 25, 2011 ∣ vol. 108 ∣ no. 43 www.pnas.org/cgi/doi/10.1073/pnas.1114746108


on the order of hundreds to thousands of microns. In our device, TECH), into a waterproof stainless steel housing with an outer
two bimorphs are aligned such that their bending axes are per- diameter of 3 mm and a rigid length of 4 cm. The 0.8-N.A. GRIN
pendicular to each other and serve as the resonant and nonreso- assembly has a total length of 7.62 mm and is protected by a
nant drivers for the lateral X-Y scanning. A commercial double- 1.4-mm o.d. stainless steel housing. The miniature size of the
clad optical fiber (DCF SM-9/105/125-20A, NuFern) is mounted scanner and lens assembly enable the distal end of the endoscope
as a cantilever onto these axes and is driven resonantly in one to have dimensions that are well suited for a minimally invasive
dimension and nonresonantly in the other. The DCF has a core, probe. Fig. 1A shows a mechanical assembly diagram of the distal
inner cladding, and an outer cladding with diameters of 9, 105, end, and Fig. 1B shows a photograph of the external protective
and 125 μm, respectively. To determine the free hanging length housing that encapsulates the miniaturized scanner and lens
of the DCF that would enable high speed resonant scanning on assembly. Within the protective housing, the double-clad fiber tip
the order of 1 kHz, we used the following equation (26): is centered laterally and separated approximately 200 μm axially
sffiffiffiffi from the 1-mm diameter back aperture of the GRIN assembly
βR E (Fig. 1C). For endoscopic imaging, a tabletop setup is used
ν¼ ; (Fig. 1D); see Materials and Methods for full setup details.
4πL2 ρ
Pulse Characterization. We precompensate the fiber dispersion
where E is the Young’s modulus (silica), ρ is the density (silica), R by imposing an appropriate quadratic chirp on the input pulses
is fiber radius, L is the fiber overhang length, and β is a constant with a rotating cylindrical lens and grating (27), which has the
that is dependent upon the boundary conditions of the cantilever advantage of allowing continuous tuning of the second-order dis-
and the vibration mode number. In our calculation, we use the persion over a large range without perturbing the spatial align-
boundary condition of a fixed-free cantilever beam and the ment. Fig. 2A shows the second-order autocorrelation traces
zeroth-order vibration mode (β ≈ 3.52). From this calculation, we
chose a fiber overhang length of approximately 9 mm for our
scanner, which has a measured resonant frequency of 1.05 kHz,

APPLIED PHYSICAL
in close agreement with the theoretical prediction of 1.1 kHz. A

SCIENCES
3-mm length piezo bimorph was sufficient to provide a fiber-tip
deflection range of over 1 mm at a peak-to-peak drive voltage
(Vpp) of 50 V.
In order to achieve large nonresonant fiber-tip deflection,
while minimizing the overall rigid length of the cantilever, we
optimized the active length of the nonresonant piezo bimorph
using the following equation:
Downloaded from https://www.pnas.org by 111.249.51.128 on October 29, 2023 from IP address 111.249.51.128.

 
2
V applied
D ¼ 2cðL1 þ 2L1 L2 Þ × ;
V spec

where D is the nonresonant fiber-tip deflection (μm), c is a pro-


portionality constant, L1 is the free length of the nonresonant
piezo bimorph, and L2 is the total length of the resonant piezo
bimorph and the overhang length of the optical fiber (i.e.,
12 mm). Using this calculation as a guide, we chose a nonresonant
piezo length of 14 mm, which theoretically predicted deflection
up to 763 μm at 200 Vpp. Our current scanner is able to achieve
approximately 650 μm nonresonant fiber-tip deflection at
200 Vpp. The underperformance in the deflection is attributed
to the small lateral dimensions of the scanner, which will slightly
reduce the electric field strength in the piezo due to the edge
effect. Although the application of relatively high voltage to the
piezo bimorphs is necessary for high fiber-tip deflection (i.e.,
200 Vpp to achieve nonresonant fiber-tip deflection up to
650 μm), we note that the electrical current within these bimorphs
is small, with estimated currents of 30 and 1 μA for the resonant
and nonresonant bimorphs, respectively.
It should be noted that in order to generate a purely linear
fiber-tip motion in our resonant axis, we incorporated a small
stiffening rod into the scanner to break the cylindrical symmetry
of the optical fiber. Additionally, we note that our scanner has
the ability to arbitrarily position the resonant axis line-scans in a
direction orthogonal to the nonresonant axis by applying a dc off-
set voltage to the nonresonant bimorph. This may be of clinical
use for measuring fast, dynamic physiological phenomena.

Three-Millimeter Outer Diameter Multiphoton Endoscope. The proto-


type miniaturized raster scanner has a width/thickness ≤1 mm, a
total length of approximately 2.6 cm, and is capable of fiber-tip Fig. 1. Three-millimeter o.d. raster scanning endoscope components and
deflection of 1.0 mm × 0.65 mm. The small width and thickness setup. (A) Mechanical assembly of the endoscope components. (B) Photo-
of the scanner allow it to be packaged, along with a commercial graph of the prototype. (C) Optical path of the internal components of
multielement GRIN lens assembly (GT-MO-080-018-810, GRIN- the distal end. (D) Tabletop endoscope imaging setup.

Rivera et al. PNAS ∣ October 25, 2011 ∣ vol. 108 ∣ no. 43 ∣ 17599
of the initial laser output as well as at 50-mW output from our
endoscope. We are able to achieve a 290-fs pulse width at 50-mW
output from the core of the DCF. The measured optical spectra
are shown in Fig. 2B. The characteristic spectrum narrowing
due to the optical nonlinearity and the normal dispersion of the
fiber is clearly visible. Fig. 2C shows the measured pulse widths
over a range of endoscope power outputs. These results are in
close agreement with numerical simulations based on the non-
linear Schrodinger’s equation (28).

Optical Resolution. The lateral and axial resolutions of our device


are characterized by the transmission image of a high-resolution
United States Air Force (USAF) test target and the axial re-
sponse of the two-photon excited fluorescence signal from a
500-nm thin film of Rhodamine B (RhB) dye, respectively. Fig. 3A
shows a transmission image of a high-resolution USAF test tar-
get, where the smallest line-width group (group 9) is displayed.
The group 9 elements have line widths ranging from 977 nm
(group 9, element 1) down to 775 nm (group 9, element 3) and
are all discernible in this image. Fig. 3B shows an inverted line-
intensity profile across the vertical lines in group 9, element 1.
Given the line widths of 977 nm and the measured peak-to-trough
ratio of 0.46, we determine a one-photon lateral resolution of Fig. 3. Lateral and axial resolution. (A) Transmission image of high-resolu-
approximately 1.1 μm full width at half maximum (FWHM), tion USAF test target. Group 9, elements 1–3 are displayed with line widths
ranging from 977–775 nm. FOVxy ¼ 10 μm × 30 μm. (B) Intensity line profile
which corresponds to a two-photon lateral resolution of approxi- across the vertical bars of group 9, element 1. Using the intensity ratios
mately 0.8 μm (FWHM). Fig. 3C shows the intensity profile pro- between the peak and trough values, we determine a one-photon lateral
duced as a result of stepping a 500-nm RhB thin film axially resolution of approximately 1.1 μm (FWHM), which corresponds to a two-
through the focal plane of the endoscope prototype. The FWHM photon lateral resolution of approximately 0.8 μm (FWHM). (C) A two-photon
axial resolution of approximately 10 μm (FWHM of the thin film response) is
determined using a RhB thin film scan.

of the intensity profile is approximately 10 μm. Using the magni-


Downloaded from https://www.pnas.org by 111.249.51.128 on October 29, 2023 from IP address 111.249.51.128.

fication of the endoscopic lenses (M ¼ 1∕4.8) and the measured


mode field diameter (MFD) of the DCF (MFD ≈ 9.1 μm at
800 nm), the theoretical two-photon lateral and axial resolutions
are approximately 0.8 μm (FWHM) and 9.4 μm (FWHM of the
thin film response), respectively. The demonstrated lateral and
axial resolutions should be sufficient for resolving subcellular
details in biological tissues.

Scan Uniformity Characterization. A raster scan pattern is superior


to a spiral or Lissajous scan pattern because of the uniformity
in its spatial scan pattern and pixel dwell time throughout the
sample. In order to demonstrate the high uniformity in our raster
scanner, Fig. 4A shows a transmission image of a 400 line-pair per
mm (LP/mm) Ronchi ruling, without any image processing. The
large deflection range in the resonant scan is evident by the ap-
pearance of the black regions on the horizontal edges of Fig. 4A,
which indicate that the fiber tip is scanning well past the 1-mm
back aperture of the GRIN lens. The visible FOVxy that is con-
tained within the GRIN lens is approximately 208 μm × 110 μm.
We measured the line widths horizontally across the Ronchi rul-
ing as shown in Fig. 4B. Over a horizontal range of approximately
110 μm, there is a line-width deviation of approximately 7.5%.
Thus, a uniform pixel dwell time over a reasonably large FOVxy
(110 μm × 110 μm) can be achieved (dashed region within
Fig. 4A).
In the Ronchi ruling image (Fig. 4A), it is apparent that the
1.25-μm-width vertical lines do not appear to be perfectly straight
throughout the entire image. This imperfection is attributed to
a slight variation in our scanner's resonant deflection, which is
dependent upon the nonresonant scan position. Although it leads
Fig. 2. Pulse characterization. (A) Autocorrelation pulse-width measure-
ments for the Ti:Sapphire output and for 50-mW endoscope output. The
to image distortions if uncorrected, such an imperfection has
measured pulse widths are approximately 100 and 290 fs, respectively. negligible impact on the sampling uniformity of the scanner.
(B) Spectra for the Ti:Sapphire output and for endoscope output at 50 mW. We observe that the variation in the resonant axis deflection is
(C) Measured pulse widths of the endoscope output as a function of output consistent, stable and repeatable between frames for given drive
power, assuming a deconvolution factor of 1.54 for a Sech2 pulse. parameters (i.e., drive frequency and applied voltage to the

17600 ∣ www.pnas.org/cgi/doi/10.1073/pnas.1114746108 Rivera et al.


APPLIED PHYSICAL
Fig. 4. Raster scan uniformity characterization. (A) Transmission image of a 400 LP∕mm Ronchi ruling (line width ¼ 1.25 μm). (B) Plot of the measured line
widths of the Ronchi ruling as a function of the horizontal pixel number across the image. A horizontal range of approximately 110 μm corresponds to a line-

SCIENCES
width deviation of 7.5%. (C) Cropped and postprocessed Ronchi ruling image from Fig. 4A (FOVxy ≈110 μm × 110 μm). The Ronchi ruling lines in C are cor-
rected to be of uniform width across the image FOVxy . (D) Cropped and postprocessed transmission image of USAF test target group 6, elements 2–3 (line
widths: 7.0 to 6.2 μm, FOVxy ≈110 μm × 110 μm).

piezos). The repeatability of the scan pattern enables us to devel- ferent depths where anatomical details such as the alveolar walls
op an image remapping algorithm to correct the original unpro- and lumens are visible; see Movie S2 for full depth scan. Fig. 5C
cessed Ronchi ruling image (see Fig. 4C for the corrected Ronchi shows five frame-averaged images at three different axial depths
ruling image), which can then be applied to correct subsequently
Downloaded from https://www.pnas.org by 111.249.51.128 on October 29, 2023 from IP address 111.249.51.128.

acquired images (e.g., Fig. 4D).


For comparison, the pixel dwell time of a typical spiral scanner
is nonuniform with differences as high as 800 times between the
center and periphery of the scan (14). It is impractical to com-
pensate for such large scan nonuniformities with power modula-
tion because of the following factors: (i) To achieve equivalent
two-photon excitation between the center and periphery of the
spiral scan, the power at the periphery must be much higher
(>10×) than that at the center; a single dispersion compensation
setup cannot provide optimal compensation over such a large
power variation because of the nonlinear pulse broadening
effects within the optical fiber (see Fig. 2C). (ii) The high power
levels required at the periphery of the spiral scan will add signif-
icant cost and complexity to the laser source, and may preclude
the use of fiber-based laser sources, which are widely considered
to be an ideal source for use in the clinical setting because of their
robustness, small footprint, and low cost.

Ex Vivo Intrinsic Fluorescence Imaging. An ideal optical diagnostic


tool will acquire high-quality images with minimal frame aver-
aging and without tissue staining or processing. The use of tissue
staining, although common for ex vivo imaging (e.g., in the
laboratory setting), is not ideal for many in vivo clinical imaging
applications; therefore, we demonstrate the potential of our
device to be used as an in vivo imaging instrument by acquiring
intrinsic TPF/SHG signal from label-free ex vivo mouse tissue. Fig. 5. TPF/SHG images of ex vivo mouse tissue. (A) Unaveraged SHG images
Because of the raster scanner’s highly uniform pixel dwell time, of mouse tail tendon at 10, 20, and 30 μm from the surface. (B) Unaveraged
minimal frame averaging was necessary in order to obtain high- intrinsic fluorescence images of mouse lung at 50, 60, and 70 μm from the
quality images; see Materials and Methods for additional image tissue surface. In B1, alveolar walls (w) and lumens (a) are clearly visible; in
acquisition details. B2, a bronchiole (b) is distinguishable. (C) Five frames averaged intrinsic fluor-
escence images of mouse colon at 35, 45, and 55 μm from the surface. In C1,
Fig. 5A shows unaveraged ex vivo images of SHG from col-
enterocytes (e) are visible; in C2, crypts (c) and goblet cells (g) are present.
lagen fibers taken from a mouse tail at three representative Scale bars, 10 μm. All images were acquired at 4.1 frames∕second, 800-nm
depths within the tissue; see Movie S1 for full depth scan. In these excitation, and the uniformly sampled FOVxy of approximately 110 μm ×
images, individual strands are discernible. Fig. 5B shows unaver- 110 μm is displayed. For the images in A, the power at the sample is approxi-
aged intrinsic fluorescence images of ex vivo mouse lung at dif- mately 30 mW; for the images in B and C, the power at the sample is 75 mW.

Rivera et al. PNAS ∣ October 25, 2011 ∣ vol. 108 ∣ no. 43 ∣ 17601
of the two-photon excited intrinsic fluorescence from ex vivo
mouse colon tissue; see Movie S3 for full depth scan. Mouse
colon tissue consists of closely packed tubular glands (crypts)
covered by a layer of columnar absorptive cells and mucus secret-
ing goblet cells. The cellular structures in Fig. 5C represent the
intrinsic fluorescence emitted from these epithelial cells. The
morphological details present in the unstained images displayed
in Fig. 5 indicate that our prototype shows potential for future
clinical use. No tissue damage was observed during the course of
the ex vivo imaging session, and the imaging conditions were
comparable to those shown by other investigators to have negli-
gible tissue mutagenicity (29).
For comparison, we imaged two-photon excited intrinsic fluor-
escence from ex vivo mouse tissue using a commercial multi-
photon microscope (Fluoview FV1000D, Olympus). Fig. 6 shows
unstained images acquired from both systems at comparable
depths within the tissue. This figure was obtained with compar-
able amounts of two-photon excited fluorescence at the sample Fig. 6. Imaging comparison between the multiphoton endoscope and a
per frame in each system. These images demonstrate that the commercial multiphoton microscope. (A) Unaveraged intrinsic fluorescence
multiphoton endoscope is capable of acquiring images of anato- images of ex vivo mouse lung. A1 shows the image acquired from the multi-
mical features that are similar in appearance to those obtained photon endoscope, and A2 shows the image acquired from the Olympus
from a conventional multiphoton microscope. multiphoton microscope. (B) Five frames averaged intrinsic fluorescence
images of ex vivo mouse colon. B1 shows the image acquired from the multi-
Conclusions and Outlook. In summary, we have demonstrated a photon endoscope, and B2 shows the image acquired from the Olympus
compact and flexible raster scanning microendoscope capable multiphoton microscope. Scale bars, 10 μm. Comparable amounts of two-
photon excited fluorescence at the sample per frame were maintained in
of acquiring TPF and SHG images at 4.1 frames∕s with a FOVxy
each system. The displayed images were acquired with 800-nm excitation
of 110 μm × 110 μm. The uniform pixel dwell time across this and a FOVxy of approximately 110 μm × 110 μm.
field of view enables our prototype to obtain ex vivo tissue images
solely from intrinsic fluorescence and SHG signal with minimal
within the tissue. Image acquisition is done with the multiphoton microscope
frame averaging. Though further experimentation is necessary to software MPScan (30) and stored as MPView files. MPView files are then ex-
confirm the probe’s effectiveness as a cancer diagnostic tool, the ported to Matlab and the public domain program Image J for data analysis
ability of our device to acquire subcellular resolution, label-free and image processing.
images at fast image acquisition rates shows great potential for
Downloaded from https://www.pnas.org by 111.249.51.128 on October 29, 2023 from IP address 111.249.51.128.

translating our endoscope into a minimally invasive probe cap- Mouse Sample Processing. Adult CD-1 mice were obtained from Charles River
able of real-time in vivo imaging in the clinical setting. Laboratories (Charles River Laboratories International, Inc.) and euthanized
by carbon dioxide asphyxiation. The organs of interest (a lung lobe and a
Materials and Methods segment of the colon) were dissected, visually inspected for abnormalities,
Endoscopic Imaging Setup. In this setup (Fig. 1D), a mode-locked Ti:Sapphire and kept in chilled PBS until imaging. A normal-appearing deflated lung
laser (MaiTai, Newport) is used as the excitation source. The femtosecond lobe was selected from a mouse and imaged at various depths, sampling
excitation pulses at 800-nm wavelength are precompensated for fiber disper-
the respiratory zone of the lung tissue (intact alveoli, see Fig. 5B). A short
sion by using a rotating cylindrical lens and grating (28) prior to being
(approximately 0.5–1 cm) segment of the mouse colon was obtained and
coupled into the core of the 1-m-long double-clad optical fiber. The distal
cut open to allow direct imaging of the inner (epithelial) lining of the colon
end of the DCF is mounted onto our resonant/nonresonant scanner, which
sweeps the fiber tip in a raster pattern across the back aperture of the multi- (see Fig. 5C). In addition, collagen fibers were removed from a mouse tail (see
element GRIN lens assembly. The resonant axis is scanned at a frequency of Fig. 5A). Each imaging sample was embedded in agarose gel, plated on a
1.05 kHz, and the nonresonant axis is scanned at 4.1 frames∕s, enabling fast standard glass microscope slide, kept immersed in PBS, and imaged within
2D lateral scanning. The GRIN assembly functions as a 0.8-N.A. water immer- 1 h of euthanasia. All animal housing and experimentation was performed
sion objective with a measured 135-μm working distance and provides a 4 in accordance with the animal welfare guidelines of the Institutional Animal
.8× demagnification (Fig. 1C), see SI Text for additional details of the GRIN Care and Use Committee.
assembly working distance characterization. For TPF and SHG imaging, the
excited signal is epi-collected through the GRIN assembly to the core and in- ACKNOWLEDGMENTS. We thank members of the Xu and Webb research
ner cladding of the DCF. Signal transmitted through the DCF is then reflected groups for discussions and technical suggestions. We thank Dr. Douglas
by a 705-nm dichroic beam splitter (FF705-Di01-25 × 36, Semrock) and deliv- Scherr and Dr. Sushmita Mukherjee of Weill Cornell Medical College for
ered to a two-channel detection system composed of two ultra bi-alkali valuable discussions and suggestions. We also thank Charlotte E. Egan, PhD,
(R7600U-200, Hamamatsu) photomultiplier tubes (PMTs). A 405-nm dichroic Research Associate, Cornell University, Department of Microbiology and
beam splitter (Di01-R405-25 × 36, Semrock) is positioned in the emission path Immunology for helping identify anatomical structures in the ex vivo tissue
at 45° between the two PMTs to separate the SHG and autofluorescence images. Our project was supported by National Institutes of Health/National
imaging channels. The ex vivo images were acquired by rigidly mounting Cancer Institute Grant R01-CA133148 and National Institutes of Health/
the endoscope, mounting tissue samples on a microscope stage, and manip- National Institute of Biomedical Imaging and Bioengineering Grant R01-
ulating the stage position at 5 μm to acquire images at different axial depths EB006736, “Development of Medical Multiphoton Microscopic Endoscopy.”

1. Zipfel WR, et al. (2003) Live tissue intrinsic emission microscopy using multiphoton- 6. Skala MC, et al. (2005) Multiphoton microscopy of endogenous fluorescence differ-
excited native fluorescence and second harmonic generation. Proc Natl Acad Sci entiates normal, precancerous, and cancerous squamous epithelial tissues. Cancer
USA 100:7075–7080. Res 65:1180–1186.
2. Mukherjee S, et al. (2009) Human bladder cancer diagnosis using Multiphoton micro- 7. Lin SJ, et al. (2006) Discrimination of basal cell carcinoma from normal dermal stroma
scopy. Proc Soc Photo Opt Instrum Eng 7161:716117-1–716117-10. by quantitative multiphoton imaging. Opt Lett 31:2756–2758.
3. Pavlova I, et al. (2010) Multiphoton microscopy as a diagnostic imaging modality for 8. Gu M, Bao HC, Li JL (2010) Cancer-cell microsurgery using nonlinear optical endo-
lung cancer. Proc Soc Photo Opt Instrum Eng 7569:756918-1–756918-7. microscopy. J Biomed Opt 15:050502-1–050502-3.
4. Wang CC, et al. (2009) Differentiation of normal and cancerous lung tissues by multi- 9. Huang X, El-Sayed IH, Qian W, El-Sayed MA (2006) Cancer cell imaging and photo-
photon imaging. J Biomed Opt 14:044034-1–044034-4. thermal therapy in the near-infrared region by using gold nanorods. J Am Chem
5. Wilder-Smith P, et al. (2004) In vivo multiphoton fluorescence imaging: A novel Soc 128:2115–2120.
approach to oral malignancy. Lasers Surg Med 35:96–103.

17602 ∣ www.pnas.org/cgi/doi/10.1073/pnas.1114746108 Rivera et al.


10. Helmchen F, Fee MS, Tank DW, Denk W (2001) A miniature head-mounted neuro- 20. Dickensheets D, Kino GS (1996) Micromachined scanning confocal optical microscope.
technique two-photon microscope: High-resolution brain imaging in freely moving Opt Lett 21:764–766.
animals. Neuron 31:903–912. 21. Hofmann U, et al. (1999) Electrostatically driven micromirrors for a miniaturized con-
11. Wu T, Ding Z, Wang K, Chen M, Wang C (2009) Two-dimensional scanning realized by focal laser scanning microscope. Proc Soc Photo Opt Instrum Eng 3878:29–38.
an asymmetry fiber canilever driven by single piezo bender actuator for optical coher- 22. Piyawattanametha W, et al. (2009) In vivo brain imaging using a portable 2.9 g
ence tomography. Opt Express 17:13819–13829. two-photon microscope based on a microelectromechanical systems scanning mirror.
12. Seibel EJ, Smithwick QYJ (2002) Unique features of optical scanning, single fiber en- Opt Lett 34:2309–2311.
doscopy. Lasers Surg Med 30:177–183. 23. Fu L, Jain A, Cranfield C, Xie H, Gu M (2007) Three-dimensional nonlinear optical en-
13. Myaing MT, MacDonald DJ, Li X (2006) Fiber-optic scanning two-photon fluorescence doscopy. J Biomed Opt 12:040501-1–040501-3.
24. Piyawattanametha W, et al. (2006) Fast-scanning two-photon fluorescence imaging
endoscope. Opt Lett 31:1076–1078.
based on a microelectromechanical systems two-dimensional scanning mirror. Opt Lett
14. Engelbrecht CJ, Johnston RS, Seibel EJ, Helmchen F (2008) Ultra-compact fiber-optic
31:2018–2020.
two-photon microscope for functional fluorescence imaging in vivo. Opt Express
25. Tang S, et al. (2009) Design and implementation of fiber-based multiphoton endo-
16:5556–5564.
scopy with microelectromechanical systems scanning. J Biomed Opt 14:034005-1–
15. Wu Y, Leng Y, Xi J, Li X (2009) Scanning all-fiber-optic endomicroscopy system for 3D
034005-7.
nonlinear optical imaging of biological tissues. Opt Express 17:7907–7915.
26. Kinsler LE, Frey AR, Coppens AB, Sanders JV (1982) Fundamentals of Acoustics (Wiley,
16. Hendriks BHW, et al. (2011) High-resolution resonant and nonresonant fiber-scanning New York), 3rd Ed.
confocal microscope. J Biomed Opt 16:026007-1–026007-8. 27. Durst ME, Kobat D, Xu C (2009) Tunable dispersion compensation by a rotating cylind-
17. Sawinski J, Denk W (2007) Miniature random-access fiber scanner for in vivo multipho- rical lens. Opt Lett 34:1195–1197.
ton imaging. J Appl Phys 102:034701-1–034701-8. 28. Boyd RW (1992) Nonlinear Optics (Academic, San Diego), pp 274–280.
18. Goetz M, et al. (2007) In-vivo confocal real-time mini-microscopy in animal models of 29. Dela Cruz JM, McMullen JD, Williams RM, Zipfel WR (2010) Feasibility of using multi-
human inflammatory and neoplastic diseases. Endoscopy 39:350–356. photon excited tissue autofluorescence for in vivo human histopathology. Biomed Opt
19. Le Harzic R, Weinigel M, Riemann I, Konig K, Messerschmidt B (2008) Nonlinear optical Express 1:1320–1330.
endoscope based on a compact two axes piezo scanner and a miniature objective lens. 30. Nguyen QT, Tsai PS, Kleinfeld D (2006) MPScope: A versatile software suite for multi-
Opt Express 16:20588–20596. photon microscopy. J Neurosci Methods 156:351–9.

APPLIED PHYSICAL
SCIENCES
Downloaded from https://www.pnas.org by 111.249.51.128 on October 29, 2023 from IP address 111.249.51.128.

Rivera et al. PNAS ∣ October 25, 2011 ∣ vol. 108 ∣ no. 43 ∣ 17603

You might also like