Vous êtes sur la page 1sur 9

Broadband miniature fiber optic ultrasound

generator
Xiaotian Zou,1 Nan Wu,2 Ye Tian,2 and Xingwei Wang1,2,*
1
Department of Biomedical Engineering and Biotechnology, University of Massachusetts, 1 University Ave., MA,
Lowell, 01854, USA
2
Department of Electrical and Computer Engineering, University of Massachusetts, 1 University Ave., MA, Lowell,
01854, USA
*
Xingwei_Wang@uml.edu

Abstract: This paper presents the design, fabrication and characterization


of a broadband miniature fiber optic ultrasound generator based on
photoacoustic (PA) ultrasound generation principle for biomedical
ultrasound imaging and ultrasound non-destructive test (NDT) applications.
A novel PA generation material, gold nanocomposite, was synthesized by
directly reducing gold nanoparticles within polydimethylsiloxane (PDMS)
through a one-pot protocol. The fiber optic ultrasound generator was
fabricated by coating the gold nanocomposite on the tip of the optical fiber.
The efficiency of the PA generation using gold nanocomposite was
increased 105 compared to using aluminum thin film and 103 compared to
using graphite mixed within epoxy. The ultrasound profile and the acoustic
distribution have been characterized. The amplitude of the generated
ultrasound signal was as high as 0.64 MPa and the bandwidth was more
than 20 MHz. This paper also demonstrated its capability for ultrasound
imaging of a tissue specimen.
©2014 Optical Society of America
OCIS codes: (060.2310) Fiber optics; (060.2350) Fiber optics imaging; (110.7170) Ultrasound.

References and links


1. A. Baerwald, S. Dauk, R. Kanthan, and J. Singh, “Use of ultrasound biomicroscopy to image human ovaries in
vitro,” Ultrasound Obstet. Gynecol. 34(2), 201–207 (2009).
2. A. J. Hunter, B. W. Drinkwater, and P. D. Wilcox, “Autofocusing ultrasonic imagery for non-destructive testing
and evaluation of specimens with complicated geometries,” NDT Int. 43(2), 78–85 (2010).
3. G. Sposito, C. Ward, P. Cawley, P. B. Nagy, and C. Scruby, “A review of non-destructive techniques for the
detection of creep damage in power plant steels,” NDT Int. 43(7), 555–567 (2010).
4. F. S. Foster, J. Mehi, M. Lukacs, D. Hirson, C. White, C. Chaggares, and A. Needles, “A new 15-50 MHz array-
based micro-ultrasound scanner for preclinical imaging,” Ultrasound Med. Biol. 35(10), 1700–1708 (2009).
5. B. Jadidian, N. M. Hagh, A. A. Winder, and A. Safari, “25 MHz ultrasonic transducers with lead-free
piezoceramic, 1-3 PZT fiber-epoxy composite, and PVDF polymer active elements,” IEEE Trans. Ultrason.
Ferroelectr. Freq. Control 56(2), 368–378 (2009).
6. K. A. Snook, C. H. Hu, T. R. Shrout, and K. K. Shung, “High-frequency ultrasound annular-array imaging. Part
I: array design and fabrication,” IEEE Trans. Ultrason. Ferroelectr. Freq. Control 53(2), 300–308 (2006).
7. E. J. Gottlieb, J. M. Cannata, C. H. Hu, and K. K. Shung, “Development of a high-frequency (> 50 MHz)
copolymer annular-array, ultrasound transducer,” IEEE Trans. Ultrason. Ferroelectr. Freq. Control 53(5), 1037–
1045 (2006).
8. X. Zou, N. Wu, Y. Tian, Y. Zhang, and X. Wang, “Polydimethylsiloxane thin film characterization using all-
optical photoacoustic mechanism,” Appl. Opt. 52(25), 6239–6244 (2013).
9. Y. Tian, N. Wu, X. Zou, H. Felemban, C. Cao, and X. Wang, “Fiber-optic ultrasound generator using periodic
gold nanopores fabricated by a focused ion beam,” Opt. Eng. 52(6), 065005 (2013).
10. E. Biagi, F. Margheri, and D. Menichelli, “Efficient laser-ultrasound generation by using heavily absorbing films
as targets,” IEEE Trans. Ultrason. Ferroelectr. Freq. Control 48(6), 1669–1680 (2001).
11. Y. Hou, J.-S. Kim, S. Ashkenazi, S.-W. Huang, L. J. Guo, and M. O’Donnell, “Broadband all-optical ultrasound
transducers,” Appl. Phys. Lett. 91(7), 073507 (2007).
12. N. Wu, Y. Tian, X. Zou, V. Silva, A. Chery, and X. Wang, “High-efficiency optical ultrasound generation using
one-pot synthesized polydimethylsiloxane-gold nanoparticle nanocomposite,” J. Opt. Soc. Am. B 29(8), 2016–
2020 (2012).

#212251 - $15.00 USD Received 20 May 2014; revised 4 Jul 2014; accepted 8 Jul 2014; published 18 Jul 2014
(C) 2014 OSA 28 July 2014 | Vol. 22, No. 15 | DOI:10.1364/OE.22.018119 | OPTICS EXPRESS 18119
13. Y. Hou, J. S. Kim, S. W. Huang, S. Ashkenazi, L. J. Guo, and M. O’Donnell, “Characterization of a broadband
all-optical ultrasound transducer-from optical and acoustical properties to imaging,” IEEE Trans. Ultrason.
Ferroelectr. Freq. Control 55(8), 1867–1877 (2008).
14. H. Won Baac, J. G. Ok, H. J. Park, T. Ling, S.-L. Chen, A. J. Hart, and L. J. Guo, “Carbon nanotube composite
optoacoustic transmitters for strong and high frequency ultrasound generation,” Appl. Phys. Lett. 97(23), 234104
(2010).
15. T. Buma, M. Spisar, and M. O’Donnell, “A high-frequency, 2-D array element using thermoelastic expansion in
PDMS,” IEEE Trans. Ultrason. Ferroelectr. Freq. Control 50(9), 1161–1176 (2003).
16. H. W. Baac, J. G. Ok, A. Maxwell, K.-T. Lee, Y.-C. Chen, A. J. Hart, Z. Xu, E. Yoon, and L. J. Guo, “Carbon-
nanotube optoacoustic lens for focused ultrasound generation and high-precision targeted therapy,” Sci. Rep. 2,
989 (2012).
17. P. K. Jain, K. S. Lee, I. H. El-Sayed, and M. A. El-Sayed, “Calculated absorption and scattering properties of
gold nanoparticles of different size, shape, and composition: applications in biological imaging and
biomedicine,” J. Phys. Chem. B 110(14), 7238–7248 (2006).
18. D. Ryu, K. J. Loh, R. Ireland, M. Karimzada, F. Yaghmaie, and A. M. Gusman, “In situ reduction of gold
nanoparticles in PDMS matrices and applications for large strain sensing,” Smart Struct. Syst. 8(5), 471–486
(2011).
19. K. Seshan, Handbook of Thin Film Deposition (William Andrew, 2012).
20. V. Pathak, V. Singh, and Y. Sanjay, “Ultrasound as a modern tool for carcass evaluation and meat processing: A
review,” Int. J. Meat Sci. 1(2), 83–92 (2011).

1. Introduction
Various studies of ultrasound generators have been conducted to meet the rising challenges of
advanced ultrasound applications such as biomedical ultrasound imaging and ultrasound non-
destructive test (NDT) [1–3]. Other than conventional piezoelectric ultrasound generators [4–
7], optical ultrasound generators, especially fiber optic ultrasound generators are attractive
candidates for many advanced ultrasound applications [1, 8, 9]. The most commonly used
mechanism to generate ultrasound signals optically is the photoacoustic (PA) principle, which
is a wide bandwidth ultrasound generation method because the pulse width of the ultrasound
can be tailored by ultra-fast lasers [10, 11]. By taking advantages of PA principle and optical
fibers, novel fiber optic ultrasound generators featuring wide bandwidth and compact size for
biomedical imaging and NDT applications in limited spaces can be achieved.
The PA principle is an optical approach to generate ultrasound signals [12]. It involves a
PA generation material which absorbs the optical energy from the laser and converts it into
localized temperature rise. The localized temperature rise will cause the expansion of the PA
material due to the thermal expansion effect. The PA material will contract when the laser is
shut off. Therefore, the expansion/contraction cycle will generate mechanical waves which
are acoustic signals. The most significant advantage of the PA principle is that the profile of
the acoustic signals is similar to the profile of the laser, which means that the pulse width of
the acoustic signals can be tailored by the laser beam. Ultra-fast lasers, such as nanosecond
lasers, have been commercially available. Therefore, the pulse width of the generated acoustic
signal can be very short (generally in nanoseconds), which leads to a wide bandwidth of
acoustic signals [10, 11, 13, 14].
The key factor of the PA generation principle is the PA generation material which absorbs
and converts the optical energy to heat and then acoustic signal. The performances of the PA
generation, such as the energy conversion efficiency and the bandwidth, rely on the PA
generation material. An ideal PA generation material should feature a high optical energy
absorption capability and a high coefficient of thermal expansion (CTE). Recently, various
studies have been exerted on developing novel materials to increase the efficiency of the PA
generation and it has been reported that materials based on polymer show higher PA
generation efficiency than metallic materials [10–15]. Graphite mixed within epoxy and
graphite mixed within PDMS were reported by Biagi’s and Buma’s group, respectively. By
mixing graphite within epoxy, Biagi’s group reported that the efficiency of the PA generation
was increased 2 orders of magnitude compared to the thin aluminum film [10]. By replacing
graphite with gold nanostructure, Hou’s group reported a broad bandwidth optical ultrasound
transducer [11, 13]. Two dimensional gold nanostructures were fabricated by nanoimprint

#212251 - $15.00 USD Received 20 May 2014; revised 4 Jul 2014; accepted 8 Jul 2014; published 18 Jul 2014
(C) 2014 OSA 28 July 2014 | Vol. 22, No. 15 | DOI:10.1364/OE.22.018119 | OPTICS EXPRESS 18120
technique and a layer of PDMS was coated above the gold nanostructure. Recently, Baac’s
work used another material, carbon nanotube composite to optically generate ultrasound
signals [14, 16].
The optical energy absorption capability of the PA generation material can be further
improved by applying noble metal nanoparticles due to their high optical energy absorption
capabilities at the plasmon resonant frequencies. It has been proved that gold nanoparticles
(Au NPs) show the maximum optical absorption energy at the wavelength of 520 nm when
the diameter is around 20 nm [17]. Therefore, the PA generation efficiency can be improved
by applying Au NPs.
In this paper, a novel fiber optic PA ultrasound generator based on gold nanocomposite
was designed, fabricated, and characterized. The fabrication process of fiber optic PA
generator was very easy to operate with relatively low cost. The gold nanocomposite was
achieved by directly reducing gold nanoparticles (Au NPs) within Polydimethylsiloxane
(PDMS) and was coated on the tip of optical fibers to generate strong ultrasound signals. The
gold nanocomposite was synthesized by a one-pot protocol [18] and it has been demonstrated
that such material features high optical energy absorption capability which makes it an
excellent material for PA generation applications [12].
This paper is organized as follows. Section 2 presents the design and fabrication
procedure of a miniature fiber optic ultrasound generator based on the gold nanocomposite.
Section 3 describes the ultrasonic pulse generation experiment using proposed generator,
based on the experimental results, the PA generation efficiency was determined. In section 4,
the ultrasonic field distribution test was characterized; the ultrasound attenuation coefficient
and directivity angle range was calculated based on the experimental findings. In section 5 of
this paper, an ultrasound image of a tissue specimen was obtained by the proposed generator,
and Section 6 concludes the paper. In summary, all those experimental results proved that the
fiber optic ultrasound generator with high energy conversion efficiency and wide bandwidth
could be used in biomedical imaging and NDT applications.
2. Methodology
2.1 Gold nanocomposite
The gold nanocomposite was prepared by following a one-pot protocol by directly mixing the
gold salt (HAuCl4·3H2O) into PDMS [12, 18]. Au NPs were reduced from the gold salt by
PDMS molecules. Briefly, the PDMS matrix was prepared by mixing the base and the curing
agent with a weight ratio of 10:3. A certain weight of the gold salt was crushed into powder
and was mixed within the prepared PDMS matrix. The gold salt/PDMS mixture was subject
to an ultrasonic bath for 30 min in ice water and then the mixture was degassed for
approximately 20 min in a vacuum chamber. During the ultrasonic bath, the color of the gold
salt/PDMS mixture turned to ruby red indicating that Au NPs with a diameter of around 20
nm were reduced and the mixture was turned into gold nanocomposite, which is consistent
with the previous study [12]. The concentration of Au NPs within the PDMS can be adjusted
by different amount of the gold salt.
The absorption spectra of the gold nanocomposite material have been presented in our
previous study [12]. Because the peak optical energy absorption wavelength of gold
nanocomposite material is at 530 nm, the optical energy absorption capability is maximized
when the wavelength of the optical irradiation is tuned at the similar wavelength. In this
paper, the wavelength of the laser is at 532 nm. Therefore, the optical energy absorption can
be maximized.
2.2 Fiber optic ultrasound generator
The fiber optic ultrasound generator was fabricated by applying the gold nanocomposite on
the tip of an optical fiber. The generator structure is illustrated in Fig. 1. A piece of multi-

#212251 - $15.00 USD Received 20 May 2014; revised 4 Jul 2014; accepted 8 Jul 2014; published 18 Jul 2014
(C) 2014 OSA 28 July 2014 | Vol. 22, No. 15 | DOI:10.1364/OE.22.018119 | OPTICS EXPRESS 18121
mode fiber (MMF) with a core diameter of 400 µm was stripped, cleaved, and the end face
was polished. The MMF was dipped into the gold nanocomposite perpendicularly and was
pulled out of the gold nanocomposite slowly. Due to the high viscosity of the gold
nanocomposite, the gold nanocomposite will be attached on the end face of the MMF.
Finally, the MMF was mounted at about 5 mm above a hot plate (set at 120 þC) while
maintaining the perpendicular position overnight to cure the gold nanocomposite.

Fiber cladding
Fiber core
Gold nanocomposite

Multi-mode fiber Gold nanocomposite


Fig. 1. The structure of the fiber optic ultrasound generator.

The microscopic picture of the fiber optic ultrasound generator coated with the gold
nanocomposite is illustrated in Fig. 2. The magnitude of the microscope was 100. The
concentration of the gold salt in the gold salt/PDMS mixture was 7.58% by weight. Due to the
surface tension of the gold nanocomposite, the gold nanocomposite formed a spherical shape
on the tip of the optical fiber. By calculating the ratio between the diameter of the optical
fiber and the thickness of the gold nanocomposite, the thickness of the gold nanocomposite
was approximately 105 μm at the thickest part.

Gold nanocomposite

105 µm
Multi-mode fiber

Fig. 2. The microscopic picture of the fiber optic ultrasound generator.

3. Ultrasonic pulse generation verification


3.1 Experimental setup
An ultrasound pulse generation experiment was performed to evaluate the performance of the
broadband fiber optic ultrasound generator. Figure 3(a) and Fig. 3(b) shows the schematic
diagram and the experimental setup, respectively. Experiments were conducted under the
water. The optical irradiation source was a 532 nm Nd:YLF nanosecond laser (Surelite-I-10,
Continuum) with a pulse width of 5 ns and a repetition rate of 10 Hz. The laser beam was
coupled into the fiber optic ultrasound generator through a coupler (F810SMA-543,
Thorlabs). In Fig. 3(c), a hydrophone (HGL-0200, Onda) with an aperture size of 200 µm was
utilized as a receiver 1 mm away from generator to collect the ultrasound signals. Once the
laser source was radiated, a trigger signal was sent out from the laser system to trigger a data
acquisition card (DAQ) (M2i.4032, Spectrum) with a sampling rate of 50 MHz. The

#212251 - $15.00 USD Received 20 May 2014; revised 4 Jul 2014; accepted 8 Jul 2014; published 18 Jul 2014
(C) 2014 OSA 28 July 2014 | Vol. 22, No. 15 | DOI:10.1364/OE.22.018119 | OPTICS EXPRESS 18122
ultrasound signal was collected by the hydrophone and the signal was transferred to the DAQ
for process and analysis.

(a)

Hydrophone

1 mm
(c)

Fiber optic
ultrasound generator
(b)
Fig. 3. Experimental setup for the ultrasonic pulse generation and ultrasonic field distribution.
(a) Schematic diagram of the experimental setup. (b) The photo of the experimental setup. (c)
Zoomed in photo to illustrate the distance between the generator and the hydrophone.

3.2 Results and discussions


Figure 4(a) shows a typical ultrasound signal that was generated by the fiber optic ultrasound
generator. The peak to peak amplitude of the ultrasonic pressure was measured to be 0.64
MPa and the distance between the hydrophone and the generator tip was approximately 1
mm. The pulse width was measured to be 160 ns. After performing the Fourier transform, the
frequency domain of the generated ultrasound signal is illustrated in Fig. 4(b) (The 0 dB
refers to the total signal power). The bandwidth of the signal was at least 20 MHz. It is
interesting to note that, the thickness of the gold nanocomposite will affect the bandwidth of
the generator [13, 14]. As the ultrasound propagated along the material, high frequency
components attenuate faster than low frequency components. Therefore, the extra thickness of
the gold nanocomposite may attenuate high frequency components of the generated
ultrasound. By taking the advantage of nanofabrication method [19], the fiber optic
ultrasound generator with the ultra-thin gold nanocomposite layer can be fabricated to achieve
the higher ultrasound bandwidth.

#212251 - $15.00 USD Received 20 May 2014; revised 4 Jul 2014; accepted 8 Jul 2014; published 18 Jul 2014
(C) 2014 OSA 28 July 2014 | Vol. 22, No. 15 | DOI:10.1364/OE.22.018119 | OPTICS EXPRESS 18123
0
0.4

Amplitude (MPa)
0.2 -20

Power (dB)
0.0 -40

-0.2
-60
(a) (b)
0.0 0.5 1.0 1.5 2.0 0 5 10 15 20
Time (μs) Frequency (MHz)

Fig. 4. Ultrasound signal generated by the fiber optic ultrasound generator. (a) The profile of a
typical generated ultrasound signal. (b) The frequency domain of the generated ultrasound
signal.

The efficiency of the PA generation can be described by the following equation [10]:
Ea
η= , (1)
Eoptical

where Eoptical is the energy of the laser and Ea is the energy of the ultrasound signal,
respectively. Ea can be estimated by the following equation [10]:
c ∞
Ea ≅ A p 2 (t )dt , (2)
B 0

where c is the sound velocity in the water; B is the bulk modulus of the water; A is the spot
area; p is the acoustic pressure.
The energy of the laser emitting the optical fiber Eoptical was measured after laser passing
through a bare polished MMF with a core diameter of 400 µm, which was 11 μJ/pulse (laser
fluence = 8.75 mJ/cm2/pulse).
The energy of the ultrasound signal is Ea = 1.92 nJ via Eq. (2). Therefore, the efficiency of
the PA generation was determined as 0.18 × 10−3. The efficiency was approximately 5 orders
of magnitude increased comparing to the PA generation efficiency by using aluminum thin
film, approximately 103 times increased comparing to using graphite mixed within epoxy
[10]. From [16], the PA transmitter using carbon nanotube composite generated the high
frequency ultrasound signal with generation efficiency of 1.4 × 10−3 (laser fluence = 42.4
mJ/cm2/pulse). The author’s PA generation efficiency was lower in one order of magnitude
comparing to using carbon nanotube composite. This is because the high frequency
component in the ultrasound signal was attenuated by the extra thickness of the gold
nanocomposite film.
4. Ultrasonic field distribution
4.1 Experimental setup
In this section we presented the experimental activity which led to better understanding of the
fiber optic ultrasonic source by characterizing its energy distribution. The ultrasonic field
produced by the fiber optic ultrasound generator obeys the physical laws of wave propagation
and it can be simply divided into many small PA point sources attached on the fiber tip, and
thus producing an interference pattern at any position in the field. The experimental setup was
similar to the ultrasonic pulse generation test. The same optical irradiation source was utilized
and the experiment was also performed under the water media. In addition, the hydrophone
was mounted on a 2-axis stepper motor stage (NRT 100, Thorlabs) to provide accurate

#212251 - $15.00 USD Received 20 May 2014; revised 4 Jul 2014; accepted 8 Jul 2014; published 18 Jul 2014
(C) 2014 OSA 28 July 2014 | Vol. 22, No. 15 | DOI:10.1364/OE.22.018119 | OPTICS EXPRESS 18124
scanning capability during the test. For every scanned position, the peak to peak ultrasonic
amplitude was recorded by DAQ after averaged 50 times. Figure 3(b) indicates the scanning
orientation.
4.2 Results and discussions
A longitudinal section of the ultrasonic field distribution was measured and the results are
illustrated in Fig. 5. The ultrasonic field was acquired within a rectangular area (5.0 mm by
4.0 mm) with the resolution of 0.1 mm by the scanning hydrophone. In Fig. 5(a), the color
map represented the ultrasonic pressure. It was noted that the fiber optic ultrasound generator
was placed at 0 mm of lateral direction and 0 mm of axial direction. In the contour map, the
focal point pressure was found to be 0.78 MPa within the position coordinate (0 mm, 1.2
mm). Moreover, at the position coordinate (0 mm, 1 mm), the pressure was found to be 0.64
MPa. This was in good agreement with the result from the ultrasonic pulse generation
experiment.
5 0 5
-40

Normalized Magnitude (dB)


Axial Position (mm)

Axial Position (mm)


4 4

Amptitude (MPa)
0.16
-30

3 0.31 3
-20
2 0.47 2

1.2 mm Focal point -10


1 0.62 1
(a) (b) 0 mm
0 0.78 0 0
-2 -1 0 1 2 -2 -1 0 1 2
Lateral Position (mm) Lateral Position (mm)
Fig. 5. Ultrasonic field distribution in a longitudinal section generated from fiber optic
ultrasound generator. (a) Pressure distribution of ultrasonic field. (b) Normalized magnitude
distribution of ultrasonic field.

3
Lateral Position : 0 mm Axial Position : 1.2 mm
0 0 mm
Normalized Magnitude (dB)

Normalized Magnitude (dB)

0
-0.27 mm 0.25 mm -6 dB
1.2 mm
-3 -8
-6 dB
-6
0.15 mm -16
2.25 mm
-9
-24
-12

-15 -32
(a) (b)
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0 -2 -1 0 1 2

Axial Position (mm) Lateral Position (mm)


Fig. 6. Extracted from Fig. 5(b): pressure distribution along the both axial direction and lateral
position from the focal point (0 mm, 1.2 mm). (a) Pressure distribution along the axial
direction. (b) Pressure distribution along the lateral direction.

In Fig. 5(b), the color map represented the normalized magnitude in decibel scale. Figure
6 was extracted from Fig. 5(b) to show the pressure distribution in the cross section along
both the axial direction and the lateral direction from the focal point (0 mm, 1.2 mm). From
Fig. 6, the focal area at −6 dB was approximately 0.52 mm by 2.10 mm.

#212251 - $15.00 USD Received 20 May 2014; revised 4 Jul 2014; accepted 8 Jul 2014; published 18 Jul 2014
(C) 2014 OSA 28 July 2014 | Vol. 22, No. 15 | DOI:10.1364/OE.22.018119 | OPTICS EXPRESS 18125
5. Ultrasound imaging
5.1 Experimental setup
To further characterize this fiber optic ultrasound generator, its ultrasound imaging
capabilities must be demonstrated. In this section, we present the experimental activity, in
which the fiber optic generator was used to image a tissue specimen. The photo of the
experimental setup is shown in Fig. 7. This experiment was performed under the water media.
The same laser was used as the optical radiation source. The specimen holder was attached
with the 2-axis stepper motor stage to provide accurate scanning capability. The hydrophone
was fixed and placed at the other side of the specimen holder. For every scanned pixel, the
ultrasonic pulse hit the specimen, penetrated through the specimen and the data of that pixel
was recorded by hydrophone.

Specimen holder
mounted on a 2-axis
stepper motor stage
Hydrophone
Tissue specimen

Fiber optic
ultrasound generator
Fig. 7. The photo of the ultrasound imaging experimental setup.

5.2 Results and discussions


In this experiment, the fiber optic ultrasound generator and hydrophone operated in a
transmission C-mode. The test specimen was constituted by a slice of pork soft tissue with the
thickness of 1 mm. The ultrasound imaging was obtained by incrementally moving the
specimen in between the fixed generator and hydrophone, while the ultrasonic wave
propagation time between the generator and the hydrophone was measured at each point.
Therefore, the speed of sound of each point in the tissue was determined and mapped into the
contour figure pixel by pixel using the following equation:
hc
v= (3)
(t − tw )c + h

where v is the speed of sound in sample, h = 1 mm is the thickness of sample, c = 1540 ms−1
is the speed of sound in water, t is the measured propagation times from generator to
hydrophone, and tw is the propagation time of the standard propagation path with same length
when there is only water between generator and hydrophone. The ultrasound image
experiment results and the photo of specimen were illustrated in Fig. 8. The resolution of
ultrasonic image was 200 µm. It was noted from [20], the ultrasound waves travel more
quickly through muscle tissues than fat tissues, therefore the measured ultrasound propagation
times through a specimen provides an indication of the tissue’s composition. Compared
between Figs. 8(a) and 8(b), it can be clearly observed that the relative propagation time at
each point (which is inversely proportional to the speed of sound) is strongly correlated to the
muscular-to-fat ratio of tissue at that point. This observation was consisted with reference
[20], which suggests the promise of C-mode ultrasound imaging for our system.

#212251 - $15.00 USD Received 20 May 2014; revised 4 Jul 2014; accepted 8 Jul 2014; published 18 Jul 2014
(C) 2014 OSA 28 July 2014 | Vol. 22, No. 15 | DOI:10.1364/OE.22.018119 | OPTICS EXPRESS 18126
For high-resolution biomedical ultrasound imaging, one difficulty in this research is the
hard-to-controlled thickness of the gold nanocomposite film on the tip of the optical fiber.
Thickness of the gold nanocomposite film affects the bandwidth of the generator. As the
ultrasound propagate along the material, high frequency components attenuate faster than low
frequency components. Therefore, the extra thickness of the gold nanocomposite may
attenuate high frequency components of the generated ultrasound. In order to accomplish the
high-frequency (> 30 MHz) and ultrahigh-frequency (> 100 MHz) biomedical ultrasound
image, future studies in this research will focus on the fabrication of ultra-thin gold
nanocomposite film layer by taking the advantage of nanofabrication method, e.g., Focused
Ion Beam (FIB) milling [9]. In addition, a picosecond laser or a femtosecond laser system
will be utilized to further improve the pulsed laser source and tailor the bandwidth of the
generator.
8 8
1600
7 7

Speed of Sound (m/s)


1568 6
6

5 5
1536

Y (mm)
Y (mm)

4 4

3 1504 3

2 2
1472
1 1

0 1440 0
1 2 3 4 5 6 7 8 1 2 3 4 5 6 7 8
(a) (b)
X (mm) X (mm)

Fig. 8. The ultrasound imaging of a slice of pork tissue: the ultrasound imaging is obtained by
moving the specimen in between the fixed generator and hydrophone. (a) The ultrasound
image of a slice of pork tissue. (b) Photo of the tissue specimen (slice of pork tissue).

6. Conclusions
In this paper, we have designed, fabricated, and characterized the first fiber optic ultrasound
generator based on PA generation technique by using gold nanocomposite as the ultrasound
generation material. An optical fiber with a core diameter of 400 μm was coated with the gold
nanocomposite. The verification experiment was performed to validate the ultrasound
generation capability. The experimental results showed that ultrasound signals with an
amplitude of 0.64 MPa was generated by the fiber optic ultrasound generator and bandwidth
was more than 20 MHz. The PA generation efficiency was approximately 5 orders of
magnitude increased comparing to using aluminum thin film and 103 times increased
comparing to using graphite mixed within epoxy.
The ultrasonic field distribution was scanned by a hydrophone attached with 2-axis
stepper motor stage. The focal point was approximately 1.2 mm away from the generator with
the pressure of 0.78 MPa. Moreover, the first ultrasound image of a tissue specimen was
obtained with the resolution of 200 µm by our proposed generator. In summary, the fiber
optic ultrasound generator could lead to the development of a new generation of ultrasonic
probes featuring high PA efficiency, wide bandwidth, easy fabrication, and miniature size.
Acknowledgments
The authors would like to thank the National Science Foundation for sponsoring this work
(CMMI: 1055358 CAREER).

#212251 - $15.00 USD Received 20 May 2014; revised 4 Jul 2014; accepted 8 Jul 2014; published 18 Jul 2014
(C) 2014 OSA 28 July 2014 | Vol. 22, No. 15 | DOI:10.1364/OE.22.018119 | OPTICS EXPRESS 18127

Vous aimerez peut-être aussi