Vous êtes sur la page 1sur 13

224

ieee transactions on ultrasonics, ferroelectrics, and frequency control, vol. 53, no. 1, january 2006

Development of a 35-MHz Piezo-Composite


Ultrasound Array for Medical Imaging
Jonathan M. Cannata, Member, IEEE, Jay A. Williams, Qifa Zhou, Timothy A. Ritter,
and K. Kirk Shung, Fellow, IEEE
AbstractThis paper discusses the development of a 64element 35-MHz composite ultrasonic array. This array was
designed primarily for ocular imaging applications, and features 2-2 composite elements mechanically diced out of a
ne-grain high-density Navy Type VI ceramic. Array elements were spaced at a 50-micron pitch, interconnected
via a custom exible circuit and matched to the 50-ohm
system electronics via a 75-ohm transmission line coaxial
cable. Elevation focusing was achieved using a cylindrically
shaped epoxy lens. One functional 64-element array was
fabricated and tested. Bandwidths averaging 55%, 23-dB
insertion loss, and crosstalk less than 24 dB were measured. An image of a tungsten wire target phantom was acquired using a synthetic aperture reconstruction algorithm.
The results from this imaging test demonstrate resolution
exceeding 50 m axially and 100 m laterally.

I. Introduction
igh frequency (> 30 MHz) ultrasound is currently
used for various imaging applications in ophthalmology [1][3], dermatology [4], [5], and small animal studies [6], [7]. At present there are several commercial ultrasound systems available for use in the 25-MHz to 50-MHz
frequency range1 . These systems rely upon single-element
transducers that are mechanically scanned in a line or arc
to form an image slice. Array systems, on the other hand,
are desired because they use electronic scanning to do so.
Arrays also lack movable parts that may be hazardous to
patients, can be steered and dynamically focused in the
image plane, and can achieve higher image frame rates.
Unfortunately at the present time commercial array systems are not yet available at frequencies above 30 MHz due
mainly to limitations in fabrication technology and equipment, as well as a lack of quality high-frequency materials
and electronics. Despite these limitations, a few investigators have successfully designed and built high-frequency
ultrasonic arrays. The eorts of these researchers are summarized next.

Manuscript received April 11, 2005; accepted July 15, 2005. The
authors would like thank the National Institutes of Health (NIH) for
providing the funding through grant # P41-EB2182.
J. M. Cannata, J. A. Williams, Q. Zhou, and K. K. Shung are
with the NIH Resource on Medical Ultrasonic Transducer Technology, Department of Biomedical Engineering, University of Southern
California, Los Angeles, CA (e-mail: cannata@usc.edu).
T. A. Ritter is with the U.S. Air Force, Keesler Air Force Base,
Biloxi, MS.
1 Systems include those by VisualSonics, Inc., Toronto, Ontario,
Canada, http://www.visualsonics.com; Ultralink LLC., St. Petersburg, FL, http://www.arcscan.com; and Capistrano Labs, Inc., San
Clemente, CA http://www.capolabs.com.

Lower-frequency arrays typically use mechanical dicing


to separate array elements, whereby elements are cut from
a plate of piezoelectric ceramic or single crystal and backlled with a polymer ller [8]. This technique generally has
been limited to arrays designed to operate in the less-than
20-MHz range. However, in recent years, several studies
have proven that it is a viable option for manufacturing
arrays with operational center frequencies up to 30 MHz
[9][11]. The most sophisticated of the arrays built was the
128-element 30-MHz 1-3 piezo-composite array developed
by Michau et al. [11]. This array was designed to have an
element-to-element spacing, or pitch, of approximately 2
in water or 100 m, with composite posts and array elements both separated using a mechanical dicing saw. Alternately, a 48-element 30-MHz array was developed using
a patented process of bonding thin plates of piezoceramic
with a carefully engineered microsphere-loaded polymer to
create a 2-2 piezo-composite matrix [12], which was then
mechanically diced to form individual array elements [13].
This array was also designed for a 100-m pitch. As an alternative to mechanical dicing, several investigators have
fabricated arrays in the 2030 MHz range using thin sheets
of piezo-polymer materials [14], [15]. Low lateral coupling
and low acoustic impedance for polyvinylidene uoride
(PVDF) and poly(vinylidene uoride-triuoroethylene)
(P(VDF-TrFE)) make these materials good choices for
high-frequency array design. Unfortunately the low capacitance of these materials and the high parasitic capacitance of typical preamplier inputs preclude the use of
suitably sized active elements for high-frequency operation
[16]. Other keress linear array designs incorporating high
dielectric piezoceramics have been evaluated but have yet
to be fabricated at high frequencies [17]. For array development at very high frequencies sol-gel deposition of thin
and thick lms of lead-zirconate-titanate (PZT) are viable
options. Lukacs et al. [18] reported the use of sol-gel deposition of PZT lms in the fabrication of 50200 MHz singleelement transducers and 4060 MHz arrays. Mechanical
separation of elements was achieved using laser dicing technology. Unfortunately the porosity of the lm produced
poor piezoelectric properties, with a thickness-mode coupling coecient (kt ) of lower than 25% and a low relative
clamped dielectric permittivity (S33 < 250). Improvements
in this technique are ongoing, with reported piezoelectric
coupling as high as 50% [19]. Sputtering thin PZT lms is
another option for fabrication of very-high-frequency arrays, but as with sol-gel deposition this method has been
shown to produce reduced piezoelectric properties when

c 2006 IEEE
08853010/$20.00 

cannata et al.: 35-mhz piezo-composite ultrasound array

225

TABLE I
Initial Design Goals for the 35-MHz Array.
Center frequency
Number of elements
Element-to-element spacing (pitch)
Elevation aperture
Elevation focus
Bandwidth (6 dB)
20 dB pulse length
Crosstalk (element-to-element)

35 MHz
64
50 m (1.2)
3 mm
10 mm
> 50%
< 120 ns
<30 dB

compared to bulk PZT [20]. Sputtered ZnO was used to


fabricate a novel 100-MHz array for nondestructive testing
[21]. This array used a sapphire lens for elevation focusing
and a single matching layer for acoustic coupling. Elements
were separated with a 100-m pitch (6.7) by wet etching
through the top chrome/gold electrode as well as the ZnO.
Unfortunately the large element pitch produced (approximately 7 in water at 100 MHz) makes this array dicult
to use in medical imaging applications. In recent years nonpiezoelectric transducers have become a viable alternative
to piezoelectric-based transducer designs. Work conducted
by Buma et al. concluded that optical generation of sound
using the thermoelastic eect is a promising alternative
to piezoelectricity for very-high-frequency two-dimensional
(2-D) arrays [22]. Alternately, capacitive micromachined
ultrasonic transducer (CMUT) fabrication techniques have
been proven eective in fabricating arrays up to 45 MHz
[23]. Lastly, several researchers have proposed array designs that circumvent the very small spatial scaling required for high-frequency linear and phased arrays [24],
[25]. Both the split aperture array designs by Talman and
Lockwood [24] and the actuated array design by Ritter et
al. [25] have been shown to be capable of producing highresolution images, and may provide excellent opportunities
for experimental studies at high frequencies in the future.
This paper describes the design, fabrication, and evaluation of a 64-element 2-2 composite 35-MHz array. Individual array elements were spaced at a 50-m pitch (1.2
in water) and were 3 mm long in the elevation direction.
Mechanical dicing was used to form the composite elements out of a bulk piece of ne-grain high-density Navy
Type VI ceramic. Elevation focusing was achieved using a
cylindrically shaped epoxy lens. Element interconnection
was made with a custom-designed exible circuit and a
quarter-wavelength, 75-, coaxial cable. The initial design
goals for this array are listed in Table I.
II. Methods
Fig. 1 displays a cutaway view of the array and all major
design components. Each array element incorporated two
piezoceramic slabs separated by an Al2 O3 -loaded epoxy
and with Cr/Au electroplating on both surfaces. The array composite was backed by a lossy conductive epoxy
which also served as the ground connection for the elements. The backed array was housed in a nonconductive

Fig. 1. A cutaway drawing of the array showing all of the major


design components (not to scale).

ceramic frame which enabled the electrical connection of


individual array elements to a exible circuit via sputtered
Cr/Au electrodes. A cylindrical acoustic lens was molded
onto the surface of the array using an unloaded epoxy.
This lens provides a xed elevation focus and acoustical
impedance matching between the composite to the load
medium, as well as protects the fragile array elements from
the environment. To provide RF shielding, the array was
encased in a brass tube capped o with a stainless steel
end piece. Only the active area of the array was exposed
through a small opening in the end piece. The metal housing assembly surrounding the array was also connected to
the coaxial cable bundle shield and, therefore, provided
continuous RF shielding from the distal array face to the
proximal connector box.
The design, fabrication, and testing of the array was
achieved using the following steps.
A. 2-2 Composite Design and Fabrication
There exist a number of techniques currently available for fabrication of piezo-polymer composites. The most
common of these is the dice and ll technique whereby
a mechanical dicing saw is used to cut kerfs into a piece of
bulk piezoceramic, the kerfs being subsequently backlled
with epoxy [26]. Unfortunately this technique can be problematic when considering the small feature sizes required
for very-high-frequency transducer operation. Other composite machining techniques available include laser ablation [18] and chemical etching [21]. Interdigital pair bonding [27] and more recently interdigital phase bonding [28]
have been determined to be eective modied dicing methods for manufacturing high-volume fraction 2-2 or 1-3 composites at frequencies up to 80 MHz. The stacking and
bonding of thin piezoceramic plates has also been shown
to be an eective method for creating high-frequency 2-2
composites [12]. Using tape-cast PZT is a viable alternative to this stack and bond technique for production of
large quantities of 2-2 composites at low cost [29]. This
technique involves the printing of carbon black ink on
thin sheets of green piezoceramic tape. The tape layers are
then stacked to form the 2-2 structure, and the carbon is
volatilized during heat treatment of the stack. The voids
left by the removed carbon can then be backlled with

226

ieee transactions on ultrasonics, ferroelectrics, and frequency control, vol. 53, no. 1, january 2006

epoxy under vacuum. Another technique that is suitable


for large-scale production is termed the lost mold technique [30]. This technique involves the pressing of a piezoceramic paste into a mold, which is usually silicon and can
be designed for various composite geometries. The mold is
removed by chemical etching. After sintering the remaining composite posts, the voids left by the mold removal are
backlled with epoxy. It should be noted, however, that at
present both the tape-cast and lost-mold techniques exhibit inferior piezoelectric properties when compared to
composites made from bulk piezoceramics. Because ease
of manufacture and use of bulk piezoceramic plates were
critical design considerations when the 35-MHz array was
developed, the dice and ll method was adopted to manufacture a 2-2 composite for this study.
The 35-MHz composite was diced using a 13-m hubbed
nickel/diamond blade (Asahi Diamond Industrial Co.,
Ltd., Tokyo, Japan) and a Tcar 864-1 (Thermocarbon,
Inc., Casselberry, FL) dicing saw. The polymer used to ll
the kerfs was a mixture of Epo-Tek 301 (Epoxy Technologies, Billerica, MA) and 1-5 m aluminum oxide (Al2 O3 )
particles (Buehler, Ltd., Lake Blu, IL). A plate of Navy
Type VI ne-grain piezoceramic (TRS600FGHD, TRS Ceramics, State College, PA) measuring 10100.5 mm was
selected and waxed to a glass carrier plate using a lowtemperature paran wax. The mechanical dicing saw was
rst used to cut kerfs in the ceramic 130 m deep at a 50m pitch. These rst cuts were subsequently lled with the
Al2 O3 -loaded epoxy (17% Al2 O3 by volume) using capillary action. A larger volume fraction of Al2 O3 was desired
for this application because it would translate to a higher
shear wave resonant frequency in the kerf. Unfortunately
a higher percent Al2 O3 made the mixture too viscous for
this application. The ller in the rst set of kerfs was left
to cure at room temperature in a dry nitrogen environment for 48 hours. The excess epoxy was then lapped o
to expose the diced ceramic. It was important to lap at
least 1020 m into the ceramic to expose the narrower
part of the kerfs. The dicing saw was then aligned to the
central portion of the exposed ceramic posts and a second
set of cuts was made to produce the nal 25-m composite pitch. The second set of diced kerfs was subsequently
lled with the same ller as before. After curing, excess
epoxy was lapped o the top of the composite. The most
uniform and smallest kerfs were produced at the deepest
dicing depths (Fig. 2). Therefore the deepest section of the
diced ceramic was chosen to produce the array composite.
The composite was lapped to a nal thickness of 50 m after it was placed in the array housing. A scanning electron
micrograph (SEM) image of a composite cross section before placement in the array housing is shown in Fig. 3. An
average composite kerf width produced was 14 m with a
net piezoceramic volume fraction of 44%.

Fig. 2. An SEM image of the diced ceramic before lapping it to the


nal composite thickness. The lighter portions of this image are the
ceramic posts whereas the darker portions are the kerf ller. The
dashed lines indicate the portion of the diced ceramic used in the
35-MHz array.

Fig. 3. An SEM image of a lapped and backed 2-2 composite. The


lighter portions of this image are the ceramic posts whereas the
darker portions are the kerf ller. The irregular textured section under the composite is the silver epoxy backing layer.

B. Finite Element Model Optimization of a Composite


Array Element

ever, a nite element model (FEM) was used to predict


composite and array performance prior to fabrication. The
relevant properties of the active and passive materials used
in the model are listed in Tables II and III, respectively.
A 2-D nite element model (PZFLEX, Wiedlinger Associates, Los Altos, CA) was rst used to generate an electrical impedance magnitude and phase plot for a 1.5 0.3mm, 50-m-thick virtual piece of fabricated composite resonating in air (Fig. 4). The series (fS ) and parallel (fP )
resonance frequencies from this model were used to determine the thickness mode coupling coecient kt for the
composite material based upon the following formula [31]:

 

fS
fP fS
kt =
tan
.
(1)
2 fP
2
fP

The nished composite was very fragile in its nal state.


It was therefore not tested prior to array fabrication. How-

The average modeled electromechanical coupling coefcient for the composite was 0.64 and the rst piezoelec-

cannata et al.: 35-mhz piezo-composite ultrasound array

227

TABLE II
Relevant Bulk Material Properties for the TRS600FGHD Piezoelectric Ceramic.1
Stiness
constants2
cE
11 (GPa)
cE
33
cE
12
cE
13
cE
44

Dielectric
constants

Stress
constants2

Other
properties

140

2
S
33 /o

1350

e15 (C/m2 )

(GPa)

121

1700

e31

(C/m2 )

5.7

(GPa)

75

2
S
11 /o
T
33 /o

3670

e33 (C/m2 )

25.8

(GPa)

90

T
11 /o

3830

(GPa)

22

20.24

Density2 (kg/m3 )

V33

7500

(m/s)

3966


k33

0.68

1 Courtesy
2 Used

of TRS Ceramics, State College, PA.


in FEM.

TABLE III
The Properties of the Passive Materials Used in the Array Design.1

Material
Epo-Tek 301
(Lens)
E-Solder 3022
(Backing layer)
Epo-Tek 301 + 17%
Al2 O3 (Kerf ller)
1 All

Density
(kg/m3 )

Vlong 2
(m/s)

Attenuationlong
(dB/mm)

Vshear 2
(m/s)

Attenuationshear
(dB/mm)

1150

2675

13.5

1270

48

3200

1850

110

1610

2710

15.9

1375

49

measurements were performed at 30 MHz [32].


and Vshear are the longitudinal and shear phase velocities, respectively.

2V
long

trically coupled lateral resonance attributed to the kerf


occurred near 55 MHz. Also, based upon the series resonance peak frequency of approximately 41 MHz, the longitudinal velocity for the modeled composite geometry was
4100 m/s. Therefore, with the ceramic and epoxy densities listed in Tables II and III, respectively, the acoustical
impedance for the modeled composite was calculated to
be approximately 16.7 MRayls.
It has been previously reported that nite element modeling can provide an accurate depiction of high-frequency
array performance [32], as well as reduce the number of
time-consuming prototype fabrication runs. For this study
FEM was used as an aide to determining the optimal materials and geometries used in the array design so that the
design goals in Table I could be met. For PZFLEX, as well
as any other FEM, the accuracy of the model results are
limited by the accuracy with which the properties of materials used in the design are measured. It would be ideal
if the complete set of properties were measured over the
entire bandwidth of the device. Unfortunately, at high frequencies it may be very dicult to characterize some materials due to the increase in attenuation associated with
the elevated frequency. An example of this problem can be
seen in Table III. It was not possible to characterize the
shear wave velocity and attenuation of the lossy conductive backing material used in the current 35-MHz array
design. Therefore, in order to eectively model the array
using FEM, a number of assumptions were made to ll
in these missing material properties. The shear wave attenuation was assumed to be four times that of the longi-

tudinal wave attenuation. This approximation was loosely


based upon the reported dierence between longitudinal
and shear attenuation for other epoxy mixtures measured
at high frequencies [33]. An approximate shear velocity
was calculated based upon the expected Poisson ratio ()
and the measured longitudinal velocity using the following
formula [34]:
Vshear
=
Vlong

0.5
.
1

(2)

A Poisson ratio of 0.37 was assumed for this material,


which is indicative of many rigid epoxies and plastics [34].
Therefore the shear velocity was assumed to be 45% that
of the longitudinal velocity reported.
A quarter-wavelength of coaxial cable can provide an
improved impedance match between a transducer and the
electronics if it possesses the proper impedance characteristics [35]. If properly designed and implemented, this
transmission line coax can serve to increase bandwidth
and sensitivity, as well as provide minor adjustments to
the center frequency of the transducer. Cable impedance
matching is well suited for use with devices operating at
high frequencies (> 20 MHz), where the length of the coax
can be approximately three meters or less. For this study, a
high-impedance (> 50 ohm) micro-coaxial cable was used
to match a high (> 200 at resonance)-impedance array element to the 50- send/receive electronics. This
technique can also be applied to large aperture singleelement transducers, which typically display an electrical

228

ieee transactions on ultrasonics, ferroelectrics, and frequency control, vol. 53, no. 1, january 2006
TABLE IV
Measured Properties for the 75- Precision Interconnect
Coaxial Cable.1
Property

PI #171-0574-XX

Characteristic impedance (Zo )


Propagation constant ()
Propagation velocity (Vp )
Resistance/unit length (r)
Capacitance/unit length (c)
Conductance/unit length (g)
Inductance/unit length (l)
Attenuation/unit length

73.8 1.6i
0.071 + 0.99i m1
2.2 108 m/s
6.77 /m
61 pF/m
674 S/m
0.33 H/m
0.61 dB/m

1 The

cable was characterized at 35 MHz. The respective cable parameters are displayed with units of meters (m), ohms (), henrys
(H), farads (F), siemens (S), and decibels (dB).
Fig. 4. Modeled electrical impedance magnitude (solid line) and
phase angle (dashed line) for a 1.5-mm 0.3-mm 0.05-mm piece
of composite resonating in air.

impedance of less than 50 , by using a low impedance


coax [36].
In order to utilize this coax in the nite element model
the cable was analyzed using the well-known transmission line equations. The rst step was to characterize the
impedance Zo and propagation constant for the coax at
35 MHz based upon the following formula for a coax length
x [37]:
Zx = Zo

[Zload + Zo T anh(x)]
,
[Zo + Zload T anh(x)]

(3)

where Zload represents the electrical impedance of the


transducer and Zx is the transformed coaxial impedance
measured on the system end of the cable.
The values for Zo and were obtained by measuring
the complex open and short-circuit impedance for a sample cable on an HP 4194 Impedance Analyzer (Agilent
Technologies, Englewood, CO). A given cable length x was
prepared for measurement, and the complex transformed
impedance was recorded for both the short-circuit and the
open-circuit Zload values. The two measured values for Zx
and (3) were then used to solve for Zo and . The distributed network representation for the 75- micro-coaxial
cable (#171-0574-XX, Precision Interconnect, Portland,
OR) used in the array design was also solved [38]. The
resultant per-unit-length values of cable propagation velocity (v), series resistance (r), series inductance (l), shunt
capacitance (c), shunt conductance (g), and attenuation
() were recorded and listed in Table IV.
Because of the moderately low composite acoustic
impedance (16.7 MRayls), only a single matching layer
was used in the array design. Based upon the formula by
DeSilets et al. [38] for broadband transducer operation,
the ideal matching layer impedance was determined to be
3.35 MRayls. Therefore the array lens epoxy (Epo-Tek 301,
Z = 3.1 MRayls) was chosen and assumed to be a single /4 matching layer in the 2-D nite element model.
From the FEM the ideal matching layer thickness was determined to be 19 m and the ideal micro-coaxial cable

length was 1.2 m. The electrical impedance of a modeled


single array element before and after adding the coaxial
cable is displayed in Fig. 5. The model was nally used
to predict the pulse/echo response from a at plate reector placed at 9.5 mm away from the array face. The echo
response, shown in Fig. 6, displays a center frequency of
34.8 MHz and 6 dB bandwidth of 53%. Once the bandwidth requirements for the array element in the nite element model were met, the next developmental step was
to fabricate a full 64-element array module as shown previously in Fig. 1.
C. Array Module Fabrication
The conductive backing material, E-solder 3022 (Von
Roll Isola, Inc., New Haven, CT), was cast on a lapped and
electroded plate of 2-2 composite material using a 2,000g
centripetal force for ten minutes. This process separated
the epoxy into loaded and unloaded layers and served to
compact the silver akes up against the composite plate.
The backed composite was cured in a dry nitrogen environment for 24 hours at room temperature and then postcured at 40 C and 50 C for 10 hours each. The elevated
temperature post-cure increased the glass transition temperature of the epoxy in order to provide a more rigid
substrate for the latter array fabrication steps. After the
backing was fully cured, it was lapped down to a thickness
of approximately 3 mm in order to remove the unloaded
epoxy layer and to ensure that the silver-loaded portion
was parallel to the composite plate. In this state the backing epoxy displayed an electrical resistance of 0.6 /mm.
The backed composite was then diced to a width of 3 mm
and length of approximately 3.7 mm, to provide at least
ve extra elements on either side of the 64-element array.
The backed composite was then cast into a rigid ceramic frame (machinable glass-mica ceramic, McMasterCarr Supply Company, Cleveland, OH) using Epo-Tek 301
epoxy. This rigid frame was used to provide support to
the composite, as well as serve as a nonconductive barrier surrounding the composite that could be used as a
surface for electrical interconnect to individual array elements [13]. The framed composite was allowed to cure at

cannata et al.: 35-mhz piezo-composite ultrasound array

229

Fig. 6. The FEM pulse-echo response for a single 35-MHz array


element. The axes on the top and right of the gure refer to the
frequency spectrum (dashed line). Also displayed are the calculated
center frequency (CF), the 6 dB bandwidth (BW), and the 20 dB
pulse length (PL) for the response.

ing the radius of curvature of the lens using the following


formula [39]:
f
Fig. 5. Modeled electrical impedance magnitude (solid line) and
phase angle (dashed line) for a single composite array element before
(top) and after (bottom) coaxial cable impedance matching.

room temperature for 48 hours in a dry nitrogen environment, and then post-cured at 40 C and then 50 C for 24
hours each. The composite and frame were then lapped
down by 10 m in thickness to ensure that the composite
posts were 50-m thick and that the epoxy bond line between composite and frame was at and contiguous. The
top and side surfaces of the framed composite were then
cleaned and sputtered with a total of 4500
A Cr/Au. Element separation was achieved by mechanically scratchdicing the top electrode layer over the composite kerfs and
ceramic frame with the prescribed 50-m spacing using the
13-m diamond/nickel hubbed blade previously described.
The opposing two electroplated sides of the ceramic frame
were also scratched-diced to separate individual element
electrodes. For this a 50-m blade was used to completely
remove a 50-m width of electroplating, leaving connections for either the odd or even element electrodes on one
side of the ceramic frame spaced at a 100-m pitch (Fig. 1).
The electroplated array was then prepared for casting of
the epoxy lens.
The epoxy lens was cast onto the array with a polished
quartz cylindrical rod (ISP Optics, Irvington, NY) as a
mold. The desired focal length f was achieved by specify-

c2 ,
c1

(4)

where is the radius of curvature of the lens, c1 is the


sound velocity in the lens material, and c2 is the sound
velocity in the medium. A quartz rod with a radius approximately equal to 4.34 mm was chosen to focus the
array at a range of 10 mm. The quartz rod was sprayed
with a mold release agent (Ease Release 200, Mann Formulated Prods., Gillete, NJ) to ease in the epoxy post-cure
removal process. As noted previously, the apex of the lens
was designed to be approximately /4 thick at the center
of the elevation aperture to aid in acoustically matching
the transducer to the load medium. To ensure this, the
cylindrical rod was centered over the elevation aperture
and oset by 19 m using shims. The lens epoxy, EpoTek 301, was degassed and ltered to remove bubbles and
impurities prior to application. A bolus of the epoxy was
allowed to wick between the cylindrical rod and array assembly. Surface tension held the liquid epoxy in place during curing. The lens was allowed to cure for two days in a
dry nitrogen environment before post-curing at 40 C for 4
hours. After curing, the edges of the lens above the ceramic
frame were lapped parallel to the array elements.
D. Array Interconnect, Housing, and Termination
The connection of individual coaxial cables to array
elements was accomplished through the use of two in-

230

ieee transactions on ultrasonics, ferroelectrics, and frequency control, vol. 53, no. 1, january 2006

Fig. 7. A picture of the nished 35-MHz array. The bottom right side
of this image shows an enlarged view of the front of the array. The
long axis of the hole cut in the stainless steel nosepiece is along the
azimuth direction.

termediate exible circuits incorporating 5-m-thick Au


traces on a 50-m-thick polyimide surface (Dynamic Research Corp., Wilmington, MA). The coaxial cable assembly (#171-0574-XX, Precision Interconnect, Portland,
OR) was attached to the exible circuit by the cable manufacturer using a proprietary low-temperature solder process. The exible circuit was connected to the framed array
by carefully aligning and bonding the ex circuit traces
to the electrodes on the two sides of the framed array.
The epoxy used for this process was Epo-Tek 301. The
newly interconnected array assembly was allowed to cure
for 48 hours at room temperature in a dry nitrogen environment before further processing. A grounding wire was
connected from the conductive epoxy array backing to the
cable assembly shield. The array assembly was housed in a
ve-part cylindrical metal and plastic housing. Individual
coaxial cables were terminated into an aluminum enclosure housing 64 individual SMB connectors. This ensured
satisfactory electrical isolation between each coaxial cable.
A photograph of the completed array is shown in Fig. 7.
E. Array Characterization
The ultimate indication of ultrasonic array performance
is its ability to form an image. However, image quality depends not only on the array but also on the system electronics and signal processing algorithms. Thus it is dicult
to compare dierent arrays based solely on image analysis. Therefore, several standard non-imaging tests were
rst performed on the 35-MHz array so that it could be
compared to arrays built prior to this study.
The electrical impedance magnitude and phase angle for
each element were measured using a HP4194 impedance
analyzer with the z-probe attachment. For this test the
array face was placed in a water bath, and the electrical
impedance magnitude and phase angle were measured over
the array pass-band.

The pulse-echo response test is the most common test


performed on ultrasonic single-element and array transducers. This test was used to measure and provide a relative comparison of array element center frequency, bandwidth, focus, pulse length, and sensitivity. A Panametrics 5900PR pulser/receiver (Panametrics, Inc., Waltham,
MA) was used to excite each element using the 1-J, 50-
settings with 20 dB total gain on receive. The array was
positioned in a degassed/deionized water bath opposite a
polished stainless steel plate reector. The time delay observed on the oscilloscope (Lecroy LC534, Chestnut Ridge,
NY) between excitation and the largest rst echo response
was recorded to determine elevation focal length. The fast
Fourier transform (FFT) function on the oscilloscope was
used to determine the frequency response of this RF echo.
The two 6 dB points of this power spectrum dened the
upper and lower band edges of the signal, and the mean
of these frequencies was recorded as the center frequency.
The percent bandwidth was calculated by dividing the frequency dierence between the two 6 dB points by the
center frequency. The amplitude of the echo signal was
recorded for relative element sensitivity comparisons. The
pulse length of the echo waveform was recorded as the
length of time between the rst and last points where the
signal was 20 dB relative to the peak.
The level of electrical and acoustical separation between
elements was determined by measuring crosstalk. For this
test the array was positioned in a degassed/deionized
water bath opposite an absorptive piece of rubber. A
Sony/Tektronix AFG2020 function generator (Tektronix,
Inc., Beaverton, OR), set in burst mode, was used to excite a representative element with the applied voltage measured as a reference. Voltages across the three nearest elements were measured and compared to this reference voltage. This process was repeated at discrete frequencies over
the array pass-band.
The azimuthal one-way directivity was measured by rotating a representative element around an axis along its
length and center. A needle hydrophone (Precision Acoustics, Dorchester, UK), placed at the elevation focus, was
used to acquire the amplitude of the time-domain response
at discrete angular positions. The Field II program was
used to simulate the directivity of a single array element
[40], and to estimate the eective element width by matching a theoretical directivity curve to the measured values.
Insertion loss was measured by exciting a representative array element with a 35-MHz voltage burst, and receiving the reected echo from a polished steel reector
placed at the elevation focus. The receive power across a
50- load was referenced to the source power delivered to
a 50- reference load and expressed in decibels. The measured value was then corrected for loss due to diraction in
the azimuth direction [41], attenuation in the water bath
(2.2 104 dB/mm-MHz2 ) [42], and reection from the
steel target (0.6 dB).
For the nal performance test, the array was used to
image a custom-made wire target phantom composed of
ve evenly spaced 20-m-diameter tungsten wires (Cali-

cannata et al.: 35-mhz piezo-composite ultrasound array

231

fornia Fine Wire Company, Grover Beach, CA). There are


currently two 16-channel analog [43] and digital [44] highfrequency beamformers available to produce real-time images at center frequencies up to 40 MHz. Unfortunately,
given the current geometry of the 35-MHz array, it was determined that 16 transmit and receive channels would not
provide adequate lateral resolution for this study. Therefore a synthetic aperture reconstruction algorithm was
used to form the image of the wire target phantom.
The 1.2- pitch is problematic for synthetic aperture
imaging because the rst-order grating lobe appears at
one-half the angular location observed in a conventional
beamforming system [45]. For the 35-MHz array without
beam steering, grating lobes are expected to occur at 57
and 28.5 for conventional and synthetic systems, respectively. To limit the eect of these secondary lobes, array elements were only used for reconstruction if they fell
within a 9 acceptance angle for a point in the formed
image. Therefore, the aperture size used for reconstruction was varied dynamically throughout the image to produce a consistent lateral resolution. Backprojection using
a monostatic approach was accomplished by delaying and
summing the time-domain contributions to each pixel according to the following formula [13]:
P (xi , zi ) =




2
we Re t (xe xo )2 + zo2 ,
c
e=1

TABLE V
Measured Properties for the 64-Element Composite Array.
Property

Value

Number of elements
Number of open elements
Number of shorted elements
Average center frequency
Highest/lowest center frequency
Average bandwidth (6 dB)
Highest/lowest bandwidth
Average sensitivity
Highest/lowest sensitivity
20 dB pulse length
Electrical impedance magnitude
(at 35 MHz)
Electrical impedance phase angle
(at 35 MHz)
Focal point
Insertion loss

64
1
0
35.3 MHz
36.5 MHz/34.2 MHz
55%
62%/49%
403 mV
444 mV/360 mV
94 ns

Measured

31 ohms
48 degrees
9.5 mm
22.8 dB

on a representative array element.

N


(5)

where xi , zi is the location of the pixel in the image plane,


e is the index of the element over the range of 1 to N
(number of elements), we is the apodization function, Re
is the time-domain response, t is the time, c is the propagation velocity, xo , zo is the location of the point in object
space, and xe is the position of the array element. The individual time-domain responses were acquired by manually
connecting each of the 64 array elements to a Panametrics
5900 pulser-receiver using the same settings as described
for the pulse-echo test setup. Image reconstruction was accomplished o-line using programs written in Matlab [25].

III. Results and Discussion


All of the individual element test results for the completed array are summarized in Table V. The array displayed only one open element and no shorted elements.
Fig. 8 shows the electrical impedance magnitude and
phase angle for a typical array element. The general trend
of this curve matches that of the FEM impedance plot
shown in Fig. 5. However, the resonance peak seen in the
FEM plot near 40 MHz, which corresponds to the piezocomposite parallel resonance, is not visible in the measured
impedance plot. The likely cause of this discrepancy is a
higher-than-desired electrical trace resistance between an
array element and a coaxial cable. Exploring the FEM further veried that increasing element trace resistance would
produce the measured result. According to the model, the

Fig. 8. Measured electrical impedance magnitude (solid line) and


phase angle (dashed line) for a representative composite array element.

consequence of this is a slight reduction of pulse sensitivity


and bandwidth.
A comparison of the pulse-echo responses for a typical
and the worst (subjectively judged by the authors) connected array element is shown in Fig. 9. The array displayed an average center frequency of 35.3 MHz with an
average 6 dB bandwidth of 55%. Both of these average measurements were consistent with the initial requirements for the array design and the nite element model.
The measured 20 dB pulse length for a typical element
of 94 ns also met the initial specications for the array.
However, this recorded pulse length is slightly longer than
expected from the FEM. This error between model and
experiment can be attributed to the observed additional
resistance in the element traces, as well as the expected
errors due to the aforementioned assumptions used in the
model. A larger increase in positive trace resistance is also
the most likely cause of the additional reduction of sensi-

232

ieee transactions on ultrasonics, ferroelectrics, and frequency control, vol. 53, no. 1, january 2006

Fig. 10. The variation of center frequency across the array.

Fig. 11. The variation of sensitivity and bandwidth across the array.
Fig. 9. Measured pulse-echo response for two 35-MHz array elements.
The waveform on the top (element #64) was deemed typical whereas
the one on the bottom (element #24) was judged to be the worst.
The axes on the top and right of the gure refer to the frequency
spectrum (dashed line). Also displayed are the calculated center frequency (CF), the 6 dB bandwidth (BW), the 20 dB pulse length
(PL), and peak-to-peak voltage (Vpp ).

tivity and pulse shape recorded for the worst active element. Overall the performance of the array was very consistent across the aperture, as shown in Figs. 10 and 11.
Crosstalk measurements indicated satisfactory but not
ideal element-to-element isolation. The measured crosstalk
for the array is shown in Fig. 12. Near the center frequency
of the array the measured crosstalk was <24 dB. There
are two likely possibilities for this higher-than-desired
crosstalk: acoustic or mechanical cross-coupling between
elements in the 2-2 composite array structure through
the backing, kerf ller and lens, and/or electrical crosscoupling caused by a competition between the high array

Fig. 12. Measured crosstalk for the array.

cannata et al.: 35-mhz piezo-composite ultrasound array

233

Fig. 13. Measured and modeled one-way directivity for a single-array


element. This gure shows three sets of hydrophone measurements
taken at a range of 10 mm, and the modeled result also taken at a
range of 10 mm using a 70-m element width.

element impedance and the impedance between successive


electrodes. Although a combination of the two sources is
likely, coupling was observed to be instantaneous, indicating that electrical crosstalk was dominant. As a comparison, Guess et al. [46] observed combined crosstalk levels
exceeding 24 dB for adjacent elements on a similarly
constructed 3.0-MHz array. However, that study showed
that, while high crosstalk levels were observed for nearest neighbors, crosstalk between next nearest neighbor elements was much lower at less than 45 dB over the passband. This was not the case for the 35-MHz array design.
In fact, the crosstalk magnitude for the next nearest neighbor element, which had a ex circuit connection next to
the source element, was comparable to the magnitude observed for the nearest neighbor. This observation further
reinforces the notion that the major source of crosstalk in
the fabricated array was electrical in nature. In the future
this issue could be overcome by improving the electrical
impedance characteristics of an array element via reduced
electrode trace resistance, and/or by increasing element
area or dielectric permittivity of the piezo-ceramic used in
the design.
The measured one-way directivity pattern for a representative array element is shown in Fig. 13. Three
data sets were taken for comparison. The higher-thandesired crosstalk observed explains why the eective element width (70 m) was 40% larger than the array pitch.
A similar result was observed by Felix et al. for a 3.5-MHz
array [47]. The null in the peak of these patterns can also
be attributed to crosstalk eects [41].
The measured insertion loss for a representative array
element was 50.3 dB. Correction for diraction in the azimuth direction was achieved using the directivity pattern
for a 70-m element. After compensation for attenuation
caused by the water and reection from the target, an
estimated insertion loss of 22.8 dB was obtained. This

Fig. 14. Synthetic aperture image of a wire phantom reconstructed


using a half-angle of 9 and no apodization or thresholding. The
dynamic range of the image is 40 dB and the display uses a linear
gray scale for mapping.

value was comparable to a similarly built, high-frequency


lens-focused transducer [36], but was 7.8 dB higher than
reported by Ritter et al. [13] for an optimized 30-MHz 22 composite array with two acoustic matching layers and
lens. However, given the simplicity of the 35-MHz array
design, lower element sensitivity is a compromise in this
more mass-producible design.
An image of the ve-wire test phantom is shown in
Fig. 14 using a linear gray scale and 40-dB dynamic range.
No apodization or thresholding was implemented during
reconstruction. Plots of the axial and lateral line spread
functions for the center wire are shown in Fig. 15. The
measured FWHM (full-width half-maximum) resolutions
were 42 m and 95 m for the axial and lateral directions,
respectively. These measurements correlate well with the
resolutions of 40 m axially and 95 m laterally predicted
from a Field II simulation and oer a slight improvement

234

ieee transactions on ultrasonics, ferroelectrics, and frequency control, vol. 53, no. 1, january 2006

<30 dB crosstalk, all of the initial array design goals were


met: a center frequency near 35 MHz, > 50% bandwidth,
and a pulse length of less than 120 ns. A wire phantom image, reconstructed using a synthetic aperture technique,
demonstrated resolutions of 95 m or less with minimal
artifacts. The results indicate that this array design, when
coupled with an adequate high-frequency beamformer, will
be suitable for clinical applications. Assessment of the array on animals and human patients is anticipated in the
near future.

References

Fig. 15. Axial (top) and lateral (bottom) line spread functions for
the center wire of the phantom.

laterally and signicant improvement axially over the resolutions reported by Ritter et al. [13] for the 30-MHz composite array. The phantom image displayed low-level sidelobes resulting from the reconstruction algorithm and the
limited number of elements used for reconstruction. These
sidelobes could be reduced in amplitude by using apodization; however, this comes at a cost of increased main lobe
width [48].

IV. Conclusions
This paper describes the development and fabrication
of a 35-MHz linear array made using mechanically diced
2-2 composite elements. With the exception of the desired

[1] G. R. Lockwood, D. H. Turnbull, D. A. Christopher, and F. S.


Foster, Beyond 30 MHz: Applications of high frequency ultrasonic imaging, IEEE Eng. Med. Biol., vol. 15, pp. 6071, 1996.
[2] F. S. Foster, C. J. Pavlin, G. R. Lockwood, L. K. Ryan, K. A.
Harasiewicz, L. Berube, and A. M. Rauth, Principles and applications of ultrasonic backscatter microscopy, IEEE Trans.
Ultrason., Ferroelect., Freq. Contr., vol. 40, pp. 608616, 1993.
[3] D. J. Coleman, R. H. Silverman, A. Chabi, M. J. Rondeau, K. K.
Shung, J. Cannata, and H. Linco, High-resolution ultrasonic
imaging of the posterior segment, Ophthalmology, vol. 111, no.
7, pp. 13441351, 2004.
[4] D. H. Turnbull, B. G. Starkoski, K. A. Harasiewicz, J. L. Semple, L. From, A. K. Gupta, D. N. Sauder, and F. S. Foster, A
40100 MHz B-scan ultrasound backscatter microscope for skin
imaging, Ultrasound Med. Biol., vol. 21, no. 1, pp. 7988, 1995.
[5] C. Passman and H. Ermert, A 100 MHz ultrasound imaging
system for dermatologic and ophthalmologic diagnostics, IEEE
Trans. Ultrason., Ferroelect., Freq. Contr., vol. 43, no. 4, pp.
545552, 1996.
[6] D. H. Turnbull, In utero ultrasound backscatter microscopy of
early stage mouse embryos, Comput. Med. Imaging Graph., vol.
23, no. 1, pp. 2531, 1999.
[7] F. S. Foster, M. Y. Zhang, Y. Q. Zhou, G. Liu, J. Mehi, E.
Cherin, K. A. Harasiewicz, B. G. Starkoski, L. Zan, D. A.
Knapik, and S. L. Adamson, A new ultrasound instrument for
in vivo microimaging of mice, Ultrasound Med. Biol., vol. 28,
no. 9, pp. 11651172, 2002.
[8] C. S. DeSilets, Transducer arrays suitable for acoustic imaging, Ph.D. dissertation, Stanford University, Stanford, CA,
May 1978.
[9] M. Lethiecq, G. Feuillard, L. Ratsimandresy, A. Nguyen-Dinh,
L. Pardo, J. Ricote, B. Andersen, and C. Millar, Miniature
high frequency array transducers based on new ne grain ceramics, in Proc. IEEE Ultrason. Symp., 1994, pp. 10091013.
[10] A. Nguyen-Dinh, L. Ratsimandresy, P. Mauchamp, R. Dufait, A.
Flesch, and M. Lethiecq, High frequency piezo-composite transducer array designed for ultrasound scanning applications, in
Proc. IEEE Ultrason. Symp., 1996, pp. 943947.
[11] S. Michau, P. Mauchamp, and R. Dufait, Piezocomposite
30 MHz linear array for medical imaging: Design challenges and
performances evaluation of a 128 element array, in Proc. IEEE
Ultrason. Symp., 2004, pp. 898901.
[12] T. A. Ritter, E. Gerber, K. K. Shung, and T. R. Shrout, Method
for manufacture of high frequency ultrasound transducers, U.S.
Patent 6 183 578, Feb. 6, 2001.
[13] T. A. Ritter, T. R. Shrout, T. Tutwiler, and K. K. Shung, A
30-MHz composite ultrasound array for medical imaging applications, IEEE Trans. Ultrason., Ferroelect., Freq. Contr., vol.
49, no. 2, pp. 217230, 2002.
[14] P. A. Payne, J. V. Hateld, A. D. Armitage, Q. X. Chen, P. J.
Hicks, and N. Scales, Integrated ultrasound transducers, in
Proc. IEEE Ultrason. Symp., 1994, pp. 15231526.
[15] S. J. Carey, C. M. Gregory, M. P. Brewen, J. M. Birch, S.
Ng, and J. Hateld, PVDF array characterization for high frequency ultrasonic imaging, in Proc. IEEE Ultrason. Symp.,
2004, pp. 19301933.
[16] S. J. Carey, C. M. Gregory, and J. V. Hateld, Electronics for
high impedance ultrasonic transducers, in World Congress of
Ultrasound 2003, Paris, France, 2003, pp. 263266.

cannata et al.: 35-mhz piezo-composite ultrasound array


[17] C. E. Morton and G. R. Lockwood, Evaluation of keress linear
arrays, in Proc. IEEE Ultrason. Symp., 2002, pp. 12571260.
[18] M. Lukacs, M. Sayer, and F. S. Foster, Single-element and linear
array transducer design for ultrasound biomicroscopy, in Proc.
Med Imaging: Ultrason. Transd. Eng., vol. 3341, 1998, pp. 272
282.
[19] F. Levassort, L. Pascal, T. H. Hue, J. Holc, T. Bove, M. Kosec,
and M. Lethiecq, High performance piezoelectric lms on substrates for high frequency imaging, in Proc. IEEE Ultrason.
Symp., 2001, pp. 10351038.
[20] R. Kline-Schoder, D. Kynor, M. Jaeger, A. Windor, and C. DeSilets, Physical vapor deposition of multi-layered PZT lms for
ultrasonic transducer fabrication, in Proc. Med. Imaging: Ultrason. Transd. Eng., vol. 3664, 1999, pp. 221228.
[21] Y. Ito, K. Kushida, K. Sugawara, and H. Takeuchi, A 100 MHz
ultrasonic transducer array using ZnO thin lms, IEEE Trans.
Ultrason., Ferroelect., Freq. Contr., vol. 42, no. 2, pp. 316324,
1995.
[22] T. Buma, M. Spicar, and M. ODonnell, Thermoelastic expansion vs. piezoelectricity of high-frequency, 2D arrays, IEEE
Trans. Ultrason., Ferroelect., Freq. Contr., vol. 50, no. 8, pp.
10651068, 2003.
[23] O. Oralkan, S. Hansen, B. Bayram, G. G. Yaralo
glu, A. S.
Ergun, and B. Khuri-Yakub, High frequency CMUT arrays
for high-resolution medical imaging, in Proc. IEEE Ultrason.
Symp., 2004, pp. 399402.
[24] J. R. Talman and G. R. Lockwood, Evaluation of the radiation
pattern of a split aperture linear phased array for high frequency
imaging, IEEE Trans. Ultrason., Ferroelect., Freq. Contr., vol.
47, no. 1, pp. 117124, 2000.
[25] T. A. Ritter, T. R. Shrout, and K. K. Shung, A high frequency ultrasound array incorporating an actuator, in Proc.
Med. Imaging: Ultrason. Transducer Eng., vol. 4325, 2001, pp.
3646.
[26] H. P. Savakas, K. A. Klicker, and R. E. Newnham, PZTepoxy piezoelectric transducers: A simplied fabrication procedure, Mater. Res. Bull., vol. 16, no. 6, pp. 677680, 1981.
[27] R. Liu, K. A. Harasiewicz, and F. S. Foster, Interdigital pair
bonding for high frequency (2050 MHz) ultrasonic composite
transducers, IEEE Trans. Ultrason., Ferroelect., Freq. Contr.,
vol. 48, no. 1, pp. 299306, 2001.
[28] J. Yin, M. Lukacs, K. Harasiewicz, and F. S. Foster, Ultra-ne
piezoelectric composites for high frequency ultrasonic transducers, in Proc. IEEE Ultrason. Symp., 2004, pp. 19621965.
[29] W. Hackenberger, S. Kwon, P. Rehrig, K. Snook, S. Rhee,
and X. Geng, 2-2 PZT-polymer composites for high frequency
(>20 MHz) ultrasound transducers, in Proc. IEEE Ultrason.
Symp., 2002, pp. 12531256.
[30] S. Cochran, A. Abrar, K. J. Fox, D. Zhang, T. W. Button,
B. Su, C. Meggs, and N. Porch, Net-shape ceramic processing as a route to ultrane scale 1-3 connectivity piezoelectric
ceramic-polymer composite transducers, in Proc. IEEE Ultrason. Symp., 2004, pp. 16821685.
[31] IEEE Standard on Piezoelectricity, ANSI/IEEE Standard 1761987, New York: IEEE, Inc., 1988.
[32] J. M. Cannata and K. K. Shung, A comparison of model and
experiment for a high frequency (35 MHz) linear ultrasonic array, in Proc. IEEE Ultrason. Symp., 2003, pp. 16581662.
[33] H. Wang, T. A. Ritter, W. Cao, and K. K. Shung, High frequency properties of passive materials for ultrasonic transducers, IEEE Trans. Ultrason., Ferroelect., Freq. Contr., vol. 48,
no. 1, pp. 7884, 2001.
[34] G. S. Kino, Acoustic Waves: Devices, Imaging, and Analog Signal Processing. Englewood Clifs, NJ: Prentice-Hall, 1987.
[35] R. E. McKeighen, Design guidelines for medical ultrasonic arrays, in Proc. Med. Imaging: Ultrason. Transd. Eng., vol. 3341,
1998, pp. 218.
[36] J. M. Cannata, T. A. Ritter, W.-H. Chen, R. H. Silverman,
and K. K. Shung, Design of ecient, broadband single-element
(2080 MHz) ultrasonic transducers for medical imaging applications, IEEE Trans. Ultrason., Ferroelect., Freq. Contr., vol.
50, no. 11, pp. 15481557, 2003.
[37] A. P. Albrecht, Transmission lines, in Electronic Designers
Handbook. L. J. Giacoletto, Ed. New York: McGraw-Hill, 1977,
pp. 8.18.78.

235
[38] C. S. DeSilets, J. D. Fraser, and G. S. Kino, Design of ecient
broadband piezoelectric transducers, IEEE Trans. Sonics Ultrason., vol. 25, no. 3, pp. 115125, 1978.
[39] K. K. Shung, M. B. Smith, and B. M. W. Tsui, Principles of
Medical Imaging. San Diego, CA: Academic, 1992.
[40] J. Jensen, Field: A program for simulating ultrasound systems, Med. Biol. Eng. Comp., vol. 34, no. 1, pp. 351353, 1996.
[41] A. R. Selfridge, The design and fabrication of ultrasonic transducers and transducer arrays, Ph.D. dissertation, Stanford University, Stanford, CA, July 1983.
[42] G. R. Lockwood, D. H. Turnbull, and F. S. Foster, Fabrication
of high frequency spherically shaped ceramic transducers, IEEE
Trans. Ultrason., Ferroelect., Freq. Contr., vol. 41, no. 2, pp.
231235, 1994.
[43] Q. Wu, R. L. Tutwiler, and K. K. Shung, Design of an analog beamformer for very high-frequency ultrasound transducer
arrays, in Proc. Med. Imaging: Ultrason. Transd. Eng., vol.
3982, 2000, pp. 142149.
[44] C.-H. Hu, K. A. Snook, X.-C. Xu, J. T. Yen, K. K. Shung, and
P.-J. Cao, FPGA based digital high frequency beamformers for
arrays, in Proc. IEEE Ultrason. Symp., 2004, pp. 13471350.
[45] P. D. Corl, P. M. Grant, and G. S. Kino, A digital synthetic focus acoustic imaging system for NDE, in Proc. IEEE Ultrason.
Symp., 1978, pp. 263266.
[46] J. F. Guess, C. G. Oakley, S. J. Douglas, and R. D. Morgan,
Cross-talk paths in array transducers, in Proc. IEEE Ultrason.
Symp., 1995, pp. 12791282.
[47] N. Felix, D. Certon, E. Lacaze, M. Lethiecq, and F. Patat, Experimental investigation of cross-coupling and its inuence on
the elementary radiation pattern in 1-D ultrasonic arrays, in
Proc. IEEE Ultrason. Symp., 1999, pp. 10531056.
[48] C. H. Frazier and W. D. OBrien, Synthetic aperture techniques
with a virtual source element, IEEE Trans. Ultrason., Ferroelect., Freq. Contr., vol. 45, no. 1, pp. 196207, 1998.

Jonathan M. Cannata (S01M04) was


born in Carson, CA on August 4, 1975. He
received his B.S. degree in Bioengineering
from the University of California at San Diego
in 1998, and his M.S. and Ph.D. degrees in
Bioengineering from the Pennsylvania State
University, University Park, PA, in 2000 and
2004, respectively.
Since 2001 Dr. Cannata has served as the
resource manager for the NIH (National Institutes of Health) Resource Center for Medical Ultrasonic Transducer Technology at Penn
State University (2001 to 2002) and currently at USC (University of
Southern California). In 2005 he was appointed the title of Research
Assistant Professor of Biomedical Engineering at USC. His current
research interests include the design, modeling, and fabrication of
high frequency single element ultrasonic transducers and transducer
arrays for medical imaging applications. Dr. Cannata is a member of
the Institute of Electrical and Electronics Engineers (IEEE).

Jay A. Williams was born on June 14,


1955, in Bellefonte, PA. He graduated from
State College Area School District of PA in
1973 and audited courses in Physics, Computer Science, Mechanical Engineering, Architecture, and Business Management at the
Pennsylvania State University, 19751977. He
had been in industry for over twenty-ve years
in technical elds such as Mass Spectrometry,
Microwave Telecommunication, Liquid Chromatography, Digital Electronics, Ultrasound,
and Information Technology. From 1990 until
2001, he worked in ultrasonics at Blatek, Inc., eight-and-a-half years
in engineering and two-and-a-half years in management, including
ve years as Network Administrator and two years as Quality System
Manager in charge of establishing their ISO9001/CGMP compliant
quality system.

236

ieee transactions on ultrasonics, ferroelectrics, and frequency control, vol. 53, no. 1, january 2006

Mr. Williams joined the NIH Resource Center for Medical Ultrasonic Transducer Technology in March, 2002 as Research Technologist, just prior to their move to the University of Southern California
in August of that year. He has also served as the webmaster for the
Resource Center website http://bme.usc.edu/UTRC since the move
to Los Angeles, CA. He has now worked in the ultrasound eld for
over fourteen years.
His research interests are in novel techniques for high-frequency
transducer/array design and fabrication, and the role information
technology can play in improving the capabilities and accessibility of
technology.

Qifa Zhou received his Ph.D. degree from the


Department of Electronic Materials and Engineering at Xian Jiaotong University, Xian,
China in 1993.
He is currently a Research Assistant Professor at the NIH Resource on Medical Ultrasonic Transducer Technology (Los Angeles,
CA) and the Department of Biomedical Engineering at University of Southern California, Los Angeles, CA. Before joining USC in
2002, he worked in the Department of Physics
at Zhongshan University of China at Guang
Zhou, the Department of Applied Physics at Hong Kong Polytechnic University at Kowloon, and the Materials Research Laboratory
at Pennsylvania State University, University Park, PA.
His current research interests include the development of ferroelectric thin lms, MEMS technology, modeling and fabrication of
high frequency ultrasound transducers and arrays for medical imaging applications. He has published more than 70 papers in this area.
He is a member of the Materials Research Society.

Timothy Ritter was born in Harrisburg, PA


on February 19, 1965. He earned his B.S. degree in Mechanical Engineering from Penn
State University in 1987, his M.S. degree in
Physics from the University of Connecticut
in 1991, and his Ph.D. degree in Bioengineering from Penn State University in 2000. From
1998 until 2001 he served as manager of the
NIH Resource Center for Medical Ultrasonic
Transducer Technology. In 2000 he was also
appointed an Assistant Professor of Bioengineering at Penn State. He accepted a commission in the U.S. Air Force in 2001 and was assigned to Keesler Air
Force Base following completion of Ocer Training School. His current assignment is in therapeutic and diagnostic medical physics. He
is a member of the American Association of Physicists in Medicine
and the Health Physics Society.

K. Kirk Shung (S73M75SM89F93)


obtained a B.S. degree in electrical engineering from Cheng-Kung University in Taiwan
at Tainan in 1968, a M.S. degree in electrical engineering from University of Missouri,
Columbia, MO in 1970 and a Ph.D. degree
in electrical engineering from University of
Washington, Seattle, WA, in 1975. He did
postdoctoral research at Providence Medical
Center in Seattle, WA, for one year before being appointed a research bioengineer holding
a joint appointment at the Institute of Applied Physiology and medicine, Seattle, WA. He became an assistant
professor at the Bioengineering Program, Pennsylvania State University, University Park, PA in 1979 was promoted to professor in 1989.
He was a Distinguished Professor of Bioengineering at Penn State
until September 1, 2002 when he joined the Department of Biomedical Engineering, University of Southern California, Los Angeles, CA,
as a professor. He has been the director of NIH Resource on Medical
Ultrasonic Transducer Technology since 1997.
Dr. Shung is a fellow of the IEEE, the Acoustical Society of America and the American Institute of Ultrasound in Medicine. He is a
founding fellow of the American Institute of Medical and Biological
Engineering. He has served for two terms as a member of the NIH
Diagnostic Radiology Study Section. He received the IEEE Engineering in Medicine and Biology Society early career award in 1985 and
coauthored a best paper published in IEEE Transactions on UFFC in
2000. He is the distinguished lecturer for the IEEE UFFC society for
20022003. He was elected an outstanding alumnus of Cheng-Kung
University in 2001.
Dr. Shung has published more than 160 papers and book chapters. He is the author of a textbook Principles of Medical Imaging
published by Academic Press in 1992. He co-edited a book Ultrasonic Scattering by Biological Tissues published by CRC Press in
1993. Dr. Shungs research interest is in ultrasonic transducers, high
frequency ultrasonic imaging, and ultrasonic scattering in tissues.

Vous aimerez peut-être aussi