You are on page 1of 9

This article has been accepted for inclusion in a future issue of this journal.

Content is final as presented, with the exception of pagination.

JOURNAL OF MICROELECTROMECHANICAL SYSTEMS 1

Effects of DC Bias Tuning on Air-Coupled


PZT Piezoelectric Micromachined
Ultrasonic Transducers
Yuri Kusano , Qi Wang , Student Member, IEEE, Guo-Lun Luo, Yipeng Lu, Member, IEEE, Ryan Q. Rudy,
Ronald G. Polcawich, Senior Member, IEEE, and David A. Horsley, Senior Member, IEEE

Abstract— In this paper, we report an air-coupled piezoelectric voltage to output pressure depends on the bias voltage and
ultrasonic micromachined transducer (PMUT) using a lead- the inverse square of the capacitor gap. Therefore, CMUTs
zirconate-titanate (PZT) piezoelectric layer. A dc bias voltage achieve high electromechanical coupling when the capacitor
applied to the PZT film controls its polarization and intrinsic
stress, tuning the frequencies of two closely-spaced resonance gap is small, on the order of 100 nm to 200 nm. However,
modes of the rectangular shaped PMUT. At an optimal bias air-coupled operation requires much larger gaps, on the order
voltage of approximately 5 V, the modes nearly overlap at of 1 to 3 microns, in order to achieve sufficient sound
230 kHz, increasing the bandwidth by a factor of 8 relative to pressure output. Maintaining high electromechanical coupling
the zero dc biased state. Measurements of the electromechanical in an air-coupled CMUT therefore typically requires high bias
coupling coefficient of the PZT film show that it is maximized at
5-6 V, agreeing with device performance experiments. Acoustic voltages of 150 V or more [3]. For example, recent work
transmission and reception were demonstrated using two identi- on air-coupled CMUTs with perforated membranes achieved
cal PMUTs by adding the optimum dc bias of 6 V to the 1.4 V improved performance and bandwidth by biasing at 250 V [4].
peak-to-peak ac voltage which is maximum within the linear In contrast, piezoelectric-MUTs (PMUTs) [5] can be uti-
displacement regime. The signal detectable range of the transmit lized without high bias voltages, resulting in simpler electronic
and receive measurement with optimum dc bias tuning was 4 cm
to 19 cm in air. [2017-0223] interfaces. A typical PMUT device vibrates in a flexural
mode to emit ultrasound into the surrounding environment,
Index Terms— Microelectromechanical devices, piezoelectric air or fluid. Taking advantage of improved availability of high-
transducers, ultrasonic, PMUT, PZT, dc bias.
quality piezoelectric films, PMUTs have recently been demon-
strated for many applications, for instance, range-finding [6],
I. I NTRODUCTION low-power 3-D ultrasonic imaging [7], and ultrasonic finger-
print sensors [8]. Among many piezoelectric MEMS devices,
U LTRASONIC transducers are widely used in various
applications including medical imaging, nondestructive
evaluation, object recognition, and range-finding [1], [2].
lead-zirconate-titanate (PZT) and aluminum nitride (AlN) are
the two most commonly-used piezoelectric materials [9], [10].
Based on MEMS technology, micromachined ultrasonic trans- While the low dielectric constant of AlN makes it the best
ducers (MUTs) have many advantages over conventional ultra- material for receiving ultrasound, PZT’s much higher piezo-
sonic transducers such as miniature size, low cost, low power electric coefficient makes it the best material for transmitting
consumption, wafer-scale fabrication, ability to create 1-D and ultrasound. Here, we study PMUTs fabricated from thin-film
2-D array structures, and easy integration with supporting PZT with the intent to reduce the required transmit voltage
electronics. from ∼30 V [7] to below 10 V, reducing the cost and
The capacitive-MUT (CMUT) is one of the best-known complexity of the supporting electronics.
MUTs. In a CMUT, the electromechanical coupling from input A second goal of this work is to increase the PMUT’s band-
width. In pulse-echo range-finding, the transducer’s bandwidth
Manuscript received September 14, 2017; revised November 10, 2017; contributes to several important performance parameters:
accepted January 5, 2018. This work was supported by Berkeley Sensor and (1) the minimum duration of the transmitted pulse which in
Actuator Center Industrial Members. Subject Editor R. Pratap. (Corresponding
author: Yuri Kusano.) turn determines the minimum measurement range (since the
Y. Kusano is with the Department of Electrical and Computer Engineering, transducer cannot receive while it is transmitting); (2) the
University of California at Davis, Davis, CA 95616 USA (e-mail: ykusano@ axial resolution, which is the minimum separation between two
ucdavis.edu).
Q. Wang, G.-L. Luo, and D. A. Horsley are with the Department of resolved objects; and (3) the RMS range noise [7]. In addition,
Mechanical and Aerospace Engineering, University of California at Davis, transducers with wide bandwidth enable more sophisticated
Davis, CA 95616 USA. signal processing schemes such as pulse encoding (e.g. chirped
Y. Lu was with the University of California at Davis, Davis, CA 95616
USA. He is now with Qualcomm, Santa Clara, CA 94720 USA. transmissions).
R. Q. Rudy and R. G. Polcawich are with the U.S. Army Research Fluid-coupled PMUTs used at high resonance frequen-
Laboratory, Adelphi, MD 20783 USA. cies (>5 MHz) often have high bandwidth (>30%) because
Color versions of one or more of the figures in this paper are available
online at http://ieeexplore.ieee.org. fluid damping is high compared to air damping. Additionally,
Digital Object Identifier 10.1109/JMEMS.2018.2797684 the higher speed of sound in fluid means that the size of a
1057-7157 © 2018 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission.
See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.

2 JOURNAL OF MICROELECTROMECHANICAL SYSTEMS

TABLE I
Y OUNG ’s M ODULUS AND L AYER T HICKNESSES OF PMUT D EVICE

fluid-coupled PMUT is usually large relative to the transmitted


wavelength. These factors allow efficient coupling in immersed
PMUTs [11], [12]. On the contrary, PMUTs operating in air
require lower frequencies (<1 MHz) and are usually smaller
than the wavelength (λ = 1.9 mm at 175 kHz), resulting
in poor coupling and low bandwidth (1%–5%) [13]. An air-
coupled PMUT with 10% bandwidth was reported [14], but
required a Helmholtz resonant tube etched into the backside
of MUT to achieve this result. Another method to extend
the bandwidth is to use multiple transducers with closely-
spaced natural frequencies so that the composite bandwidth
is the sum of the individual transducer bandwidths; this
approach has been demonstrated with an array that contains
five different membranes of slightly different dimensions [15].
Here, we use a rectangular membrane design and, through
appropriate selection of the length-to-width ratio, demonstrate Fig. 1. (a) Optical image, (b) schematic top view and (b) cross-section of
that two vibration modes with closely-spaced resonance fre- the air-coupled PZT PMUT.
quencies can be achieved in a single membrane [2]. Because
these modes have different sensitivities to piezoelectric stress,
we are able to tune the frequency spacing of the two modes by
adjusting a dc bias voltage applied across the PZT layer. At the
optimum bias point, a 12% bandwidth is achieved without the
need for a Helmholtz resonator.

II. PMUT D ESIGN


An optical top view image, a schematic top view and a
schematic cross-section of a 1200 μm × 230 μm rectangular
shaped PMUT is shown in Fig. 1(a)-(c). Table 1 lists the
elastic modulus and layer thicknesses of the materials in the
device [16]–[18]. Note that only the inner top electrodes were
Fig. 2. PMUT fabrication process flow: deposition and patterning of (a) PZT,
used in this work. bottom electrode, and top electrode, (b) Al2 O3 and Au/Pt/Cr, (c) sacrificial
photoresist and thick Au layer, (d) photoresist removal and Si release through
A. Fabrication frontside XeF2 etching.

A simplified fabrication process flow is shown in Fig. 2.


The fabrication process starts by growing a 500 nm layer of top electrode of the device. The second etch removes the PZT
SiO2 on a silicon wafer via thermal oxidation. Titanium is and bottom platinum from the field, stopping on the silicon
sputtered with a thickness of 30 nm and oxidized in a furnace dioxide layer. The bottom electrode is accessed by a third ion-
to create titanium dioxide which creates a good template for mill step followed by a PZT wet-etch using a combination of
the subsequently sputtered, (111)-textured platinum electrode HCl:HF:H2 O (120:1:240) to remove any residual PZT. The
with a thickness of 100 nm [19]. The oriented platinum wafers are annealed at 500°C in an oxygen environment to
forms a templating layer for a 475 nm thick, chemical- mitigate any physical etch and hydrogen damage from the ion
solution deposited, (001)-textured PZT film with a Zr/Ti ratio milling steps.
of 52/48 [20]. The PZT is then capped with a 100 nm thick With the actuator stack patterned, a tri-layer Au/Pt/Cr
iridium oxide electrode that is sputter deposited and furnace (730nm/20nm/20nm) is deposited using electron beam evap-
annealed in oxygen at 650 °C for 30 minutes. Next, the actu- oration and patterned via liftoff, Fig. 2(b). This gold layer is
ator stack is patterned using three argon ion milling steps. used for contact pads for electrical probe testing. The silicon
In the first etch, the iridium oxide is patterned to define the dioxide layer is then patterned by reactive ion etching to define
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.

KUSANO et al.: EFFECTS OF DC BIAS TUNING ON AIR-COUPLED PZT PMUTs 3

Fig. 3. SEM image of (a) top view (b) cross-section of the air-coupled
PZT PMUT.

the etch holes for release. Next, a sacrificial photoresist layer


is defined, and a thick gold layer is deposited via electron
beam evaporation and liftoff, Fig. 2(c). This gold layer bridges
from the top electrode to the metal interconnect on the silicon
dioxide in the field without shorting the top and bottom
electrodes. The sacrificial photoresist layer is removed with an
oxygen plasma to form the Au air bridge. Finally, the wafer
is exposed to XeF2 , which isotropically etches the underlying
silicon, releasing the device, Fig. 2(d).
Fig. 3 shows SEM images of the top view and cross-section
of the PMUT following XeF2 release. As shown in the figure,
the top electrode is divided into two sections, separated by
etch-holes that perforate through the PZT/SiO2 layer. The total
3 μm thick gold layer at the left and right edge of the PMUT
Fig. 4. (a) Measured bipolar dielectric constant εr33 and tan δ ver-
stiffens these edges such that variations in the Si undercut in sus bias voltage demonstrating the polarization switching. (b) Comparing
the XeF2 etch have minimal effect on the PMUT’s resonance transverse piezoelectric coefficient e31, f and dielectric constant εr33 of
frequency [10]. Small etch holes were designed along the the PZT 52/48 thin film versus bias voltage. (c) Calculated figure of
merit (e31, f )2 /ε0 εr33 .
center of the membrane to enable the membrane to be released
through a front side etch, since this process allows closer
spacing of PMUTs than processes that require a through-wafer coefficient is inherent to all ferroelectric materials and is
etch [21], [22]. As demonstrated in recent work on air-coupled related to the polarization reversal mechanisms within the
CMUTs, small perforations in the MUT membrane have a ferroelectric material. In particular, e31, f diminishes as the
negligible effect on the output pressure [23]. voltage is increased above 3V. While this reduction is due
to many factors, one important factor is that the piezoelectric
B. PZT Film Characterization coefficient is proportional to the dielectric constant. Once the
The dielectric constant, εr33 , and corresponding loss tan- film is fully polarized, reductions in the dielectric constant
gent, tan δ, of the PZT film were measured as a function tend to reduce the piezoelectric coefficients. [24], [25]. As the
of bias voltage over a range from 0V to 10 V, Fig. 4(a). nonlinear relationship with voltage is unique to ferroelectric
The transverse thin-film piezoelectric coefficient e31, f and materials such as PZT, non-ferroelectric piezoelectric materials
dielectric constant εr33 are compared in Fig. 4(b). The laser such as AlN will not exhibit these characteristics. Moreover,
Doppler vibrometer (LDV) was used to measure the deflection the ability to tune the bandwidth by modulating the stress in
of a cantilever beam under varying bias voltage. The change the piezoelectric films through bias voltage will be limited.
in PZT stress that would be necessary to create the deflection- A figure of merit (FOM) for transducer performance is the
bias profile was extracted, and then e31, f was determined transverse electromechanical coupling coefficient kt2 , which is
by differentiating with respect to the electric field. Note that proportional to (e31, f )2 /ε0 εr33 [GPa] [26]. The FOM, plotted
the film reaches maximum piezoelectric coefficient at 3 V, in Fig. 4(c), reaches a maximum of 8 GPa at 6 V, and
corresponding to an electric field of 63 kV/cm. At voltages is roughly constant over the range from 5 V to 8 V. The
above 5V the piezoelectric coefficient is seen to decrease. peak FOM does not occur at the voltage that maximizes the
The dielectric constant is maximum at 1 V, after which piezoelectric coefficient. Instead, it peaks at a much higher
it decreases steadily with voltage. The voltage-dependence voltage where the dielectric constant is lower. Note that the
of the dielectric constant, dielectric loss, and piezoelectric values in Fig. 4 were collected from a particular wafer which
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.

4 JOURNAL OF MICROELECTROMECHANICAL SYSTEMS

Fig. 7. Frequency response of the PZT PMUT measured via LDV in air
with dc bias voltages of 0 V, +6 V, and +10 V. The inset shows a quarter of
Fig. 5. Measured I-V curve of a PMUT device with a 0.475 μm thickness
the membrane with the two mode shapes from FEM simulation.
PZT film, where breakdown voltage was 35 V and breakdown field was
760 kV/cm. The inset optical image shows the top view of Au air-bridge
collapse that occurred at the breakdown. III. R ESULTS
A. Frequency Response
The frequency response of the PMUT in air was measured
via LDV showing the presence of two vibration modes with
closely spaced natural frequencies near 180 kHz. The LDV
measurements were done at a single point of the membrane
indicated on Fig. 1(b). The measurement location was chosen
on the top electrode surface to avoid the etching holes and the
transparent SiO2 surface. A dc bias voltage applied to pole
the PZT layer creates in-plane piezoelectric stress that enables
the frequencies of the two modes to be tuned. As the dc bias
voltage was varied between ±10 V, a significant change in
the resonant frequencies, displacement sensitivities, and 3-dB
bandwidth was clearly observed for the two resonance modes.
In Fig. 7, the frequency response of the two resonance modes
Fig. 6. Measured profile of the PMUT under dc bias voltages of 0 V, +6 V, with dc bias voltages of 0 V, +6 V, +10 V respectively
and +10 V. The plot shows the cross-section height profile across the width are shown. The FEM-simulated mode shapes of the two
of the rectangular PMUT membrane.
modes ((1,1) mode and (1,3) mode) are shown in the inset.
contained the devices used for the following characterizations. The two modes independently have 2.4 kHz bandwidth with
displacement sensitivity of approximately 2000 nm/V with
The same FOM trends were observed among the wafers,
small ac voltage (−10 dBm) and without any dc bias voltage.
although the FOM for this specific wafer was not the highest.
Current-voltage (I-V) measurements were conducted to As a result of increasing the dc bias to +6 V, the highest
3-dB bandwidth of 20.3 kHz with displacement sensitivity
identify the breakdown voltage of the PZT film. The break-
down voltage was measured to be 35 V, corresponding to a of 5000 nm/V were observed. The bias voltage of 6 V
coincides with the value when the kt2 is maximized, Fig. 4(b).
breakdown field of 760 kV/cm as Fig. 5 shows. The inset
For a rectangular membrane clamped at all four boundaries,
of Fig. 5 shows that, when the applied voltage reaches the
breakdown, the Au bridge interconnecting top electrode and the stress-free resonance frequency of the (n, m) mode f nm can
be calculated from [29] and [30]:
the probing pad collapsed.    
The static shape of the PMUT membrane was measured πt E m 2  n 2
using a Wyko optical profilometer system. The cross-section f nm =   + (1)
4 6 1 − υ2 ρ a b
height profile across the shorter width of the rectangular
PMUT membrane is shown in Fig. 6, where the height at where t is the total membrane thickness, E, ν, and ρ are
SiO2 surface is set to zero. At 0 V bias, the center of the the average Young’s modulus, Poisson’s ratio and mass den-
membrane is deflected 3.6 μm above the membrane anchors sity, and a, b are the length and width of the membrane
due to the large residual stress gradient through the composite (a  b). Here, we used a = 1200 μm, b = 230 μm
stack [27], [28]. The membrane’s curvature is reduced as the resulting in f 11 = 121 kHz and f13 = 136 kHz. FEM
applied bias voltage is increased and at +10 V the center of simulation was used to increase the accuracy by modeling the
the membrane moves downwards by approximately 1.6 μm precise structure, resulting in stress-free frequencies f11 =
relative to the unbiased state. 134 kHz and f 13 = 155 kHz, corresponding to the mode
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.

KUSANO et al.: EFFECTS OF DC BIAS TUNING ON AIR-COUPLED PZT PMUTs 5

shapes shown in the inset of Fig. 7. Because residual stress


has a strong effect on the resonant frequencies of a thin
membrane (∼1 μm thick), simulations were also performed
using estimates of the residual stresses in the membrane layers:
−300 MPa (compressive) in the thermal SiO2 layer, 700 MPa
and 200 MPa (tensile) in the Pt and PZT layers [27], [28].
Under the above stressed condition, the FEM-simulated natural
frequencies increased to f 11 = 175 kHz and f13 = 194 kHz,
close to the experimentally-observed frequencies. Note that,
because the PZT and SiO2 layers have nearly equal thickness,
the contributions of their stresses nearly cancel, and the
dominant effect is residual stress in the 0.1 μm Pt electrode.
We also simulated the stress in the PZT layer required to
achieve the 60 kHz frequency shift observed in dc bias tuning
experiments– a 60 kHz increase in the resonant frequency
was observed when the PZT stress was increased by 40 MPa
in simulation, comparable to the 45 MPa value predicted
using the maximum value of the piezoelectric coefficient
(e31, f,max = 7.5 C/m2 ) multiplied by the electric field at 3V
bias (60 kV/cm).
Fig. 8(a), (b) shows the dependence of the PMUT resonant
frequency and displacement sensitivity when tuning the dc
bias voltage between −10 V and +10 V. Color-filled triangles
correspond to the first resonance mode and open triangles
correspond to the second mode observed. In addition, upward
pointing triangles present the points when sweeping up from
−10 V to +10 V, while downward pointing triangles present
the points when sweeping down from +10 V to −10 V.
The frequencies of the two modes are nearly overlapping
with certain dc bias voltages applied, especially −5 V / +6 V
while increasing and −5 V / +5 V while decreasing. From
the same measurement results, the dependence of the 3-dB
bandwidth on dc bias was analyzed and is shown in Fig. 8(c).
Wide 3-dB bandwidth above 20 kHz was achieved with
dc bias near ±5 V when the two closely-spaced modes
have overlapping bandwidths. For other dc bias voltages, the
3-dB bandwidth was only few kHz to 10 kHz since the two
resonance peaks were separated and only the first resonance
mode’s bandwidth was taken into account. Compared to the
non-biased condition, displacement sensitivity increased by a
factor of 2.5 and the 3-dB bandwidth increased by a factor Fig. 8. Dependence of (a) PMUT resonant frequency and (b) displacement
of 8 through dc bias tuning. sensitivity at resonance of the 1st peak () and the 2nd peak () on dc
bias tuning. (c) Corresponding 3dB bandwidth due to dc bias tuning.

B. Time Response
Increasing the bandwidth by overlapping multiple modes were used to prevent repoling of the PZT thin film. Fig. 9(a)
allows the transmission time to be shorter and hence reduces shows the driving voltage and the PMUT time-response mea-
the minimum measurable range and improves the axial resolu- sured at zero dc bias using the resonant frequency of 185 kHz.
tion in time-of-flight (ToF) measurements [2], [15]. At the bias Fig. 9(b) shows the optimum tuning result with +5 V dc bias
conditions where the bandwidth is maximum, the PMUT is using the resonant frequency of 240 kHz. The frequency of the
capable of transmitting shorter pulses for ToF measurements, driving pulse train was increased from 185 kHz to 240 kHz
where the speed of the wave, ultrasound in this case, deter- because the resonance frequency of the PMUT shifted due to
mines the distance to the target. The time response in air was the dc bias applied, as shown in Fig. 8(a). Comparing the two
demonstrated from pulsed ring-down measurement via LDV. time-response plots, the decay time constant τ significantly
The PMUT was excited using a function generator with a decreases from 0.28 ms to 0.05 ms.
20-cycle square wave at the resonant frequency for the two The pulse transmissions were simulated with a model of
different dc bias conditions. The drive voltage amplitude was two second-order systems in series, corresponding to the
200 mVpp for 0 Vdc and 50 mVpp for +5 Vdc to produce two resonance modes. This model was chosen based on an
similar displacement. In both cases, unipolar drive voltages empirical fit of the measured frequency response at the various
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.

6 JOURNAL OF MICROELECTROMECHANICAL SYSTEMS

Fig. 11. Membrane displacement as a function of ac driving voltage


Fig. 9. Experimentally measured pulsed-response via LDV. The PMUT amplitude, demonstrating the linear response up to 1.4 Vpp with fixed dc bias
device was excited at the resonant frequencies corresponding to (a) 0 V dc of 6 V. The inset shows the time response of the membrane displacement for
bias and (b) +5 V dc bias, demonstrating five-fold reduction of the decay different ac voltages.
time constant.

for 0 V dc bias and 0.05 ms for +5 V dc bias, which coincide


with the experimental results.
The approximately five-fold reduction in decay time is equal
to the increase in the bandwidth, from roughly 4 kHz at
0 V bias, to 20 kHz at +5 V bias. The minimum pulse
transmission time is approximately 4τ = 0.2 ms. In pulse-
echo measurements, the first echoes that can be received must
arrive later than the end of pulse transmission. Therefore, this
time defines the minimum range for pulse-echo measurements,
rmin = 4τ c/2 = 3.4 cm at +5 V bias, compared to a minimum
range of 19 cm at 0 V bias.

C. Maximum Linear Operating Range


Ideally, using a short and high-power pulse for the transmit
signal would give the highest ultrasound output. However,
as reported previously, a fully-clamped membrane shows an
undesired limitation in dynamic range due to nonlinearity at
Fig. 10. Simulated pulse transmission by a model with a cascaded 2nd-order high drive voltages [6], [7]. At the limit of the linear operating
system representing each of the two resonance modes measured via LDV
at (a) 0 Vdc bias and (b) +5 Vdc bias matching the experimental results range, increasing the voltage no longer produces a proportion-
in Fig. 9. ate increase in ultrasound output. Fig. 11 shows the measured
displacement amplitude with various excitation voltages for
bias voltages, in order to validate that the observed pulse-input the 6 Vdc biased PMUT, with the inset showing some of the
time-response was consistent with the observed changes in the pulsed time-responses. The displacement scales linearly with
frequency response. The normalized transfer function of the voltage up to a maximum displacement amplitude of 1.2 μm
cascaded second-order systems can be expressed as at 1.4 V peak-to-peak (Vpp) ac voltage. Considering this
result, experiments from the next section are using 1.4 Vpp for
ω12 ω22
Q1 Q2
driving ac voltage.
TF(s) = ω1 × ω2 (2)
s2 + Q1 s + ω12 s2 + Q2 s + ω22
D. Distance Sensor
where ω1 , ω2 are the two resonant frequencies (in rad/s) and To demonstrate the range-finding concept [31], [32], pulsed
Q 1 , Q 2 are the corresponding quality factors. Using the quality transmit and receive measurements were performed using two
factors and resonant frequencies measured in the frequency- identical PMUT devices, each wire-bonded to a printed circuit
domain LDV measurements (Fig. 7), the PMUT time-response board (PCB). The basic experimental setup and the result are
was simulated using Eq. (2), with the results shown in Fig. 10. shown in Fig. 12. The receive PMUT was placed 4 cm apart
The input pulse waveforms were replicated from the measure- from the transmit PMUT within an acrylic tube to reduce
ments except both amplitudes were normalized to 1. The decay the noise from the surrounding environment, Fig. 12(a). The
time-constant of the simulated pulse transmission was 0.3 ms tube was lined with an acoustically-absorbing foam in order to
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.

KUSANO et al.: EFFECTS OF DC BIAS TUNING ON AIR-COUPLED PZT PMUTs 7

Fig. 13. SNR versus distance between the transmit and receive PMUTs at
230 kHz compared with the theoretical model predicting linear spreading and
absorption loss.

loss, the ratio of the received pressure pr x to the transmitted


pressure pt x is
pr x a
G= = G ac · 10−α D (3)
pt x 4D
where G ac is the acoustic gain (=1), a is the effective
Fig. 12. (a) Experimental setup; transmit (TX) and receive (RX) PMUTs membrane length, D is the distance, α = 3.61 × 10−5 f −
are 4 cm apart in a foam-lined acrylic tube. (b) Transmitted signal sent 0.985 [dB/m] is the attenuation coefficient, and f is the
from function generator to TX PMUT, and received signal measured with frequency of the ultrasound wave. The attenuation coeffi-
RX PMUT, demonstrating the detection of ultrasound pulses on the 1st and
2nd round-trips. cient α increases with frequency and humidity [34]. Under
the condition of 60% relative humidity and room temperature,
prevent reflections from the sidewall. A burst signal was sent α is 7.3 dB/m at 230 kHz. At 20 cm distance, for example, the
from a function generator to the transmit PMUT, producing an spreading term is −77 dB and absorption loss term is −6 dB,
ultrasound wave which propagated acoustically to the receive generating total attenuation G of −83 dB.
PMUT which converted the ultrasound pressure to a charge Fig. 13 shows the measured SNR for free-space transmit-
output. The charge was detected using a low-noise charge receive measurements conducted between two PMUTs sepa-

amplifier (gain 10 V/pC, noise 90×10−21 C/ Hz) connected rated by 4 cm to 19 cm. The PMUTs were tuned optimally
to the receive PMUT and the out-of-band noise was filtered to 6 V dc bias at the 230 kHz operating frequency. The
out by post processing. theoretical limit of the detection of 12 dB SNR was nearly
Applying 6 V dc bias in addition to 1.4 Vpp drive voltage 45 cm, however, the maximum range measured was 19 cm due
at resonance frequency of 230 kHz for 5 pulse transmission to the limitation in the experimental setup. The measurements
was successfully detected at the receive PMUT, Fig. 12(b). in this section were conducted in free-space without any foams
The first ultrasound burst was received at tr x = 0.116 ms to absorb the surrounding noise, resulting in increasing the
which matches the travel distance of 343 m/s ×tr x = 3.98 cm. path-loss and noise detected at the receive PMUT. On the
The detected time of the second echo was three times of tr x other hand, the minimum detectable range was 4 cm in
agreeing with the expected travel distance. The estimated this experiment thanks to the wide bandwidth and fast time
bandwidth calculated from the rise time trise = 0.014 [ms] response achieved through dc bias tuning. For comparison,
was BW = 0.35/trise = 25 [kHz] similar to measured value via the zero biased PMUT could not detect signal at any range
LDV shown in Fig. 7. The ringdown time tringdown = 0.05 [ms] because of the overlap of the transmission signal feedthrough
was equivalent to the transmission decay time in Fig. 9(b). and the received signal at short range (less than 10 cm), while
at further ranges, the signal was too small to detect.
E. Path Loss and Free-Space Measurements
The ultrasonic wave spreads and attenuates while traveling IV. C ONCLUSION
through the air. Because it is small relative to the acoustic We have presented a method to increase bandwidth and
wavelength, the PMUT is a nearly-omnidirectional radiator, reduce transmission time of air-coupled PZT PMUTs. Through
causing a linear spreading of pressure with distance [6]. dc bias tuning at ±5 V, where the FOM of electromechan-
In addition, at 230 kHz, ultrasound is significantly absorbed ical coupling coefficient, (e31, f )2 /ε0 εr33 , is maximum, two
by air [33]. From the two factors of spreading and absorption adjacent resonance modes almost overlapped, resulting in
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.

8 JOURNAL OF MICROELECTROMECHANICAL SYSTEMS

significantly improved displacement sensitivity and 3-dB [16] Y. Li and Z. Zeng, “Elastic properties of transition metal dioxides:
bandwidth compared to zero dc bias condition. Consequently, XO2 (X = Ru, Rh, Os, andIr),” Int. J. Mod. Phys. C, vol. 19, no. 8,
pp. 1269–1275, Aug. 2008.
the time-response measurement results showed that the decay [17] S. Yagnamurthy, I. Chasiotis, J. Lambros, R. G. Polcawich,
time constant was reduced nearly five-fold by adding +5 V dc J. S. Pulskamp, and M. Dubey, “Mechanical and ferroelectric behavior
bias to the driving voltage. An optimum tuning of the dc bias of PZT-based thin films,” J. Microelectromech. Syst., vol. 20, no. 6,
pp. 1250–1258, Dec. 2011.
also increased the transmit and receive sensitivity, allowing [18] A. Adami, M. Decarli, R. Bartali, V. Micheli, N. Laidani, and
in-air transmit and receive measurement to be performed. The L. Lorenzelli, “Mechanical characterization of thin TiO2 films by
proposed dc bias tuning of the PMUT achieving wideband means of microelectromechanical systems-based cantilevers,” Rev. Sci.
Instrum., vol. 81, no. 1, p. 15109, Jan. 2010.
and short pulse demonstrated a good potential to improve the [19] D. M. Potrepka, G. R. Fox, L. M. Sanchez, and R. G. Polcawich,
minimum range and axial resolution for in-air range-finding “Pt/TiO2 growth templates for enhanced PZT films and MEMS devices,”
applications. in Proc. MRS, vol. 1299. Jan. 2011, p. mrsf10-1299-s02-04. [Online].
Available: https://doi.org/10.1557/opl.2011.53
[20] L. M. Sanchez et al., “Optimization of PbTiO3 seed layers and Pt
ACKNOWLEDGMENT metallization for PZT-based piezoMEMS actuators,” J. Mater. Res.,
vol. 28, no. 14, pp. 1920–1931, 2013.
The authors acknowledge the contributions of Brian Power [21] D. E. Dausch, J. B. Castellucci, D. R. Chou, and O. T. von Ramm,
and Joel Martin of the US Army Research Laboratory (ARL), “Theory and operation of 2-D array piezoelectric micromachined ultra-
sound transducers,” IEEE Trans. Ultrason., Ferroelect., Freq. Control,
and Steven Isaacson of General Technical Services. vol. 55, no. 11, pp. 2484–2492, Nov. 2008.
[22] F. Akasheh, J. D. Fraser, S. Bose, and A. Bandyopadhyay, “Piezoelec-
R EFERENCES tric micromachined ultrasonic transducers: Modeling the influence of
structural parameters on device performance,” IEEE Trans. Ultrason.,
[1] Y. Kusano, Q. Wang, R. Q. Rudy, R. G. Polcawich, and D. A. Horsley, Ferroelect., Freq. Control, vol. 52, no. 3, pp. 455–468, Mar. 2005.
“Wideband air-coupled PZT piezoelectric micromachined ultrasonic [23] N. Apte, K. K. Park, and B. T. Khuri-yakub, “Experimental evaluation
transducer through DC bias tuning,” in Proc. 30th IEEE Int. Conf. Micro of CMUTs with vented cavities under varying pressure,” in Proc. IEEE
Electro Mech. Syst. (MEMS), Jan. 2017, pp. 1204–1207. Int. Ultrason. Symp. (IUS), Jul. 2013, pp. 1724–1727.
[2] Y. Lu et al., “Broadband piezoelectric micromachined ultrasonic trans- [24] A. L. Kholkin, E. K. Akdogan, A. Safari, P.-F. Chauvy, and N. Setter,
ducers based on dual resonance modes,” in Proc. 28th IEEE Int. Conf. “Characterization of the effective electrostriction coefficients in ferro-
Micro Electro Mech. Syst. (MEMS), Jan. 2015, pp. 146–149. electric thin films,” J. Appl. Phys., vol. 89, no. 12, pp. 8066–8073,
[3] I. O. Wygant et al., “50 kHz capacitive micromachined ultrasonic Jun. 2001.
transducers for generation of highly directional sound with parametric [25] N. Setter et al., “Ferroelectric thin films: Review of materials, properties,
arrays,” IEEE Trans. Ultrason. Ferroelect. Freq. Control, vol. 56, no. 1, and applications,” J. Appl. Phys., vol. 100, no. 5, p. 051606, Sep. 2006.
pp. 193–203, Jan. 2009. [26] P. Muralt, “PZT thin films for microsensors and actuators: Where do
[4] N. Apte, K. K. Park, A. Nikoozadeh, and B. T. Khuri-Yakub, “Band- we stand?” IEEE Trans. Ultrason., Ferroelect., Freq. Control, vol. 47,
width and sensitivity optimization in CMUTs for airborne applications,” no. 4, pp. 903–915, Jul. 2000.
in Proc. IEEE Int. Ultrason. Symp., Sep. 2014, pp. 166–169. [27] J. S. Pulskamp, A. Wickenden, R. Polcawich, B. Piekarski, M. Dubey,
[5] P. Muralt and J. Baborowski, “Micromachined ultrasonic transducers and G. Smith, “Mitigation of residual film stress deformation in mul-
and acoustic sensors based on piezoelectric thin films,” J. Electroceram., tilayer microelectromechanical systems cantilever devices,” J. Vac. Sci.
vol. 12, nos. 1–2, pp. 101–108, 2004. Technol. B, Microelectron. Process. Phenom., vol. 21, no. 6, p. 2482,
[6] R. J. Przybyla et al., “In-air rangefinding with an aln piezoelectric 2003.
micromachined ultrasound transducer,” IEEE Sensors J., vol. 11, no. 11, [28] E. Zakar et al., “Stress analysis of SiO2 /Ta/Pt/PZT/Pt stack for MEMS
pp. 2690–2697, Nov. 2011. application,” in Proc. 12th IEEE Int. Symp. Appl. Ferroelectr. (ISAF),
[7] R. J. Przybyla, H. Y. Tang, A. Guedes, S. E. Shelton, D. A. Horsley, and vol. 2. Jul./Aug. 2000, pp. 757–759.
B. E. Boser, “3D ultrasonic rangefinder on a chip,” IEEE J. Solid-State [29] A. W. Leissa, “Vibration of plates,” NASA, Washington, DC, USA,
Circuits, vol. 50, no. 1, pp. 320–334, Jan. 2015. Tech. Rep. SP-160, 1969.
[8] Y. Lu et al., “Ultrasonic fingerprint sensor using a piezoelectric micro- [30] A. K. Mitchell and C. R. Hazell, “A simple frequency formula
machined ultrasonic transducer array integrated with complementary for clamped rectangular plates,” J. Sound Vibrat., vol. 118, no. 2,
metal oxide semiconductor electronics,” Appl. Phys. Lett., vol. 106, pp. 271–281, Oct. 1987.
no. 26, p. 263503, Jun. 2015. [31] R. Przybyla et al., “A micromechanical ultrasonic distance sensor
[9] P. Muralt et al., “Piezoelectric micromachined ultrasonic transducers with >1 meter range,” in Proc. 16th Int. Solid-State Sensors, Actuat.
based on PZT thin films,” IEEE Trans. Ultrason., Ferroelect., Freq. Microsyst. Conf. (TRANSDUCERS), Jun. 2011, pp. 2070–2073.
Control, vol. 52, no. 12, pp. 2276–2288, Dec. 2005. [32] M. Shin, J. S. Krause, P. DeBitetto, and R. D. White, “Acoustic
[10] S. Shelton et al., “CMOS-compatible AlN piezoelectric micromachined Doppler velocity measurement system using capacitive micromachined
ultrasonic transducers,” in Proc. IEEE Int. Ultrason. Symp., Sep. 2009, ultrasound transducer array technology,” J. Acoust. Soc. Amer., vol. 134,
pp. 402–405. no. 2, pp. 1011–1020, Aug. 2013.
[11] Y. Lu, H.-Y. Tang, S. Fung, B. E. Boser, and D. A. Horsley, “Pulse- [33] D. T. Blackstock, Fundamentals of Physical Acoustics. Hoboken, NJ,
echo ultrasound imaging using an AlN piezoelectric micromachined USA: Wiley, 2000.
ultrasonic transducer array with transmit beam-forming,” J. Micro- [34] H. E. Bass, L. C. Sutherland, A. J. Zuckerwar, D. T. Blackstock, and
electromech. Syst., vol. 25, no. 1, pp. 179–187, Feb. 2016. D. M. Hester, “Atmospheric absorption of sound: Further developments,”
[12] Q. Wang, Y. Lu, S. Mishin, Y. Oshmyansky, and D. A. Horsley, “Design, J. Acoust. Soc. Amer., vol. 97, no. 1, pp. 680–683, Jan. 1995.
fabrication, and characterization of scandium aluminum nitride-based
piezoelectric micromachined ultrasonic transducers,” J. Microelectro-
mech. Syst., vol. 26, no. 5, pp. 1132–1139, Oct. 2017. Yuri Kusano received the B.E. degree in applied
[13] M. D. Williams, B. A. Griffin, T. N. Reagan, J. R. Underbrink, and physics from Keio University, Japan, in 2015, and
M. Sheplak, “An AlN MEMS piezoelectric microphone for aeroacoustic the M.S. degree in electrical and computer engi-
applications,” J. Microelectromech. Syst., vol. 21, no. 2, pp. 270–283, neering from the University of California at Davis,
2012. Davis, CA, USA, in 2017, where she is currently
[14] S. Shelton, O. Rozen, A. Guedes, R. Przybyla, B. Boser, and pursuing the Ph.D. degree with the Department of
D. A. Horsley, “Improved acoustic coupling of air-coupled microma- Electrical and Computer Engineering, with a focus
chined ultrasonic transducers,” in Proc. IEEE 27th Int. Conf. Micro on piezoelectric MEMS devices for ultrasonic sensor
Electro Mech. Syst. (MEMS), Jan. 2014, pp. 753–756. applications. She is currently a Graduate Student
[15] A. Hajati et al., “Three-dimensional micro electromechanical system Researcher with the Berkeley Sensor and Actua-
piezoelectric ultrasound transducer,” Appl. Phys. Lett., vol. 101, no. 25, tor Center, University of California at Davis. Her
p. 253101, 2012. research interests include piezoelectric MEMS sensors and actuators.
This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.

KUSANO et al.: EFFECTS OF DC BIAS TUNING ON AIR-COUPLED PZT PMUTs 9

Qi Wang (S’15) received the B.S. degree in Ronald G. Polcawich (M’07–SM’16) received the
mechanical engineering and automation from the B.S. degree in materials science and engineer-
Nanjing University of Aeronautics and Astronautics, ing from Carnegie Mellon University in 1997,
Nanjing, China, in 2011, and the M. S. degree and the M.S. degree in materials and the Ph.D.
in nanomechanics from Tohoku University, Sendai, degree in materials science and engineering from
Japan, in 2013. Penn State University in 1999 and 2007, respec-
He is currently a Graduate Student Researcher tively. He is currently a Program Manager with
with the Berkeley Sensor and Actuator Center, Uni- DARPA, Microsystems Technology Office and cur-
versity of California at Davis, where he is currently rently on detail to DARPA from the Micro & Nano
pursuing the Ph.D. degree in mechanical engineering Materials & Devices Branch, U.S. Army Research
with the Department of Mechanical and Aerospace Laboratory, Adelphi, MD, USA, while at DARPA,
Engineering. His research interests include piezoelectric thin films and MEMS he is leading research projects in advanced materials processing combined
sensors and actuators. with micromechanics for small scale robotics and developing new materials
processes, device designs, and integration approaches required to push inertial
and aiding sensor performance for enabling trusted, heterogeneous position,
navigation, and timing (PNT) systems.
He was the Team Lead for PiezoMEMS Technology, ARL, with a focus on
Guo-Lun Luo received the double-major B.S. developing component technologies to enable cognitive RF communication
degree of physics/aeronautics and astronautics engi- and radar systems and MEMS inertial and aiding sensors to provide PNT
neering from National Cheng Kung University solutions for SWAP-C constrained platforms. He currently holds 16 patents,
in 2011, and the M.S. degree in mechanical engi- has four patent applications pending review, and has authored over 100 journal
neering from National Tsing Hua University in 2013. and proceedings articles, and authored three book chapters on fabrication and
In 2013, he interned with Qualcomm, Taiwan, for design of piezoelectric MEMS devices using PZT thin films. His research
mirasol display. From 2014 to 2016, he joined interests include materials processing of lead-zirconate-titanate (PZT) thin
Asia Pacific Microsystems Inc., where he was a films, MEMS fabrication, RF components, MEMS actuator technologies,
Senior Research &Development Engineer and a mm-scale robotics, MEMS inertial sensors, and sensors for aiding inertial sys-
Project Manager for process development of MEMS tems. He and his colleagues were the recipients of the 2006 Department of the
devices. He is currently pursuing the Ph.D. degree Army Research and Development Achievement Award for Piezoelectric RF
in mechanical and aerospace engineering with the University of California at MEMS Switch Using PZT Thin Films and the 2009 U.S. Army Research Lab-
Davis. His research interests include piezoelectric MEMS devices. oratory Engineering Award for ground-breaking work on Piezoelectric MEMS.
He received the 2012 Presidential Early Career Award for Scientists and
Engineers and the 2015 IEEE UFFC Ferroelectrics Young Investigator Award.
Dr. Polcawich is currently a member of the IEEE Ferroelectrics Committee
and the Technical Program Committee IV Applications of Ferroelectrics,
Yipeng Lu received the B.S. degree in materi- served as an elected member of the IEEE Ultrasonics, Ferroelectrics, and
als science and engineering from Jilin University, Frequency Control Administrative Committee from 2014 to 2016. Since
Changchun, China, in 2007, the M.S. degree in 2016, he has been the Chair of the UFFC Membership Committee. He is
microelectronics from Shanghai Jiao Tong Univer- on the Technical Advisory Committee for the PiezoMEMS Workshop,
sity, Shanghai, China, in 2010, and the Ph.D. degree co-organized the 2013 meeting in Washington, DC, USA, and co-organized
in mechanical engineering from the University of the 2018 meeting in Orlando, FL, USA.
California at Davis (UCD), CA, USA, in 2015. Prior
to joining the Berkeley Sensor and Actuator Center
at UCD, as a Graduate Student Researcher, he was David A. Horsley (M’97–SM’17) received the B.S.,
a Digital Hardware Engineer with Huawei in 2011. M.S., and Ph.D. degrees in mechanical engineer-
He is with Qualcomm, as a Staff Engineer. His ing from the University of California at Berkeley,
research interests include MEMS sensors and actuators. Berkeley, CA, USA, in 1992, 1994, and 1998,
respectively. He held research and development posi-
tions with DiCon Fiberoptics, Inc., Hewlett Packard
Laboratories, and OniX Microsystems, Inc., and
Ryan Q. Rudy received the B.S.E. and M.S.E. was with the Faculty at UC Davis. In 2013, he
degrees in mechanical engineering from the Univer- co-founded Chirp Microsystems to commercialize
sity of Michigan, Ann Arbor, MI, USA, in 2009 and MEMS ultrasonic transducers. He is currently a
2010, respectively, and the Ph.D. degree in mechan- Professor with the Department of Mechanical and
ical engineering from the University of Maryland Aerospace Engineering, University of California at Davis, Davis, CA, USA,
at College Park, College Park, MD, USA, in 2014, and also an Adjunct Professor with the Department of Mechanical Engineer-
with a focus on miniaturized ultrasonic motors. ing, University of California at Berkeley, Berkeley, CA, USA, and has been
He is currently a Mechanical Engineer with the the Co-Director of the Berkeley Sensor and Actuator Center since 2005.
Micro & Nano Materials & Devices Branch, His research interests include microfabricated sensors and actuators with
U.S. Army Research Laboratory, Adelphi, MD, applications in ultrasonics and physical sensors. He was a recipient of the NSF
USA. His current research interests include piezo- CAREER Award, the UC Davis College of Engineering’s Outstanding Junior
electric MEMS, specifically piezoelectric electromechanical resonators and Faculty Award, and the NSF I/UCRC Association’s Alexander Schwarzkopf
filters. Award.

You might also like