You are on page 1of 10

284

JOURNAL OF MICROELECTROMECHANICAL SYSTEMS, VOL. 21, NO. 2, APRIL 2012

Monolithic Integration of Pressure Plus Acceleration


Composite TPMS Sensors With a Single-Sided
Micromachining Technology
Jiachou Wang, Xiaoyuan Xia, and Xinxin Li

AbstractThis paper concerns the development of a single-side


micromachined tire-pressure monitoring system (TPMS) sensor
for automobiles, which is with a piezoresistive pressure sensor
and a cantilevermass piezoresistive accelerometer monolithically
integrated in a 1.6 mm 1.5 mm sized (111)-silicon chip.
Single-wafer-based front-side silicon micromachining and metal
electroplating technologies are employed to fabricate the device.
Specially designed releasing trenches along 111 orientation are
constructed to form the hexagonal pressure-sensitive diaphragm
and the postsealed vacuum reference cavity. The fabrication of the
accelerometer is also based on a hexagonal diaphragm that is latterly cut into suspended cantilevers and seismic mass. To achieve
a high sensitivity, a high-density copper thick film is selectively
electroplated to significantly increase the mass. The performance
of the 115-g-ranged accelerometer is measured, exhibiting a sensitivity of 99.9 V/g (under 3.3-V supply), nonlinearity of 0.45%
FS, and the noise floor of better than 0.2 g. The 750-kPa-ranged
pressure-sensor sensitivity is measured as 0.108 mV/kPa (under
3.3-V supply), with the nonlinearity error smaller than 0.1% FS
and the temperature coefficient of sensitivity as 0.19%/ C FS
before compensation. The noise floor of the pressure-sensor output signal is 0.15 kPa. The zero-point temperature coefficient
is tested as 0.11%/ C FS and 0.024%/ C FS for the accelerometer and the pressure sensor, respectively. Fabricated with
the low-cost front-side micromachining technique, the small-sized
TPMS sensors are promising in practical applications and volume
production.
[2011-0192]
Index TermsAccelerometer, front-side monolithic micromachining, piezoresistance, pressure sensor, sensor.

I. I NTRODUCTION

ITH the development of silicon micromachining


technology [1][5], silicon pressure sensors and
accelerometers have been separately developed and widely
used in industry and consumer application due to their tiny size,
low cost, and linear output property [6]. Recently, along with
Manuscript received June 23, 2011; revised October 11, 2011; accepted
November 13, 2011. Date of publication December 28, 2011; date of current
version April 4, 2012. This work was supported in part by the Chinese 973
Program under Grant 2011CB309503, in part by the National Natural Science
Foundation of China Projects under Grants 60725414 and 91023046, and in
part by the National S&T Major Project under Grant 2009ZX02038. The work
of X. Li was supported by the Korean World Class University project under
Grant R32-2009-000-20087-0. Subject Editor D. D. Cho.
The authors are with the State Key Laboratory of Transducer Technology
and Science and Technology on Microsystem Laboratory, Shanghai Institute
of Microsystem and Information Technology, Chinese Academy of Sciences,
Shanghai 200050, China (e-mail: xxli@mail.sim.ac.cn).
Color versions of one or more of the figures in this paper are available online
at http://ieeexplore.ieee.org.
Digital Object Identifier 10.1109/JMEMS.2011.2178117

the electronic market extended into automotive products such as


tire-pressure monitoring system (TPMS), great research effects
have been motivated again to develop monolithically integrated
multifunctional silicon piezoresistive sensors that feature
high reliability, low costs, and volume fabrication capability
[7]. It deserves to point out that, in TPMS application, only
using a pressure sensor to monitor the tire pressure is not
adequate. In order to enable the embedded battery long-term
work for several years, a microprocessor-controlled energysaving working mode needs an accelerometer to detect the
wheel-rotation-induced centrifugal acceleration. Only when
the centrifugal acceleration is detected that the pressure sensor
will be waken up for tire-pressure monitoring. Therefore, a
pressure + acceleration multifunctional sensor is needed for
the TPMS applications.
Compared with the hybrid packaging scheme for realizing
simultaneous detection of pressure and acceleration, monolithically integrated multifunctional sensor scheme has the
advantages of smaller size, lower cost, higher reliability, and
higher fabrication throughput. However, monolithic fabrication
of multifunctional sensor is relatively complex, and the compatibility issue in fabricating different sensing elements should be
well solved at first. With the rapid development of silicon micromachining technology, to date, the fabrication technologies
for some multifunctional sensors have been under exploration,
and more and more monolithic multifunctional sensors have begun to be reported [6], [8]. Most previously reported multifunctional sensors that are fabricated with surface-micromachining
techniques usually use deposited thin films such as polysilicon
film [9], [10] and silicon nitride film [6], [11], where SiN film
was used to form pressure-sensitive diaphragm and polysilicon film was used for sensing piezoresistors. Such surfacemicromachining techniques have the following disadvantages.
The piezoresistive sensitivity of polysilicon is much lower than
that of single-crystalline silicon [12]. Moreover, deposited SiN
film suffers from residual stress that severely influences the
sensor performance. In general, bulk-silicon-micromachining
or bulk/surface combined micromachining techniques are
more suitable for fabrication of high-performance mechanical
sensors.
A monolithically integrated pressure accelerometer and temperature composite sensor was proposed by Wang et al. [6] and
fabricated in a 2.5 2.5 mm2 single chip by using bulk/surface
combined micromachining fabrication. The accelerometer design was based on heat-convection-sensing mechanism. The

1057-7157/$26.00 2011 IEEE

WANG et al.: MONOLITHIC INTEGRATION OF PRESSURE PLUS ACCELERATION COMPOSITE TPMS SENSORS

accelerometer is not suitable for the long-term energy-saving


TPMS applications as the sensor consumes a relatively high
power for thermal convection. Moreover, a closed cavity with
a relatively large volume is needed for the heat convection
that makes difficulties in reducing device size and decreasing
fabrication cost.
Xu et al. [8] reported a multifunctional sensor fabricated
with double-sided micromachining process and wafer-bonding
technique. Two-step anisotropic etching from the wafer backside was processed to form the pressure diaphragm and the
seismic mass, where two layers of mask were used. Then,
wafer-bonding process was used to seal the pressure-reference
cavity and form the sensor chip frame [13][17]. Finally, the
movable detecting beam structures are released by using deep
reactive ionic etch (RIE) from the silicon wafer front side. Since
the bonding frame and the anisotropic-etching-formed inclined
sidewalls occupied a large device size, the sensor chip size is
quite large, and the fabrication cost cannot be very low. Moreover, the residual stress resulting from the bonding structure
easily causes instability of the sensor, particularly in varied
environmental temperature. With the double-side micromachining scheme, the backside etching formed thickness values of
the diaphragm, and the beams are not easy to be with the wellcontrolled uniformity, due to the wafer thickness nonuniformity
and the nonuniformity in deep etching rate within the whole
wafer. Therefore, high fabrication yield, high uniformity, and
low-cost composite TPMS sensors are somewhat difficult to
be volume manufactured by using the wafer-bonding-based
double-sided micromachining processes.
To solve the aforementioned problems induced by either the
double-sided bulk-micromachining or the conventional surfacemicromachining techniques, herein, we develop a novel singlewafer-based front-side micromachining technique to fabricate
on-chip integrated multifunctional TPMS sensor. The proposed
techniques are the further development work on the basis of
a surface and bulk micromachining (SBM) process that was
proposed by Prof. Chos group for fabricating advanced micromachining structures like comb-drive electrostatic actuators,
capacitive inertial sensors, etc. [18][21]. The SBM process can
be employed to fabricate a lot of surface micromachined structures, where single-crystalline silicon is used as the structural
layers instead of polysilicon or silicon nitride thin films. In this
way, the thin-film residual stress problem can be avoided, and
the higher aspect ratio silicon structures facilitate elimination of
mechanical crosstalk signal, thereby enhancing micromechanical performance. This paper further develops the techniques
for fabrication of typical piezoresistive-sensing structures for
TPMS multifunctional sensors, where large-area hexagonalshaped silicon diaphragms can be formed by defining a series of
small openings along 211 orientation. For underetch releasing
large-area diaphragms, the design rules for the layout of the
small openings are given. After the opening holes are sealed
in vacuum, a piezoresistive absolute pressure sensor and a
piezoresistive accelerometer are designed and monolithically
integrated for TPMS applications. Such single-side micromachined piezoresistive sensors have advantages of small chip
size, high fabrication yield, low cost, and more compatibility
with semiconductor foundry process for volume production.

285

Fig. 1. Three-dimensional schematic of our proposed TPMS composite sensor, with front-side micromachined hexagonal diaphragm on top of a vacuumsealed reference cavity, two cantilevers, and a Si/Cu seismic mass.

In the following sections, a technical description on design,


fabrication, and characterization of the composite sensor will
be given sequentially.
II. D ESIGN OF THE TPMS C OMPOSITE S ENSOR
A. Sensor Configuration
Fig. 1 shows the 3-D schematic of our proposed monolithic
TPMS sensor structure that consists of an accelerometer and a
pressure sensor.
The accelerometer sensor is with the cantilevermass structure first fabricated into a hexagonal diaphragm that is latterly
etched into suspended cantilevers and seismic mass. Since the
hexagonal diaphragm for the accelerometer is formed simultaneously with the pressure-sensing diaphragm, the structural
thickness of the two parts has to be identical. In order to achieve
a high sensitivity for the pressure sensor, the thickness of the
diaphragm is designed as only 10 m. Such a thin silicon
diaphragm is with a very small seismic mass that leads to a very
low sensitivity. Herein, we employ electroplated thick copper
film to significantly increase the mass and the sensitivity to
acceleration, as the density of Cu is much higher than that of
silicon. To achieve a high sensitivity, high-density copper thick
film is selectively electroplated on the surface of the mass location to largely increase the mass. The gap between the seismic
mass and the substrate is embedded into the single silicon chip
rather than being formed by siliconsilicon (or siliconglass)
bonding. Compared to the conventional cantilevermass accelerometer structure, herein, the gap can be controlled by
fabrication process to optimize the squeeze-film damping and
overrange protection capability. Two piezoresistors are located
near the roots of the two cantilevers, respectively, and interconnected to form a half-bridge readout circuit (together with other
two reference resistors) for acceleration detection.
For the design of the pressure sensor, we directly use the
simultaneously formed single-crystal silicon hexagonal-shaped
diaphragm as the pressure-sensitive diaphragm. Compared with

286

JOURNAL OF MICROELECTROMECHANICAL SYSTEMS, VOL. 21, NO. 2, APRIL 2012

Fig. 2. Hexagonal-shaped silicon diaphragm formation scheme in (111) wafer. (a) Lateral underetch and bottom-release evolution is top view illustrated.
(b) Lateral underetch evolutions along AA cross section and BB  cross section are schematically shown.

the traditional bulk-micromachined pressure sensor structure


reported in [7], [14][16], herein, the reference-pressure cavity
beneath the diaphragm is embedded into the silicon substrate by
using lateral underetch. The underetch is along 110 orientation with all the boundary walls as {111} planes. Four piezoresistors are optimally designed at the maximum stress locations
of the diaphragm to form a full-sensitive Wheatstone bridge.
To design p-type piezoresistance in (111) silicon, Eulers
angle-based coordinate-system transformation for the piezoresistive coefficient is used that was theoretically presented in
[12]. Based on analytical calculation for the piezoresistance
in different crystal orientations, we make clear that both the
longitudinal and the transverse piezoresistive coefficients are
44 /2 and 44 /6, respectively. The two piezoresistive coefficient values are always unchanged for any crystal orientation in
(111) plane. Moreover, the shearing piezoresistive coefficient is
always zero [22]. According to the homogeneous piezoresistive
properties, the design for the piezoresistors mainly relies on the
location where maximum stress (in fact, the difference between
the longitudinal and transverse directions) can be induced by
deflection of the hexagonal diaphragm.

b must be larger than the distance w. Using trigonometry


calculation, b can be expressed as
b = tan(19.47 ) h w.

According to (1), by optimizing the gap w or the etching


depth h, the square area between the two adjacent trenches
can be completely released by lateral underetch along 211
orientation. Based on (111)-silicon anisotropic etching theory
[18][21], [23], now, we are concerned with the evolvement of
the etching front of the trench 1 and trench 3 (or the trench 2 and
trench 4) along the 110 orientation. If the lateral underetch
evolutions developed from the bottom side of the trench 1 (or
trench 2) and the upper side of the trench 3 (or trench 4) do not
intersect each other, the etch process should stop at these points
of O1 , O2 , O3 , and O4 that are shown in Fig. 2. To release the
square area between the trench 1 and trench 3 (or the trench 2
and trench 4), the distance n must be smaller compared to the
length of 2m + w to guarantee intersection of the underetch.
Thereby, the following relationship among the distance n, the
length m, and the distance w should be satisfied as
n [i m + (i 1) w] tan(30 )

B. Formation Mechanism of the Hexagonal Diaphragm


Fig. 2 shows the formation mechanism of the hexagonalshaped diaphragm in (111)-wafer where aqueous alkaline
anisotropic etchant is used. Four trenches with identical dimensions and a certain depth are etched using deep RIE. Then, lowstress silicon nitride layer is formed to cover and protect the
vertical sidewalls. As is shown in the cross section A A in
Fig. 2, the lateral underetch evolution between trench 1 and
trench 2 (or between trench 3 and trench 4) along the 111
orientation is bounded by {111} planes. In order to completely
release the square area between the two trenches, the distance

(1)

(2)

where i is the number of the trenches in one row along 111


orientation. In the case shown in Fig. 2, i = 2. According to
(1) and (2), various sizes of the hexagonal-shaped diaphragms
can be formed by the lateral underetch where {111} planes are
used as the etching stop sidewalls, and the lateral underetch can
excavate through the bottom of a certain number of trenches
that are laid out into a row aligned to 111 orientation. After
etching through from the structure bottom, a hexagonal-shaped
cavity structure is formed beneath the diaphragm with all the
boundary walls as {111} planes (shown with the BB  cross
section in Fig. 2).

WANG et al.: MONOLITHIC INTEGRATION OF PRESSURE PLUS ACCELERATION COMPOSITE TPMS SENSORS

287

than that of the cantilever-end loaded mass and can be ignored,


M can be expressed as




(a1 +b2 )2
a21 tan 30 (h2 Si +h3 Cu ) (4)
M a2 b2 +
2
where a2 and h2 are the seismic-mass length and thickness,
Si and Cu are the density of single-crystal silicon and the
electroplated copper, respectively, and h3 is the thickness of the
copper thick film. Since the Si/Cu mass is much thicker than
the cantilever, deflection of the mass is reasonably negligible.
Therefore, the maximal deflection wmax occurs at the far end
of the mass (i.e., x = a1 + b2 ), with the value as
wmax (a1 +b2 )
=


Ma 
12(a1 +0.5b2 )2 15(a1 +0.5b2 )a1 +5a21 a1 . (5)
3
Eb1 h1
The generated stress on the cantilever surface can be as
T (x) = Eh w (x) =

Fig. 3. ACES simulation results of the hexagonal diaphragms for the pressure
sensor (denoted as -a series) and the beammass region of the accelerometer
(denoted as -b series). Step (1): Etching initial from the rows of releasing
trench that are designed according to analysis. Step (2): Etched shapes at some
moment of midway. Step (3): Final shapes of the formed hexagonal diaphragms.

Based on the analysis shown earlier, two rows of releasing


trenches, which are along 111 orientation, are designed for
the pressure-sensitive diaphragm, and six rows of releasing
trenches are designed for the even larger beammass area of
the accelerometer. The professional etching software of ACES
is used to simulate the evolution of the etching front and the
formation process of the etched micromechanical structures,
with the results shown in Fig. 3. The dark regions denote the
released hexagonal-shaped diaphragm.
C. Design of the Accelerometer
Please refer to the accelerometer configuration schematically
shown in Fig. 1. Based on mechanical analysis [12], the deflection w(x) along the sensing direction of the cantilever (i.e., the
normal direction to the (111) wafer) can be expressed as

w(x) =

M a(3a1 + 1.5b2 x)x2


Eb1 h31

(3)

where a is the externally applied acceleration; a1 , b1 , and h1


are the cantilever length, width, and thickness, respectively; b2
is the seismic-mass width; E is Youngs modulus of silicon; and
M is the total effective mass of the cantilevermass structure.
Since the self-mass of the narrow cantilevers is much smaller

3M a(a1 + 0.5b2 x)
b1 h21

(6)

where h is the distance between the surface of the cantilever


and the neutral plane (i.e., the middle plane) of the cantilever.
The maximum stress for piezoresistance is located near the root
of the cantilever. The designed piezoresistor pattern is shown
in Fig. 1. With the half-bridge circuit, the accelerometer output
voltage is
Vout

344 M aVsupply
=
4lp b1 h21

lp
(a1 + 0.5b2 x) dx
0

344 M a
=
(2a1 + b2 lp )Vsupply
8b1 h21

(7)

where 44 is the piezoresistive coefficient component [12],


Vsupply is the supplied voltage on the half-bridge circuit, and
lp is the piezoresistor length. The sensitivity can be analyzed as
S=

344 M
(2a1 + b2 lp )Vsupply .
8b1 h21

(8)

The natural frequency of the accelerometer is expressed


as [16]

2Eb1 h31
1
(13a21 + 6.5a1 b2 + b22 ). (9)
f0 =
2
4(2a1 + b2 )
M a1
In our design of the accelerometer, the cantilever length,
width, and thickness are finalized as 80, 34, and 10 m, respectively. The seismic-mass length, width, and thickness are 601,
460, and 10 m, respectively. The copper thick-film thickness is
20 m, and the piezoresistor length is 60 m. For a 3.3-V power
supplied on the half-bridge circuit, the sensitivity is designed
as 0.103 mV/g under 3.3-V supply, the natural frequency is
6.93 kHz, and the measured range is 115 g. The pressure sensor
waking-up threshold is normally around 20 g that is according
to the car speed of 25 km/h and is also dependent on the
wheel size.

288

JOURNAL OF MICROELECTROMECHANICAL SYSTEMS, VOL. 21, NO. 2, APRIL 2012

Fig. 4. Schematic top-view piezoresistor layout on the hexagonal diaphragm.

According to (8), the sensitivity of the accelerometer would


be only 0.013 mV/g if there is no electroplated 20-m-thick
copper to reinforce the seismic mass (i.e., only the thin silicon
as the mass). The sensitivity of the accelerometer with the highdensity copper is 7.9 times as that of the accelerometer without
the copper.
D. Design of the Pressure Sensor
According to the formation mechanism of the hexagonalshaped diaphragm, two rows of releasing trenches are designed
for the pressure sensor [see Fig. 3(a)]. The piezoresistor layout
of the hexagonal diaphragm is shown in the top-view schematic
in Fig. 4. Since an analytical modeling for the hexagonaldiaphragm pressure sensor is difficult, finite-element-analysis
software of ANSYS is used to study the stress distribution on
the diaphragm. According to the piezoresistive pressure sensor
design rules given in [12], the locations with maximum difference between longitudinal stress component and transverse
stress component are suited for laying out the piezoresistors for
highest sensitivity.
Fig. 5 shows the ANSYS-simulated stress distribution on the
upper surface of the diaphragm for various lengths and widths.
The input pressure is set as 450 kPa. For given diaphragm width
and thickness of w = 56 m and h = 10 m, the simulation
results in Fig. 5(a) show that, when the diaphragm length
increases from 300 to 400 m, the stress distributions along
the straight line aa in Fig. 4 are always symmetrical. The
maximum tensile stress in x-direction is at the center of the
diaphragm, with the values increasing from 20 to 32.5 MPa.
About 17.0-MPa, maximum compressive stress occurs at both
terminals of the diaphragm. The maximum tensile stress in
y-direction is also at the center of the diaphragm, with the
value increasing from 27 to 43 MPa. It is worthy pointing out
that, at the location where x-direction stress equals zero, the
y-direction tensile stress will increase with the diaphragm
length of l. Similarly shown in Fig. 5(b), for given length and
thickness of l = 300 m and h = 10 m, when the diaphragm
width w increases from 56 to 156 m, the top-surface stress
along bb in Fig. 4 also shows a symmetrical distribution.
Along the line aa or the line bb in Fig. 4, at the locations of zero x-direction stress, the corresponding y-direction

Fig. 5. Simulated stress distribution on the top surface of the hexagonal


diaphragm. (a) Along the straight line aa in Fig. 4, the x- and y-direction
normal stress distributions are shown. (b) Along the straight line bb in Fig. 4,
the normal stress distributions in x- and y-directions are shown.

stress along the line bb is compressive [see Fig. 5(b)] with
the absolute value being almost equal to that of the tensile
y-direction stress along the line aa [see Fig. 5(a)].
It is well known that, the thinner the diaphragm, the greater
the stress and the higher the sensitivity to pressure. However,
when the diaphragm is too thin, the linearity of the sensor
will be deteriorated. In our design, the diaphragm thickness is
designed the same as the cantilever thickness of 10 m. Consequently, the sensitivity of the sensor is a function of the length l
and the width w. According to the ANSYS simulated stress,
four piezoresistors are optimally designed at the maximumstress locations that are shown in Fig. 4. Two piezoresistors
(R1 and R3 ) are located near both sides of the diaphragm and
are along the line aa , where the x-direction stress is almost
zero and the y-direction stress gets the maximum tensile value.
The other two piezoresistors (R2 and R4 ) are located near both
sides of the diaphragm along the line bb , where the x-direction
stress is also near zero and the y-direction stress gets the maximum compressive value. Under applied pressure from the top
side of the diaphragm, R1 and R3 will increase the resistance
value, whereas R2 and R4 will decrease the resistance value.
The four resistors form a fully sensitive Wheatstone bridge.
With the designed geometric parameters shown in Table I and

WANG et al.: MONOLITHIC INTEGRATION OF PRESSURE PLUS ACCELERATION COMPOSITE TPMS SENSORS

TABLE I
D ETAILED D IMENSIONS FOR P RESSURE S ENSOR

Fig. 6. Main steps of the single-wafer-based front-sided micromachining


process for the TPMS sensor.

the ANSYS simulated stress data, the pressure sensitivity is


calculated as 0.112 mV/kPa under 3.3-V supply.
III. FABRICATION
The TPMS composite sensor is fabricated with a novel
single-wafer-based single-sided bulk-micromachining technology. The fabrication process flow is shown in Fig. 6 and
detailedly described as follows.
1) The initial substrate is single-side polished n-type (111)
4-in silicon wafers (double-side polished wafer is not
needed as all the micromachining processes are performed from the front side), with a resistivity of
310 cm and a thickness of 450 m. After thermal
oxidization, photolithographic steps are conducted from
the front side of the wafer for patterning piezoresistors
on a photoresist layer. Then, buffered hydrofluoric acid is
used to etch SiO2 , with the photoresist as etching mask.
Piezoresistors are formed by boron ion implantation followed by drive-in process. The doping concentration is
controlled at the level of about 1 1019 /cm3 .

289

2) Low-pressure chemical vapor deposition (LPCVD) lowstress silicon nitride layer and silicon dioxide layer are
sequentially deposited and patterned. The deposited passivation layers are used as mask for following silicon deep
RIE and the final aqueous alkaline lateral underetch for
the bottom release of the diaphragms. Then, deep RIE is
performed to a depth of 10 m to define the thickness of
the diaphragms. The trench opening width is 4 m.
3) A 0.2-m-thick low-stress nitride film and a 0.2-mthick TEOS oxide film are sequentially deposited with
LPCVD to protect the vertical surface of the silicontrench sidewalls from following anisotropic wet etching. These passivation layers are then anisotropically dry
etched by using RIE to expose bare silicon at the bottom
surface of the trenches, while the passivation layers on the
vertical sidewalls are retained. Then, silicon deep RIE is
processed again to deepen the trenches. The newly deepened depth equals the pressure-reference cavity height,
i.e., the releasing gap by following underetch.
4) The wafer is dipped into a 40% 50 C aqueous KOH (if
alkaline contaminant should be strictly avoided, TMAH
etchant can also be used) for about 9 h to complete the
bottom release by lateral underetching. Without protection of the passivation layer, the lower part of the trench
sidewalls will be underetched along the lateral direction,
i.e., 110 direction, while the top part trench sidewalls
will not be etched due to the passivation layer there.
5) Conformal LPCVD polysilicon is 597 C deposited to
seal the pressure sensor reference cavity by refilling
the narrow trenches. During the LPCVD process, the
pressure in the furnace is merely hundreds of millitorr.
After deposition, the wafers are cooled down to room
temperature, and the sealed pressure in the cavity will
become even lower and can be considered near a vacuum.
The vacuum is a good reference for the absolute pressure sensor. After deposition, high-temperature annealing
should be processed to eliminate stress of the polysilicon.
6) After the excessive polysilicon at the wafer surface is
stripped using RIE, electric contact holes are opened, and
an aluminum film is sputtered, patterned, and sintered for
piezoresistance interconnection.
7) After a TiW/Au seed layer (50 nm/100 nm in thickness) is sputtered, 20-m-thick photoresist (AZ9260) is
spin coated and patterned. With the photoresist layer as
mask, 20-m-thick copper is selectively electroplated on
the surface of the seismic-mass area of the accelerometer. After the photoresist being stripped by acetone, the
Cu and TiW seed layers are sequentially removed by
5% H2 SO4 + 1% H2 O2 and pure H2 O2 , respectively.
Then, the photoresist is spray coated and patterned. With
the photoresist as etching mask, the mechanically deflectable structure of the accelerometer is freed by sequentially using RIE and deep RIE.
8) After the single-sided micromachining processes for the
monolithic composite sensor chip, a wafer-level prepackaging step is implemented to facilitate following TPMS
module packaging. The device wafer is flipped and
aligned with a silicon cap wafer that was already formed

290

JOURNAL OF MICROELECTROMECHANICAL SYSTEMS, VOL. 21, NO. 2, APRIL 2012

Fig. 7. Infrared images of the fabricated hexagonal diaphragms. (a) Two rows
of releasing trenches along 111 orientation are designed, from which the
hexagonal-shaped pressure-sensitive diaphragm is finally formed. (b) Eight
rows of 111-oriented releasing trenches can be clearly seen that is for the
formation of the large diaphragm for the accelerometer structure.

Fig. 9.

Fig. 8. Images of the fabricated TPMS sensor. (a) Top-view infrared image
showing the composite sensor chip. (b) SEM image of the accelerometer,
with the cantilever and the Si/Cu seismic mass clearly shown. (c) SEM image
showing the cut cross-sectional view of the pressure-sensor diaphragm and the
reference cavity. (d) Photograph of the TPMS sensor after capping and dicing.
The sensor is much smaller than the match bud.

separately by using anisotropic wet etch. After a benzocyclobutene thin film is spray coated on the cap-wafer
surface as adhesion media layer, wafer-level silicon-tosilicon adhesive bonding is implemented in a Karl Suss
SB6 bonder. With heating and curing steps, the waferlevel capping prepackaging is carried out under purified
N2 gas environment. More detailed information about the
wafer-level prepackaging process can be obtained from
[24]. The prepackaging can effectively protect the moving
structure of the accelerometer from damage or contaminant during the following saw-dicing and modulepackaging steps. At the pressure-sensor region and the
wire-bonding pad location, a window has been opened
throughout the cap wafer to expose the sensing diaphragm
to external pressure and the aluminum pads for wire
bonding.
Fig. 7 shows the infrared images showing the fabricated
hexagonal-shaped diaphragms for the pressure sensor and the
beammass region of the accelerometer, respectively. The release etching formed cavities are with the same shapes and
dimensions compared to the simulation results shown in Fig. 3.
The images in Fig. 8 show the fabricated TPMS sensor.
Fig. 8(a) is a top-view infrared image of the fabricated com-

Schematic of the measurement setup for accelerometer.

posite sensor. The dark-gray regions denote the hexagonal


diaphragm of the pressure sensor, and the black region shows
the electroplated copper thick-film seismic mass. The closeup view of the accelerometer is shown in the SEM image in
Fig. 8(b). The pressure sensor is with the diaphragm, and the
embedded reference cavity beneath it is shown in the cut crosssectional view of the SEM image in Fig. 8(c). The hermetic
performance of the sealed reference cavity has been tested
by using helium leak detectors (Varian autotest 947), showing
a tightness of better than 109 atm cc/s. After wafer-level
capping and saw dicing, the TPMS sensor is shown in Fig. 8(d).
The dimension of the TPMS sensor is as small as 1.6
1.5 0.8 mm3 that is much smaller compared with the match
bud in Fig. 8(d). Compared with the conventional monolithicintegrated multifunctional TPMS sensor [6], [8], the chip size
of the developed device in this paper is about three-fifths of that
fabricated by Wang et al. and only one-third of that fabricated
by Xu et al. The smaller sensor size means large device number
in one wafer and lower fabrication cost per sensor.
IV. C HARACTERIZATION AND D ISCUSSION
A precise centrifugal-acceleration generation/testing system (Y53200, Electronic Technology Company Ltd.) is used
for the measurement of the accelerometer integrated in the
TPMS sensor. Fig. 9 schematically shows the measurement
setup. The accelerometer is fixed at the centrifuge wheel.
The rotation-generated centrifugal acceleration is adjustable by
the centrifuge control system (Y53200, Electronic Technology
Company Ltd.). DC 3.3-V power (INSTEK GPS-3303C) is
supplied to the half-sensitive Wheatstone bridge of the accelerometer. The output signal of the accelerometer is directly
fed into a digital multimeter (Agilent 34401A) for readout.
Without any amplification, the tested output voltage versus
applied acceleration is shown in Fig. 10. Within the 115-g measured range, the sensitivity of the half-sensitive Wheatstonebridge sensor is tested as about 99.9 V/g under 3.3-V supply,
which agrees well with the designed value of 103.4 V/g. By
using the best fitting-line method [12], the nonlinearity error of

WANG et al.: MONOLITHIC INTEGRATION OF PRESSURE PLUS ACCELERATION COMPOSITE TPMS SENSORS

Fig. 10. Tested accelerometer output voltage versus applied acceleration.

Fig. 11. Tested accelerometer zero-point offset versus temperature.

the accelerometer is measured as about 0.45% FS. Compared


to the sensitivity, the noise floor of the accelerometer output
signal is evaluated being lower than 0.2 g.
Fig. 11 shows the tested accelerometer zero-point offset
versus temperature change within the range of 40 C90 C
that is required from the TPMS application. With the definition
of temperature coefficient of offset (TCO) [18], the TCO of the
accelerometer is 0.11%/ C FS for the whole 115-g range.
Moreover, the accelerometer is dynamically tested by a
vibration exciter system. By adjusting the frequency of the
vibration exciter, the resonant frequency of the accelerometer
is tested as about 6.8 kHz that has a good agreement with the
designed value of 6.93 kHz.
In TPMS application, the accelerometer is used as a switching device triggered by the wheel rotation. When the vehicle
speed is higher than 25 km/h, the centrifugal acceleration
signal will wake up the tire-pressure measurement system, in
which way the TPMS module can work in an energy-saving
mode. Therefore, very high accuracy is not needed for the
accelerometer. The testing obtained performance of our fabricated accelerometers can well meet the requirements of TPMS
application.
Druck DPI-104 pressure gauge and PV211 handheld pump
system is used to test the performance of the pressure sensor.
At a room temperature of 20 C, the measured output voltage
versus applied pressure is shown in Fig. 12, resulting in a
sensitivity of 0.108 mV/kPa under 3.3-V power supply for
the Wheatstone bridge. The full measured range of absolute
pressure is 750 kPa. The measured sensitivity is slightly lower

291

Fig. 12. Tested pressure sensor output versus applied pressure.

Fig. 13. SEM image showing the close-up view of the bottom surface of the
diaphragm.

than the designed 0.112 mV/kPa. The discrepancy in sensitivity


is probably due to the fabrication tolerance. For example,
during the cavity vacuum-sealing process, the trench refilling
polysilicon is quite conformal and may also deposit on the
bottom surface of the diaphragm, which may lead to a little bit
thicker pressure-sensor diaphragm than the designed thickness
of 10 m. Fig. 13 shows the close-up view of the bottom surface
of the single-crystal silicon diaphragm, where the light-graycolored polysilicon thin layer coating on the bottom surface is
clearly observable. The thickened pressure-sensitive diaphragm
may lower the sensitivity. Based on the results in Fig. 12, a
lower than 0.1% nonlinearity error of the pressure sensor is
generally obtained that is defined by the best straight-line fitting
method.
The zero-point offset and full-scale sensing output of the
pressure sensor as functions of temperatures are tested with
the results shown in Fig. 14. The temperature coefficient of
zero-point offset is as low as 0.024%/ C FS within the
range from 20 C to 120 C, where the full measure range
is 750 kPa. Then, the temperature coefficient of sensitivity is
measured as 0.19%/ C. The measured reasonable results in
the temperature drift of the piezoresistive pressure sensor also
indicates that the two rows of opened holes embedded in the
silicon diaphragm do not cause severe structural stress on the
diaphragm and, hence, have no obvious negative influence to
the sensor stability.
Since the pressure sensor is the core detection component
in the TPMS application, which should accurately monitor the

292

JOURNAL OF MICROELECTROMECHANICAL SYSTEMS, VOL. 21, NO. 2, APRIL 2012

underetching released from the front side of the (111) wafer.


In order to achieve a high sensitivity of the accelerometer,
high-density copper thick film is selectively electroplated to
largely increase the seismic mass. Compared with the previous composite sensors, double-sided or multiwafer-bondingstructure-based process is avoided, and therefore, the composite
sensor size after capping prepackaging is reduced to 1.6 mm
1.5 mm 0.8 mm. Satisfactory testing results are obtained for
the pressure sensor and the accelerometer that well meet the
performance requirements of automobile TPMS application.
With the advantages of single-side micromachining processes,
ultrasmall device size, and low-cost batch-fabrication capability, the developed high-performance TPMS sensors are promising in practical volume production and applications.
Fig. 14. Tested zero-point offset and sensitivity of the pressure sensor in terms
of temperature change.

R EFERENCES
TABLE II
P ERFORMANCE C OMPARISON B ETWEEN P RESENT
S ENSOR AND P REVIOUSLY P UBLISHED O NES

real-time variation of the tire pressure to ensure the safe driving


of vehicles, there is a fundamental requirement to the overall
precision of the pressure measurement. The tested repeatability
error and hysteresis error of the fabricated pressure sensor are
0.32% FS and 0.15% FS, respectively. With the nonlinearity
error taken into consideration, the overall accuracy error of the
pressure sensor [12] is evaluated as 0.37% FS. Moreover, the
signal noise floor of the pressure sensor is about 0.02% FS for
750 or 0.15 kPa. Table II shows the performance comparison
between present sensor and previously published devices. According to the testing results for Infeneon SP-30 sensor that
is provided by Shanghai Hangsheng Automotive Electronics
Company, Ltd., our sensor specifications generally equal to that
of the market available product.
V. C ONCLUSION
Monolithic-integrated TPMS composite sensors have been
proposed and fabricated with a novel single-wafer-based singleside silicon bulk-micromachining and electroplating technologies. The fabrication of the accelerometer and pressure sensors
is all based on hexagonal-shaped diaphragms that are laterally

[1] J. Wang and X. Li, Monolithic-integrated silicon bulk-micromachined


accelerometer and pressure-sensor for tire-pressure-monitoring-system
(TPMS) application, in Proc. TRANSDUCERS, Beijing, China,
Jun. 59, 2011, pp. 703706.
[2] P. Prem and S. Sazuo, Complex three-dimensional structures in Si{111}
using wet bulk micromachining, J. Micromech. Microeng., vol. 19,
no. 10, pp. 19, Aug. 2009.
[3] D. Xu, B. Xiong, and Y. Wang, Design, fabrication and characterization of a front-etched micromachined thermopile for IR detection,
J. Micromech. Microeng., vol. 20, no. 11, pp. 110, Nov. 2010.
[4] B. Kim, J. Kim, and J. Kim, A highly manufacturable large area array MEMS probe card using electroplating and flipchip bonding, IEEE
Trans. Ind. Electron., vol. 56, no. 4, pp. 10791085, Apr. 2009.
[5] R. N. Dean and A. Luque, Applications of microelectromechanical systems in industrial processes and services, IEEE Trans. Ind. Electron.,
vol. 56, no. 4, pp. 913925, Apr. 2009.
[6] Q. Wang, X. Li, T. Li, M. Bao, K. Zhang, X. Ge, W. Zhou, and H. Bao,
A novel monolithically integrated pressure, accelerometer and temperature composite sensor, in Proc. TRANSDUCERS, Denver, CO, 2009,
pp. 11181121.
[7] Y. Zhang, B. Liu, L. Liu, Z. Zhang, Z. Tan, H. Lin, and T. Ren, Design,
fabrication and characterization of novel piezoresistive pressure microsensor for TPMS, in Proc. IEEE ASSCC, 2006, pp. 443446.
[8] J. Xu, Y. Zhao, Z. Jiang, and J. Sun, A monolithic silicon multi-sensor
for measuring three-axis acceleration, pressure and temperature, J. Mech.
Sci. Technol., vol. 22, no. 4, pp. 731739, Apr. 2008.
[9] H. Guckel, Surface micromachined pressure transducers, Sens. Actuators A, Phys., vol. 28, no. 2, pp. 133146, Jul. 1991.
[10] L. Lin, H. Chu, and Y. Lu, A simulation program for the sensitivity and
linearity of piezoresistive pressure sensors, J. Microelectromech. Syst.,
vol. 8, no. 4, pp. 514522, Dec. 1999.
[11] B. Bae, B. Flachsbart, K. Park, and M. Shannon, Design optimization of a piezoresistive pressure sensor considering the output signal-tonoise ratio, J. Micromech. Microeng., vol. 14, no. 12, pp. 15971607,
Dec. 2004.
[12] M. H. Bao, Micro Mechanical Transducers: Pressure Sensors, Accelerometers and Gyroscopes. Amsterdam, The Netherlands: Elsevier,
2000.
[13] P. W. Barth, F. Pourahmadi, R. Mayer, J. Poydock, and K. Petersen, A
monolithic silicon accelerometer with integral air damping and overrange
protection, in Proc. IEEE Solid State Sens. Actuator Workshop, 1988,
pp. 3538.
[14] H. Jin, D. Lu, S. Shen, and M. Bao, High linearity piezoresistance
micromachining accelerometer with low range, J. Funct. Mater. Devices,
vol. 9, no. 2, pp. 135151, 2001.
[15] Y. Zhang, B. Liu, L. Liu, Z. Zhang, Z. Tan, and H. Lin, Design and
fabrication characterization of novel piezoresistive pressure microsensor
for TPMS, in Proc. IEEE Solid-State Circuits, 2006, pp. 443446.
[16] T. Chou, C. Chu, C. Lin, and K. Chiang, Sensitivity analysis of packaging
effect of silicon-based piezoresistive pressure sensor, Sens. Actuators A,
Phys., vol. 152, no. 1, pp. 1938, May 2009.
[17] L. Chen and W. Cheng, Wafer-level chip scale packaging for piezoresistive pressure sensors using a dry-film shielding approach, Sens. Actuators
A, Phys., vol. 152, no. 2, pp. 261266, Jun. 2009.

WANG et al.: MONOLITHIC INTEGRATION OF PRESSURE PLUS ACCELERATION COMPOSITE TPMS SENSORS

[18] S. W. Lee, S. J. Park, D. I. Cho, and Y. S. Oh, Surface/Bulk micromachining (SBM) process and deep trench oxide isolation method for MEMS,
in IEDM Tech. Dig., 1999, pp. 701704.
[19] S. W. Lee, S. J. Park, J. P. Kim, S. C. Lee, and D. D. Cho, Surface/
bulk micromachined single-crystalline-silicon micro-gyroscope,
J. Microelectromech. Syst., vol. 9, no. 4, pp. 557567, Dec. 2000.
[20] S. J. Park, D. H. Kwak, H. H. Ko, T. Y. Song, and D. D. Cho, Selective
silicon-on-insulator (SOI) implant: A new micromachining method without footing and residual stress, J. Micromech. Microeng., vol. 15, no. 9,
pp. 16071613, Sep. 2005.
[21] S. J. Park, S. W. Lee, S. W. Yi, and D. D. Cho, Mesa-supported, singlecrystal microstructures fabricated by the surface/bulk micromachining
process, Jpn. J. Appl. Phys., vol. 38, no. 7A, pp. 42444249, Jul. 1999.
[22] J. Wang and X. Li, A high-performance dual-cantilever high-shock
accelerometer single-sided micromachined in (111) silicon wafers,
J. Microelectromech. Syst., vol. 19, no. 6, pp. 15151520, Dec. 2010.
[23] H. Seidel, L. Csepregi, A. Heuberger, and H. Baumgrtel, Anisotropic
etching of crystalline silicon in alkaline solutions, J. Electrochem. Soc.,
vol. 137, no. 11, pp. 36123626, Nov. 1990.
[24] K. Zhang, W. Jiang, and X. Li, Wafer-level sandwiched packaging for
high-yield fabrication of high-performance MEMS inertial sensor, in
Proc. IEEE MEMS, 2008, pp. 814817.

Jiachou Wang was born in Fujian Province, China,


in 1979. He received the B.S. and M.S. degrees
in mechanical engineering from Harbin University
of Science and Technology, Harbin, China, in 2002
and 2005, respectively, and the Ph.D. degree in
mechatronics engineering from Harbin Institute of
Technology, Harbin, in 2008.
In 2008, he became an Assistant Researcher and is
currently a Research Associate with the Shanghai Institute of Microsystem and Information Technology,
Chinese Academy of Sciences, Shanghai, China. His
research interests include MEMS fabrication and design, micro-/nanosensors,
transducers and actuators, robotics and automation, and micro- and nanopositioning and manipulation.

293

Xiaoyuan Xia received the Ph.D. degree in microelectronics and solid-state electronics from the
Shanghai Institute of Microsystem and Information Technology, Chinese Academy of Sciences,
Shanghai, China, in 2009.
Since 2009, she has been engaged as a full-time
Research Staff member with the State Key Laboratory of Transducer Technology, Shanghai Institute
of Microsystem and Information Technology, for the
investigation of MEMS/NEMS fabrication technologies and resonant cantilever sensors.

Xinxin Li received the B.S. degree in semiconductor physics and devices from Tsinghua University,
Beijing, China, in 1987, and the Ph.D. degree in
microelectronics from Fudan University, Shanghai,
China, in 1998.
He was a Research Engineer with Shenyang Institute of Instrumentation Technology, Shenyang,
China, for five years. He was also with The
Hong Kong University of Science and Technology,
Kowloon, Hong Kong, as a Research Associate and
with Nanyang Technological University, Singapore,
as a Research Fellow. He then joined Tohoku University, Sendai, Japan, as
a nonpermanent Lecturer (Center of Excellence Research Fellowship). Since
2001, he has been a Professor with the Shanghai Institute of Microsystem and
Information Technology, Chinese Academy of Sciences, Shanghai, where he
currently serves as the Director of the State Key Laboratory of Transducer
Technology. He has invented about 60 patents and published about 200 papers
in refereed journals and conference proceedings. He is on the Editorial Board
of the Journal of Micromechanics and Microengineering. For a long period of
time, his research interests have been in the fields of micro-/nanosensors and
MEMS/NEMS.
Prof. Xinxin Li served as a Technical Program Committee member for IEEE
MEMS 2008 and 2011, IEEE Sensors from 2002 to 2011, and Transducers
2009 and 2011.

You might also like