Sie sind auf Seite 1von 14

IEEE SENSORS JOURNAL, VOL. 7, NO.

12, DECEMBER 2007

1639

Zero-Rate Output and Quadrature Compensation in


Vibratory MEMS Gyroscopes
Mikko Saukoski, Student Member, IEEE, Lasse Aaltonen, Student Member, IEEE, and
Kari A. I. Halonen, Member, IEEE

AbstractIn this paper, issues related to the zero-rate output


(ZRO) of a vibratory microgyroscope are studied. Different
sources of the ZRO are discussed and how the effect of each source
can be minimized and their stability improved is described. The
effects of synchronous demodulation and electrostatic quadrature
compensation performed with a dc voltage on the final ZRO are
analyzed. Ways to implement the control loop for electrostatic
quadrature compensation performed with a dc voltage are described, concentrating on a case where the compensation voltage
is generated with a digital-to-analog converter and the controller
techis digital. In particular, extending the resolution with
niques is studied. The experimental work shows the feasibility of
the implemented quadrature compensation loop and analyzes the
ZRO sources of one particular gyroscope implementation. How to
perform the ZRO measurements in such a way that the various
sources can be distinguished from each other is also described.

61

Index TermsAngular velocity, gyroscopes, microelectromechanical systems (MEMS), offset, quadrature error, sensors,
stability, zero-rate output (ZRO).

I. INTRODUCTION
NGULAR VELOCITY sensors have applications in areas
such as inertial navigation, automotive safety, and stability control systems, platform stabilization, including picture
stabilization in camcorders and cameras, and robotics, to name
but a few. In particular, automotive chassis control systems such
as Electronic Stability Program/Electronic Stability Control
(ESP/ESC) have created a rapidly growing market for lowto medium-priced, medium-accuracy angular velocity sensors
which could be realized with micromachining [1], [2].
Traditionally, angular velocity has been measured with a
rotating wheel gyroscope, which is based on the conservation
of angular momentum. This has been replaced by fiber-optic
and ring laser gyroscopes in precision applications. (Strictly
speaking, the term gyroscope refers to the traditional rotating
wheel device, whereas the term angular velocity sensor refers
to angular rate measurement devices in general. However, they
are used interchangeably in the literature, and the convention
will be followed here as well.) Optical gyroscopes are the most
accurate angular velocity sensors available at the moment, and

Manuscript received June 29, 2007; revised August 21, 2007; accepted August 25, 2007. This work was supported in part by VTI Technologies, Vantaa,
Finland, and in part by the Finnish Funding Agency for Technology and Innovation (TEKES). The associate editor coordinating the review of this paper and
approving it for publication was Prof. William C. Tang.
The authors are with SMARAD-2/Electronic Circuit Design Laboratory, Helsinki University of Technology, FI-02150 Espoo, Finland (e-mail:
mhs@ecdl.tkk.fi; ljaalton@ecdl.tkk.fi; karih@ecdl.tkk.fi).
Digital Object Identifier 10.1109/JSEN.2007.908921

are used, for instance, in inertial navigation systems. However,


they are too expensive and too bulky for most of the aforementioned applications. [3]
Almost all the micromechanical angular velocity sensing elements reported in the literature have been vibratory gyroscopes,
which are based on the transfer of energy between two oscillation modes by the Coriolis effect. Different structures that have
been utilized include tuning forks, vibrating beams and plates,
and vibrating rings. The possible manufacturing techniques include bulk and surface silicon micromachining, different hybrid silicon technologies, quartz micromachining, and electroplating. The actuation mechanisms that have been used include
electrostatic (capacitive), piezoelectric, and electromagnetic actuation. The sensing methods include capacitive, piezoelectric,
piezoresistive, and optical sensing [3].
One of the most significant nonidealities in angular velocity
sensors is the zero-rate output (ZRO). The ZRO is the signal
that the sensor outputs when no angular velocity is applied [3].
It resembles the offset in electronic circuits, and can also be
referred to as the offset signal or the zero-rate offset.
If the ZRO of a gyroscope were constant, it would be possible to remove it from the final output signal with calibration,
either analog or digital. Therefore, as long as it does not limit the
dynamic range of the system, the magnitude of the ZRO is not
important as such, but rather its stability over time and environmental variations, such as variation of temperature. In practice,
the sources of the ZRO are typically not well controlled and,
hence, it is often necessary to minimize the absolute values in
order to minimize the variation. [4]
The sources of the ZRO can be identified as residing in either
the sensor element, the interface electronics, or an unintended
interaction between the two due to packaging anomalies that
may be unstable with sensor or environmental vibration, or with
temperature. This paper does not address sensor packaging. One
important nonideality of the sensor element which is closely
related to the ZRO is the mechanical quadrature signal [3]. It
is caused by manufacturing imperfections and manifests itself
as an output signal from the micromechanical element when
no angular velocity is applied. This signal can be considerably
larger than the Coriolis signal produced by the full-scale angular
velocity [5]. However, the mechanical quadrature signal has a
90 phase shift compared with the Coriolis signal, and it can
hence be distinguished using synchronous demodulation. Then,
the maximum tolerable mechanical quadrature signal depends
again on the dynamic range, together with the achievable phase
accuracy of the demodulation.
Often, the mechanical quadrature signal can be so large that it
would limit the dynamic range of the system. The degree of re-

1530-437X/$25.00 2007 IEEE

1640

IEEE SENSORS JOURNAL, VOL. 7, NO. 12, DECEMBER 2007

quired phase accuracy in the demodulation can also be impractically strict. However, it is widely known that an effective way
to cancel this signal is to apply a dc voltage to properly placed
electrodes in the sensor element [5][7]. There are also several
other ways to compensate or cancel the mechanical quadrature
signal, either electrostatically or purely electronically. These
methods will be discussed briefly in Section IV-A1. Mechanical
adjustment or trimming of the sensor element is also possible,
but is highly dependent on the actual design and manufacturing
process and will not be addressed in this paper.
This paper will focus on these two nonidealities and the relationships between them. Different sources of the ZRO that have
been reported in the literature will be discussed, and they are
given a uniform representation. Ways to minimize the effect of
these sources on the final ZRO will be considered. How the electrostatic quadrature compensation performed with a dc voltage
affects the ZRO will be analyzed. Different ways to implement
the control loop for the compensation will be presented, concentrating on a case where the compensation voltage is generated
with a digital-to-analog converter (DAC), and the controller is
digital [4], [8].
This paper is organized in an introductory part formed by this
section together with Section II, and three other parts formed by
Sections IIIV. First, in the other half of the introductory part, in
Section II, the general properties of a vibratory microgyroscope
will be presented.
The first of the three other parts is formed by Section III. In
this section, various implementations of the quadrature compensation loop will be studied. The focus will be on a case where
the compensation voltage is generated with a DAC, and the controller is digital. In particular, extending the resolution with
techniques will be studied.
The second part is formed by Section IV. Here, different
sources of the ZRO will be discussed, and how they manifest
themselves in the final ZRO after the synchronous demodulation
will be shown. The effect of the electrostatic quadrature compensation performed with a dc voltage will also be analyzed.
The third part is formed by Section V. Here, experimental results will be given to support the theory that has been presented.
This section is divided into two parts, such that the first part
examines different implementations of the quadrature compensation loop and the second part studies the ZRO components of
one particular microgyroscope.
Finally, in Section VI, the paper ends with some concluding
remarks.
II. VIBRATORY MICROGYROSCOPES
A vibratory microgyroscope is composed of two orthogonal
mechanical resonators. A schematic drawing of such a system
without any nonidealities is shown in Fig. 1. In the figure, the
four rollers are used to express the fact that when the mass
is moving along the -direction, the -directional spring and
and
have no effect on the behavior, and vice
damper
versa. Thus, the system comprises two 1-degree-of-freedom res, the damper
, and the
onators, one formed by the spring
, the damper
, and
mass , and the second by the spring
the mass . Boths springs are assumed to be massless.
In the configuration depicted in Fig. 1, the masses of both
- and -directional resonators are identical and equal to .

Fig. 1. A schematic drawing of a vibratory microgyroscope without any


nonidealities.

However, in order not to limit the generality of the following


analysis, the mass of the -directional resonator will be denoted
, and that of the -directional resonator with
.
with
The 2-D equation of movement (EoM) for the system can be
written as [9]
(1)
where
, and are the mass, damping, spring, and
Coriolis force matrices and the excitation force vector, respectively; and are the displacements of the - and -directional
resonators, and the dots denote first and second derivatives with
respect to time.
, and into (1), it can be written
By substituting
as

(2)
Here, the terms
, and
are defined in Fig. 1.
and
were already defined above. The terms
The masses
and
represent the nonproportional damping, and
and
the anisoelasticity [9]. These terms have been added in
order to take the nonidealities of the mechanical system into acis the angular velocity of the system about
count. Finally,
and
are external forces used to excite the
the axis, and
resonators.
Now, on the basis of (2), the operation of a vibratory
gyroscope can be understood as follows. The -directional
resonator, often called the primary resonator or the drive
resonator, is driven to oscillation at its resonance frequency
which is also known as the operating frequency, using external excitation . The oscillation amplitude
is controlled and hence the -directional displacement can be
written as
(3)
is the vibration amplitude, determined by an external
where
controller.
By solving the -directional EoM with the assumption that
, the movement of the -directional resonator, often

SAUKOSKI et al.: ZERO-RATE OUTPUT AND QUADRATURE COMPENSATION IN VIBRATORY MEMS GYROSCOPES

1641

called the secondary resonator or the sense resonator, can be


written as

(4)
is the magnitude of the transfer
Here,
function of the secondary resonator from force to displacement
, defined as
at the operating frequency
(5)
and

is the phase shift at the same frequency, defined as

Fig. 2. A continuous-time feedback loop for quadrature signal compensation.

(6)
where
is the ratio between the resonance frequencies of the primary and secondary resonators, the latter defined
, and
is the quality factor of the secas
. Here, it is
ondary resonator, defined as
assumed that the angular velocity signal is narrowband, i.e., the
and do not vary significantly over the signal
values of
bandwidth.
Let us first consider the ideal case, where the off-diagonal
. Now, the secondary resonator remains
terms
is applied
at rest unless a -directional angular velocity
to the device. In that case, the Coriolis effect couples the movement from the primary resonator to the secondary resonator. The
movement of the secondary resonator can be written as
(7)
From (7), it can be seen that the secondary resonator movement contains the amplitude-modulated angular velocity infor.
mation around the carrier frequency
Next, let us consider the effect of the off-diagonal terms in the
causes an output signal cor2-D EoM. The anisoelastic term
. This output
responding to an angular velocity
signal has a 90 phase shift when compared with the actual Coriolis signal, and hence is referred to as the mechanical quadrature signal. (The term mechanical quadrature signal is used
to distinguish it from other signals which will be introduced
later and are also in quadrature with respect to the Coriolis
signal.) On the other hand, the nonproportional damping term
causes an output signal corresponding to an angular ve. This signal is in-phase with the Coriolis
locity
signal. Whereas the Coriolis signal can be distinguished from
the mechanical quadrature signal by phase-coherent demodulation, it is impossible to differentiate the Coriolis signal from the
output inflicted by nonproportional damping.
The analysis is also applicable to vibratory gyroscopes with
torsional resonators by replacing the masses with moments of
inertia, displacements with angles, and forces with torques.
Depending on the value of , the gyroscope can be said to be
,
operated either in low-pass mode or mode-matched. If
the element operates in the low-pass mode, indicating that the
operating frequency is below the resonance frequency of the
secondary resonator. The sensitivity of the gyroscope element
,
can be increased by reducing . At the extreme, when

.
the gyroscope is said to be mode-matched, i.e.,
, increasing both
Now, the Coriolis signal is amplified by
sensitivity and resolution when compared with low-pass mode.
However, this would lead to an impaired open-loop bandwidth,
together with possible issues with gain stability and linearity and
the phase instability of the output signals.
For the analysis in the following sections, it is assumed that
the element is operated in the low-pass mode, in open-loop con. Further, it will be assumed that
and the
figuration
. If, for example,
mode separation are high enough for
and
, then
, a value that is small
for certain
enough to be considered negligible. The required
falls rapidly when is increased.
III. QUADRATURE SIGNAL COMPENSATION
As was shown in Section II, the off-diagonal terms in the
spring matrix of the 2-D EoM (2) cause a mechanical quadrature signal to the secondary resonator. This section will present
different ways to implement the control loop for electrostatic
quadrature compensation performed with a dc voltage [5][7].
The focus will be on a case where the compensation voltage is
generated with a DAC, and the controller is digital [4], [8].
A. Continuous-Time Compensation
A straightforward continuous-time quadrature compensation
loop is presented in Fig. 2. If the feedback part is ignored for
is inflicted
a while, the force
by the primary resonator movement, as described in Section II.
The force excites the secondary resonator with the transfer
, causing the uncompensated quadrature
function
movement
. The resulting signal goes through the
and is synchrodisplacement-to-voltage conversion
nously down converted to dc. Then, it is filtered to the desired
bandwidth with a low-pass filter, yielding the output signal
.
In the feedback part, the signal is first brought to a controller
, which outputs the voltage for the quadrature compensation electrodes. In the micromechanical element, the compensation voltage is converted to an electrostatic force and modin a proper phase. Finally,
ulated to the operating frequency
. The
the result is summed to the uncompensated force
includes the dynamics from the
transfer function
compensation voltage to the resulting force. In the real device,

1642

Fig. 3. A feedback loop with A/D and D/A conversion for quadrature signal
compensation.

the conversion from voltage to force and the further up converoccur simultaneously. Depending on the sensor elsion to
ement, the relationship between the quadrature compensation
voltage and the amplitude of the force generated can be linear,
or it can have a higher order dependence, such as a quadratic
one.
Typically, the quadrature signal needs to be zeroed at dc, requiring the controller to have a pole in the origin. The simplest solution would be a single integrator with
. Because of the low-pass mode operation, the compensation bandwidth is most likely limited by the low-pass filter
after the demodulation. If the bandwidth needs to be increased
beyond this limitation, a more complex controller with zeroes
that provide phase lead at higher frequencies might be required.
In this case, a complete PID (proportional-integrator-differentiator) controller can be used, for instance.
B. Compensation Using DAC With Digital Controller
When the quadrature signal is converted to the digital domain and the controller is implemented with digital signal processing (DSP), the loop takes the form shown in Fig. 3. In the
figure, the signal is A/D converted before demodulation, and
down converted and filtered in the digital domain. The A/D converter (ADC) could also be located after the low-pass filter, in
which case the demodulation is performed in the analog domain. The demodulation can also be performed using subsampling techniques.
A fundamental issue in the digital controller is that because
of the limited resolution of the feedback DAC, the output signal
can exhibit oscillations between two values even when
the actual linear compensation loop is stable. This can happen if
there is an integrator in the loop and the signal at the controller
input changes by more than one least-significant bit (LSB) when
the DAC output changes by one LSB. Now, it might occur that
the controller input cannot be brought to zero with any DAC
inteoutput level. The integrator in the controller
grates the residual signal until the DAC output changes by one
LSB. Then, the integrator output starts moving in the opposite
direction, until it returns to the original value, after which the
oscillation cycle starts over again. Because of the phase shifts
and in
and
, respectively, the oscillation cou, and causes
ples to the Coriolis output, attenuated by

IEEE SENSORS JOURNAL, VOL. 7, NO. 12, DECEMBER 2007

a spurious angular velocity signal whose spectral content deand


pends on the uncompensated quadrature movement
the exact loop gain.
-type converter [4], [8], [10], then with a
If the ADC is a
dc input signal, the input voltage can be revealed with unlimited
resolution by integrating it for a long enough time. The same
, if the converter is a bandpass
applies to an input signal at
ADC ([10, Ch. 5]). This means that the DAC output will
always oscillate between two values.
If the DAC resolution is sufficient, so that a change of one
LSB at the DAC output causes a change of less that one LSB
-ADC, the
at the controller input, then in the case of a nonintegrated in
can be kept at zero, and
signal
the output stays stable. With simulations, it can be shown that
ADC, as
an identical resolution requirement applies for a
changes below one LSB at the ADC output are dithered to noise
modulation. Since the gain from the DAC
because of the
output to the controller input is typically high, this leads to a
very strict resolution requirement in the DAC.
A solution for the stringent resolution requirement is to inmodulation. Now, the rescrease the DAC resolution with
olution of the DAC output at dc is not limited by the converter
itself, but rather by the word length at the modulator input (controller output). By setting the word length to be sufficiently long,
the spurious components can be brought as low as desired, or
completely removed.
A block diagram of the configuration is shown in Fig. 4. In
the figure, the actual DAC is preceded with an interpolator, in, and a noise-shaping loop (NL). These
terpolation filter
DAC ([10, Ch. 7]). The
building blocks form a second-order
figure also shows the word lengths and the sampling rates at different parts. The limiter inside the NL prevents an overflow at
the summer output from occurring in case the input signal grows
too high.
As the controller output signal can be considered a dc signal,
the interpolation filter can usually be removed, and the interpolation performed with a simple zero-order hold. The design of
DAC involves choosing the word lengths and
,
the
the interpolation factor , and the number of final quantization
levels . Parameters and
are used to set the quantization
to a desired level. Parameter
noise in the output signal
is used to achieve a sufficient resolution, as described above.
is chosen to be such that the NL functions propFinally,
erly without the summer output being clipped.
An important consideration is to prevent the quantization
noise at the DAC output from causing the final signal-to-noise
ratio (SNR) to deteriorate. This noise can couple to the Coriolis signal through two paths. The first is the noise in the
actual compensation voltage, which couples as a result of
the phase errors in synchronous demodulation, as described
earlier. To make this noise source low enough, parameters
and
are used. The second path is direct coupling through
parasitic capacitances, in the case the detection is performed
capacitively. This noise source can be eliminated by setting
, where is an integer greater than zero.
Now, because of the zero-order hold at the DAC output, the
continuous-time output spectrum has a zero at
, and the
noise coupling is thus minimized.

SAUKOSKI et al.: ZERO-RATE OUTPUT AND QUADRATURE COMPENSATION IN VIBRATORY MEMS GYROSCOPES

Fig. 4. A feedback loop with

1643

61 DAC for quadrature signal compensation.

IV. ZERO-RATE OUTPUT (ZRO)


In this section, the fundamental sources of the ZRO signal
in a micromechanical vibratory gyroscope with capacitive excitation and detection will be discussed [4]. Special attention
will be paid to the stability of each source. Possible methods to
minimize the effect of different sources will be described. How
the synchronous demodulation and the electrostatic quadrature
compensation performed with a dc voltage affect the final ZRO
will also be described [4].
The discussion will be kept at a general level, with applicability to any vibratory gyroscope fulfilling the necessary assumptions. The final significance of each source is dependent
on the actual system, and can be determined only after the design is known. As an example case, one gyroscope design will
be experimentally analyzed in Section V.
A. Sources of ZRO
The sources of the ZRO can be identified as residing in either the sensor element or the interface electronics. The sources
residing in the sensor element can be further divided into two
groups, ones resulting from the off-diagonal terms in the 2-D
EoM (2) and ones resulting from other nonidealities.
As described in Section II, the sources resulting from the offdiagonal terms in the 2-D EoM are 1) the mechanical quadrature signal [5], [9], [11], [12] and 2) the signal resulting from
the nonproportional damping [9]. The sources related to other
nonidealities of the sensor element are 3) the direct electrical
cross-coupling of the primary resonator excitation signal to the
secondary resonator detection [12] and 4) the direct excitation
of the secondary resonator by the primary resonator excitation
signal [11], [13]. In the electronics, a possible source of the ZRO
is 5) the cross-coupling of the primary signal and different clock
signals to the secondary resonator detection [4].
Throughout the following analysis, it will be assumed that the
voltage signal used to excite the primary resonator is
(8)

and
are the dc and ac components of the exciwhere
tation signal, respectively. As the electrostatic force is proportional to the square of the voltage, the resulting primary resonator motion is then

(9)
where
is the primary resonator transfer function from
excitation voltage to displacement. Further, the gain and phase
shift of the secondary resonator detection circuit (displaceand ,
ment-to-voltage converter) are assumed to be
respectively.
1) Mechanical Quadrature Signal: The mechanical quadrature signal [5], [9], [11], [12] is generated when the primary resonator movement couples directly to the secondary resonator as
a result of the nondiagonal terms in the spring matrix of the
2-D EoM. The magnitude of the force generated is directly proportional to the primary resonator displacement. Hence, the mechanical quadrature signal at the output of the detection circuit
can be written as
(10)
where is a parameter that determines the magnitude of the
coupling. With the parameters defined in Section II, is equal
to
.
Four different ways to handle the mechanical quadrature
signal can be identified. First, it can be removed or brought to
a desired level at the sensor element by either careful design
[14], [15] or by tight process control and by postmanufacturing
trimming and screening.
Second, it can be compensated electronically by injecting a
signal with the same amplitude but opposite phase to the input
of the secondary resonator detection circuit, effectively cancelling the mechanical quadrature signal. This might be difficult
to achieve, as both the amplitude and the phase of the feedback
signal need to be carefully controlled. Although the amplitude

1644

IEEE SENSORS JOURNAL, VOL. 7, NO. 12, DECEMBER 2007

can be controlled in a feedback loop, correct phasing needs to


rely on tight phase control in the electronics. [14]
Third, the same effect as with electronic compensation can be
achieved by exciting the secondary resonator electrostatically
with a force that has an opposite phase compared with the mechanical quadrature signal. The same limitations apply as with
electronic compensation. [14]
Finally, a fourth option is to compensate the mechanical
quadrature signal electrostatically by bringing a dc voltage
to properly placed compensation electrodes. This generates a
compensating force with a frequency
, in proper phase to
cancel the force caused by anisoelasticity [4][7].
The first, third, and fourth methods cancel the actual quadra, whereas the second method cancels the
ture movement
quadrature signal after it is converted to charge. Further, the
third and fourth methods require an additional set of electrodes
to apply the electrostatic compensation force, whereas the first
and second methods do not require the electrodes.
In practice, the first method is always applied in the gyroscope
design. Then, if the remaining quadrature signal still needs to be
cancelled, the other three methods can be applied. The choice
of different methods is always dependent on the required performance and cost.
When the mechanical quadrature signal is compensated either electronically or electrostatically, the compensation affects
the ZRO caused by the other sources through phase shifts. The
interactions resulting from electrostatic compensation with a dc
voltage (the fourth method described above) will be analyzed in
detail in Section IV-B.
2) Nonproportional Damping: If there are off-diagonal
terms in the damping matrix
in the 2-D EoM, a force proportional to the velocity
of the primary
resonator will be generated [9]. At the output of the detection
circuit, this signal can be expressed as

(11)
where is a parameter that determines the magnitude of the
nonproportional damping. Again, with the parameters defined
.
in Section II, is equal to
The only way to minimize this source is to minimize the parameter by mechanical design. To keep the source stable together with the gain
, needs to be kept stable.
3) Electrical Cross-Coupling in Sensor Element: In electrical cross-coupling [12], the primary resonator excitation
signal couples to the secondary resonator detection circuit
through stray capacitances. Because only the time-varying
can couple through capacitances,
component
the cross-coupled signal at the output of the detection circuit
can be written as
(12)
where is a parameter that determines the magnitude of the
cross-coupling.
There are several methods to avoid electrical cross-coupling.
First, and most obviously, the stray capacitances should be made
as small as possible. Next, in the case of differential detection,

Fig. 5. Possible sources of electrical cross-coupling in the sensor element.

the cross-coupling to both inputs should be as symmetrical as


possible, in order to have the coupled signal as a common-mode
component in the input and, hence, rejected by the detection
circuit. Additionally, if the excitation is differential, both signals
should cross-couple symmetrically to the detection circuit so
that they cancel each other.
Third, the primary resonator excitation and secondary resonator detection can be separated in the frequency domain. This
,
is possible either by exciting the primary resonator at
which is made possible by the quadratic relationship between
the excitation voltage and force, or by modulating the secondary
signal to a higher frequency for detection [16].
Fourth, the primary resonator excitation and secondary resonator detection can be separated in the time domain as well.
This means that the excitation and detection are performed in
separate phases, leading to a discrete time implementation. Finally, the source of cross-coupling could be eliminated by using
some other form of excitation than electrostatic, such as magnetic [17].
To keep the output signal caused by cross-coupling constant,
should be kept constant. In other words, both
the product
the strength and the source of cross-coupling should be as stable
as possible throughout the full range of operating conditions.
Fig. 5 shows the electrical cross-coupling path through the cato the input of a detection circuit implemented as
pacitor
a charge integrator [4]. It also illustrates another, related issue
that should be considered. Here, the excitation capacitance
and the secondary resonator detection capacitance
have a
common middle electrode, biased through a parallel combinaand
. Now, depending on the biasing impedance
tion of
, the voltage
can vary with
, as a result of the primary resonator excitation
frequency
signal or time-dependent capacitances. The phase of the voltage
and the source of the variance. The
variation is dependent on
error signal caused by this effect can be reduced by making the
biasing impedance small enough. The other methods described
above, including differential excitation and detection, and timeand frequency-domain multiplexing are also applicable.
4) Direct Excitation of Secondary Resonator: The direct excitation of the secondary resonator by the primary resonator excitation signal [11], [13] (called direct motion coupling in
[13] means that as a result of nonidealities in the sensor eleused to excite the primary resment, the voltage signal
onator also excites a motion in the secondary resonator through
in the 2-D EoM (2). As in the primary resonator, the
the term

SAUKOSKI et al.: ZERO-RATE OUTPUT AND QUADRATURE COMPENSATION IN VIBRATORY MEMS GYROSCOPES

magnitude of the force created and, hence, also the amplitude


.
of the cross-coupled signal is relative to the product
Therefore, the magnitude of this signal at the output of the detection circuit can be written as

(13)
where is a parameter that determines the magnitude of direct
excitation.
The only ways to eliminate the effect of direct excitation are
either to ensure the orthogonality of the primary resonator excitation force and the secondary resonator motion by process
control or by postprocess trimming, or by exciting the primary
resonator in such a way that the secondary resonator cannot
respond to it directly. It should be observed that if the secondary resonator movement rises from the primary resonator
movement through anisoelasticity, the resulting signal will be
in quadrature, whereas the signal resulting from direct excitation is in-phase with the Coriolis signal.
To keep the direct excitation force and, hence, the output
signal constant, the term should be constant, together with
. As the latter term varies in order to conthe product
trol the vibration amplitude of the primary resonator, the term
needs to be minimized until the required level of ZRO stability
also needs to be kept constant
is achieved. The gain
in order to keep the output signal caused by direct excitation
stable.
The direct excitation plagues not only vibratory gyroscopes
with capacitive excitation, but also those with piezoelectric excitation. In [18], the sources of the parasitic secondary signal in a
tuning fork gyroscope with piezoelectric excitation and piezoresistive detection are analyzed. In the reference, direct excitation
of secondary movement is referred to as actuation unbalance.
5) Cross-Coupling in Electronics: Due to parasitic stray capacitances or inductances in the electronics, various signals can
couple to the secondary resonator detection circuit and inflict
ZRO [4]. The source of coupling can be either the signal from
the primary resonator or a clock signal that is synchronized to
the primary resonator. The contribution of the clock signals, referred to the output of the secondary resonator detection circuit,
can be written as

(14)
and
describe the magnitudes of the cross-coupled
where
signals. Parameter refers to a signal that is in-phase with the
Coriolis signal, and to a signal that is in quadrature with the
Coriolis signal.
The cross-coupling primary resonator output signal, again referred to the output of the secondary resonator detection circuit,
can be written as

(15)

1645

and
describe the magnitudes of cross-coupling.
where
Again, refers to the cross-coupled signal that is in-phase with
the Coriolis signal, and to the signal that is in quadrature with
the Coriolis signal. If identical detection circuits are used for
both the primary and secondary resonators, the primary signal
.
is 90 phase shifted relative to the Coriolis signal, and
This source of coupling can be significant if the primary and
secondary resonator detection circuits are located close to each
other, which needs to be the case if good matching of the components in the circuits is required.
B. Effects of Synchronous Demodulation and Quadrature
Compensation
In the previous section, the different sources of the ZRO were
expressed at the output of the secondary resonator detection circuit in (10)(15). However, because of the synchronous demodulation, these equations do not directly represent the magnitude of the final resulting ZRO. Furthermore, the electrostatic
quadrature compensation performed with a dc voltage affects
the ZRO signal of the gyroscope. In this section, these two effects will be analyzed.
1) Synchronous Demodulation: After the signals of
(10)(15) are summed at the output of the detection circuit, they enter the demodulator. The signal at the input of the
demodulator is then

(16)
Next, the signal is down converted into two components, the
in-phase component, which also carries the Coriolis information, and the quadrature component. One should be careful in
order not to confuse the quadrature component in the demodulator output with the mechanical quadrature signal, which is
only one part of the quadrature component, albeit a significant
one.
The down conversion is performed by multiplying with
for the in-phase component and with
for
the quadrature component, and by filtering the signal with a
low-pass filter. Assuming that there is a further gain of two after
the filtering, the resulting output signals from the demodulator
are

(17)

1646

IEEE SENSORS JOURNAL, VOL. 7, NO. 12, DECEMBER 2007

and

from the actual mechanical quadrature signal


present, it is nulled, irrespective of the phase shift
. However, if the other terms given in (18) are present,
they cause a fraction of the mechanical quadrature signal to be
left after the compensation, so that the sum is equal to zero. The
remaining mechanical quadrature signal is then
(18)

for the in-phase and quadrature components, respectively. Now,


(17) gives the ZRO when the mechanical quadrature signal is
not compensated [4], under the assumptions made earlier.
To gain more insight into the result, (17) needs to be reduced to input angular velocity. This can be done by dividing
, leading to
by
(20)
By substituting this into (17), the ZRO signal in the presence of
the quadrature compensation can be written as

(19)
has also been applied. From (19),
where the assumption
it can be seen which parameters affect the input-referred ZRO
and how it can be minimized.
In a practical implementation, the mechanical quadrature
can be so high [4], [5] that the degree to
signal
) needs to be controlled to yield
which the phase shift (or
good ZRO stability is impractically strict. If the mechanical
quadrature signal is, for example, 10 times larger than the
full-scale Coriolis signal, then a phase shift of 0.06 causes
the ZRO inflicted by the mechanical quadrature signal to rise
to 1% of the full-scale Coriolis signal. This also means that
a tiny variation in the phase shift, or a variation in the gain
, causes significant variations in the ZRO. Additionally, a mechanical quadrature signal ten times higher than the
full-scale Coriolis signal could already limit the dynamic range
of the secondary resonator detection circuit.
2) Quadrature Compensation: Next, the effect of the electrostatic quadrature compensation performed with a dc voltage
is analyzed. The compensation is performed in a closed-loop
configuration, as described in Section III. The detailed implementation of the feedback loop can be ignored for the purpose
of this analysis. A steady-state operation will be assumed, i.e.,
it will be assumed that there are no transients. This is justified,
as the ZRO is a dc signal by definition. It will also be assumed
that the quadrature compensation at dc is perfect, i.e., there is an
ideal integrator in the feedback loop. As the dynamic operation
will not be analyzed, the detailed controller parameters are not
important as such, as long as they ensure a stable operation of
the feedback loop.
When applied, the quadrature compensation drives the signal
given in (18) to zero. When there are no other components apart

(21)
From (21), it can be seen that while the term
has been removed, the effect of the quadrature compensation is
present through the sine times tangent and cosine times tangent
terms that have appeared in the equation compared with (17).
The error terms are not proportional to the magnitude of the
;
uncompensated mechanical quadrature signal
they depend only on the phase shift terms and .
Now, as it is assumed that the element is operated in low-pass
, then with reasonable values of the phase shift
mode with
, the terms involving the sine times the tangent in (21) are rendered negligible compared with the terms involving the cosine.
Furthermore, the last and the third last term are completely cancelled. The ZRO can then be written as

(22)
The equation shows the components that affect the final ZRO
after quadrature compensation. They are the signal caused by
nonproportional damping, the signal electrically cross-coupled
at the sensor element, the direct excitation, and any cross-coupled clock signals or a component of the signal from the primary
resonator that are in-phase with the Coriolis signal. Further, the
phase shift determines the magnitude of the ZRO. If is small,
then as a result of the cosine function, the ZRO is not very sensitive to variations in . If, for example, varies between 1 and

SAUKOSKI et al.: ZERO-RATE OUTPUT AND QUADRATURE COMPENSATION IN VIBRATORY MEMS GYROSCOPES

1647

Fig. 6. A block diagram of the angular velocity sensor including the bulk micromachined sensor element, the analog part implemented with a custom IC, and the
digital signal processing part implemented with an FPGA chip.

2 , the variation in the ZRO is less than 500 ppm. The ZRO stability is also dependent on the stability of the gain
.
After reducing (22) to input angular velocity, it can be written
as

(23)
and have been
Compared with (19), the terms involving
removed. Otherwise, identical methods can be applied to minimize the input-referred ZRO.
V. EXPERIMENTAL RESULTS
The theory presented in Sections III and IV was studied experimentally by using a developed version of the MEMS angular
velocity sensor originally presented in [4] and [8]. The system
comprises a bulk micromachined sensor element, analog interface electronics implemented with a custom IC, and DSP implemented with a field-programmable gate array (FPGA) chip. A
block diagram of the system is shown in Fig. 6. The actual micromechanical sensor element is shown in Fig. 7. The operating
is
kHz, and the mode separation
frequency
.
In the system, the gyroscope output signals are converted into
ADCs ([10], Ch.
the digital domain with bandpass (BP)
5) after analog signal processing. The analog signal processing
comprises capacitance-to-voltage (C/V) conversion of the signals using continuous-time charge-sensitive amplifiers (CSAs),

Fig. 7. Left: Schematic drawing of the structure of the sensor element. Right:
Scanning electron microscope image of the middle (structural) wafer. (Pictures
courtesy of VTI Technologies, Vantaa, Finland.)

filtering of the signals, and level normalization of the signals


with an attenuator (Attn) in the primary and a variable-gain amplifier (VGA) in the secondary channel. Next, the signals are
phase-coherently demodulated to in-phase (I) and quadrature
(Q) components. Then, they are filtered and decimated to reduce the oversampling ratio and to achieve the final resolution.
Hz and the
The final decimated sampling frequency is
bandwidth is 40 Hz. The full-scale angular velocity is 100 /s.
The middle-electrode biasing impedance is made negligible by
in Fig. 5.
using a large external
The analog IC was encapsulated in a 120lead ceramic carrier
and it was combined with the sensor element on a printed circuit
board (PCB). The DSP part was realized with an FPGA chip
mounted on a separate PCB. The two PCBs were connected
together with ribbon cables.
This section is divided into two separate parts. First, in
Section V-A, two different implementations of the quadrature

1648

Fig. 8. Quadrature signal when the compensation is performed with a 7-bit


DAC illustrating the oscillation between two levels as a result of the limited
DAC resolution.

IEEE SENSORS JOURNAL, VOL. 7, NO. 12, DECEMBER 2007

Fig. 9. Spectrum of the resulting Coriolis output when the quadrature signal
varies between two levels as a result of limited compensation resolution. (7281, sampling frequency 125 Hz,
point FFT, Kaiser window with
/s).

= 13

100

FS =

compensation loops described in Section III will be compared,


showing how the limited DAC resolution causes oscillation
between two levels and how this affects the angular velocity
techoutput of the sensor. It will also be shown how the
niques described in Section III can be applied to eliminate the
oscillation. Second, in Section V-B, the ZRO sources of the
presented sensor will be studied. During the ZRO measurements, the quadrature compensation performed with the latter
implementation presented in Section V-A is used.
A. Quadrature Signal Compensation
The uncompensated mechanical quadrature signal of the
sensor element was measured as being 20 times the full-scale
Coriolis signal, so the need for compensation was evident.
In the first implementation, the quadrature compensation was
performed, as shown in Fig. 3. The controller was implemented
as a single integrator. The DAC is a 7-bit high-voltage (HV)
DAC with an LSB size of 30 mV, yielding a full-scale output
V [19]. Coarse tuning of the DAC is used to set the
of
output around a desired level. The measured quadrature signal
is shown in Fig. 8. As can be seen, the signal exhibits oscillation between two levels, with a difference of about 130 LSBs.
The average of the quadrature signal is zero. The spectrum
of the corresponding Coriolis signal can be seen in Fig. 9.
The spectral components indicated with triangles are caused
by the time-varying quadrature signal and the phase error in
synchronous demodulation.
The result indicates that the DAC resolution needs to be
increased by at least 8 bits in order to remove the spurious
components. This was done by implementing the quadrature
DAC (Fig. 4).
compensation next with a second-order
The controller was still implemented as a single integrator.
bits and
The resolution of the controller output was
the frequency of the interpolated signal 500 Hz. This gives an
. The interpolation was performed
interpolation factor of
with a zero-order hold. The same 7-bit HV DAC was used as
.
the output D/A converter, yielding
Fig. 10 shows a simulated output spectrum of the DAC. It
can be seen that there is a notch at 10 kHz, corresponding to

61

Fig. 10. Simulated output spectrum from the


DAC used for quadrature
compensation. Top: Overview of the spectrum. Bottom: Spectrum around the
operating frequency ! , with the horizontal line indicating the output band, sampling frequency
width. (400 000-point FFT, Kaiser window with
400 kHz,
: V).

FS = 1 92

= 13

the operating frequency. This means that cross-coupling should


have no effect on the SNR. The output bandwidth is indicated in

SAUKOSKI et al.: ZERO-RATE OUTPUT AND QUADRATURE COMPENSATION IN VIBRATORY MEMS GYROSCOPES

Fig. 11. Quadrature signal when the compensation is performed with a


DAC.

61

1649

Fig. 13. Spectrum of the resulting Coriolis output when the quadrature signal is
compensated with a dc voltage. (7281-point FFT, Kaiser window with
,
/s).
sampling frequency 125 Hz,

FS = 100

= 13

of the primary resonator excitation signal. The term


can easily be found by changing
and letting the primary
resonator amplitude control loop to keep the product
constant.
The measurement of the term
is not as
also affects
straightforward, as varying the product
the primary resonator vibration amplitude and, hence, the other
ZRO sources. However, the measurement can be performed by
modulating the dc voltage with a low-frequency [21], so that
the exciting voltage in (8) becomes

Fig. 12. Spectrum of the resulting Coriolis output when a


for compensation. (7281-point FFT, Kaiser window with
/s).
frequency 125 Hz,

FS = 100

61 DAC is used
= 13, sampling

the figure. Fig. 11 shows the measured quadrature signal. Now,


the oscillation observed in Fig. 8 has been removed. Fig. 12
shows a measured output spectrum of the Coriolis signal. For
comparison, a spectrum with the quadrature compensation performed with an external dc voltage source is shown in Fig. 13.
The spectral components at approximately 0.3 Hz are caused
by the time-varying primary signal. By comparing Figs. 12 and
13, it can be seen that the compensation has no effect on the
final SNR. This has been verified also with Allan variance analysis [20].
B. ZRO
Next, the various ZRO sources in the aforementioned system
were experimentally studied. A challenge is to find suitable
measurement methods which can distinguish between the various sources. By examining (22), it can be recognized that the
and the direct excitation
electrical cross-coupling
can
of the secondary resonator
and
be found by varying the dc and ac components

(24)
The frequency
can be freely chosen, as long as it lies within
the final output bandwidth of the angular velocity sensing
system and the phase shift of the system at that frequency is
known. From (24), it can be seen that there are no electri,
cally cross-coupling components apart from the signal at
whereas, because the electrostatic force is proportional to the
square of the exciting voltage, there are force components at
and
. Because of the high-quality factor of
the primary resonator, the primary movement excited by these
components is negligible, whereas the secondary movement
found.
can be measured and the parameter
and
Then, the terms
are still left. The sum of the last two terms can
be found by disconnecting the input of the secondary resonator
detection circuit from the sensor element, after which only the
signals cross-coupled in the electronics appear at the output. The
can be distinguished from clock signals by
term
letting it vary with time with a known frequency and measuring
the cross-coupling of the time-dependent component. Then,
can be solved by a simple subtraction operation. Finally,
the first term can be found by subtracting all the other components from the total ZRO.

1650

IEEE SENSORS JOURNAL, VOL. 7, NO. 12, DECEMBER 2007

TABLE I
MEASURED ZRO COMPONENTS, REDUCED TO INPUT

Fig. 14. Measured ZRO as a function of the peak-to-peak amplitude V


the square-wave primary resonator excitation signal.

of

, only the absolute value of


of the varying dc component
this component could be resolved, but not its sign. The absolute
value is 3.9 /s, accounting for 5.3% of the total ZRO. Finally,
the ZRO caused by nonproportional damping was evaluated by
subtracting all the other components from the total ZRO. The
/s or
/s, depending on the sign of the
result is either
component caused by direct excitation. This accounts for either
4.4% or 6.1% of the total ZRO.
The measured ZRO components are shown in Table I. All
values are calculated from five repeated measurements, except
, which is evaluated
the electrical cross-coupling
on the basis of line fitting. First, the signal cross-coupled in the
electronics was measured to be 62.1 /s, accounting for 84.5% of
the total ZRO. (All the percentages are calculated as ratios between the absolute value of each component and the total ZRO.
Therefore, they do not add up to 100%.) The parameter was
measured as being 16 dB higher than the parameter . This is
an expected result, as described in Section IV-A5.
Next, the electrical cross-coupling of the primary resonator
excitation signal was measured. Since the sensor element and
the interface IC were combined on a PCB, the ZRO proved to
be extremely sensitive to any disturbances. However, by keeping
the configuration completely intact during the measurement, the
electrical cross-coupling could be reliably resolved, although
it would change immediately if the measurement setup were
altered.
The measured ZRO as a function of the peak-to-peak amof the square-wave primary resonator excitation
plitude
signal is shown in Fig. 14. The figure also shows a fitted line,
together with the correlation coefficient , and the norm of the
residuals . The ZRO caused by the electrical cross-coupling is
.
then evaluated on the basis of the line and the nominal
The result is 10.8 /s, accounting for 14.7% of the total ZRO.
The total ZRO was measured in conjunction with the previous
measurement. Then, the ZRO caused by the direct excitation
was
of the secondary resonator
measured. As it was not possible to reliably control the phase

VI. CONCLUSION
In this paper, issues related to the ZRO of a vibratory microgyroscope were studied. Different sources of the ZRO that have
been reported in the literature were discussed, and they were
given a uniform representation. The discussed sources were: a
mechanical quadrature signal; nonproportional damping; electrical cross-coupling in the sensor element; direct excitation of
the secondary resonator, and cross-coupling in the electronics.
Attention was paid to the stability of each source. Methods
to minimize the effect of different sources were described,
whenever possible. The effects of synchronous demodulation
and electrostatic quadrature compensation performed with a dc
voltage on the final ZRO were analyzed.
Various ways to implement a control loop for the electrostatic quadrature compensation were presented, concentrating
on a case where the compensation is performed with a dc voltage
generated with a DAC and the controller is digital. In particular,
modulation
how to extend the resolution of the DAC by
and, hence, avoid the spurious components in the output signal
resulting from limited resolution was studied.
In the experimental work, a digitally controlled quadrature
compensation loop was implemented, first with a 7-bit HV
DAC with the same HV D/A conDAC, then with a 15-bit
verter as the output DAC. How the spurious components can be
removed without compromising the final SNR was shown.
Next, the ZRO sources for one particular MEMS gyroscope
implementation were studied experimentally. It was found that
cross-coupling in the electronics is the most important ZRO
source in this implementation. This is because the primary and

SAUKOSKI et al.: ZERO-RATE OUTPUT AND QUADRATURE COMPENSATION IN VIBRATORY MEMS GYROSCOPES

TABLE II
DEFINITIONS OF THE MOST SIGNIFICANT SYMBOLS
USED IN THE ZRO EQUATIONS

1651

A. Blomqvist, P. Klemetti, and H. Brachkov are acknowledged


for numerous valuable discussions. T. Salo is also acknowlADCs used in the
edged for implementing the bandpass
system.
REFERENCES

secondary resonator detection circuits were located right next


to each other in order to achieve good matching of their phase
shifts [4]. Another major source of the ZRO was the cross-coupling of the primary resonator excitation signal to the secondary
resonator detection circuit. This source was found to be extremely sensitive to variations in the measurement setup, as the
sensor element and the interface IC were combined on a PCB.
This needs to be taken into account when designing the final
packaging for the gyroscope, in order to keep the ZRO stable
over the time.
In the present implementation, the middle-electrode biasing
in
impedance was rendered negligible by using a large
Fig. 5. This was made possible by the PCB implementation.
In a fully integrated solution, this error source needs to be
analyzed more carefully.
modulation
As a possible extension in the future, the
could also be applied to the primary resonator excitation. Currently, the primary resonator output signal also exhibits variation as a consequence of a discrete number of possible control
.
steps in the excitation signal level
APPENDIX
The most significant symbols used in the ZRO equations are
collected in Table II for convenience.
ACKNOWLEDGMENT
The authors wish to thank VTI Technologies for the initial
ideas on the system architecture and for providing the sensor
elements, together with assistance and equipment in the rate
table measurements. In particular, K. Trmlehto, T. Salo,

[1] A. M. Madni, L. E. Costlow, and S. J. Knowles, Common design


techniques for BEI GyroChip quartz rate sensors for both automotive
and aerospace/defense markets, IEEE Sensors J., vol. 3, no. 5, pp.
569578, Oct. 2003.
[2] R. Neul, U.-M. Gmez, K. Kehr, W. Bauer, J. Classen, C. Dring,
E. Esch, S. Gtz, J. Hauer, B. Kuhlmann, C. Lang, M. Veith, and R.
Willig, Micromachined angular rate sensors for automotive applications, IEEE Sensors J., vol. 7, no. 2, pp. 302309, Feb. 2007.
[3] N. Yazdi, F. Ayazi, and K. Najafi, Micromachined inertial sensors,
Proc. IEEE, vol. 86, no. 8, pp. 16401659, Aug. 1998.
[4] M. Saukoski, L. Aaltonen, T. Salo, and K. Halonen, Readout and
control electronics for a microelectromechanical gyroscope, in Proc.
IEEE Instrum. Measurement Technol. Conf., Sorrento, Italy, Apr. 2006,
pp. 17411746.
[5] W. A. Clark, R. T. Howe, and R. Horowitz, Surface micromachined
Z-axis vibratory rate gyroscope, in Proc. Tech. Dig. Solid-Stale
Sensor and Actuator Workshop, Hilton Head Island, SC, Jun. 1996,
pp. 283287.
[6] P. Ward, Electronics for Coriolis force and other sensors, U.S. Patent
5 672 949, Sep. 30, 1997.
[7] A. Sharma, M. F. Zaman, and F. Ayazi, A 0.2 /hr micro-gyroscope
with automatic CMOS mode matching, in Proc. IEEE Int. Solid-State
Circuits Conf. Dig. Tech. Papers, San Francisco, CA, Feb. 2007, pp.
386387.
[8] M. Saukoski, L. Aaltonen, T. Salo, and K. Halonen, Integrated readout
and control electronics for a microelectromechanical angular velocity
sensor, in Proc. Eur. Solid-State Circuits Conf., Montreux, Switzerland, Sep. 2006, pp. 243246.
[9] A. S. Phani, A. A. Seshia, M. Palaniapan, R. T. Howe, and J. A. Yasaitis, Modal coupling in micromechanical vibratory rate gyroscopes,
IEEE Sensors J., vol. 6, no. 5, pp. 11441152, Oct. 2006.
[10] R. Schreier and G. C. Temes, Understanding Delta-Sigma Data Converters.. New York: Wiley, 2005.
[11] H. Xie and G. K. Fedder, Integrated microelectromechanical gyroscopes, J. Aerosp. Eng., vol. 16, no. 2, pp. 6575, Apr. 2003.
[12] M. S. Weinberg and A. Kourepenis, Error sources in in-plane silicon
tuning fork MEMS gyroscopes, J. Microelectromech. Syst., vol. 15,
no. 3, pp. 479491, Jun. 2006.
[13] H. Xie and G. K. Fedder, Fabrication, characterization, and analysis
of a DRIE CMOS-MEMS gyroscope, IEEE Sensors J., vol. 3, no. 5,
pp. 622631, Oct. 2003.
[14] J. A. Geen, A path to low cost gyroscopy, in Proc. Tech. Dig. SolidState Sensor and Actuator Workshop, Hilton Head Island, SC, Jun.
1998, pp. 5154.
[15] J. A. Geen and D. W. Carow, Micromachined Gyros, U.S. Patent 6
505 511, Jan. 14, 2003.
[16] L. Aaltonen, M. Saukoski, and K. Halonen, Upconverting capacitance-to-voltage converter for readout of a micromechanical
gyroscope, in Proc. IEEE Norchip Conf., Linkping, Sweden, Nov.
2006, pp. 267270.
[17] M. Lutz, W. Golderer, J. Gerstenmeier, J. Marek, B. Maihfer, S.
Mahler, H. Mnzel, and U. Bischof, A precision yaw rate sensor
in silicon micromachining, in Proc. Int. Conf. Solid-State Sens.
Actuators, Chicago, IL, Jun. 1997, pp. 847850.
[18] S. Gnthner, M. Egretzberger, A. Kugi, K. Kapser, B. Hartmann, U.
Schmid, and H. Seidel, Compensation of parasitic effects for a silicon
tuning fork gyroscope, IEEE Sensors J., vol. 6, no. 3, pp. 596604,
Jun. 2006.
[19] L. Aaltonen, M. Saukoski, and K. Halonen, On-chip digitally tunable
high voltage generator for electrostatic control of micromechanical devices, in Proc. IEEE Custom Integrated Circuits Conf., San Jose, CA,
Sep. 2006, pp. 583586.
[20] Recommended Practice for Inertial Sensor Test Equipment, Instrumentation, Data Acquisition and Analysis, IEEE Standard 1554, 2005.
[21] L. Aaltonen, P. Rahikkala, M. Saukoski, and K. Halonen, Continuous
time interface for 1:5 g closed-loop accelerometer, in Proc. IEEE
Int. Conf. IC Design Technol., Austin, TX, May 2007, pp. 187190.

1652

IEEE SENSORS JOURNAL, VOL. 7, NO. 12, DECEMBER 2007

Mikko Saukoski (S06) was born in Savukoski,


Finland, in 1978. He received the M.Sc. degree in
electrical engineering from the Helsinki University
of Technology (TKK), Espoo, Finland, in 2004.
Currently, he is working towards the Ph.D. degree
in electrical engineering at the Electronic Circuit
Design Laboratory, TKK.
His main research interests are electronics design
for microelectromechanical sensors and actuators,
and low-voltage, low-power analog circuit design.

Lasse Aaltonen (S07) was born in Turku, Finland,


in 1980. He received the M.Sc. and Lic.Sc. degrees
in electrical engineering from Helsinki University
of Technology (TKK), Espoo, Finland, in 2004 and
2006, respectively. Currently, he is working towards
the Ph.D. degree at the Electronic Circuit Design
Laboratory, TKK.
His main research interest is design of electronics
for microelectromechanical sensors.

Kari A. I. Halonen (M02) received the M.Sc.


degree in electrical engineering from Helsinki University of Technology, Espoo, Finland, in 1982, and
the Ph.D. degree in electrical engineering from the
Katholieke Universiteit Leuven, Leuven, Belgium,
in 1987.
Since 1988, he has been with the Electronic
Circuit Design Laboratory, Helsinki University of
Technology. Since 1993, he has been an Associate
Professor, and since 1997, a Full Professor at the
Faculty of Electrical Engineering and Telecommunications. He became the Head of the Electronic Circuit Design Laboratory in
1998. He is author or coauthor over 200 international and national conference
and journal publications on analog integrated circuits. He specializes in CMOS
and BiCMOS analog integrated circuits, particularly for telecommunication
applications.
Dr. Halonen has been awarded the Beatrice Winner Award at the ISSCC02
Conference in 2002. He is a TPC member of ESSCIRC and ISSCC. He has been
an Associate Editor of the IEEE TRANSACTIONS ON CIRCUITS AND SYSTEMS I,
a Guest Editor for the IEEE JOURNAL OF SOLID-STATE CIRCUITS, and was the
Technical Program Committee Chairman for the European Solid-State Circuits
Conference in 2000.

Das könnte Ihnen auch gefallen