You are on page 1of 22

8 Interface Circuits for

Capacitive MEMS Gyroscopes


Hongzhi Sun and Huikai Xie

CONTENTS
8.1 Operations of MEMS Gyroscopes........................................................................................ 161
8.1.1 Coriolis Effect............................................................................................................ 161
8.1.2 Excitation of the Drive Mode.................................................................................... 165
8.1.3 Matched versus Unmatched Modes........................................................................... 167
8.2 Read-Out Circuits.................................................................................................................. 168
8.2.1 Continuous-Time Sensing.......................................................................................... 168
8.2.1.1 Open-Loop Amplifiers................................................................................ 169
8.2.1.2 Transimpedance Amplifiers........................................................................ 172
8.2.2 Discrete-Time Sampling............................................................................................ 174
8.2.3 Discussions................................................................................................................ 177
8.3 Considerations for the Nonidealities..................................................................................... 178
8.3.1 Quadrature Error....................................................................................................... 178
8.3.2 Direct-Coupled Motions............................................................................................ 179
8.3.3 Phase Issues in the Drive Loop................................................................................. 180
8.4 Summary............................................................................................................................... 180
References....................................................................................................................................... 181

8.1  OPERATIONS OF MEMS GYROSCOPES


8.1.1  Coriolis Effect
Most of the commercially available microelectromechanical system (MEMS) gyroscopes for con-
sumer electronics are based on the vibratory concept [1–3], in which case the rotary rate is detected
through the so-called Coriolis effect. The Coriolis effect is intuitively illustrated in Figure 8.1. A
plate with an ideal smooth surface is rotating at a rate Ω, with respect to an inertial frame of refer-
ence, for example, the Earth, and a particle A starts to move toward point B at t0 with a constant
velocity v, whose direction points from A to B (Figure 8.1a). The plate is referred to as a rotating
frame of reference. In the inertial frame of reference, since the surface has no frictions, the particle
will move straightly with a constant velocity until it arrives at the other side of the plate at t0 + Δt
(Figure 8.1b). However, observed from point B, which is static in the rotating frame of reference, the
trajectory of the particle is curved, as shown in Figure 8.1c, as there is an extra force applied on the
particle. This effect is called the Coriolis effect. Consequently, the extra force is called the Coriolis
force and the acceleration caused by the Coriolis force is called the Coriolis acceleration. In this
chapter, the output signal of an MEMS gyroscope due to the Coriolis effect is generally called the
Coriolis signal.

The Coriolis effect can also be derived mathematically. If a particle has an observed velocity v ′

in a rotating frame of reference, which is rotating with an angular velocity Ω in an inertial frame of

161
162 MEMS: Fundamental Technology and Applications

(a) (b) (c)

B
A B A B
Ω Ω A

FIGURE 8.1  The trajectories of a moving particle in an inertial frame of reference and a rotating frame of
reference: (a) t = t0; (b) t = t0 + Δt; and (c) observed from B.

reference, its velocity observed from the inertial frame of reference is given by

  d   d      
v =  r  =  r  + Ω × r = v′ + Ω × r (8.1)
Downloaded by [Universiti Malaysia Perlis] at 02:42 01 February 2016

 dt  I  dt  R


where r is the position vector of the particle referred to the inertial frame of reference, and the
subscripts I and R represent the derivative calculations in the inertial and the rotating frames of
reference, respectively. For the validity of the derivation in general cases, all the variables in this
equation are in the format of vectors. Similarly, the accelerations in two frames of references satisfy

 d   d   
 dt v  =  dt v  + Ω × v (8.2)
I R

By substituting Equation 8.1 into Equation 8.2 and letting

 d  d 
a =  v and a ′ =  v ′
 dt  I  dt  R

the acceleration in the inertial frame of reference is expressed as

        
a = a ′ + 2(Ω × v ′ ) + Ω × (Ω × r ) + arot × r (8.3)

 
where a ′ is the acceleration of the particle referred to the rotating frame and arot the rotation accel-
eration. For a particle that has no acceleration in the inertial frame of reference, it is apparent that

a=0

By multiplying the mass on both sides of Equation 8.3, the net force in the rotating frame of
reference is given by
       
ma ′ = −2 m(Ω × v ′ ) − 2 mΩ × (Ω × r ) − 2 marot × r (8.4)

 
The term −2m(Ω × v ′ ) is the Coriolis force, whose direction is orthogonal to both the rotation
and the local velocity, and whose amplitude is proportional to the product of the rotary rate and the
    
local velocity. The term mΩ × (Ω × r ) is called the centrifugal force, and the term marot × r is due
to the rotary acceleration of the noninertial frame of reference. It can be proved later that the other
Interface Circuits for Capacitive MEMS Gyroscopes 163

two terms are out of the band of the Coriolis signal. Therefore, only the Coriolis force is focused on
at the beginning of analysis. Then, Equation 8.4 is reduced to

  
ma ′ = −2 m(Ω × v ′ ) (8.5)

So, the Coriolis effect causes a local linear acceleration that is proportional to the rotary rate,
and the acceleration can be detected by an accelerometer. In a vibratory gyroscope, the moving part
with a certain mass is called the proof mass.
Figure 8.2 shows a simplified mechanical model of a vibratory gyroscope that is used to detect the
rotation that is perpendicular to the paper plane. A frame is connected to a fixed wall through a spring
and a damper in the vertical direction, while the movements in the horizontal direction are limited
by the friction-free rollers. Inside the frame, a proof mass is connected to the frame in the same way
except that its movements are restricted within the horizontal direction. The gyroscope has two opera-
Downloaded by [Universiti Malaysia Perlis] at 02:42 01 February 2016

tion modes: the drive mode and the sense mode. In the drive mode, the frame is driven by external
forces, for example, electrostatic force and piezoelectric force, to move vertically. Owing to the support
of the rollers inside the frame, the proof mass moves together with the frame, so the value of Cs does
not change. In the sense mode, the Coriolis force induced by the rotation is in the horizontal direction.
Since the frame is not movable horizontally, the movements of the center proof mass cause the changes
in Cs, which is measured by the interface circuits. The governing equations of both modes are

d2 x dx
md + bd + kd x = Fd (8.6)
dt 2 dt

d2 y dy dx
ms + bs + k s y = 2 ms Ω (8.7)
dt 2 dt dt

Drive mode
Frame kd bd

bs

Ω
Sense mode Cs
ks

Cd

Rollers

FIGURE 8.2  Simplified model of a capacitive MEMS vibratory gyroscope.


164 MEMS: Fundamental Technology and Applications

In Equations 8.6 and 8.7, m, b, and k represent mass, damping coefficient, and spring constant,
respectively, and the subscripts d and s distinguish the drive mode and the sense mode, respectively.
Fd is the driving force, which is assumed sinusoidal, given by

Fd (t) = Fd sin (ω t) (8.8)

By solving Equation 8.6, we obtain the function of the displacements in the driving mode:

Fd
x (t ) = sin(w t − j d ) (8.9)
kd (1 − (w /w d )2 )2 + (w /Qdw d )2

where
Downloaded by [Universiti Malaysia Perlis] at 02:42 01 February 2016

w /w d
j d = arctan (8.10)
Qd [1 − (w /w d )2 ]

Similarly, the movements in the sense mode can be obtained by solving Equation 8.7 and then
substituting Equation 8.9 into the result, given by

2 Fd msw Ω(t ) 1 sin(w t − j d − j ds )


y(t ) = ⋅ (8.11)
kd ks (1 − (w /w d ) ) + (w /Qdw d )
2 2 2
(1 − (w /w s )2 )2 + (w /Qsw s )2

and

w /w s w /w d
j ds = j s − j d = arctan − arctan (8.12)
Qs [1 − (w /w s ) ]
2
Qd [1 − (w /w d )2 ]

In Equations 8.9 through 8.12, Q and ω are the quality factor and the natural resonant frequency,
respectively, and the following equations exist:

miw i ki
Qi = and w i = , i = d, s
bi mi

According to Equation 8.11, the rotation rate is modulated to the drive-mode frequency in the
Coriolis signal. Therefore, the rotation rate can be obtained from the amplitude of the Coriolis sig-
nal. The frequency of the drive mode is usually in the order of kHz, while the rotation rate is at most
several Hz, so the other terms in Equation 8.4 are out of the band of the Coriolis force.
The sensor shown in Figure 8.2 is capacitive because the mechanical signal is transduced into the
electrical domain through the change of the capacitor. The electrical signal can be voltage, current,
frequency, duty cycle [4–7], etc., depending on the architecture of the interface circuits. There are
many examples that use other mechanisms, such as piezoresistive [8,9], etc., as the bridge. However,
the capacitive approaches are the most popular in both academia and industries, owing to the advan-
tages of high sensitivity, good linearity, and low power consumption. Therefore, the discussion in
this chapter will focus on the circuitry design for capacitive MEMS gyroscopes.
Interface Circuits for Capacitive MEMS Gyroscopes 165

8.1.2  Excitation of the Drive Mode


The proof mass of a capacitive gyroscope is driven by an electrostatic force in the drive mode. To
illustrate how the electrostatic force works, Figure 8.3 shows two capacitors in series, where the two
outer electrodes are fixed and the central electrode is movable in the horizontal direction. The two
fixed plates are connected with two out-of-phase AC signals, Vac+ and Vac−, respectively, and the
central plate is biased with a DC voltage Vdc. The potential energy stored in each capacitor is

1 eA
E = ⋅ ⋅ (Vac − Vdc ) 2 (8.13)
2 g

where ε is the dielectric constant, A the overlap area, and g the distance between the two plates.
Since the electrostatic force tends to change the overlap area, the force between the two electrodes
Downloaded by [Universiti Malaysia Perlis] at 02:42 01 February 2016

is derived as

∂E 1 ew
Fe = − =− ⋅ ⋅ (Vac − Vdc )2 (8.14)
∂l 2 g

where w is the width of the electrode plate. For the plate driven by two differential signals, as shown
in Figure 8.3, the net force is given by

ew
Fnet = Fe + − Fe − = 2 ⋅ VacVdc (8.15)
g

It can be found by comparing Equations 8.14 and 8.15 that the differential driving cancels the
second-order harmonics in Equation 8.14 and achieves better linearity of the electrostatic force, so
it is desirable in the sensor design. So, the transfer function of the drive mode can be derived from
Equations 8.6 and 8.15 as

X (s) 2ewVdc 1
H (s) = = ⋅ 2 (8.16)
Vac (s ) g s + (w d /Qd )s + Qd2

The magnitude of Equation 8.16 achieves the peak value at ωd, which is the natural resonant
frequency of the system, and so does the Coriolis acceleration in response to the same rotation.

l Fixed

g
Movable

+
Vac+ –
Vbias Vac–
Fixed

FIGURE 8.3  Electrostatic force between two fixed plates and a horizontally movable plate.
166 MEMS: Fundamental Technology and Applications

Therefore, the proof mass should be driven at its resonance in the drive mode. The amplitude of the
displacement at the resonance is given by
|Fnet | ⋅ Qd
xmax = (8.17)
kd
The values of the quality factors in the drive mode are usually in the order of 102 for the gyro-
scopes working in the atmosphere, and the value of Q can be as high as tens of thousands for those
devices with vacuum packaging. The bandwidth of the mode, on the other hand, is adversely pro-
portional to Q, as shown below:
wd
BW = (8.18)
Q
Downloaded by [Universiti Malaysia Perlis] at 02:42 01 February 2016

So, the bandwidth may be as low as around 10 Hz for the devices at the atmosphere and even less
than 1 Hz for those in vacuum [4]. This means that it is very hard to predefine a fixed resonant fre-
quency for a batch of devices, if not impossible, considering the variations during fabrications and the
environmental factors. In a modern design, the resonant frequency is found by a self-oscillation loop
in which the MEMS structure works as a loop filter. Figure 8.4 shows the block diagram of a possible
choice of the whole interface circuits [4,5]. The drive-mode movements of the proof mass are first
detected by the read-out circuits, whose details will be discussed in Section 8.2. After being ampli-
fied by a controlled gain with a proper phase shift, the signal is fed back to the driving electrodes to
close the loop. The Barkhausen criterion should be satisfied to start up the oscillation, stated by

|HF/V (s)Hx/F (s)H V/x (s)G| = 1


(8.19)
∠ HF/V (s)Hx/F (s)H V/x (s)G = 0

Note that ∠ Hx/F (s) = −90° at the natural resonant frequency, so the electronics in the self-oscil-
lation loop should generate a 90° phase lead to meet Equation 8.19. Since an oscillation without any

|V|set

+
Automatic gain
control loop
Hctrl(s)

Interface circuitry Vctrl

Drive mode
Hv/x(s) G(Vctrl) |Vdrive|
Hx/F (s)

Vdrive
Phase
HF/V (s)
tuning

Rotation Sense mode Low-pass Rotary rate


Hv/x(s) Gain
Hx/F (s) filter
Demodulation

FIGURE 8.4  Block diagram of the whole system.


Interface Circuits for Capacitive MEMS Gyroscopes 167

a­ mplitude control will go to saturation, which means serious nonlinearities, another gain control
loop is present to keep the control signal at the proper level.

8.1.3  Matched versus Unmatched Modes


It has been proved that both the drive mode and the sense mode of a capacitive MEMS vibratory
gyroscope are second-order systems, which means that each mode has its own resonant frequency.
The two frequencies can either be identical to “match” the two modes or be different to leave the
two modes “unmatched.” The design of the two frequencies is very important, as it is directly
related to the performance of a sensor and the architecture of the electronics system.
We first examine the case in which the two modes are not matched and the drive-mode resonant
frequency is located at the pass band of the sense mode. The amplitude of the Coriolis-induced
displacement is given by
Downloaded by [Universiti Malaysia Perlis] at 02:42 01 February 2016

2 Fd msw Ω
|xCoriolis | = ⋅ Qd (8.20)
kd ks

According to Equations 8.10 and 8.12, φd = 90° and φds = −90°, so the Coriolis signal is in phase
with the driving signal. The signal in the self-oscillation loop can be used directly to demodulate
the Coriolis signal for the rotation rate.
Since the frequency response of the sense mode is flat at the pass band, the sensitivity of a sen-
sor barely changes as the drive-mode resonant frequency drifts. This means that the sensor has
stable sensitivity against environmental fluctuations and fabrication variations. The bandwidth of
the detectable rotation signal is

BWgyro = fs _ 3dB − fd (8.21)


where fs_3dB is the frequency at which the magnitude of the sense-mode transfer function curves up
by 3 dB and fd is the resonant frequency of the drive mode, as shown in Figure 8.5a.
If the drive mode and the sense mode are matched, their resonant frequencies are exactly
the same, so the proof mass resonates in both modes. According to Equations 8.11 and 8.16, the

(a) (b)
|H(s)| Drive mode |H(s)| Drive mode
Sense mode Sense mode

BWgyro

3 dB 3 dB

BWgyro

fd fs f (Hz) fs f (Hz)
fd

FIGURE 8.5  Bandwidth of (a) unmatched and (b) matched modes.


168 MEMS: Fundamental Technology and Applications

d­ isplacement at the resonant frequency is Q times larger than that in the pass band, where Q is the
quality factor. Therefore, the amplitude of the Coriolis-induced displacement is derived as

2 Fd msw Ω
|xCoriolis | = ⋅ Qd ⋅ Qs (8.22)
kd ks

The amplitude of the Coriolis signal is considerably larger than those sensors without mode
matching, considering the number of Qs can be from hundreds in the atmosphere, up to as high as
thousands in vacuum. Therefore, the mode matching improves the sensitivities significantly.
The main drawback of the mode-matched gyroscope is their narrow bandwidth. Since the
Coriolis acceleration is around the resonant frequency of the sense mode, the bandwidth of the
detectable rotation rate is determined by the quality factor of the sense mode:
Downloaded by [Universiti Malaysia Perlis] at 02:42 01 February 2016

BWgyro = fs/2Qs (8.23)

So, the bandwidth of the sensor is severely limited compared to the unmatched case. Moreover,
it is very hard to make the two modes by themselves, even though the parameters of both modes
are well designed, because the bandwidths of both modes are so narrow that a little environmental
change or a mismatch happening during fabrication will ruin the match. Usually, extra circuits
and electrodes on sensors are necessary to tune the resonant frequency dynamically to keep the
mode match. The technologies of tuning the mechanical structures are beyond the scope of this
chapter, and interested readers can refer to the literatures about mode-matching gyroscopes for
more details.
Moreover, both φd and φds in Equations 8.10 and 8.12 are equal to 90° when the two modes are
matched, so there is a 90° difference in phase between the driving signal and the Coriolis signal.
The driving signal is available from the self-oscillation loop, and extra circuitry might be necessary
to generate the quadrature signal to demodulate the Coriolis signal for the rotation rate.
On the basis of the aforementioned discussion, the gyroscopes with matched modes have higher
sensitivities, lower bandwidths, and possibly more complicated circuitry than the unmatched ones.
So, the mode-matching technologies are usually employed on the high-end sensors that need to be
sensitive to low rotary rate while the cost is not an important issue.

8.2  READ-OUT CIRCUITS


The read-out circuits work to convert the capacitive change to electrical signals that can be further
processed, such as voltage, current, frequency, and so on [4–7]. The electrical signals can come
either from discrete-time sampling [10] or from continuous-time transduction [4,5,11]. This section
will summarize the commonly used read-out circuits for capacitive MEMS gyroscopes.

8.2.1  Continuous-Time Sensing


The capacitance changes of the MEMS gyroscopes redistribute the charge stored on the sensing
capacitors and thus resulting in voltage or current variation. Correspondingly, the continuous-time
read-out circuits can be divided into these two major categories: voltage and current-sensing ampli-
fiers, the former of which amplify the voltage signal directly, whereas the latter of which convert
the current to voltage with certain amplification. In this section, we will discuss the voltage-sensing
circuits first, followed by the current-sensing circuits.
Let us look at the simplified model of the interface between the sensing capacitors and the
read-out circuits in Figure 8.6. The two variable capacitors with the initial capacitance of Cs0 are
biased with two AC signals, vm and −vm, and their common node is connected to the circuits. Their
Interface Circuits for Capacitive MEMS Gyroscopes 169

vm

Cs0 + ΔCs

is
vs

Cs0 – ΔCs Zin

−vm
Downloaded by [Universiti Malaysia Perlis] at 02:42 01 February 2016

FIGURE 8.6  Interface between the variable capacitors and the interface circuits.

capacitances are changed during rotation by ΔCs in absolute value but with opposite signs, and
ΔCs ≪ Cs0. Zin is the lumped input impedance of the read-out circuits. It can be easily derived that

∆Cs  1 
vs = v when Z in   (8.24)
Cs 0 m  jw Cs 0 

where ω is the radian frequency of vm.


So, the voltage signal at the input node is proportional to the capacitance change caused by the
Coriolis force. Similar results can be obtained if the bias voltage signals applied at the sensing
capacitors are DC instead of AC, except that the ω in Equation 8.24 becomes the drive-mode fre-
quency, which is much lower than the AC modulation signal; therefore, more stringent requirement
is applied to Zin to satisfy Zin ≫ (1/jωCs0).
As Zin drops down, vs starts to be attenuated as more current flows through Zin. The extreme con-
dition happens when Zin ≈ 0, in which case the voltage at the common electrode is grounded, and
the current flowing out of the node is given by
is = jωΔCsvm (8.25)
It is interesting to note that the net current is also proportional to the capacitance change. The
current-sensing circuits deal with the current and convert it to voltage signal with proper level.
Similar to the voltage-sensing case, the DC bias will lead to the same equation, and the value of ω
will refer to the resonant frequency of the drive mode.
According to the aforementioned analysis, the input impedance determines whether the voltage
sensing or the current sensing is applicable to the Coriolis signal detection. The actual interface
circuits are not exactly one of the two extreme cases. For the voltage sensing, open-loop architec-
tures are usually employed for their high input impedance; however, the parasitic parameters of the
circuits and the interconnections make the value of Zin finite, which attenuates the amplitude of vs
at the input nodes of the circuits. For the current sensing, on the other hand, the transimpedance
amplifiers are usually employed for the virtual ground they can provide, but the finite loop gain of
the op amp causes their input impedance to be nonzero. The simplified schematics of an open-loop
amplifier and a transimpedance amplifier are shown in Figure 8.7.

8.2.1.1  Open-Loop Amplifiers


For the open-loop amplifier shown in Figure 8.7a, the input transistor is biased at a proper DC
level through Zbias to ensure the right operation conditions. The AC impedance of Zbias should be
large enough so that the AC signal from the sensing capacitors is not affected. The commonly used
170 MEMS: Fundamental Technology and Applications

(a) (b) ⎯⎯
vm √4kBTRf
vn_modu vm
RL in_R vn_modu
Cs0 + ΔCs Cgd Zf Rf
Cs0 + ΔCs
vn_ f vn_them vout
vs Cf
gm

Zbias
+
Cs0 + ΔCs Cstray Cgs
Cs0 + ΔCs Cp
+
– ishot
Vbias vn_opamp
–vm –vm
Downloaded by [Universiti Malaysia Perlis] at 02:42 01 February 2016

FIGURE 8.7  Simplified schematics and noise models of (a) open-loop amplifier and (b) transimpedance
amplifier.

components include diodes, diode-connected MOSFETs, resistors with switching technologies, etc.
In this chapter, a reverse-biased diode will be chosen as an example for further analysis. The AC
signal at the input node is then given by

2 ∆Cs
vs = v (8.26)
2Cs 0 + C p m

where Cp is the parasitic capacitance coming from the input transistors and the stray capacitance of
the interconnections between the sensors and the circuits. Considering the Miller effect, the value
of Cp is calculated as

Cp = Cstray + Cgs + (1 + gm R L ) Cgd (8.27)

where gm is the transconductance of the input transistor, R L is the load resistance of the amplifier,
and Cgs and Cgd are the gate-to-source and the gate-to-drain capacitance, respectively. It is noted
that the Miller effect increased the parasitic capacitance, which attenuates the input signal more
seriously. One possible solution is adding a cascode transistor at the drain of the input transistor, to
reduce its output resistance, but the cost is smaller headroom for the output signal. The gain of the
open-loop amplifier is

G = gm R L (8.28)

so the output signal of the open-loop amplifier is

2 ∆Cs
vout = v g R (8.29)
2Cs 0 + C p m m L

Noise is the most important issue that affects a gyroscope’s performance, so the noise analy-
sis is necessary for all amplifiers. For open-loop architectures, the major noise sources include the
Brownian noise from the mechanical structure, the flicker noise and the thermal noise from the
amplifier, the shot noise from the DC-biasing diodes, and the noise injected from the modulation
signal source. The noise from the mechanical structures is white noise, whose spectral density is

Fm2 = 4 kBTb (8.30)



Interface Circuits for Capacitive MEMS Gyroscopes 171

The unit of Fm2 is N2/Hz, and kB, T, and b are the Boltzmann constant (1.38 × 10−23 J/K), the abso-
lute temperature, and the viscous damping coefficient, respectively. The Brownian noise is deter-
mined by the mechanical structure design and thus contributes the same to all interface circuits. So,
the electronic noise should be paid the most attention in the noise analysis.
The spectral density of the shot noise existing in the diodes for DC biasing can be written as

2
ishot = 2qI (8.31)

where q is the charge of a single electron and I the DC current. In a well-biased circuit, the DC leak-
age current is very small, so the shot noise is negligible.
The noise from the modulation signal source can inject into the input nodes of the read-out
amplifier; however, in fully differential architectures, the two input nodes are accepting almost the
Downloaded by [Universiti Malaysia Perlis] at 02:42 01 February 2016

same amount of noise, so most of the injected noise cancels out each other. Therefore, fully differ-
ential architectures are preferred to minimize the effect of the injected noise.
The flicker noise of a MOSFET arises from the extra energy states at the interface between the
silicon and the oxide layer. The flicker noise of a single transistor is modeled as a voltage source in
series with the gate, as shown in Figure 8.7a, and its spectral density is

K 1
vn2 _ f = ⋅ (8.32)
Cox WL f

where K is the process-dependent constant, Cox the gate capacitance density, and W and L are the
width and the length of the transistor, respectively. It is noticed that the power spectrum density of
the flicker noise is proportional to 1/f, so it is also called 1/f noise. As discussed above, vm can be
either AC or DC. If vm is DC, the input signal is the base-band Coriolis signal in the order of kHz, so
the flicker noise is the dominant noise source. Equation 8.32 suggests that increasing the transistor
size will reduce the flicker noise. However, a larger transistor contributes larger parasitic capaci-
tance, which attenuates the signal, so it might or might not improve the signal-to-noise ratio (SNR).
The transistor size for the optimum SNR is calculated from the following analysis.
The equation of SNR is given by

vs2
SNR = (8.33)
vn2 _ f ⋅ ∆f

The length of the transistor can take the minimum allowed value, considering all practical issues
like offset etc., because short transistor is faster while consuming the same power, as discussed later.
The optimized transistor width can be derived by letting


(SNR ) = 0 (8.34)
∂W

Substitute Equations 8.26, 8.32, and 8.33 into Equation 8.34, and the optimized width should
satisfy that

Cgs + (1 + gm R L) Cgd = 2Cs + Cstray (8.35)

The thermal noise is due to random fluctuation of carriers, and for a MOSFET, it mainly comes
from the channel. Thermal noise is a white noise, so it has uniform spectrum over all frequencies.
172 MEMS: Fundamental Technology and Applications

The thermal noise has much less power spectral density than the flicker noise does at low fre-
quency, but it will take over at high frequency. The frequency at which the flicker noise and the
thermal noise equals in power spectral density is called the corner frequency, and it is around tens
to hundreds of kHz for modern CMOS technologies. The Coriolis acceleration can be modulated
to a higher frequency in AC detection, in which case vm is an AC signal instead of DC. So, the dom-
inant noise source would be the thermal noise as long as the carrier frequency is higher than the
corner frequency, and the noise level is lower than the flicker noise at the base band. When mod-
eled as a voltage source in series with the gate, the spectral density of the thermal noise is given by

8 1
vn2 _ them = k T⋅ (8.36)
3 B gm

where gm is the transconductance of the transistor and is given by


Downloaded by [Universiti Malaysia Perlis] at 02:42 01 February 2016

gm = 2 mCox   I D
W
 L (8.37)

where µ is the mobility of the carriers and ID the drain current of the MOSFET. Equation 8.37 sug-
gests that the thermal noise can be reduced by increasing the transistor width or the bias current.
Increasing the width will lead to a similar issue on SNR optimization as in the case of the flicker
noise, and it can be proved that the optimized width satisfies

1
Cgs + (1 + gm RL )Cgd = (2Cs + Cstray ) (8.38)
3

According to Equations 8.36 and 8.37, the thermal noise can be reduced by increasing the power
consumption. However, since vn _ them ∝ I −1/ 4 , the power consumption needs to be increased by 16
times to reduce the noise voltage by a factor of 2. It is very inefficient, especially when considering
the more and more stringent requirements of modern consumer electronics on low power.
One drawback of the AC detection is that any initial mismatches of the sensing capacitors will
be modulated to the carrier frequency and be mixed with the Coriolis signal. Although the capaci-
tance mismatches and the Coriolis signal is distinguishable in frequency, it is still possible that the
mismatch signal can saturate the interface circuits or affect the linearity, considering the fact that
the Coriolis signal is weak compared to the initial mismatch. The other cost of the AC detection is
the higher power consumption. The gain–bandwidth product of an amplifier is given by

gm 2mCox (W /L )I D
GBW = = (8.39)
CL CL

Higher operation frequency requires higher bandwidth and thus higher power consumption.

8.2.1.2  Transimpedance Amplifiers


The transimpedance amplifier (TIA) [4,11], on the other hand, creates a virtual ground for the cur-
rent from the sensor and converts the current to voltage. Depending on the configuration, the tran-
simpedance can be resistive or capacitive. The amplifiers with resistive configuration are also called
transresistance amplifiers (TRAs). In some literatures, TIAs are referred only to those with resistive
transimpedance. The capacitive configuration is also called transcapacitance amplifier (TCA) or
Interface Circuits for Capacitive MEMS Gyroscopes 173

charge-sensitive amplifier (CSA). To avoid confusions, TIAs and CSAs will be used in this chapter
to call the transimpedance amplifiers with resistive and capacitive configurations, respectively.
We will talk about the TIAs first. The output signal of a TIA is calculated by multiplying the
current from the sensor with the transimpedance, revealed by

vout = is Zf = jωΔCsRf vm (8.40)

Here ω is the radian frequency of the modulated Coriolis signal if vm is AC, while it is the base-
band frequency if vm is DC. It is noticed that the output has a 90° phase shift, and its amplitude
is frequency-dependent. In other words, the TIA works as a differentiator. Similar to the mode-
matched sensors mentioned above, a quadrature clock is required to demodulate the Coriolis signal.
The feedback resistor can be implemented with a long MOSFET working in the linear region. For
the stability of the feedback, an extra capacitor Cf in parallel with the resistor is necessary, and the
capacitance should satisfy
Downloaded by [Universiti Malaysia Perlis] at 02:42 01 February 2016

1
Cf  (8.41)
jw R f

to meet the characteristics of the transfer function.


For a CSA whose transimpedance is capacitive, the output signal becomes

∆Cs
vout = is Z f = v (8.42)
Cf m

Compared to Equation 8.40, the output in Equation 8.42 is no longer dependent on frequency,
and the output signal is proportional to the ratio of the capacitance change over the feedback capaci-
tance. The CSA works as an integrator, so any DC current will accumulate on Cf. To prevent the
CSA from being saturated by the DC current, a resistor in parallel with Cf is necessary, and the
resistance should satisfy

1
Rf  (8.43)
jw C f

to meet the characteristics of the transfer function.


The feedback resistors of TIAs and CSAs contribute noises that are not negligible. The equiva-
lent noise at the output of a TIA is

vn _ fb = 4 kBTR f (8.44)

The equivalent noise referred to the capacitance change is given by

4 k B T /R f
Cn _ fb = (8.45)
jw vm

It can be proved that Equation 8.45 also works for CSAs. Therefore, large feedback resistors
are always beneficial for the noise performance of the current-sensing amplifiers. The feedback
resistance of a CSA is so large that its noise contribution is negligible. However, a trade-off exists
between the feedback resistance and the linear range of a TIA.
174 MEMS: Fundamental Technology and Applications

Assume that the total noise of the op amp referred to the input node is written as vn _ opamp , so the
output noise is given by

vn _ out = vn _ opamp ⋅ (1 + jw (Cs + C p )Z f ) + Cn _ fb ⋅ jw vm Z f (8.46)


By comparing Equation 8.46 with Equations 8.40 and 8.42, it can be observed that the TIA and
CSA have the noise gain larger than the signal gain. The SNR at the output is

2
 jw ∆Cs vm Z f 
SNR out =  (8.47)
 vn _ opamp ⋅ (1 + jw (Cs + C p )Z f ) + Cn _ fb ⋅ jw vm Z f 

Downloaded by [Universiti Malaysia Perlis] at 02:42 01 February 2016

Compared to the open-loop amplifiers that have the same gain of the signal and the input-
referred noise, the current-sensing amplifiers have lower SNR at the output. It is not a big issue when
ωΔCsZf  ≫ 1 and Cp is well controlled because the extra noise gain is negligible. However, the SNR
will be degraded significantly if the signal gain is close or less than unity, which happens on those sen-
sors with ultra-small sensing capacitance, for example, some gyroscopes integrated on CMOS chips.
Similar to the open-loop amplifiers, the dominant noise of TIAs and CSAs can be the flicker
noise or the thermal noise, depending on the operation frequency. The issues of initial mismatch and
higher power consumption at high carrier frequency also exist.
Since larger transistor size can reduce the noise level and the SNR simultaneously, there should
exist an optimum size for the best SNR. Skipping the tedious analysis, the equations of the transistor
size for optimized SNR are

Cgs + Cgd = Cs + Cstray (8.48)

for flicker noise and

1
Cgs + Cgd = (C + Cstray ) (8.49)
3 s

for thermal noise, respectively. For the ease of analysis, the signal gain is assumed much larger than
unity, so the term “1” in Equation 8.47 can be neglected. The term Cs should change to 2Cs for dif-
ferential capacitor pairs.

8.2.2  Discrete-Time Sampling


The discrete-time read-out circuits sample the signal under certain frequency, which is usually
much higher than the Coriolis signal, for further processing. The discrete-time read-out circuits
usually refer to the switched-capacitor (SC) circuits, which have been widely used on capacitive
MEMS sensors [10]. A typical SC amplifier is schemed in Figure 8.8.
In Figure 8.8, vref is a DC bias voltage applied at the sensing capacitors and Cs is the sensing
capacitor of the gyroscope. When Φ1 is high and Φ2 is low, the feedback capacitor C2 is discharged
and the voltage vref is stored at Cs. Since the clock frequency of Φ is much higher than the Coriolis
signal, the capacitance of Cs is considered constant during the cycle. When Φ2 is high and Φ2 is low,
the left side of Cs is grounded, whereas the other side is still virtually grounded, so Cs is discharged
and its stored charge is transferred to as C2, so the voltage signal sampled at the dropping edge of
Φ1 is transferred to the load capacitance with the gain of Cs/C2.
Interface Circuits for Capacitive MEMS Gyroscopes 175

Φ1

SW3

Φ1 C2
Vref – Φ2
Φ1
SW1 vout
Cs + SW4
Φ2
SW2 Φ2
C1

FIGURE 8.8  Schematic of a typical SC amplifier.

Cs
vout = ⋅V
Downloaded by [Universiti Malaysia Perlis] at 02:42 01 February 2016

(8.50)
C2 ref

If the circuit is configured fully differential, and the capacitances at the moment of sampling are
Cs + ΔCs and Cs − ΔCs, respectively, the differential output is given by

2 ∆Cs
vout = ⋅ Vref (8.51)
C2

Therefore, the capacitance variation is converted to voltage signal. The noises of the switched-
capacitor front ends mainly come from the switch noise, the noise from the op amp, as well as the
noise folded from high frequency owing to aliasing.
A transistor working as a switch is modeled as a resistor when they are on, so they contribute
thermal noise to the circuits. The power of the noise that is sampled to a capacitor is calculated by
integrating the noise spectral density across the capacitor over all the frequencies, as stated in

+∞

vn2 _ sw = ∫v 2
n_c df (8.52)
−∞

The analysis of the noises from all switches is tedious and case by case, depending on the posi-
tions of the switches and the switching sequences. Regarding the circuits shown in Figure 8.8, the
dominant switch noise comes from SW3, and the power of the noise at the output can be calculated
with the method described in [12].

kBT
vn2 _ sw = (8.53)
C2

It is interesting to notice that the noise power due to the switches is independent of the on-resis-
tance of the switches. Equation 8.53 suggests that the switch noise is reduced when the capacitor
size is increased. However, large C2 will reduce the signal gain, and the SNR will be attenuated
seriously when the signal gain drops below unity because the noise gain is always larger than unity.
Simple analysis of the circuit shown in Figure 8.8 reveals that the circuit suffers from the flicker
noise and the offset of the op amp because the base-band frequency of the Coriolis signal is low.
This problem can be solved by two approaches, chopper stabilization (CHS) and correlated double
sampling (CDS) [13]. In the CHS approach, the reference voltage changes its polarity in every two
adjacent sampling cycles, so the Coriolis signal is modulated to the sampling frequency, which
176 MEMS: Fundamental Technology and Applications

Φ′1

SW3
Φ1
Φ1 C2
– C1
Vref Φ ′1
SW1 vout
Cs
+
SW2 Φ2
Φ1
vos SW4 Φ2

FIGURE 8.9  Schematic of an example of the SC circuits with correlated double sampling.

is much higher than the base-band frequency. An example of the CDS approaches is shown in
Downloaded by [Universiti Malaysia Perlis] at 02:42 01 February 2016

Figure 8.9 [14], where vos represents the DC offset and low-frequency flicker noise because both
of them are regarded constant during sampling. In the sampling phase, the offset and the low-
frequency noise from the op amp are presampled and are stored onto a load capacitor. In Φ2, the net
charge flowing from Cs to C2 is Vref Cs, and the output of the op amp becomes
Cs (8.54)
vcds = vos + Vref
C2

The potential of the output node is given by

Cs (8.55)
vout = Vref
C2

Therefore, the offset and the low-frequency noise are cancelled by CDS. Based on the offset-
cancelling techniques, the noise contributed by the op amp is mainly thermal noise.
Another issue coming with the increased bandwidth is called noise folding, which is illustrated
in Figure 8.10. According to the sampling theorem, the sampled signal in the frequency domain is

+∞
1
X s (w ) =
T ∑ X (w − nw
s s ) (8.56)
n =−∞

Signal Pass band

Noise

2 fs fs 0 fs 2fs
Sampling

Signal Noise

2fs fs 0 fs 2fs

FIGURE 8.10  Illustrations of noise folding.


Interface Circuits for Capacitive MEMS Gyroscopes 177

Ts and ωs represent the sampling period and the sampling frequency, respectively. It can be observed
that the noise at the frequencies of the higher-order harmonics of the sampling signals will be folded
back to the base band and be added into the total noise. The equation assumes an infinite bandwidth,
which is not true in the real case. So, the number of the harmonics, n, is determined by the band-
width of the amplifier in the feedback loop. In SC circuits, the bandwidth of the amplifier is usually
several times larger than that of the signal to ensure that the circuit behaves properly. Assuming the
bandwidth of the amplifier is BW, then

= Nfs (8.57)
BW 

The spectrum of the signal and the noise will be

+N
1 fs f
vs ( f ) =
Ts ∑ v ( f − nf ),
0 s −
2
< f < s
2
Downloaded by [Universiti Malaysia Perlis] at 02:42 01 February 2016

n =− N

1 (8.58)
= v (f)
Ts 0

and
+N
1
ns ( f ) =
Ts ∑ n ( f − nf )
0 s (8.59)
n =− N

The power of the signal and of the noise are

1
Ss ( f ) = S (8.60)
Ts2 0

and

1
Ns = NN 0 (8.61)
Ts2

So, the SNR after sampling is reduced by a factor of N:

Ss 1 S0
= ⋅ (8.62)
Ns N N0

8.2.3  Discussions
The main advantages of the discrete-time interface circuitry are the robustness and the good com-
patibility with other discrete-time signal-processing blocks, for example, the signal-delta A/D con-
verters. Since the virtual ground is available, no extra biasing circuitry is necessary and the signal is
less sensitive to the parasitic parameters. Because of the kT/C switching noise and the noise folding,
however, the noise performance of the switched-capacitor circuits is worse than the continuous-time
counterparts.
The continuous-time circuits avoid the extra noises existing in the sampling process, but the A/D
converter may cost more area and/or complicates the whole designs. The choice of the architecture
depends on the applications. Because of the possible attenuation caused by the parasitic capacitance,
178 MEMS: Fundamental Technology and Applications

the open-loop amplifiers are more suitable for the monolithic solutions, which integrate both the
sensors and the interface circuits on the same chips and thus have low parasitics. Owing to the limi-
tation of the fabrication technologies, the sensing capacitance of the monolithic gyroscopes is usu-
ally much smaller than that of those with separate device dies. The gain of the open-loop amplifiers
is independent on the sensing capacitance, so high gain is achievable to avoid the SNR degradation
of the transimpedance amplifiers.
The two-chip solutions that have two separate dies for MEMS structures and circuits, however,
have larger sensing capacitance and larger parasitic capacitance from the bonding pads. The current-
sensing amplifiers can achieve high gain easily, so the SNR degradation is no longer a big issue.
Moreover, the virtual ground provided by the negative feedback prevents the signal attenuation caused
by the parasitics. Therefore, the current-sensing amplifiers are more suitable for the two-chip solutions.
Another advantage of the TIA is the inherent 90° phase shift. Recall that the self-oscillation loop
in the drive mode requires a total 90° phase shift over the whole loop, and the TIAs can be used
as the front ends to avoid extra phase shifter. Furthermore, the transimpedance of a TIA is usually
Downloaded by [Universiti Malaysia Perlis] at 02:42 01 February 2016

implemented with a long MOSFET, whose equivalent resistance is tunable by the gate voltage, so
the TIA used in the driving loop can simplify the design of the automatic gain control loop.

8.3  CONSIDERATIONS FOR THE NONIDEALITIES


In the foregoing discussion in this chapter, we assume that a gyroscope is an ideal second-order
system in both the drive mode and the sense mode, and the two modes do not interfere with each
other. In fact, it is not the truth. Various nonidealities exist in real gyroscopes, and some of them
may affect the performance of a gyroscope so significantly that they have to be considered during
the electronics design. This section will discuss these important nonidealities.

8.3.1  Quadrature Error


The first important nonideality to discuss is the quadrature error. Owing to the defects and/or varia-
tions of the fabrication process, imperfections such as asymmetries and anisoelasticities may hap-
pen to the MEMS structures, which couple movements into the sense mode when the proof mass
is driven by an external force. The extra movements are in phase with those in the drive mode. In
comparison, according to Equation 8.12, the Coriolis signal is orthogonal to the drive-mode dis-
placement, so the error signal is a quadrature in phase of the Coriolis signal. For the ease of analysis,
the mixed signal out of the interface circuits is written as

vmix (t) = ACoriolis cos ωt + Aquadrature sin ωt (8.63)

Since the drive-mode movement is usually three to four orders stronger in amplitude than that
induced by the Coriolis effect, even a small coupling factor will cause the quadrature error to be more
significant than the Coriolis signal. It can be imagined that the real Coriolis signal will be easily over-
whelmed if the quadrature error is left without any processing. Since the frequency of the quadrature
error is the same as that of the Coriolis signal, it cannot be removed by filtering. One solution to
remove the quadrature error is the so-called synchronous demodulation. If the demodulation signal is

m(t) = cos (ωt) (8.64)

thus the signal after demodulation is given by

vdm (t ) = vmix (t )m(t )


1 1
= A (1 + cos 2w t) + Aquadrature sin 2w t (8.65)
2 Coriolis 2
Interface Circuits for Capacitive MEMS Gyroscopes 179

After passing through a low-pass filter, only the Coriolis signal will be retained. The chal-
lenges applied by the quadrature error on the interface circuit design are: first, although the
quadrature error is removable with signal-processing technologies, it is not distinguishable from
the desired signal in the front ends, so the front-end amplifier should have enough dynamic range
to prevent being saturated by the quadrature error. The gain of the front ends could be limited,
and the power consumption could be higher to keep the same gain bandwidth, especially for the
open-loop designs. Second, the phase delay of the electronics must be controlled precisely. If the
phase of the mixed signal is delayed by Δφ, the signals before and after the demodulation are
given by

vmix (t) = ACoriolis cos(ωt + Δφ) + Aquadrature sin(ωt + Δφ) (8.66)

and
Downloaded by [Universiti Malaysia Perlis] at 02:42 01 February 2016

1 1
vdm (t ) = ACoriolis [cos ∆j + cos(2w t + ∆j )] + Aquadrature [sin(2w t + ∆j ) + sin ∆j ] (8.67)
2 2

After the low-pass filter, the quadrature error retained in the final result is

1
vqe = A sin ∆j (8.68)
2 quadrature

There will be a nonzero output even when there are no rotations, and the error is called the zero-
rate output (ZRO). The bandwidth of a front end needs to be sufficient for small phase delay, and the
cost would be higher power consumption and/or more noise folding.
Another method for cancelling the quadrature error is adding extra driving fingers to the sense
mode, so that an electrostatic force that counters the quadrature error can be applied to the sense
mode directly. Alternatively, the quadrature error cancellation can be achieved by injecting an elec-
tric signal that is out of the phase of the quadrature error into the interface circuits of the sense
mode. These methods, however, require extra feedback loop and thus the optimization of the system
level, so they are out of the scope of this chapter.

8.3.2  Direct-Coupled Motions


We assume that all the springs of the gyroscope are ideally 1D, so the proof mass moves only in
the directions of the springs. However, the springs of real devices are realized with beams, so the
stiffness in the other two dimensions is not infinite. Therefore, off-axis movements exist when the
proof mass is driven along a direction. Since the spring constants in the off-axis directions are much
higher than that of the in-axis direction, the resonant frequency of the cross-axis mode is much
higher than the in-axis mode. According to Equation 8.10, the cross-axis motions are orthogonal
to the drive mode, and thus are in phase with the Coriolis signal, so the coupled movements of this
type are also called direct coupling.
The method of phasing out the quadrature error with synchronous demodulation does not
apply to the direct coupling signal because the direct coupling is not distinguishable even in phase
from the Coriolis signal. However, the dynamic cancellation with external mechanical/electrical
signal, as used for the quadrature error cancellation, is probably one solution. If the quadrature
error has been removed completely, the ZRO of a gyroscope indicates the extent of the direct cou-
pling. The ZRO can work as a reference to control the amplitude of the external mechanical/elec-
trical signals to cancel the effects of the direct-coupled motions, as discussed for the cancellation
of the quadrature error. For the interface circuits without the dynamic cancellation technologies,
180 MEMS: Fundamental Technology and Applications

the front-end circuitry needs to have large dynamic range to tolerate the signal amplitude due to
the direct-coupled motions.

8.3.3 Phase Issues in the Drive Loop


As discussed in this chapter, the resonance of the drive mode is excited by a self-oscillation loop,
and the criteria on amplitude and phase should be satisfied together to start up the oscillation. In this
section, the effects caused by the nonideal phase issue existing in the electronics will be examined.
If all the circuits in the self-oscillation loop generate a total phase delay of Δφ, to satisfy the
phase relationship in Equation 8.19, the phase of the transfer function of the drive mode is given by

∠Hx/F (s) = −90° + Δφ (8.69)

The resonant frequency of the drive mode is then shifted by


Downloaded by [Universiti Malaysia Perlis] at 02:42 01 February 2016

tan(∆j )
∆f = ⋅ fd (8.70)
Qd

Considering the high quality factor of the drive mode, the frequency drifts very slightly and the
self-oscillation is still excitable. However, the phase that the driving force leads the Coriolis signal
is also shifted by Δφ. Since the reference clock for synchronous demodulation comes from the driv-
ing signal of the self-oscillation loop, which is in phase with the driving force, the amplitude of the
Coriolis signal after demodulator becomes

1
Vdm = A cos( ∆j ) (8.71)
2 Coriolis

and the quadrature error is given by

1
vqe = A sin(∆j ) (8.72)
2 quadrature

The quadrature error then contributes to the ZRO. The effect of quadrature error removal of
the synchronous demodulation is degraded by the extra phase delay existing in the driving loop.
Therefore, the phase of the circuitry in the self-oscillation loop needs to be controlled precisely. The
bandwidth of the interface circuitry should be sufficiently large, after careful balance among the
phase delay, the power, and the noise folding.

8.4 SUMMARY
This chapter introduces the basic knowledge of the interface circuits for capacitive MEMS gyro-
scopes, along with the principles of the gyroscopes, and the practical considerations of the nonide-
alities. The basic architectures of both discrete-time- and continuous-time-sensing technologies are
discussed in this chapter, and the noise of both cases are analyzed in details. All circuits have their
own advantages and disadvantages, so there is no “ideal option” that fits all sensor designs. The
choice of the architecture of the interface circuits should be the one after balancing all the factors,
including but not limited to sensor technologies, noise, power, and cost. The technologies of the
signal detection electronics are developing very fast, and many novel designs are not included in this
chapter, so interested readers can refer to the up-to-date literatures for more information.
Interface Circuits for Capacitive MEMS Gyroscopes 181

REFERENCES
1. ISZ-505 Single-axis z-gyro product specification, Sunnyvale, CA, InverSense Inc. 2011.
2. LY330ALH MEMS motion sensor: high performance ± 300 dps analog yaw-rate gyroscope. Switzerland,
ST Microelectronics. 2010.
3. ADXRS450 High performance digital output gyroscope. Norwood, MA, Analog Devices, 2011.
4. A. Sharma, M. F. Zaman, and F. Ayazi, A 104-dB dynamic range transimpedance-based CMOS ASIC for
tuning fork microgyroscopes, IEEE J. Solid-State Circuits, 42(8), 1790–1802, 2007.
5. H. Sun, K. Jia, X. Liu, G. Yan, Y. W. Hsu, R. M. Fox, and H. Xie, A CMOS-MEMS gyroscope interface
circuit design with high gain and low temperature dependence, IEEE Sens. J., 11(11), 2740–2748, 2011.
6. Z. Ignjatovic and M. F. Bocko, An interface circuit for measuring capacitance changes based upon capac-
itance-to-duty cycle (CDC) converter, IEEE Sens. J., 5(6), 403–410, 2005.
7. C. Hierold, A. Hhildebrandt, U. Naher, T. Scheiter, B. Mensching, M. Steger, and R. Tielert, A pure
CMOS micromachined integrated accelerometer, in Proc. IEEE Micro Electro Mechanical Systems
Workshop (MEMS ’96), San Diego, CA, 174–179, 1996.
8. N. Yazdi, F. Ayazi, and K. Najafi, Micromachined inertial sensors, Proc. IEEE, 86(8), 1640–1659, 1998.
Downloaded by [Universiti Malaysia Perlis] at 02:42 01 February 2016

9. R. Voss, K. Bauer, W. Ficker, T. Gleissner, W. Kupke, M. Rose, S. Sassen, J. Schalk, H. Seidel, and E.
Stenzel, Silicon angular rate sensor for automotive applications with piezoelectric drive and piezoresis-
tive read-out, 9th Int. Conf. Solid-State Sensors and Actuators, Chicago, IL, 879–882, 1997.
10. X. Jiang, J. I. Seeger, M. Kraft, and B. E. Boser, A monolithic surface micromachined Z-axis gyroscope
with digital output. Digest of Technical Papers. 2000 Symposium on VLSI Circuits, Honolulu, 16–19,
2000.
11. J. A. Geen, S. J. Sherman, J. F. Chang, and S. R. Lewis, Single-chip surface micromachined integrated
gyroscope with 50°/h Allan deviation, IEEE J. Solid-State Circuits, 37(12), 1860–1866, 2002.
12. R. Schreier, J. Silva, J. Steensgaard, and G. C. Temes, Design-oriented estimation of thermal noise in
switched-capacitor circuits, IEEE Trans. Circuits Syst. I, 52(11), 2358–2368, 2005.
13. C. C. Enz and G. C. Temes, Circuit techniques for reducing the effects of opamp imperfections:
Autozeroing, correlated double sampling, and chopper stabilization, Proc. IEEE, 84(11), 1584–1614,
1996.
14. N. Wongkomet and B. E. Boser, Correlated double sampling in capacitive position sensing circuits for
micromachined applications, in Proc. IEEE Asia-Pacific Conf. Circuits Syst., Chiangmai, Thailand,
November 1998, pp. 723–726.
Downloaded by [Universiti Malaysia Perlis] at 02:42 01 February 2016

You might also like