You are on page 1of 12

IEEE TRANSACTIONS ON CIRCUITS AND SYSTEMS–I: REGULAR PAPERS, VOL. 64, NO.

4, APRIL 2017 753

A Chopper Instrumentation Amplifier With Input


Resistance Boosting by Means of Synchronous
Dynamic Element Matching
Federico Butti, Massimo Piotto, and Paolo Bruschi

Abstract— In this work, we propose a method to increase resistor trimming and can accomplish chopper modulation
the parasitic input resistance caused by application of chop- with minimal changes [11], [12].
per modulation to indirect current feedback instrumentation A weakness of ICF in-amps with respect to their volt-
amplifiers. The result is obtained by applying dynamic element
matching to the input and feedback ports at the same frequency age feedback counterpart is that the input resistance is not
as chopper modulation. The proposed approach requires effective increased by feedback. This is a serious drawback when
offset ripple rejection and equalization of the input and feedback combined with chopper modulation, which introduces a para-
common mode voltages. An in-amp architecture that meets both sitic switched-capacitor resistance across the input terminals.
requirements and embodies the proposed input resistance boost- This resistance is due to the periodic charging and discharg-
ing method is described. Experimental verification is provided by
means of a prototype designed and fabricated using the 0.32 μm ing of the input capacitance caused by the input modula-
CMOS devices of the STMicroelectronics BCD6s process. The tor [13]. With typical input capacitance values and chopper
amplifier operates with a 3.3 V supply voltage and a total current frequencies, the parasitic resistance can fall below 1 M.
absorption of 170 μA. An input impedance in excess of 1 G Considering that the output resistance of integrated sensors
has been measured at a chopper frequency of 20 kHz. The input (e.g., thermoelectric micro-transducers) may be of the order
referred voltage noise density is 18 nV/sqrt(Hz) with a flicker
corner of 0.2 Hz and 200 Hz bandwidth. of several tens of k, the resulting input attenuation sig-
Index Terms— Chopper amplifier, current feedback, input
nificantly degrades gain accuracy. As it will be explained
resistance, instrumentation amplifier. in next section, compensation of the input attenuation dur-
ing the sensor calibration phase may be a non-optimal
I. I NTRODUCTION solution.
In this work, we propose a method to increase the input
I NSTRUMENTATION amplifiers (in-amps) represent the
ideal solution for interfacing sensors whose output signal
is a differential voltage [1]. The continuous progress in the
impedance of ICF chopper instrumentation amplifiers by sim-
ply using a Dynamic Element Matching (DEM) approach,
field of integrated sensors and the growing interest in fully consisting of periodically swapping the input and feedback
integrated systems for bio-potential monitoring are motivating signals at the transconductor inputs (port swapping). A similar
the design of new in-amps [2]–[7], with particular emphasis technique was originally proposed as a solution for the gain
on reduced area, low power consumption and high common error caused by input transconductor mismatch [14]. How-
mode rejection ratio (CMRR). In most cases, the signal ever, in earlier implementations, the DEM frequency was a
bandwidth extends to very low frequencies or even includes submultiple of the chopper frequency. As it will be shown
DC. In CMOS circuits, this dictates the adoption of dynamic in next section, a large increase of the input impedance can
approaches to cancel the input offset voltage and reduce flicker be achieved only if DEM and chopping are synchronous
noise [8], [9]. Among these techniques, chopper modulation, (SCPS: synchronous chopping and port swapping). For this
being immune to noise fold-over, represents an optimal choice method to be applicable, the output voltage should be free
in terms of power consumption vs. noise trade-off. A very from ripple; furthermore, the common mode voltages of the
popular architecture is the indirect current feedback (ICF) input and feedback signals should be as close as possible.
in-amp [10], which reaches high CMRR’s with no need of To reject all contributions to the output ripple, an amplifier
with a second order low pass frequency response has been
Manuscript received July 28, 2016; revised October 28, 2016; accepted designed [15], [16], while the input and feedback common
November 19, 2016. Date of publication December 21, 2016; date of current mode voltages are equalized by means of a closed loop
version March 27, 2017. This paper was recommended by Associate Editor
A. Worapishet. approach [17]. So far, experimental demonstration of the
F. Butti is with Marvell Semiconductor, Pavia, Italy and also with Dialog mentioned techniques was not provided yet. In addition, the
Semiconductor, Livorno, Italy (e-mail: Federico.Butti@diasemi.com). input impedance boosting effect of SCPS is described in this
M. Piotto is with IEIIT-PISA, CNR, Pisa, Italy (e-mail:
massimo.piotto@ieiit.cnr.it). paper for the first time.
P. Bruschi is with the Dipartimento di Ingegneria dell’Informazione, Experimental validation of the proposed ideas is provided
University of Pisa, Italy (e-mail: p.bruschi@iet.unipi.it). by means of detailed measurements performed on a prototype
Color versions of one or more of the figures in this paper are available
online at http://ieeexplore.ieee.org. designed using the 0.32 μm, 3.3 V CMOS devices of the
Digital Object Identifier 10.1109/TCSI.2016.2633384 STMicroelectronics process BCD6s (Bipolar-CMOS-DMOS).
1549-8328 © 2016 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission.
See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.

Authorized licensed use limited to: Indian Institute of Technology Hyderabad. Downloaded on December 31,2022 at 09:08:27 UTC from IEEE Xplore. Restrictions apply.
754 IEEE TRANSACTIONS ON CIRCUITS AND SYSTEMS–I: REGULAR PAPERS, VOL. 64, NO. 4, APRIL 2017

Fig. 2. Charge transfer from the vin port during a 1-to-0 clock transition for
the traditional chopper configuration. The same charge transfer occurs during
the opposite clock transition.

approach. The chopper frequency is the result of several trade-


Fig. 1. Standard indirect current feedback (ICF) architecture. The table on
offs: high chopper frequencies improve flicker noise rejection
the bottom introduces the signal definitions used throughout the paper. and allow for wider amplifier bandwidth; whereas low chopper
frequencies reduce power consumption and result in lower
residual offset [8]. Optimal chopper frequencies for low offset
II. P RINCIPLE OF O PERATION in-amps are of the order of a few tens of kHz.
As far as the parasitic input resistance [13] is concerned,
A. Standard ICF Topology With Chopper Modulation
the mechanism involved is illustrated in Fig. 2 where the input
A typical fully differential chopper stabilized ICF in-amp is part of the circuit in Fig. 1 is represented, assuming C pin =
shown in Fig. 1. Switch matrices Sin , Sfb and So , controlled by C p f b = C p . At each clock transition, the input differential
the clock signal ck (frequency fck , 50 % duty-cycle), transmit capacitance undergoes a total voltage variation equal to 2v in ,
the signal through the solid lines when ck = 1 and through so that the charge drawn from the input source is 2C p v in .
the dashed ones when ck = 0. The input, feedback and output Considering that identical charge transfers occur at the two
signals are characterized by v in , v f b and v out differential opposite clock transitions, the equivalent input resistance is
mode voltages and VC M I , VC M F and VC M O common mode given by:
voltages, respectively. Block β is a resistive voltage divider 1
Rin−chop = (3)
such that v f b = βv out , with β = R1 /(R1 + R2 ), while with 4 f ch C p
Gmi and Gmf we indicate both the input transconductors and The input resistance forms a voltage divider with the out-
their transconductances, defined as the ratio of the output put resistance of the signal source, introducing an unwanted
differential current over the input differential voltage. The attenuation that degrades the amplifier gain accuracy.
input differential capacitances of Gmi and Gmf are C pin and In principle, the input resistance given by (3) can be
C p f b , respectively. Amplifier OP and capacitors C form a increased by reducing either the chopper frequency or the input
Miller integrator. device area (i.e., C p ). Unfortunately, this increases the flicker
Keeping ck fixed to a constant logic value (clock turned noise contribution to the residual noise in the baseband. In fact,
off), the circuit in Fig. 1 behaves like a first order low pass the maximum tolerable input attenuation sets a lower limit to
filter with cut-off frequency: the achievable noise figure [18]. The flicker noise coefficients
1 βG m f of a typical CMOS process are such that, for a given sensor
fC = (1) output resistance, it is possible to size the Cp fch product in
2π C
order to obtain, at the same time, an input noise density
If the DC loop gain of the circuit is much greater than comparable to the thermal noise of the sensor and an input
one, then the DC closed loop gain of the ICF in-amp can be attenuation of a few percent. This allows compensation of
approximated by: the input attenuation during sensor calibration with negligible
G mi 1 resolution loss. Such an operation is less effective when the
A= (2)
Gm f β in-amp is included in a versatile sensor interface that has to
Since, by design, G mi = G m f , the nominal gain is 1/β. provide optimal performances with sensor output resistances
When the clock is activated, v in and v f b are modulated by Sin spread over an interval of more than a decade. In this case, the
and Sfb and demodulated by So , resulting virtually unchanged, risk is to have too much noise at the lower end of the resistance
while the output offset and flicker noise of Gmi and Gmf are interval or too much attenuation at the higher end. Possible
processed only by So , which shifts them to higher frequencies dependence of sensor output resistance on temperature can
where they can be removed by a low pass filter (not shown in also make calibration critical.
Fig. 1). In particular, modulation of the offset voltage produces
a periodic waveform with frequency f ch (offset ripple), which B. Proposed Input Resistance Boosting Technique
is difficult to reject with fully integrated filters and represents Fig. 3 shows the proposed input switching mecha-
one of the most serious drawback of the chopper modulation nism (SCPS). At the one-to-zero clock transition, the

Authorized licensed use limited to: Indian Institute of Technology Hyderabad. Downloaded on December 31,2022 at 09:08:27 UTC from IEEE Xplore. Restrictions apply.
BUTTI et al.: CHOPPER INSTRUMENTATION AMPLIFIER WITH INPUT RESISTANCE BOOSTING 755

Fig. 3. Charge transfer from the vin port during a clock transition for the
proposed approach.

signals v in and v f b = βVout are swapped together. At the


same time, the inverting and non-inverting terminals are also
swapped, producing chopper modulation of both signals. For
the sake of simplicity, let us consider that both v in and Fig. 4. Effects caused by application of DEM when the input and feedback
v out are constant signals. Then, at any clock transition, the common mode voltages are different. (a) Input transconductors with common
mode capacitances Ccm ; (b) input differential pair with relevant capacitances;
charge sourced by v in is equal to C p (v in − βv out ). In an ideal (c) variation of M1-2 Vgs mean value across a DEM period.
ICF amplifier (infinite DC loop gain), v out = v in /β, so that
charge transfer is zero and the input resistance is infinite.
The analysis can be easily extended to non-DC input signal,
Ripple is reflected in the feedback signal, degrading v in
provided that the spectrum of v in is confined to frequencies
vs. v f b matching, hence reducing the effectiveness of the
much lower than f ch , so that we can still consider v in and
impedance boosting mechanism. Rejection of the output ripple
v out to be constant across a clock period. In these conditions,
by simply cascading an active filter is likely to introduce
charge transfer over a clock period is 2C p (v in − βv out ).
significant offset and flicker noise contributions, unless the
Indicating the amplifier closed-loop frequency response with
filter occupies much more area than the amplifier, or off-chip
H ( f ), the average current in a clock period becomes:
capacitors are used [19]. Feedback loops requiring demodula-
2C p v in [1 − β H ( f )]. Hence, the input impedance is given by:
tion of the individual ripple components have been demon-
1 strated to be very effective [20]–[22]. Other solutions that
Z in ( f ) = . (4)
2 [1 − β H ( f )] f ch C p make use of switched capacitor filters [23], [24] to reject the
Comparing this expression with (3), it is apparent that the offset ripple require sampling of the amplifier output voltage
impedance boosting is effective for signals well within the and, consequently, introduce noise fold-over. Combinations of
amplifier bandwidth, where β H ( f ) is close to one, while the auto-zero and chopper techniques, implemented with a ping-
actual advantage gradually decreases as the amplifier cut-off pong architecture [25], are much less prone to offset ripple
frequency is approached. artifacts, but are less effective in terms of power consumption
Note that swapping the input and feedback signals is equiva- and compactness. Digital trimming of the input pair can be
lent to swapping the input transconductors, similar to the DEM used when the focus is on compactness and less effective ripple
technique proposed in [14]–[16] as a method to reduce the rejection is tolerable [26]. In this work, ripple rejection was
impact of G mi /G m f mismatch on the gain accuracy. However, accomplished by embedding a second order low-pass filtering
if f D E M is a submultiple of the f ch (e.g., f D E M = fch /4 function into the amplifier [15], [16].
in [14]) part of the chopper transitions is still accomplished As far as the effect of input common mode voltage mis-
by simply inverting the connections on the input and feedback match is concerned, Fig. 4 illustrates the phenomena involved.
capacitance, just as shown in Fig. 2. These chopper transitions, First, at each DEM transition, the common mode compo-
which are not accompanied by a DEM swap, still draw a full nents of the input capacitances, indicated with Ccm in Fig. 4(a),
charge packet equal to 2v in C p . The only effect is reducing experience a voltage variation just equal to VC M I − VC M F .
the frequency by which full charge packets are drawn. For As a result, an input bias current equal to 2Ccm f D E M (VC M I −
example, with f D E M = f ch /2 the input resistance is only VC M F ) is drawn from the input source. Second, input common
doubled. With lower f D E M / f ch ratios, as in [14], the effects mode alternation caused by port swapping alters the operating
are even smaller. It is then apparent that the condition to obtain point of the input transconductors. Let us consider the case of
a much larger input resistance-boosting factor is to combine an n-type input pair, as shown in Fig. 4(b), and suppose, for
each chopper transition to a DEM transition, i.e., f D E M should simplicity, that the input differential voltage is very small, so
be equal to f ch . that VG S1 ∼ VG S2 . At any DEM transition, VG S1 and VG S2
Application of the impedance boosting approach requires undergo a variation with respect to their rest value (VG S Q ),
that two conditions are fulfilled: i) the ripple superimposed given by:
on v out is negligible; ii) input and feedback common mode VG S = VC M
2C G S
(5)
voltages coincide. 2C G S + C S−gnd

Authorized licensed use limited to: Indian Institute of Technology Hyderabad. Downloaded on December 31,2022 at 09:08:27 UTC from IEEE Xplore. Restrictions apply.
756 IEEE TRANSACTIONS ON CIRCUITS AND SYSTEMS–I: REGULAR PAPERS, VOL. 64, NO. 4, APRIL 2017

FB includes a resistive voltage divider and a “common-mode


differential amplifier” (CMDA), forming a local feedback loop
that causes VC M F to track VC M I .
The amplifier ideal behavior in terms of differential com-
ponents is represented by the block diagram of Fig. 5(b). In
the rest of this section, we will demonstrate this equivalence
by finding expressions for ω01 , ω02 and βeff and introducing
the main sources of non-ideality. Let us start by analyzing
block FB. The CMDA produces an output voltage, V H , given
by:
V H = AC M D (VC M I − VC M F ) (6)
with AC M D  1. Nominally, R1 A = R1B = R1 and R2 A =
R2B = R2 . We have distinguished the upper part of the divider
(R1 A , R2 A ) from the lower one (R1B , R2B ) in order to take
into account an unwanted common-mode to differential-mode
coupling that depends on the mismatch between the upper
and lower transfer function of the divider. It is convenient
to express these functions in terms of mean value β and
mismatch β:
Fig. 5. (a) Schematic view of the proposed architecture. (b) Ideal block R1 A β R1B β
≡β+ ; ≡β− (7)
diagram where differential signals are represented by a single wire for R1 A + R2 A 2 R1B + R2B 2
simplicity.
Considering that the effect of Smf is swapping the output
terminals, so that common mode components are unaffected,
where VC M is the difference between the input and feedback the following expression can be derived:
common mode voltage. The VG S variation is recovered after β
a time that can be a significant fraction of the DEM period. VC M F = βVC M O + (1 − β) V H ± v out (8)
4
Fig. 4(c) shows a sketch of a possible VG S behavior in the Combining (8) with (6), we get:
case that VC M I > VC M F . Since the transconductance of
the stage monotonically increases with VG S , the input signal (1−β) AC M D βVC M O ± v out β/4
VC M F = VC M I + .
is processed on average by a higher transconductance than 1 + (1 − β) AC M D 1 + (1 − β) AC M D
the feedback one. The opposite occurs if VC M I < VC M F . (9)
According to (2), a gain error is introduced. Due to the For (1 − β)AC M D  1, VC M F tends to VC M I , regardless
configuration of the resistive voltage divider universally used of VC M O and v out , obtaining the desired common mode
for the feedback path (see Fig. 1), VC M F = VC M O , so that
equalization.
this error is tolerable only for VC M I values close to VC M O . Thus, in order to find the response of the FB block to
For VC M O –VC M I exceeding a few hundred mV, temporary
differential signals, we can assume for simplicity that VC M F is
shutdown of one of the input transconductors is likely to occur
perfectly equalized to VC M I , obtaining the following equation:
at each DEM transition, exacerbating the adverse effect on the
amplifier gain. v f b = βe f f v out m(t) + εC M (10)
The problems illustrated in Fig. 4 have been encountered
where βeff and εCM are defined as:
also in [14], where bootstrapping of M1 and M2 substrates  
to reduce the input common mode capacitances was proposed β 2
βe f f ≡ 1 + β∼
= β;
as an effective solution. In this work, we solved this issue by 4β (1 − β)
forcing VC M F to track VC M I by means of a local feedback β β
loop, which was already proposed in [17] as a way to obtain εC M ≡ (VC M I − VC M O ) (11)
β 1−β
an ICF in-amp with rail-to-rail input ranges.
while m(t), as customary for chopper amplifiers, indicates
a dimensionless unity-amplitude square waveform of fre-
III. P ROPOSED A MPLIFIER A RCHITECTURE
quency f ch . The term εCM , which represents the mentioned
A. Description of the Architecture common-mode to differential mode coupling, is an unwanted
The proposed amplifier, shown in Fig. 5(a), is formed side-effect of the proposed common mode equalization mech-
by the three fully differential blocks, indicated with INT1, anism. Placing modulator Sm f before the divider allows for
INT2 and FB. INT1 is a Gm-C integrator, based on the simple frequency separation of the spurious term εCM from
two-input-port composite transconductor OTA1. INT2 is a the useful signal (βe f f v out ), since only the latter is modulated
two-input integrator, based on a Gm-C-Opamp architecture. by m(t).
Switch arrays Sps , S0 and Smf are controlled by a clock In order to analyze integrator INT1, it is useful to consider
signal of frequency f ch and 50 % duty-cycle. Feedback block that composite transconductor OTA1 is equivalent to the two

Authorized licensed use limited to: Indian Institute of Technology Hyderabad. Downloaded on December 31,2022 at 09:08:27 UTC from IEEE Xplore. Restrictions apply.
BUTTI et al.: CHOPPER INSTRUMENTATION AMPLIFIER WITH INPUT RESISTANCE BOOSTING 757

transconductors Gmi and Gmf of Fig. 1 when the following


substitutions are done:
G m Ai Gm A f
G mi ≡ G mb ; G m f ≡ G mb (12)
Gm L Gm L
Note that transconductors GmAi and GmAf , loaded by GmL ,
act as a two-port pre-amplifier for Gmb . By design, GmAi and
GmAf are identical. Indicating their nominal value with GmA
the pre-amplification factor is GmA /GmL .
Switch array Sps produces chopper modulation and port
swapping (DEM) at the same frequency f ch = f D E M . The
only difference with respect to the scheme of Fig. 3 is Fig. 6. Block diagram representation of the proposed in-amp with main error
that v f b is alternated with v in , but not modulated, since, as sources.
explained above, the feedback signal is already modulated
by Smf .
where the characteristic angular frequency ωc , quality factor
Considering (10), and indicating OTA1 total output noise
Q and DC gain A0 are given by:
and offset currents with i on1 and i os1, respectively, the differ-
ential mode current at the output of S0 can be written as: βe f f ω01 1
⎧  ωc = 2π f c = βe f f ω01 ω02 ; Q = ; A0 =

⎪ G mi v in − G m f βe f f v out + εC M + i os1 + i on1 ω02 βe f f

⎨ for ck = 1 (20)
i o1 = 

⎪ G m f v in − G mi βe f f v out − εC M − i os1 − i on1

⎩ Since a Butterworth-like response has been adopted,
for ck = 0 f c coincides with the -3 dB cut-off frequency of the amplifier.
(13) The effect of all error sources on v out is represented by the
following expression:
Defining G m1 and G m1 as:
G mi + G m f v outd (s) = (v d1B + v d1C m(t)) H L P (s) + v n2 H B P (s) (21)
G m1 ≡ ; G m1 ≡ G mi − G m f ; (14)
2 where H B P (s) is a band pass transfer function given by
the output current i o1 can be rewritten as
s/(ωc Q)

 H B P (s) = (22)
i o1 = G m1 v in − βe f f v out + v d1B + v d1C m(t) (15) s 2 /ωc2 + s/ (ωc Q) + 1
where the baseband and modulated error sources, v d1B and Note that |H B P | reaches a maximum value of one at
v d1C , respectively, are defined as: ω = ωc , in contrast with |H L P | whose maximum value
G m1 is A0 . For A0  1, this contributes to relax INT2 noise
v d1B ≡εC M (16) constraints. Furthermore, H B P rejects v n2 DC components,
2G m1
making chopper modulation not necessary for INT2.
G m1 
v d1C ≡ v in + βe f f v out − εC M + v io1 + v n1 As far as INT1 is concerned, its equivalent input noise
2G m1 sources are processed by the overall amplifier response H L P .
(17)
It is possible to estimate the order of magnitude of the
where v n1 = i on1 /G m1 and v io1 = i os1/G m1 are the OTA1 baseband component v d1B by using (11) and (16) with a
input referred noise and offset voltages. conservative value of 10−2 for both β/β and G m1 /G m1
Neglecting the error sources v d1B and v d1C in (15), and con- and assuming a nominal amplifier gain of 200 (i.e., β =
sidering that INT2 is a standard, two inputs, fully differential 1/200). The v d1B value found in this way is lower than 1 μV
Gm-Op-amp integrator, it can be easily shown that the ideal for a VC M I − VC M O mismatch up to 2 V, making this error
block diagram of Fig. 5(b) actually represents the amplifier source negligible for most applications.
differential mode response when the following expressions are Voltage v d1C includes various components that are modu-
used for the unity gain angular frequencies of the integrators: lated by m(t) producing different contributions. OTA1 noise
G m1 G m2 voltage v n1 results in a flat noise density [8] roughly equal to
ω01 = ; ω02 = (18) v n1 spectral density at f = f ch . This is the dominant con-
C1 C2
tribution to the input noise voltage in the amplifier passband.
Considering also error sources v d1B and v d1C , given by
All the other v d1C components are constant (v io1 ) or quasi-
(16) and (17), and indicating with v n2 the sum of INT2 input-
constant over a clock period, provided that the input signal
referred noise and offset voltages, the block diagram of Fig. 6
is limited to frequencies much smaller than f ch . Modulation
can be easily derived from the ideal system of Fig. 5(b).
of these terms by m(t) is the origin of the output ripple.
The system depicted in Fig. 6 implements a second order
In particular, modulation of v io1 is the cause of the offset ripple
low pass filter with transfer function given by:
while the term proportional to G m1 /G m1 turns into the DEM
v out A0 ripple. Differently from previous works, such as [14], there
H L P (s) ≡ = 2 2 (19)
v in s /ωc + s/ (ωc Q) + 1 is an additional ripple source (εCM ), which is proportional

Authorized licensed use limited to: Indian Institute of Technology Hyderabad. Downloaded on December 31,2022 at 09:08:27 UTC from IEEE Xplore. Restrictions apply.
758 IEEE TRANSACTIONS ON CIRCUITS AND SYSTEMS–I: REGULAR PAPERS, VOL. 64, NO. 4, APRIL 2017

to the input/output common mode mismatch through (11).


To estimate the order of magnitude of total ripple amplitude, it
is reasonable to assume the value 10−2 for all mismatch factors
(e.g., G m1 /G m1 ) and 1 mV for v io1 . Considering a DC gain
of 200 (β = 0.005), an input voltage (v in ∼ = βVout ) of 10 mV
and a VC M I − VC M O mismatch of 1 V, then the input ripple
amplitude is at most a few mV. Note that, differently from
previous works using different DEM and chopper frequencies,
all the ripple contributions mentioned above are modulated
at f ch .
Effective ripple rejection is obtained by exploiting the
second order low-pass transfer function of the amplifier.
By setting f c two decades below f ch , the ripple is amplified by
a factor that is 80 dB smaller than the passband gain. Referring
the output ripple to the input port using the passband gain,
the ripple equivalent amplitude is smaller than 1 μV, which
is a level comparable with amplifiers using the best ripple
suppression approaches [14], [20], [21].
Note that, if the ripple is actually reduced to the values
cited above, only the baseband components can be considered.
In these conditions, (15) clearly indicates that v in and the feed-
back component βeff v out are processed by the same transcon-
ductance G m1 . Thanks to this property, for high enough a loop
gain, v in = βeff v out , even in the presence of G mi /G m f mis-
match. This guarantees that the resistance boosting mechanism
introduced by port swapping and described in Section II-B is
robust against mismatch of the input transconductors.
To achieve this result with a chopper frequency around
a few tens of kHz f c should be a few hundred Hz. Fig. 7. Elements of the OTA1 composite transconductor: (a) Schematic
views of the GmAi ,-GmAf - GmL preamplifier; (b) Gmb transconductor with
Using (18), (19) and (20) with feasible on-chip capacitance embedded chopper modulator So, split into So1 and So2.
values for C1 and C2 , it can be easily shown that G m1 and G m2
should be of the order of a few μSiemens. This has a negative
effect on the input equivalent noise voltage in the passband, B. Topology of the Main Blocks
which, as mentioned above, is dominated by OTA1 input noise
The simplified schematic view of the input preamplifier is
density at f = f ch . The problem arises from the relationship
shown in Fig. 7(a), where the role of G m Ai and G m A f is
between the transconductance (G m ) and the input PSD (power
played by nominally identical differential pairs M1, M2 and
spectral density) of the thermal noise voltage (SV T H ) which,
M3, M4, respectively (input devices), while the function of
for a single stage transconductor, is given by:
G m L is accomplished by the MO1, MO2 pair (load devices).
1 The architecture is that of a telescopic amplifier where IL , I0 ,
SV T H = m F · 4k B T (23)
Gm IG and I2 represent cascode current sources. The input devices
are biased in deep subthreshold to obtain the best trade-off
where k B is the Boltzmann constant, T the absolute temper-
between input thermal noise and current consumption. On the
ature and m F a topology dependent factor, typically greater
contrary, the load devices operate in strong inversion, with an
than one [27]. With G m1 of the order of√1 μS, the input noise
as large as possible overdrive voltage, in order to maximize
density of OTA1 would exceed 100 nV/ Hz, which is too high
the output range. As a result, the preamplifier gain is given
for many sensor interfacing applications. With a single stage
by:
transconductor, noise reduction can be obtained only through a gm−in I0 (VG S − VT H )0
G m increase, followed by a proportional capacitance increase A1 = = (24)
gm−out nVT 2I2
(to keep ωc constant), leading to excessive area occupation.
The two-stage architecture of OTA1 shown in Fig. 5(a) allows where VT is the thermal voltage (= k B T /q), gm−in , gm−out
for a better area vs. noise trade-off [15]. In practice, the are the transconductances of the input and load devices,
input noise PSD is reduced by a factor equal to the pre- respectively, (VG S −VT H )0 is the overdrive voltage of the load
amplifier gain with respect to a single stage having the same devices and n is the sub-threshold slope factor of the input
transconductance. The preamplifier adds a noise contribution devices. Transconductor Gmb , shown in Fig. 7(b), is based
that, fortunately, can be minimized with no side effects on on a folded cascode architecture where the voltage to current
the amplifier cut-off frequency. Notice that pre-amplification conversion is operated by the pseudo-differential pair M5-6,
is not necessary for INT2 transconductors when A0  1, due providing acceptable linearity over a wider input differential
to the more relaxed noise constraints mentioned earlier. range with respect to conventional pairs.

Authorized licensed use limited to: Indian Institute of Technology Hyderabad. Downloaded on December 31,2022 at 09:08:27 UTC from IEEE Xplore. Restrictions apply.
BUTTI et al.: CHOPPER INSTRUMENTATION AMPLIFIER WITH INPUT RESISTANCE BOOSTING 759

TABLE I
M AIN D ESIGN PARAMETERS

TABLE II
D IMENSIONS OF S ELECTED D EVICES

Due to (25), G m2 is given by gm9 /2. On the other hand, (26)


Fig. 8. (a) Complete INT2 schematic view; (b) CMDA architecture. is exploited to avoid a CMFB circuit for INT2. Considering
the OP large common-mode-to-common mode gain and that,
due to feedback connection of Fig. 5, Vcm B is equal to VC M O ,
Capacitors Cz compensate for the positive zero due to M5- the latter simply tracks Vcm A , which is stabilized by the CMFB
6 gate-drain capacitance, thus reducing the phase lag at the circuit of the previous block (INT1).
chopper frequency. The overdrive voltage of the input pair, Finally, Fig. 8(b) shows the CMDA amplifier. The structure
fixed to a large value (0.72 V) to obtain a wide input range, is is that of a conventional two-stage op-amp, with duplicated
controlled by the preamplifier output common mode voltage. input differential pairs. Inverting (−) and non-inverting termi-
On the other hand, the overdrive voltage of the common gate nals (+) are connected as in Fig. 5. The class AB output
devices M23-24 was set to a small value (100 mV) in order stage provides enough impulsive current to drive the large
to obtain a gm large enough. In this way, it has been possible input capacitance of the preamplifier in presence of fast
to obtain a flat frequency response well beyond the chopper input common mode transients. The bias chain on the left of
frequency. Common mode voltages of both the preamplifier Fig. 8(b) provides a quiescent current that exhibits the proper
and transconductor Gmb are stabilized by conventional CMFB dependence on Vdd required to obtain an acceptable PSSR
circuits. Chopper modulators So1 and So2 perform the opera- from this kind of topology [28].
tion represented by So in Fig. 5.
Fig. 8(a) shows INT2 schematic view, where the role IV. E XPERIMENTAL R ESULTS AND D ISCUSSION
of the two transconductors (Gm2 ) is played by M9-10 and The prototype was fabricated using the 0.32 μm, 3.3 V
M11-12 pairs. The op-amp (OP in Fig. 5) is implemented by CMOS subset of the Bipolar-CMOS-DMOS BCD6s process
the simple M13-14 pseudo-differential common source stage. (STMicroelectronics). The nominal DC gain ( A0 = R2 /R1 +1)
Considering that IT = 2I B by design, the differential and com- was set to 200 and the target input noise density to
mon mode components of the transconductor output currents, 20 nV/sqrt(Hz) for a chopper frequency of 20 kHz. In order
indicated with I2n , I2 p in the figure, can be approximated by to obtain an effective ripple rejection, a cut off frequency
the following expressions: f c = 200 Hz with a Butterworth type response was chosen.
gm9  The main parameters that appear in the block diagram of
I2 p − I2n = Vdi f f B − Vdi f f A (25)
2 Fig. 5 are summarized in Table I, while the dimensions of
I2 p + I2n gm9 selected MOSFETs are reported in Table II. High resistivity
= (Vcm B − Vcm A ) (26)
2 2 polysilicon resistors are used in the feedback voltage divider,
where gm9 = gm10 = gm11 = gm12 , Vdi f f A and Vdi f f B are the while all capacitors are of polysilicon/gate oxide/n-implant
input differential voltages of port A and port B, respectively type. Chopper modulators Sps and Smf are implemented using
(see Fig. 5) and Vcm A and Vcm B are their respective common complementary p-n pass gates in order to maintain a low
mode voltages. resistance over a wide voltage range. Modulators So1 and So2

Authorized licensed use limited to: Indian Institute of Technology Hyderabad. Downloaded on December 31,2022 at 09:08:27 UTC from IEEE Xplore. Restrictions apply.
760 IEEE TRANSACTIONS ON CIRCUITS AND SYSTEMS–I: REGULAR PAPERS, VOL. 64, NO. 4, APRIL 2017

Fig. 9. Photograph of the test chip area occupied by the in-amp. The
chip layout is superimposed on the photo in order to represent the circuit
devices, buried under the metal dummies. Blocks are as follows: S: Sps ; switch Fig. 10. (a) Output differential voltage as a function of input differential
array CM: CMDA amplifier; PRE: preamplifier; GMB: Gmb transconductor, voltage for VC M I = 1.4 V; (b) output common mode voltage as a function
VD: feedback voltage divider (R1 ,R2 ), BC: Bias circuits; CAP: C1 and C2 of input differential voltage for various VC M I values.
capacitors.

auxiliary circuits used for diagnostic purposes. These blocks


use only n-MOSFET and p-MOSFET, respectively. The are visible in the lower part of Fig. 9.
size of Sps devices, reported in Table II (same size for The circuit operates with supply voltages in the range
n and p MOSFETS), has been set to a value much larger 2.7-3.6 V with a total current consumption of 170 μA,
than the minimum in order to reduce their thermal noise. including the oscillator. Most of this current is used by the
A similar choice has been operated for the Smf switches, preamplifier (60 %) and OP (24 %). The output common mode
but the reason here was to minimize the effect on the gain, voltage was set to 1.4 V, to comply with the input range of G m2
since the switch series resistance adds up to R2 A and R2B transconductors. Unless differently specified, VC M I was set
resistances. Simulations showed that the Smf series resistance to 1.4 V.
caused a gain increase of about 0.8 % with respect to the The DC input-output characteristics have been measured
theoretical value (R2 /R1 + 1). using a parameter analyzer (HP 4145B). Frequency responses
OTA1 stage was designed considering the input noise volt- were acquired using a waveform generator (Agilent 33220A)
age PSD at f ch , as the main specification, since it sets the and a 16 bit 2-channel digitizer (Pico Technology Ltd, mod.
equivalent input noise voltage PSD in the baseband of the ADC216), both controlled by a personal computer. The same
in-amp. The OTA1 input noise was equally distributed between digitizer, combined with a programmable anti-alias filter, was
contributions from Gmb and from the preamplifier. used for noise measurements. A differential to single-ended
The former contribution was controlled by using the rela- (and vice versa) conversion board with calibrated attenua-
tively high pre-amplification gain reported in Table I. On the tions was used to interface the mentioned instruments to the
other hand, sizing of the preamplifier was aimed to make all amplifier. All experiments were performed with Vdd = 3.3 V
contributions from devices other than the input MOSFETs and, unless differently specified, VC MI = VC M O = 1.4 V,
negligible. Then, the input devices were pushed to deep f ch = 20 kHz.
subthreshold to optimize the gm /I D ratio and, consequently, The output differential voltage is plotted in Fig. 10(a)
the thermal noise vs. bias current trade-off. This justifies the as a function of the input differential voltage. The small
large M1-M4 aspect ratios reported in Table II. The non- signal gain, estimated by means of a linear fit in the
minimum length was chosen to increase the device area and, [−2 mV, +2mV] interval, is 201.2. The discrepancy with
consequently, reduce flicker noise contribution, which was not respect to the design value (200) is due to the men-
negligible even at the chopper frequency. More details on tioned effect of Smf series resistance. The linearity error
the heuristic approach adopted for the preamplifier design are is less than 0.5 % of full scale in the interval of input
reported in [16]. The clock signal was provided by an internal voltages ± 10 mV, corresponding to an output swing of
relaxation oscillator operating at 2 f ch , followed by a fre- nearly ± 2 V. The linearity error is less than 500 ppm of
quency divider (a single T-type flip-flop). In this way, a clock full scale for |v in | < 5mV.
at frequency f ch and precise 50 % duty-cycle is obtained. Fig. 10(b) shows the dependence of VC M O on the input
An external resistor was used to tune the oscillator frequency. differential voltage, for various VC M I values. The curves are
An optical micrograph of the amplifier is shown in Fig. 9, practically coincident for VC M I in the range 0.7-2.2 V. Lower
where the main blocks are specified. The circuit layout was or higher values alter the correct operation of the preamplifier.
aligned and superimposed on the actual micrograph to repre- A significant dependence of VC M O on the input differential
sent the circuit devices, otherwise hidden under planarization voltage starting for |v in | > 5 mV can also be observed.
dummies. The total area occupied by the in-amp is 0.57 mm2. This phenomenon, which is identical for negative v in values,
The test chip includes also a clock generator and other occurs when the positive-going output terminal exceeds the

Authorized licensed use limited to: Indian Institute of Technology Hyderabad. Downloaded on December 31,2022 at 09:08:27 UTC from IEEE Xplore. Restrictions apply.
BUTTI et al.: CHOPPER INSTRUMENTATION AMPLIFIER WITH INPUT RESISTANCE BOOSTING 761

Fig. 11. DC gain relative variation as a function of the input common mode
voltage. The gain for VC M I = 1.4 V is taken as the reference.

input range of the corresponding p-differential pairs of INT2,


disrupting the mechanism by which VC M O tracks the output
common mode voltage of the previous stage.
Fig. 11 shows the effect of VC M I on the DC gain. The
relative variation with respect to the value at VC M I = 1.4 Fig. 12. Magnitude (a) and phase (b) response of the amplifier with different
signal source resistances (RS ), compared with the ideal 2nd order low pass
fall within ±0.2 % over a VC M I interval even wider than that Butterworth response.
suggested for correct operation on the basis of Fig. 10(b).
Gain variations exceeding 50 % were observed in simulations
performed over the same VC M I range without common mode
equalization. These variations were ascribed to the phenom-
enon illustrated in Fig. 4(b) and (c). The experimental result
shown in Fig. 11 indirectly demonstrates the effectiveness of
the input common mode equalization circuit. The relatively
wide bandwidth of the input common mode equalization
loop (325 kHz, from simulations) guarantees that the low
sensitivity of the gain to the input common mode voltage is
maintained even when the latter include large AC components.
Experimental verification of this property was obtained by
superimposing a large sinusoidal waveform (up to 600 mV
peak-to-peak) on VC M I while the input differential signal was
set to 10 Hz, with magnitudes as large as to produce a peak-
to-peak output voltage in the range 0.4-1 V.
In these conditions, varying the frequency of the common
mode signal up to 50 kHz did not produce significant differ- Fig. 13. (a) Magnitude of the input impedance: measurements (points) and
ential mode gain variations, while intermodulation products theoretical behavior (lines) for two chopper frequencies. The input resistances
that would have been obtained with a conventional chopper amplifier and same
were more than 60 dB below the output main tone (10 Hz). input capacitances are indicated by the horizontal lines. (b) Maximum relative
The amplifier magnitude and phase frequency response is difference (absolute value) measured over input resistance measurements
shown in Fig. 12(a) and 12(b), respectively. The response was performed over a set of three different samples.
acquired for different values of the source resistance (R S ),
simply implemented by means of two identical resistors of A deeper insight into this point was obtained by estimating the
value R S /2 placed in series to each input terminal (balanced input impedance from the data of Fig. 12, using the following
configuration). The curve for R S = 0 shows a slight difference equation for the input attenuation:
with respect to the ideal Butterworth behavior. This discrep- (v out /v s ) R S =R Z in
ancy, which is clearly not critical for the performance of the = (27)
(v out /v s ) R S =0 R + Z in
amplifier, is probably due to the relatively high sensitivity of
ω0 and Q to process variations. where R S is the generic output resistance of the signal source
A more interesting feature is the dependence of the magni- and R = 20 M. The results (experimental points) are
tude on the source resistance. At low frequencies, all curves presented in Fig. 13(a) for two chopper frequencies. The
converge to the same asymptotic value. Considering that the magnitude of Z in is compared with theoretical prediction
preamplifier input capacitance is 14 pF (estimated by AC calculated using (4) with C p = 14 pF and a fit of the
simulations), with a conventional chopper ICF architecture the experimental response for H ( f ).
effective input resistance for f ch = 20 kHz would have been A good agreement between experimental and theoretical
around 890 k. This value, combined with R S = 20 M, data can be observed. As the frequency increases, the Z I N
would result in an attenuation of nearly 27 dB, which is appar- boosting factor progressively decreases, producing the excess
ently not visible in Fig. 12. This demonstrates the beneficial attenuation and phase delay visible in Fig. 12 for Rs= 0.
effect of the proposed SCPS approach on the input resistance. Fig. 13(b) shows the maximum relative difference between

Authorized licensed use limited to: Indian Institute of Technology Hyderabad. Downloaded on December 31,2022 at 09:08:27 UTC from IEEE Xplore. Restrictions apply.
762 IEEE TRANSACTIONS ON CIRCUITS AND SYSTEMS–I: REGULAR PAPERS, VOL. 64, NO. 4, APRIL 2017

TABLE III
A MPLIFIER S PECIFICATIONS C OMPARED W ITH S TATE OF A RT

modulators [29]. The average DC value of the bias currents


was 300 pA, with an offset component around 50 pA. The esti-
mated bias current noise density was 60 fA/sqrt(Hz). Current
noise contribution is negligible for the case R S = 1 M.
Periodical AC noise simulations revealed that the visible
residual flicker noise contribution originates from devices
M21-24 of the Gmb transconductor, which, as already stated,
are not affected by chopper modulation. The slope change
occurring at high frequencies can be understood considering
that, for f > f c , H B P decreases with half the slope of H L P ,
making INT2 contribution dominant in the upper frequency
interval. This explanation is supported by simulations pre-
sented in [15].
The Total Harmonic Distortion (THD) is shown in Fig. 15
Fig. 14. Output noise voltage spectral density measured with three different as a function of the input amplitude. At frequencies much
input source resistances, indicated for each curve. Dashed lines represent the lower than f c , distortion keeps below 1 % even for input
passband noise contribution from only the source resistance, for the cases levels exceeding the 5 mV limit (10 mVp−p ), set by the VC M O
RS = 1 M and 40 M. The background noise of the noise measurement
set-up was less than 30 nV/sqrt(Hz) over the whole displayed frequency range. control mechanism. Distortion get worse as the frequency
increases, as shown by the 90 Hz curve, where the 1 % THD
level is crossed with just a 5 mVp−p input stimulus. The dis-
Z in measurements performed on three different samples. Dif- tortion increase with frequency occurs when, due to the H L P
ferences are smaller than 10 % across nearly the whole ampli- response, the feedback signal does not effectively compensate
fier passband, demonstrating the robustness of the resistance the input signal and the preamplifier output voltage increases,
boosting approach. exceeding its maximum swing. Mitigation of this effect could
The output noise spectral density is shown in Fig. 14 for be obtained by reducing the pre-amplifier gain at the cost of
three different source resistances. Considering the case for a less area-efficient implementation of INT1 [15].
RS = 0, the flat region in the amplifier passband corresponds Table III shows a list of specifications of the proposed
to an input referred noise voltage of 18 nV/sqrt(Hz), while amplifier, compared with five recent works. The total output
the flicker corner frequency is around 0.2 Hz. The spectral ripple was estimated summing up the contribution of the first
density measured for R S = 40 M exhibits a considerable three harmonics of f ch . The input referred ripple reported
excess noise that has been ascribed to the contribution of in the table is obtained by dividing the output ripple mag-
the input bias currents due to charge injection from the input nitude by the passband gain A0 . The input offset voltage and

Authorized licensed use limited to: Indian Institute of Technology Hyderabad. Downloaded on December 31,2022 at 09:08:27 UTC from IEEE Xplore. Restrictions apply.
BUTTI et al.: CHOPPER INSTRUMENTATION AMPLIFIER WITH INPUT RESISTANCE BOOSTING 763

cut-off frequency of the amplifier acts as an anti-alias filter


for the typically wide noise bandwidth of the sensors, allowing
direct connection to a low sampling-rate ADC. The amplifier is
also sufficiently compact to be integrated into relatively small
systems on a chip.
ACKNOWLEDGMENT
The authors wish to thank STMicroelectronics for fabricat-
ing the test chip described in this work.
R EFERENCES
Fig. 15. Total Harmonic Distortion (THD) as a function of input signal
peak-to-peak amplitude for three different frequencies. [1] R. Wu, J. H. Huijsing, and K. A. A. Makinwa, Precision Instrumentation
Amplifiers and Read-Out Integrated Circuits. New York, NY, USA:
Springer, 2013.
gain standard deviations have been estimated over a set of [2] G. T. Ong and P. K. Chan, “A power-aware chopper-stabilized instrumen-
5 samples. Measurements performed varying the temperature tation amplifier for resistive wheatstone bridge sensors,” IEEE Trans.
Instrum. Meas., vol. 63, no. 9, pp. 2253–2263, Sep. 2014.
over a 20 ◦ C–100 ◦ C interval with a Peltier Cell cryostat [3] R. Wu, Y. Chae, J. H. Huijsing, and K. A. A. Makinwa, “A 20-b
showed that the total average offset drift was 5 nV/◦ C, while ±40-mV range read-out IC with 50-nV offset and 0.04% gain error
the DC gain drift was 12.5 ppm / ◦ C. for bridge transducers,” IEEE J. Solid-State Circuits, vol. 47, no. 9,
pp. 2152–2163, Sep. 2012.
Among the amplifiers included in Table III, the proposed [4] R. F. Yazicioglu, P. Merken, R. Puers, and C. Van Hoof, “A 60 μW

architecture exhibits the highest input resistance. Input resis- 60 nV/ Hz readout front-end for portable biopotential acquisition
tance data are not provided in [25] and [26], but both use systems,” IEEE J. Solid-State Circuits, vol. 42, no. 5, pp. 1100–1110,
May 2007.
a standard configuration for the chopper modulators, so that [5] R. F. Yazicioglu, S. Kim, T. Torfs, H. Kim, and C. V. Hoof,
an input resistance not exceeding a few tens of M can “A 30 μW analog signal processor ASIC for portable biopotential signal
be expected. The relatively high input resistance reported monitoring,” IEEE J. Solid-State Circuits, vol. 46, no. 1, pp. 209–223,
Jan. 2011.
in [30] is obtained by pre-charging the input capacitance with [6] T. Denison, K. Consoer, W. Santa, A. Avestruz, J. Cooley, and A. Kelly,

a buffered version of the input signal before each chopper “2 μW 100 nV/ Hz chopper-stabilized instrumentation amplifier for
transition. This very simple approach may introduce offset chronic measurement of neural field potentials,” IEEE J. Solid-State
Circuits, vol. 42, no. 12, pp. 2934–2945, Dec. 2007.
and 1/f noise contributions from the buffers, even with pre- [7] N. Verma, A. Shoeb, J. Bohorquez, J. Dawson, J. Guttag, and
charging times much smaller than the chopper period. This is A. P. Chandrakasan, “A micro-power EEG acquisition SoC with inte-
proven by the relatively high flicker corner reported in [30], grated feature extraction processor for a chronic seizure detection
system,” IEEE J. Solid-State Circuits, vol. 45, no. 4, pp. 804–816,
which can become a critical point when trying to reduce the Apr. 2010.
input noise density to levels similar to the other amplifier in [8] C. C. Enz and G. C. Temes, “Circuit techniques for reducing the effects
the table. The input noise vs. supply current trade-off obtained of op-amp imperfections: Autozeroing, correlated double sampling, and
chopper stabilization,” Proc. IEEE, vol. 84, no. 11, pp. 1584–1614,
with the proposed circuit and expressed through the noise Nov. 1996.
efficiency factor (NEF) is comparable to architecture of similar [9] C. Menolfi and Q. Huang, “A low-noise CMOS instrumentation ampli-
complexity [14], [21]. Better NEF and area occupation figures fier for thermoelectric infrared detectors,” IEEE J. Solid-State Circuits,
vol. 32, no. 7, pp. 968–976, Jul. 1997.
are obtained by means of a simpler architecture [26], which, [10] B. J. van den Dool and J. H. Huijsing, “Indirect current feed-
on the other hand, shows less effective ripple rejection and back instrumentation amplifier with a common-mode input range that
lacks both DEM and input resistance boosting functions. includes the negative roll,” IEEE J. Solid-State Circuits, vol. 28, no. 7,
pp. 743–749, Jul. 1993.
[11] K. A. Ng and P. K. Chan, “A CMOS analog front-end IC for portable
V. C ONCLUSIONS EEG/ECG monitoring applications,” IEEE Trans. Circuits Syst. I, Reg.
Papers, vol. 52, no. 11, pp. 2335–2347, Nov. 2005.
We have shown that alternating the input and feedback [12] J. F. Witte, J. H. Huijsing, and K. A. A. Makinwa, “A current-
port of an ICF amplifier at the same frequency as chopper feedback instrumentation amplifier with 5 μV offset for bidirectional
modulation produces a dramatic increase of the switched- high-side current-sensing,” IEEE J. Solid-State Circuits, vol. 43, no. 12,
pp. 2769–2775, Dec. 2008.
capacitor parasitic input resistance associated to the input [13] Q. Fan, J. H. Huijsing, and K. A. A. Makinwa, “Input characteristics
chopper modulator. The proposed approach has been applied of a chopped multi-path current feedback instrumentation amplifier,” in
to a modified ICF chopper in-amp capable of self-filtering Proc. 4th Int. Workshop Adv. Sensors Interfaces (IWASI), Brindisi, Italy,
Jun. 2011, pp. 61–66.
all ripple contributions exploiting its second order low-pass [14] R. Wu, J. H. Huijsing, and K. A. A. Makinwa, “A current-feedback
transfer function. The amplifier embodies a feedback loop instrumentation amplifier with a gain error reduction loop and 0.06%
that equalizes the input and feedback common mode voltages, untrimmed gain error,” IEEE J. Solid-State Circuits, vol. 46, no. 12,
pp. 2794–2806, Dec. 2011.
which is a prerequisite for the application of the proposed [15] F. Butti, P. Bruschi, M. Dei, and M. Piotto, “A compact instrumentation
resistance boosting approach. Measurements performed on amplifier for MEMS thermal sensor interfacing,” Analog Integr. Circuits
the prototype showed that the input impedance is increased Process., vol. 72, pp. 585–594, Sep. 2012.
[16] F. Butti, P. Bruschi, and M. Piotto, “A chopper modulated low noise
by more than two orders of magnitude for frequencies two instrumentation amplifier for MEMS thermal sensors interfacing,” in
decades below the amplifier cut-off frequency. This makes Proc. PRIME, Trento, Italy, Jul. 2011, pp. 133–136.
the amplifier particularly interesting for interfacing sensors [17] F. Del Cesta, A. N. Longhitano, P. Bruschi, and M. Piotto, “A novel
architecture for current-feedback instrumentation amplifiers with rail-
having small bandwidths, which is a frequent condition with to-rail input range,” in Proc. PRIME, Grenoble, France, Jun./Jul. 2014,
a large variety of MEMS sensors. In these cases, the low pp. 1–4.

Authorized licensed use limited to: Indian Institute of Technology Hyderabad. Downloaded on December 31,2022 at 09:08:27 UTC from IEEE Xplore. Restrictions apply.
764 IEEE TRANSACTIONS ON CIRCUITS AND SYSTEMS–I: REGULAR PAPERS, VOL. 64, NO. 4, APRIL 2017

[18] M. Dei, P. Bruschi, and M. Piotto, “Design of CMOS chopper amplifiers Federico Butti received the Laurea degree in elec-
for thermal sensor interfacing,” in Proc. PRIME, Jun. 2008, pp. 205–208. tronic engineering from the University of Pisa,
[19] J. H. Nielsen and E. Bruun, “A CMOS low-noise instrumentation Italy, in July 2009 and the Ph.D. degree in elec-
amplifier using chopper modulation,” Analog Integr. Circuits Process., tronic, computer and telecommunication engineering
vol. 42, no. 1, pp. 65–76, 2005. in June 2013. His main area of interest is the
[20] R. Wu, K. A. A. Makinwa, and J. H. Huijsing, “A chopper current- design of analog integrated circuits. In 2013 he
feedback instrumentation amplifier with a 1 mHz 1/f noise corner and joined Marvell Semiconductor, Pavia, Italy, where
an AC-coupled ripple reduction loop,” IEEE J. Solid-State Circuits, he has been involved in the design of multi-GHz
vol. 44, no. 12, pp. 3232–3243, Dec. 2009. √ PLLs and circuits for SerDes applications. Since
[21] Q. Fan, J. H. Huijsing, and K. A. A. Makinwa, “A 21 nV/ Hz chopper- December 2016 he has been with Dialog Semicon-
stabilized multi-path current-feedback instrumentation amplifier with ductor, Livorno, Italy.
2 μV offset,” IEEE J. Solid-State Circuits, vol. 47, no. 2, pp. 464–475,
Feb. 2012.
[22] Y. Kusuda, “Auto correction feedback for ripple suppression in a chopper
amplifier,” IEEE J. Solid-State Circuits, vol. 45, no. 8, pp. 1436–1445,
Aug. 2010. Massimo Piotto received his Laurea degree in
[23] A. Bilotti and G. Monreal, “Chopper-stabilized amplifiers with a track- electronic engineering from the University of
and-hold signal demodulator,” IEEE Trans. Circuits Syst. I, Fundam. Pisa, Italy, in 1996 and his Ph.D. degree in
Theory Appl., vol. 46, no. 4, pp. 490–495, Apr. 1999. electronic, computer and telecommunication engi-
[24] M. Belloni, E. Bonizzoni, A. Fornasari, and F. Maloberti, “A micropower neering in 2000. Since December 2001 he
chopper—CDS operational amplifier,” IEEE J. Solid-State Circuits, has been a researcher of the section of Pisa
vol. 45, no. 12, pp. 2521–2529, Dec. 2010. of the “Istituto di Elettronica e di Ingegne-
[25] M. A. P. Pertijs and W. J. Kindt, “A 140 dB-CMRR current-feedback ria dell’Informazione e delle Telecomunicazioni”—
instrumentation amplifier employing ping-pong auto-zeroing and chop- National Research Council (IEIIT-CNR). His main
ping,” IEEE J. Solid-State Circuits, vol. 45, no. 10, pp. 2044–2056, research interests concern MEMS-based smart sen-
Oct. 2010. sors, microelectronic, and nanoelectronic devices
[26] I. Akita and M. Ishida, “A chopper-stabilized instrumentation amplifier and technologies
using area-efficient self-trimming technique,” Analog Integr. Circuits
Process., vol. 81, pp. 571–582, Aug. 2014.
[27] G. Efthivoulidis, L. Toth, and Y. P. Tsividis, “Noise in Gm-C filters,”
IEEE Trans. Circuits Syst. II, Analog Digit. Signal Process., vol. 45,
no. 3, pp. 295–302, Mar. 1998. Paolo Bruschi received the Laurea degree in elec-
[28] R. Gregorian and G. C. Temes, Analog MOS Integrated Circuits for tronic engineering from the University of Pisa, Italy,
Signal Processing. New York, NY, USA: Wiley, 1986, p. 214. in 1989, and, in 1992, the Perfezionamento degree,
[29] J. Xu, Q. Fan, J. H. Huijsing, C. Van Hoof, R. F. Yazicioglu, and equivalent to the Ph.D, from the S. Anna University
K. A. A. Makinwa, “Measurement and analysis of current noise in School of Pisa. In 1993 he joined the Dipartimento
chopper amplifiers,” IEEE J. Solid-State Circuits, vol. 48, no. 7, di Ingegneria dell’Informazione of the University of
pp. 1575–1584, Jul. 2013. Pisa, where he is currently an Associate Professor.
[30] H. Chandrakumar and D. Marković, “A 2 μW 40 mVpp linear- His main area of interest is the design of mixed
input-range chopper- stabilized bio-signal amplifier with boosted input signal integrated circuit and MEMS sensor.
impedance of 300 M and electrode-offset filtering,” in Proc. ISSCC,
San Francisco, CA, USA, Feb. 2016, pp. 96–97.

Authorized licensed use limited to: Indian Institute of Technology Hyderabad. Downloaded on December 31,2022 at 09:08:27 UTC from IEEE Xplore. Restrictions apply.

You might also like