Buffer Amplifiers, Gain Blocks and Instrumentation Amplifiers

OTHER LINEAR CIRCUITS
CHAPTER 2: OTHER LINEAR CIRCUITS
SECTION 2.1: BUFFER AMPLIFIERS
SECTION 2.2: GAIN BLOCKS
SECTION 2.3: INSTRUMENTATION AMPLIFIERS
2.1
2.5
2.7
IN-AMP DEFINITIONS
2.7
OP AMP/IN-AMP FUNCTIONAL DIFFERENCES
2.8
SUBTRACTOR OR DIFFERENCE AMPLIFIER
THE THREE OP AMP INSTRUMENTATION AMPLIFIER
TOPOLOGY
2.8
2.12
PRECISION SINGLE-SUPPLY COMPOSITE IN-AMP
2.15
THE TWO OP AMP INSTRUMENTATION AMPLIFIER TOPOLOGY 2.18
IN-AMP DC ERROR SOURCES
2.22
IN-AMP NOISE SOURCES
2.26
IN-AMP BRIDGE AMPLIFIER ERROR BUDGET ANALYSIS
2.28
IN-AMP INPUT OVERVOLTAGE PROTECTION
2.29
SECTION 2.4: DIFFERENTIAL AMPLIFIERS
SECTION 2.5: ISOLATION AMPLIFIERS
2.31
2.33
ANALOG ISOLATION TECHNIQUES
2.33
AD210 3-PORT ISOLATOR
2.34
MOTOR CONTROL ISOLATION AMPLIFIER
2.35
OPTIONAL NOISE REDUCTION POST FILTER
2.36
TWO-PORT ISOLATOR
2.36
SECTION 2.6: DIGITAL ISOLATION TECHNIQUES
2.39
AD260/AD261 HIGH SPEED LOGIC ISOLATORS
2.40
iCOUPLER TECHNOLOGY
2.42
ADuM1100 ARCHITECTURE: A SINGLE-CHANNEL DIGITAL
ISOLATOR
ADuM130X/ADuM140X: MULTICHANNEL PRODUCTS
SECTION 2.7: ACTIVE FEEDBACK AMPLIFIERS
SECTION 2.8: LOGARITHMIC AMPLIFIERS
SECTION 2.9: HIGH SPEED CLAMPING AMPLIFIERS
SECTION 2.10: COMPARATORS
USING OP AMPS AS COMPARATORS
2.42
2.46
2.49
2.53
2.59
2.65
2.65
SPEED
2.71
OUTPUT CONSIDERATIONS
2.71
BASIC LINEAR DESIGN
SECTION 2.10: COMPARATORS (cont.)
INPUT CIRCUITRY
SECTION 2.11:
SECTION 2.12:
SECTION 2.13:
SECTION 2.14:
ANALOG MULTIPLIERS
RMS TO DC CONVERTERS
PROGRAMMABLE GAIN AMPLIFIERS
AUDIO AMPLIFIERS
2.75
2.77
2.83
2.87
2.95
AMPLIFIERS
2.95
VCAs (VOLTAGE CONTROLLED AMPLIFIERS)
2.98
LINE DRIVERS AND RECEIVERS
2.101
AUDIO LINE RECEIVERS
2.101
AUDIO LINE DRIVERS
2.103
CLASS-D POWER AMPLIFIERS
2.105
SECTION 2.15: AUTO-ZERO AMPLIFIERS
2.119
CHOPPER AMPLIFIERS
2.119
AUTO-ZERO AMPLIFERS IMPROVE ON CHOPPERS
2.121
IMPLEMENTATION
2.123
OPERATION DESCRIPTION
2.124
REFERENCES
2.126
OTHER LINEAR CIRCUITS
BUFFER AMPLIFIERS
CHAPTER 2: OTHER LINEAR CIRCUITS
SECTION 2.1: BUFFER AMPLIFIERS
In the early days of high speed circuits, simple emitter followers were often used as high
speed buffers. The term buffer was generally accepted to mean a unity-gain, open-loop
amplifier. With the availability of matching PNP transistors, a simple emitter follower
can be improved, as shown below in Figure 2.1A. This complementary circuit offers firstorder cancellation of dc offset voltage, and can achieve bandwidths greater than
100 MHz. Typical offset voltages without trimming are usually less than 50 mV, even
with unmatched discrete transistors.
+VS
+VS
VIN
R2
VIN
–VS
VOUT
VOUT
+VS
–VS
(A) HOS-100
R1
–VS
(B) LH0033
Figure 2.1: Early Open-Loop Hybrid Buffer Amplifiers:
(A) HOS-100 Bipolar, (B) LH0033 FET Input
If high input impedance is required, a dual FET can be used as an input stage ahead of a
complementary emitter follower, as shown in Figure 2.1B. This form of the buffer circuit
was implemented by both National Semiconductor Corporation as the LH0033, and by
Analog Devices as the ADLH0033.
Circuits such as these achieved bandwidths of about 100 MHz at fairly respectable levels
of harmonic distortion, typically better than –60 dBc. However, they suffered from dc
and ac nonlinearities when driving loads less than 500 Ω.
2.1
BASIC LINEAR DESIGN
One of the first totally monolithic implementations of these functions was the Precision
Monolithics, Inc. BUF03 shown below in Figure 2.2 (see Reference 1). PMI is now a
division of Analog Devices. This open-loop IC buffer achieved a bandwidth of about
50 MHz for a 2 V peak-to-peak signal.
The BUF03 circuit is interesting because it demonstrates techniques that eliminated the
requirement for the slow, bandwidth-limited vertical PNP transistors associated with
most IC processes available at the time of the design (approximately 1979).
+VS
1.7mA
10.4mA
Q7
J2
Q9
Q1
VIN
Q6
VOUT
J5
230µA
J1
Q5
J6
3.4mA
Q8
1.7mA
230µA
Q11
Q10
–VS
Figure 2.2: BUF03 Monolithic Open-Loop Buffer—1979 Vintage
One of the problems with all the open-loop buffers discussed thus far is that although
high bandwidths can be achieved, the devices discussed don’t take advantage of negative
feedback. Distortion and dc performance suffer considerably when open-loop buffers are
loaded with typical video impedance levels of 50 Ω, 75 Ω , or 100 Ω. The solution is to
use a properly compensated wide bandwidth op amp in a unity-gain follower
configuration. In the early days of monolithic op amps, process limitations prevented this,
so the open-loop approach provided a popular interim solution.
Practically all unity-gain-stable voltage or current feedback op amps can be used in a
simple follower configuration. Usually, however, the general-purpose op amps are
compensated to operate over a wide range of gains and feedback conditions. Therefore,
bandwidth suffers somewhat at low gains, especially in the unity-gain noninverting
mode, and additional external compensation is usually required.
2.2
OTHER LINEAR CIRCUITS
BUFFER AMPLIFIERS
+
–
Figure 2.3: Simple Unity-Gain Monolithic Buffers
A practical solution is to compensate the op amp for the desired closed-loop gain, while
including the gain setting resistors on-chip, as shown in Figure 2.4. Note that this form of
op amp, internally configured as a buffer, may typically have no feedback pin. Also,
putting the resistors and compensation on-chip also serves to reduce parasitics.
There are a number of op amps optimized in this manner. Roy Gosser’s AD9620 (see
Reference 2) was probably the earliest monolithic implementation. The AD9620 was a
1990 product release, and achieved a bandwidth of 600 MHz using ±5 V supplies. It was
optimized for unity gain, and used the voltage feedback architecture. A newer design
based on similar techniques is the AD9630, which achieves a 750 MHz bandwidth.
+
–
RF
CF
Figure 2.4: Frequency Compensated Buffer
The BUF04 unity gain buffer (see Reference 3) was released in 1994 and achieves a
bandwidth of 120 MHz. This device was optimized for large signals and operates on
supplies from ±5 V to ±15 V. Because of the wide supply range, the BUF04 is useful not
only as a standalone unity-gain buffer, but also within a feedback loop with a standard op
amp, to boost output.
Although the common definition of a buffer is unity gain device, sometimes the term is
used for a circuit with a gain of 2. Closed-loop buffers with a gain of 2 find wide
applications as transmission line drivers, as shown below in Figure 2.5. The internally
configured fixed gain of the amplifier compensates for the loss incurred by the source and
load termination. Impedances of 50 Ω, 75 Ω, and 100 Ω are popular cable impedances.
The AD8074/AD8075 500 MHz triple buffers are optimized for gains of 1 and 2,
2.3
BASIC LINEAR DESIGN
respectively. The dual AD8079A/AD8079B 260 MHz buffer is optimized for gains of 2
and 2.2, respectively.
+
RS
ZO
–
RF
CF
RG
AD8074
Triple
Voltage Feedback
G=1
BW = 500MHz
VS = ±5V
RL
RS = RL = ZO
AD8075
Triple
Voltage Feedback
G=2
BW = 500MHz
VS = ±5V
AD8079A/B
Dual
Voltage Feedback
G = 2 / 2.2
BW = 260MHz
VS = ±5V
Figure 2.5: Fixed-Gain Video Transmission Line Drivers
In implementing a high speed unity-gain buffer with a voltage feedback op amp, there
will typically be no resistor required in the feedback loop, which considerably simplifies
the circuit. Note that this isn’t a 100% hard-and-fast rule, however, so always check the
device data sheet to be sure. A unity-gain buffer with a current feedback op amp will
always require a feedback resistor, typically in the range of 500 Ω to 1000 Ω. So, be sure
to use a value appropriate to not only the basic part, but also the specific power supplies
in use.
2.4
OTHER LINEAR CIRCUITS
GAIN BLOCKS
SECTION 2.2: GAIN BLOCKS
While the op amp allows gain to be set with external resistors, there are a group of
circuits that are designed to operate at a fixed gain. These parts are typically RF
components. They also are typically designed to be operated in a 50 Ω environment, with
the inputs and outputs matched internally. Often the gain blocks are available in several
gain settings.
For example, the AD8354 RF gain block is a fixed-gain amplifier with single-ended input
and output ports whose impedances are nominally equal to 50 Ω over the frequency range
100 MHz to 2.7 GHz. Consequently, it can be directly inserted into a 50 Ω system with
no impedance matching circuitry required. The input and output impedances are
sufficiently stable versus variations in temperature and supply voltage that no impedance
matching compensation is required.
Figure 2.6: AD8352 20 dB RF Gain Block
Differential input and output gain blocks are also available. An example of a differential
input, single-ended output device is the AD8129. See Figure 2.7
AD8129/
AD8130
8
–IN
7
+VS
PD 3
6
OUT
REF 4
5
FB
+IN
1
–VS 2
+
Figure 2.7: AD8129/AD8130 Differential Input, Single-Ended Output Gain Block
2.5
BASIC LINEAR DESIGN
Fully differential input and output devices are also available, such as the AD8350. See
Figure 2.8.
IN+
1
ENBL
2
+
–
VCC 3
OUT+
4
8
IN–
7
GND
6
GND
5
OUT–
AD8350
Figure 2.8: AD8350 Differential In/Differential Out Gain Block
2.6
OTHER LINEAR CIRCUITS
INSTRUMENTATION AMPLIFIERS
SECTION 2.3: INSTRUMENTATION AMPS
The instrumentation amp is primarily used to amplify small differential voltages in the
presence of (typically) larger common-mode voltages.
In-Amp Definitions
An in-amp is a precision closed-loop gain block. It has a pair of differential input
terminals, and a single-ended output that works with respect to a reference or common
terminal, as shown in Figure 2.9. The input impedances are balanced and high in value,
9
typically ≥10 Ω. Again, unlike an op amp, an in-amp uses an internal feedback resistor
network, plus one (usually) gain set resistance, RG. Also unlike an op amp is the fact that
the internal resistance network and RG are isolated from the signal input terminals. In
amp gain can also be preset via an internal RG by pin selection, (again isolated from the
signal inputs). Typical in amp gains range from 1 to 1000.
The in-amp develops an output voltage which is referenced to a pin usually designated
REFERENCE, or VREF. In many applications, this pin is connected to circuit ground, but
it can be connected to other voltages, as long as they lie within the rated compliance
range of the in-amp. This feature is especially useful in single-supply applications, where
the output voltage is usually referenced to mid-supply (i.e., +2.5 V in the case of a +5 V
supply).
RS/2
COMMON
MODE
VOLTAGE
VCM
ΔRS
~
_
VSIG
2
+
IN-AMP
GAIN = G
+
~
RG
+
~
_
VSIG
_
2
RS/2
VOUT
VREF
~
COMMON MODE ERROR (RTI) =
VCM
CMRR
Figure 2.9: The Generic Instrumentation Amplifier (In-Amp)
2.7
BASIC LINEAR DESIGN
In order to be effective, an in-amp needs to be able to amplify microvolt-level signals,
while simultaneously rejecting volts of common-mode (CM) signal at its inputs. This
requires that in amps have very high common-mode rejection (CMR). Typical values of
in-amp CMR are from 70 dB to over 100 dB (at dc), with CMR usually improving at
higher gains.
It is important to note that a CMR specification for dc inputs alone isn’t sufficient in most
practical applications. In industrial applications, the most common cause of external
interference is 50 Hz/60 Hz ac power-related noise (including harmonics). In differential
measurements, this type of interference tends to be induced equally onto both in-amp
inputs, so the interference appears as a CM input signal. Therefore, specifying CMR over
frequency is just as important as specifying its dc value. Note that imbalance in the two
source impedances will degrade the CMR of some in amps. Analog Devices fully
specifies in-amp CMR at 50 Hz/60 Hz, with a source impedance imbalance of 1 kΩ.
Op Amp/In-Amp Functionality Differences
An op amp is a general-purpose gain block— user-configurable in myriad ways using
external feedback components of R, C, and (sometimes) L. The final configuration and
circuit function using an op amp is truly whatever you make of it.
In contrast to this, an instrumentation amp (in-amp) is a more constrained device in terms
of functioning, and also the allowable range(s) of operating gain. People also often
confuse in-amps as to their function, calling them “op amps.” But the converse is seldom
(if ever) true. It should be understood that an in-amp is not just a special type op amp; the
function of the two devices is actually fundamentally different.
Perhaps a good way to differentiate the two devices is to remember that an op amp can be
programmed to do almost anything, by virtue of its feedback flexibility. In contrast to
this, an in-amp cannot be programmed to do just anything. It can only be programmed for
gain, and then over a specific range. An op amp is configured via a number of external
components, while an in-amp is configured by either one resistor, or by pin-selectable
taps for its working gain.
Subtractor or Difference Amplifiers
A simple subtractor or difference amplifier can be constructed with four resistors and an
op amp, as shown in Figure 2.10. It should be noted that this is not a true in-amp, but it is
often used in applications where a simple differential to single-ended conversion is
required. Because of its popularity, this circuit will be examined in more detail, in order
to understand its fundamental limitations before discussing true in amp architectures.
2.8
OTHER LINEAR CIRCUITS
INSTRUMENTATION AMPLIFIERS
R1
R2
V1
_
VOUT
1+
CMR = 20 log10
+
R1'
R2'
V2
VOUT = (V2 – V1)
R2
R1
Kr
REF
Where Kr = Total Fractional
Mismatch of R1/ R2 TO
R1'/R2'
R2
R1
R2
R2'
=
CRITICAL FOR HIGH CMR
R1
R1'
EXTREMELY SENSITIVE TO SOURCE IMPEDANCE IMBALANCE
0.1% TOTAL MISMATCH YIELDS
≈
66dB CMR FOR R1 = R2
Figure 2.10: Op Amp Subtractor or Difference Amplifier
There are several fundamental problems with this simple circuit. First, the input
impedance seen by V1 and V2 isn’t balanced. The input impedance seen by V1 is R1, but
the input impedance seen by V2 is R1' + R2'. The configuration can also be quite
problematic in terms of CMR, since even a small source impedance imbalance will
degrade the workable CMR. This problem can be solved with well-matched open-loop
buffers in series with each input (for example, using a precision dual op amp). But, this
adds complexity to a simple circuit, and may introduce offset drift and nonlinearity.
The second problem with this circuit is that the CMR is primarily determined by the
resistor ratio matching, not the op amp. The resistor ratios R1/R2 and R1'/R2' must
match extremely well to reject common-mode noise— at least as well as a typical op amp
CMR of ≥100dB. Note also that the absolute resistor values are relatively unimportant.
Picking four 1% resistors from a single batch may yield a net ratio matching of 0.1%,
which will achieve a CMR of 66 dB (assuming R1 = R2). But if one resistor differs from
the rest by 1%, the CMR will drop to only 46 dB. Clearly, very limited performance is
possible using ordinary discrete resistors in this circuit (without resorting to hand
matching). This is because the best standard off-the-shelf RNC/RNR style resistor
tolerances are on the order of 0.1% (see Reference 1).
In general, the worst-case CMR for a circuit of this type is given by the following
equation (see References 2 and 3):
⎡1 + R 2 / R1⎤
CMR (dB) = 20 log ⎢
,
⎣ 4 Kr ⎥⎦
Eq. 2-1
2.9
BASIC LINEAR DESIGN
where Kr is the individual resistor tolerance in fractional form, for the case where four
discrete resistors are used. This equation shows that the worst-case CMR for a tolerance
build-up for four unselected same-nominal-value 1% resistors to be no better than 34 dB.
A single resistor network with a net matching tolerance of Kr would probably be used for
this circuit, in which case the expression would be as noted in the figure, or:
⎡1 + R 2 / R1⎤
CMR (dB) = 20 log ⎢
⎥⎦
Kr
⎣
Eq. 2-2
A net matching tolerance of 0.1% in the resistor ratios therefore yields a worst-case dc
CMR of 66 dB using Equation 2-2, and assuming R1 = R2. Note that either case assumes
a significantly higher amplifier CMR (i.e., > 100 dB). Clearly for high CMR, such
circuits need four single-substrate resistors, with very high absolute and TC matching.
Such networks using thick/thin-film technology are available from companies such as
Caddock and Vishay, in ratio matches of 0.01% or better.
In implementing the simple difference amplifier, rather than incurring the higher costs
and PCB real estate limitations of a precision op amp plus a separate resistor network, it
is usually better to seek out a completely monolithic solution.
VCM = ±270V for VS = ±15V
Figure 2.11: High Common-Mode Current Sensing
Using the AD629 Difference Amplifier
An interesting variation on the simple difference amplifier is found in the AD629
difference amplifier, optimized for high common-mode input voltages. A typical currentsensing application is shown in Figure 2.11. The AD629 is a differential-to-single-ended
amplifier with a gain of unity. It can handle a common-mode voltage of ±270 V with
supply voltages of ±15 V, with a small signal bandwidth of 500 kHz.
2.10
OTHER LINEAR CIRCUITS
INSTRUMENTATION AMPLIFIERS
The high common-mode voltage range is obtained by attenuating the noninverting input
(pin 3) by a factor of 20 times, using the R1–R2 divider network. On the inverting input,
resistor R5 is chosen such that R5||R3 equals resistor R2. The noise gain of the circuit is
equal to 20 [1 + R4/(R3||R5)], thereby providing unity gain for differential input voltages.
Laser wafer trimming of the R1–R5 thin film resistors yields a minimum CMR of 86 dB
@ 500 Hz for the AD629B. Within an application, it is good practice to maintain
balanced source impedances on both inputs, so dummy resistor RCOMP is chosen to equal
to the value of the shunt sensing resistor RSHUNT.
2.11
BASIC LINEAR DESIGN
The Three Op Amp Instrumentation Amplifier Topology
For the highest precision and performance, the three op amp instrumentation amplifier
topology is optimum for bridge and other offset transducer applications where high
accuracy and low nonlinearity are required (Figure 2.12).
+
+
R2'
R3'
A1
VSIG
~
2 _
_
_
R1'
VCM
VOUT
RG
A3
R1
+
~
_
+
VSIG
~
2
_
R3
R2
A2
+
CMR ≤ 20log
GAIN × 100
% MISMATCH
2R1
VOUT = VSIG • R3 1 +
RG
R2
IF R2 = R3, G = 1 +
VREF
+ VREF
2R1
RG
Figure 2.12: The Three Op Amp In-Amp
Resistor RG sets the overall gain of this amplifier. It may be internal, external, or
(software or pin-strap) programmable, depending upon the particular in-amp. In this
configuration, CMR depends upon the ratio matching of R3/R2 to R3'/R2'. Furthermore,
common-mode signals are only amplified by a factor of 1 regardless of gain (no
common-mode voltage will appear across RG, hence, no common-mode current will flow
in it because the input terminals of an op amp will have no significant potential difference
between them).
As a result of the high ratio of differential to CM gain in A1-A2, CMR of this in-amp
theoretically increases in proportion to gain. Large common-mode signals (within the A1A2 op amp headroom limits) may be handled at all gains. Finally, because of the
symmetry of this configuration, common-mode errors in the input amplifiers, if they
track, tend to be canceled out by the subtractor output stage. These features explain the
popularity of this three op amp in-amp configuration—it is capable of delivering the
highest performance.
The classic three op amp configuration has been used in a number of monolithic IC inamps (see References 8 and 9). Besides offering excellent matching between the three
internal op amps, thin film laser trimmed resistors provide excellent ratio matching and
gain accuracy at much lower cost than using discrete precision op amps and resistor
2.12
OTHER LINEAR CIRCUITS
INSTRUMENTATION AMPLIFIERS
networks. The AD620 (see Reference 10) is an excellent example of monolithic IC in
amp technology. A simplified device schematic is shown in Figure 2.13 below.
+VS
49.4kΩ
RG = G – 1
VB
_
+
_
+
A1
A2
10kΩ
10kΩ
_
10kΩ
A3
+
10kΩ
Q1
400Ω
24.7kΩ
24.7kΩ
Q2
VREF
400Ω
RG
–IN
VO
+IN
–VS
Figure 2.13: The AD620 In-Amp Simplified Schematic
The AD620 is a highly popular in-amp and is specified for power supply voltages from
±2.3 V to ±18 V. Input voltage noise is only 9 nV/√Hz @ 1 kHz. Maximum input bias
current is only 1nA, due to the use of superbeta transistors for Q1 - Q2.
Overvoltage protection is provided, in part, by the internal 400 Ω thin-film current-limit
resistors in conjunction with the diodes connected from the emitter-to-base of Q1 and Q2.
The gain G is set with a single external RG resistor, as noted by equation 2-3.
G = (49.4kΩ/RG) + 1
Eq. 2-3
As can be noted from this expression and Fig. 2-13, the AD620 internal resistors are
trimmed so that standard 1% or 0.1% resistors can be used to set gain to popular values.
Single-supply operation of the three op amp in-amp requires an understanding of the
internal node voltages. Figure 2.14 below shows a generalized diagram of the in-amp
operating on a single +5 V supply. The maximum and minimum allowable output
voltages of the individual op amps are designated VOH (maximum high output) and VOL
(minimum low output) respectively.
Note that the gain from the common-mode voltage to the outputs of A1 and A2 is unity.
It can be stated that the sum of the common-mode voltage and the signal voltage at these
2.13
BASIC LINEAR DESIGN
outputs must fall within the amplifier output voltage range. Obviously this configuration
cannot handle input common-mode voltages of either zero volts or +5 V, because of
saturation of A1 and A2. The output reference is positioned halfway between VOH and
VOL to allow for bipolar differential input signals.
VCM +
+
R2'
+
_
R1'
VOH=4.9V
VOL=0.1V
_
VOUT
RG
A3
R1
~
+
VSIG
~
2
_
R2'
A1
VSIG ~
2 _
VCM
GVSIG
2
VOH=4.9V
VOL=0.1V
VOH=4.9V
VOL=0.1V
+
_
VOUT = GVSIG + VREF
R2
R2
A2
VREF = 2.5V
+
2R1
G = 1+
RG
VCM –
GVSIG
2
Figure 2.14: Three Op Amp In-Amp Single +5V Supply Restrictions
While there is a number of good single-supply in amps, such as the AD627, the highest
performance devices are still among those specified for traditional dual-supply operation,
i.e., the just-discussed AD620. For certain applications, even such devices as the AD620,
which has been designed for dual supply operation, can be used with full precision on a
single-supply power system.
2.14
OTHER LINEAR CIRCUITS
INSTRUMENTATION AMPLIFIERS
Precision Single-Supply Composite In-Amp
One way to achieve both high precision and single-supply operation takes advantage of
the fact that many popular sensors (e.g. strain gauges) provide an output signal which is
inherently centered around an approximate mid-point of the supply voltage (and/or the
reference voltage). Taking advantage of this basic point allows the inputs of a signal
conditioning in-amp to be biased at “mid-supply.” As a consequence of this step, the
inputs needn’t operate near ground or the positive supply voltage, and the in-amp can still
be used with all its precision.
Under these conditions, an AD620 dual-supply in-amp referenced to the supply midpoint
followed by a rail-to-rail op amp output gain stage provides very high dc precision.
Figure 2.15 illustrates one such high performance in-amp, which operates on a single
+5 V supply.
This circuit uses the AD620 as a low cost precision in-amp for the input stage, along with
an AD822 JFET-input dual rail-to-rail output op amp for the output stage, comprised of
A1 and A2. The output stage operates at a fixed gain of 3, with overall gain set by RG.
+5V
VSIG
_ 2
~
VCM =
+2.5V
~
VSIG
2
+
+
10µF
+
_
0.22µF
P1
5kΩ
47kΩ
R3
24.9kΩ
AD620
_
RG
+
~
10Hz
NOISE
FILTER
0.1µF
R1
REF
_
75.0kΩ
R2
VOUT
A2
+
10mV TO 4.98V
+
A1, A2 = 1/2 AD822
49.9kΩ
R4
A1
_
VREF
+2.5V
1µF
Figure 2.15: A Precision Single-Supply Composite In-Amp
with Rail-to-Rail Output
In this circuit, R3 and R4 form a voltage divider which splits the supply voltage
nominally in half to +2.5 V, with fine adjustment provided by a trimming potentiometer,
P1. This voltage is applied to the input of A1, an AD822 voltage follower, which buffers
it and provides a low impedance source needed to drive the AD620’s reference pin as
2.15
BASIC LINEAR DESIGN
well as providing the output reference voltage VREF. Note that this feature allows a
bipolar VOUT to be measured with respect to this +2.5 V reference (not to GND). This is
despite the fact that the entire circuit operates from a single (unipolar) supply.
The other half of the AD822 is connected as a gain-of-3 inverter, so that it can output
±2.5 V, “rail-to-rail,” with only ±0.83 V required of the AD620. This output voltage level
of the AD620 is well within the AD620’s capability, thus ensuring high linearity for the
front end.
The general gain expression for this composite in amp is the product of the gain of the
AD620 stage, and the gain of inverting amplifier:
⎛ 49 .4 k Ω
⎞ ⎛ R2 ⎞
+ 1⎟ ⎜
GAIN = ⎜
⎟.
⎝ RG
⎠ ⎝ R1 ⎠
Eq. 2-4
For this example, an overall gain of 10 is realized with RG = 21.5 kΩ (closest standard
value). The table shown in Figure 2.16 summarizes various RG gain values, and the
resulting performance for gains ranging from 10 to 1000.
In this application, the allowable input voltage on either input to the AD620 must lie
between +2 V and +3.5 V in order to maintain linearity. For example, at an overall circuit
gain of 10, the common-mode input voltage range spans 2.25 V to 3.25 V, allowing room
for the ±0.25 V full-scale differential input voltage required to drive the output ±2.5 V
about VREF.
CIRCU IT
GAIN
RG
(Ω)
VOS, RTI
(μV)
TC VOS, RTI
(μV/°C)
10
21.5k
1000
1000
< 50
600
30
5.49k
430
430
< 50
600
100
1.53k
215
215
< 50
300
300
499
150
150
< 50
120
1000
149
150
150
< 50
30
NONLINEARITY BAND WIDTH
(kHz)**
(ppm) *
* Nonlinearity Measured Over Output Range: 0.1V < VOUT < 4.90V
** Without 10Hz Noise Filter
Figure 2.16: Performance Summary of the +5V Single-Supply
AD620/AD822 Composite In-Amp
2.16
OTHER LINEAR CIRCUITS
INSTRUMENTATION AMPLIFIERS
The inverting configuration was chosen for the output buffer to facilitate system output
offset voltage adjustment by summing currents into the A2 stage buffer’s feedback
summing node. These offset currents can be provided by an external DAC, or from a
resistor connected to a reference voltage.
To reduce the effects of unwanted noise pickup, a filter capacitor is recommended across
A2’s feedback resistance to limit the circuit bandwidth to the frequencies of interest. This
capacitor forms a first-order low-pass filter with R2. The corner frequency is 10 Hz as
shown, but this may be easily modified. The capacitor should be a high quality film type,
such as polypropylene.
2.17
BASIC LINEAR DESIGN
The Two Op Amp Instrumentation Amplifier Topology
The circuit shown in Figure 2.17 is referred to as the two op amp in-amp. It is particularly
applicable in single-supply systems. Dual IC op amps are used in most cases for good
matching, such as the OP297 or the OP284. Most often a rail-to-rail op amp is indicated.
The resistors are often a thin film laser trimmed array, possibly on the same chip. The in
amp gain can be easily set with an external resistor, RG. Without RG, the gain is simply
1 + R2/R1. In a practical application, the R2/R1 ratio is chosen for the desired minimum
in-amp gain.
V2
+
V1
A
_
A1
V1
VOUT
A2
+
R1'
_
V2
R1
R2'
C
R2
VREF
RG
R2
2R2
G = 1 + R1 + R
G
R2
2R2
VOUT = ( V2 – V1) 1 + R1 + R
G
+ VREF
R2 = R2'
R1
R1'
CMR ≤ 20log
GAIN × 100
% MISMATCH
Figure 2.17: The Two Op Amp Instrumentation Amplifier
The input impedance of the two op amp in-amp is inherently high, permitting the
impedance of the signal sources to be high and unbalanced. The dc common-mode
rejection is limited by the matching of R1/R2 to R1'/R2'. If there is a mismatch in any of
the four resistors, the dc common-mode rejection is limited to:
⎡ GAIN × 100 ⎤
CMR ≤ 20 log ⎢
⎥.
⎣ %MISMATCH ⎦
Eq. 2-5
Notice that the net CMR of the circuit increases proportionally with the working gain of
the in-amp, an effective aid to high performance at higher gains.
IC in-amps are particularly well-suited to meeting the combined needs of ratio matching
and temperature tracking of the gain-setting resistors. While thin film resistors fabricated
2.18
OTHER LINEAR CIRCUITS
INSTRUMENTATION AMPLIFIERS
on silicon have an initial tolerance of up to ±20%, laser trimming during production
allows the ratio error (not absolute value) between the resistors to be reduced to 0.01%
(100 ppm). Furthermore, the tracking between the temperature coefficients of the thin
film resistors is inherently low and is typically less than 3 ppm/ºC (0.0003%/ºC).
When dual supplies are used, VREF is normally connected directly to ground. In singlesupply applications, VREF is usually connected to a low impedance voltage source equal
to one-half the supply voltage. The gain from VREF to node “A” is R1/R2, and the gain
from node “A” to the output is R2'/R1'. This makes the gain from VREF to the output equal
to unity, assuming perfect ratio matching. Note that it is critical that the source
impedance seen by VREF be low, otherwise CMR will be degraded.
V2
+
V1
VOH=4.9V
VOL=0.1V
+
A1
A
_
R1
10kΩ
VREF
VREF =
2.5V
VOH + VOL
= 2.5V
2
VOH=4.9V
VOL=0.1V
_
R1
R2
10kΩ
10kΩ
10kΩ
R2
VOUT
A2
V1,MIN ≥
1 (G – 1)V + V
OL
REF
G
≥ 1.3V
V1,MAX ≤ 1 (G – 1)VOH + VREF ≤ 3.7V
G
V2 – V1 MAX
≤
VOH – VOL
G
≤ 2.4V
Figure 2.18: Two Op Amp In-Amp Single-Supply Restrictions
for Vs = +5V, G = 2
One major disadvantage of the two op amp in-amp design is that common-mode voltage
input range must be traded off against gain. The amplifier A1 must amplify the signal at
V1 by 1 + R1/R2. If R1 >> R2 (a low gain example in Figure 2.18), A1 will saturate if the
V1 common-mode signal is too high, leaving no A1 headroom to amplify the wanted
differential signal. For high gains (R1<< R2), there is correspondingly more headroom at
node “A,” allowing larger common-mode input voltages.
The ac common-mode rejection of this configuration is generally poor because the signal
path from V1 to VOUT has the additional phase shift of A1. In addition, the two amplifiers
are operating at different closed-loop gains (and thus at different bandwidths). The use of
a small trim capacitor “C” as shown in Fig. 2.17 can improve the ac CMR somewhat.
2.19
BASIC LINEAR DESIGN
A low gain (G = 2) single-supply two op amp in-amp configuration results when RG is
not used, and is shown above in Figure 2.18. The input common-mode and differential
signals must be limited to values which prevent saturation of either A1 or A2. In the
example, the op amps remain linear to within 0.1 V of the supply rails, and their upper
and lower output limits are designated VOH and VOL, respectively. These saturation
voltage limits would be typical for a single-supply, rail-rail output op amp (such as the
AD822, for example).
Using the Fig. 2.18 equations, the voltage at V1 must fall between 1.3 V and 2.4 V to
prevent A1 from saturating. Notice that VREF is connected to the average of VOH and VOL
(2.5 V). This allows for bipolar differential input signals with VOUT referenced to +2.5 V.
A high gain (G = 100) single-supply two op amp in-amp configuration is shown below in
Figure 2.19. Using the same equations, note that voltage at V1 can now swing between
0.124 V and 4.876 V. VREF is again 2.5 V, to allow for bipolar input and output signals.
V2
+
V1
VOH=4.9V
VOL=0.1V
+
A1
A
_
R1
VOUT
A2
VOH=4.9V
VOL=0.1V
_
R1
R2
10kΩ
10kΩ
990kΩ
R2
990kΩ
VREF
VREF =
V1,MIN ≥
2.5V
VOH + VOL
= 2.5V
2
1 (G – 1)V + V
OL
REF
G
≥ 0.124V
V1,MAX ≤ 1 (G – 1)VOH + VREF ≤ 4.876V
G
V2 – V1 MAX
≤
VOH – VOL
G
≤ 0.048V
Figure 2.19: Two Op Amp In-Amp Single-Supply Restrictions
for Vs = +5V, G = 100
All of these discussions show that the conventional two op amp in-amp architecture is
fundamentally limited, when operating from a single power supply. These limitations can
be viewed in one sense as a restraint on the allowable input CM range for a given gain.
Or, alternately, it can be viewed as limitation on the allowable gain range, for a given CM
input voltage.
2.20
OTHER LINEAR CIRCUITS
INSTRUMENTATION AMPLIFIERS
In summary, regardless of gain, the basic structure of the common two op amp in-amp
does not allow for CM input voltages of zero when operated on a single-supply. The only
route to removing these restrictions for single-supply operation is to modify the in-amp
architecture.
2.21
BASIC LINEAR DESIGN
In-Amp DC Error Sources
The dc and noise specifications for in-amps differ slightly from conventional op amps, so
some discussion is required in order to fully understand the error sources.
The gain of an in-amp is usually set by a single resistor. If the resistor is external to the
in-amp, its value is either calculated from a formula or chosen from a table on the data
sheet, depending on the desired gain.
Absolute value laser wafer trimming allows the user to program gain accurately with this
single resistor. The absolute accuracy and temperature coefficient of this resistor directly
affects the in-amp gain accuracy and drift. Since the external resistor will never exactly
match the internal thin film resistor tempcos, a low TC (<25ppm/°C) metal film resistor
should be chosen, preferably with a 0.1% or better accuracy.
Often specified as having a gain range of 1 to 1000, or 1 to 10,000, many in-amps will
work at higher gains, but the manufacturer will not guarantee a specific level of
performance at these high gains. In practice, as the gain-setting resistor becomes smaller,
any errors due to the resistance of the metal runs and bond wires become significant.
These errors, along with an increase in noise and drift, may make higher single-stage
gains impractical. In addition, input offset voltages can become quite sizable when
reflected to output at high gains. For instance, a 0.5 mV input offset voltage becomes 5 V
at the output for a gain of 10,000. For high gains, the best practice is to use an in-amp as
a preamplifier, then use a post amplifier for further amplification.
In a pin-programmable-gain in-amp such as the AD621, the gain-set resistors are
internal, well matched, and the device gain accuracy and gain drift specifications include
their effects. The AD621 is otherwise generally similar to the externally gainprogrammed AD620.
The gain error specification is the maximum deviation from the gain equation.
Monolithic in amps such as the AD624C have very low factory-trimmed gain errors, with
its maximum error of 0.02% at G = 1 and 0.25% at G = 500 being typical for this high
quality in-amp. Notice that the gain error increases with increasing gain. Although
externally connected gain networks allow the user to set the gain exactly, the temperature
coefficients of the external resistors and the temperature differences between individual
resistors within the network all contribute to the overall gain error. If the data is
eventually digitized and presented to a digital processor, it may be possible to correct for
gain errors by measuring a known reference voltage and then multiplying by a constant.
Nonlinearity is defined as the maximum deviation from a straight line on the plot of
output versus input. The straight line is drawn between the end-points of the actual
transfer function. Gain nonlinearity in a high quality in-amp is usually 0.01% (100 ppm)
or less, and is relatively insensitive to gain over the recommended gain range.
2.22
OTHER LINEAR CIRCUITS
INSTRUMENTATION AMPLIFIERS
The total input offset voltage of an in-amp consists of two components (see Figure 2.20).
Input offset voltage, VOSI, is the input offset component that is reflected to the output of
the in-amp by the gain G. Output offset voltage, VOSO, is independent of gain.
At low gains, output offset voltage is dominant, while at high gains input offset
dominates. The output offset voltage drift is normally specified as drift at G = 1 (where
input effects are insignificant), while input offset voltage drift is given by a drift
specification at a high gain (where output offset effects are negligible).
The total output offset error, referred to the input (RTI), is equal to VOSI + VOSO/G. Inamp data sheets may specify VOSI and VOSO separately, or give the total RTI input offset
voltage for different values of gain.
RS/2
ΔRS
VOSI
RG
~
~
~
VSIG
2
VOSO
IB+
VSIG
2
IN-AMP
GAIN = G
~
VOUT
IB–
VCM
VREF
RS/2
IOS =
IB+ – IB–
OFFSET (RTI) =
VOSO
G
+ VOSI + IBΔRS + IOS(RS + ΔRS)
OFFSET (RTO) = VOSO + G VOSI + IBΔRS + IOS(RS + ΔRS)
Figure 2.20: In-Amp Offset Voltage Model
Input bias currents may also produce offset errors in in-amp circuits (Fig. 2.20, again). If
the source resistance, RS, is unbalanced by an amount, ΔRS, (often the case in bridge
circuits), then there is an additional input offset voltage error due to the bias current,
equal to IBΔRS (assuming that IB+ ≈ IB– = IB). This error is reflected to the output, scaled
by the gain G.
The input offset current, IOS, creates an input offset voltage error across the source
resistance, RS+ΔRS, equal to IOS(RS+ΔRS), which is also reflected to the output by the
gain, G.
In-amp common-mode error is a function of both gain and frequency. Analog Devices
specifies in-amp CMR for a 1 kΩ source impedance unbalance at a frequency of 60 Hz.
2.23
BASIC LINEAR DESIGN
The RTI common-mode error is obtained by dividing the common-mode voltage, VCM,
by the common-mode rejection ratio, CMRR.
Figure 2.21 shows the CMR for the AD620 in-amp as a function of frequency, with a
1 kΩ source impedance imbalance.
160
140
120
G = 1000
G = 100
G = 10
CMR (dB)
100
G=1
80
60
40
20
0
0.1
1
10
100
1k
FREQUENCY (Hz)
10k
100k
1M
Figure 2.21: AD620 In-Amp Common-Mode Rejection (CMR)
vs. Frequency for 1kΩ Source Imbalance
Power supply rejection (PSR) is also a function of gain and frequency. For in-amps, it is
customary to specify the sensitivity to each power supply separately, as shown in Figure
2.22 for the AD620. The RTI power supply rejection error is obtained by dividing the
power supply deviation from nominal by the power supply rejection ratio, PSRR.
Because of the relatively poor PSR at high frequencies, decoupling capacitors are
required on both power pins to an in-amp. Low inductance ceramic capacitors (0.01 µF to
0.1 µF) are appropriate for high frequencies. Low ESR electrolytic capacitors should also
be located at several points on the PC board for low frequency decoupling.
2.24
OTHER LINEAR CIRCUITS
INSTRUMENTATION AMPLIFIERS
Figure 2.22: AD620 In-Amp Power Supply Rejection (PSR) vs. Frequency
Note that these decoupling requirements apply to all linear devices, including op amps
and data converters. Further details on power supply decoupling are found in Chapter 7.
Now that all dc error sources have been accounted for, a worst-case dc error budget can
be calculated by reflecting all the sources to the in-amp input, as is illustrated by the table
of Figure 2.23.
ERROR SOURCE
RTI VALUE
Gain Accuracy (ppm)
Gain Accuracy × FS Input
Gain Nonlinearity (ppm)
Gain Nonlinearity × FS Input
Input Offset Voltage, VOSI
VOSI
Output Offset Voltage, VOSO
VOSO ÷ G
Input Bias Current, IB, Flowing in ΔRS
IBΔRS
Input Offset Current, IOS, Flowing in RS
IOS(RS + ΔRS)
Common Mode Input Voltage, VCM
VCM ÷ CMRR
Power Supply Variation, ΔVS
ΔVS ÷ PSRR
Figure 2.23: In-Amp DC Errors Referred to the Input (RTI)
It should be noted that the dc errors can be referred to the in-amp output (RTO), by
simply multiplying the RTI error by the in-amp gain.
2.25
BASIC LINEAR DESIGN
In-Amp Noise Sources
Since in-amps are primarily used to amplify small precision signals, it is important to
understand the effects of all the associated noise sources. The in-amp noise model is
shown in Figure 2-24.
There are two sources of input voltage noise. The first is represented as a noise source,
VNI, in series with the input, as in a conventional op amp circuit. This noise is reflected to
the output by the in-amp gain, G. The second noise source is the output noise, VNO,
represented as a noise voltage in series with the in-amp output. The output noise, shown
here referred to as VOUT, can be referred to the input by dividing by the gain, G.
There are also two noise sources associated with the input noise currents IN+ and IN–.
Even though IN+ and IN– are usually equal (IN+ ≈ IN– = IN), they are uncorrelated, and
therefore, the noise they each create must be summed in a root-sum-squares (RSS)
fashion. IN+ flows through one half of RS, and IN– the other half. This generates two noise
voltages, each having an amplitude, INRS/2. Each of these two noise sources is reflected
to the output by the in-amp gain, G.
RS/2
~
~
VSIG
2
VSIG
2
VNI
RG
~
+
VNO
IN+
IN-AMP
GAIN = G
IN–
~
REF
_
VCM
VREF
RS/2
IF IN+ = IN–
NOISE (RTI) =
NOISE (RTO) =
BW •
BW •
VNO2
G2
+ VNI2 +
IN2RS2
2
I 2R 2
VNO2 + G2 VNI2 + N S
2
BW = 1.57 × IN-AMP Bandwidth @ Gain = G
Figure 2.24: In-Amp Noise Model
2.26
VOUT
OTHER LINEAR CIRCUITS
INSTRUMENTATION AMPLIFIERS
The total output noise is calculated by combining all four noise sources in an RSS
manner:
2R 2 I
2R 2 ⎞
⎛
I
NOISE ( RTO ) = BW VNO 2 + G 2 ⎜ VNI 2 + N + S + N − S ⎟
⎜
⎟
4
4
⎝
⎠
Eq. 2-6
If IN+ = IN– = IN,
⎛
I 2R 2
NOISE ( RTO ) = BW VNO 2 + G 2 ⎜ VNI 2 + N S
⎜
2
⎝
⎞
⎟
⎟
⎠
Eq. 2-7
The total noise, referred to the input (RTI) is simply the above expression divided by the
in-amp gain, G:
NOISE ( RTI ) = BW
VNO 2 ⎛⎜
I 2R 2 ⎞
+ VNI 2 + N S ⎟
⎜
⎟
2
G2
⎝
⎠
Eq. 2-8
In-amp data sheets often present the total voltage noise RTI as a function of gain. This
noise spectral density includes both the input (VNI) and output (VNO) noise contributions.
The input current noise spectral density is specified separately.
As in the case of op amps, the total in-amp noise RTI must be integrated over the
applicable in-amp closed-loop bandwidth to compute an rms value. The bandwidth may
be determined from data sheet curves that show frequency response as a function of gain.
Regarding this bandwidth, some care must be taken in computing it, as it is often not
constant bandwidth product relationship, as is true with VFB op amps. In the case of the
AD620 in-amp family, for example, the gain-bandwidth pattern is more like that of a
CFB op amp. In such cases, the safest way to predict the bandwidth at a given gain is to
use the curves supplied within the data sheet.
2.27
BASIC LINEAR DESIGN
In-Amp Bridge Amplifier Error Budget Analysis
It is important to understand in-amp error sources in a typical application. Figure 2.25
shows a 350 Ω load cell with a full-scale output of 100 mV when excited with a 10 V
source. The AD620 is configured for a gain of 100 using the external 499 Ω gain-setting
resistor. The table shows how each error source contributes to a total unadjusted error of
2145 ppm. Note, however, that the gain, offset, and CMR errors can all be removed with
a system calibration. The remaining errors—gain nonlinearity and 0.1 Hz to 10 Hz
noise— cannot be removed with calibration and ultimately limit the system resolution to
42.8 ppm (approximately 14-bit accuracy).
This example is of course just an illustration, but should be useful regarding the
importance of addressing performance-limiting errors such as gain nonlinearity and LF
noise.
+10V
VCM = 5V
499Ω
RG
+
AD620B
–
REF
G = 100
350Ω, 100mV FS
LOAD CELL
AD620B SPECS @ +25°C, ±15V
VOSI + VOSO/G = 55µV max
IOS = 0.5nA max
Gain Error = 0.15%
Gain Nonlinearity = 40ppm
0.1Hz to 10Hz Noise = 280nVp-p
CMR = 120dB @ 60Hz
MAXIMUM ERROR CONTRIBUTION, +25°C
FULLSCALE: VIN = 100mV, VOUT = 10V
VOS
55µV ÷ 100mV
550ppm
IOS
350Ω × 0.5nA ÷ 100mV
1.8ppm
Gain Error
0.15%
1500ppm
Gain
Nonlinearity
40ppm
40ppm
CMR Error
120dB
1ppm × 5V ÷ 100mV
50ppm
0.1Hz to 10Hz
1/f Noise
280nV ÷ 100mV
2.8ppm
Total
Unadjusted
Error
≈ 9 Bits Accurate
2145ppm
Resolution
Error
≈ 14 Bits Accurate
42.8ppm
Figure 2.25: AD620B Bridge Amplifier DC Error Budget
2.28
OTHER LINEAR CIRCUITS
INSTRUMENTATION AMPLIFIERS
In-Amp Input Overvoltage Protection
In their typical application as interface amplifiers for data acquisition systems, in-amps
are often subjected to input overloads, i.e., voltage levels in excess of the full scale for the
selected gain range. The manufacturer’s “absolute maximum” input ratings for the device
should be closely observed. As with op amps, many in-amps have absolute maximum
input voltage specifications equal to ±VS.
In some cases, external series resistors (for current limiting) and diode clamps may be
used to prevent overload, if necessary (see Figure 2.26). Some in-amps have built-in
overload protection circuits in the form of series resistors. For example, the AD620 series
have thin film resistors, and the substrate isolation they provide allows input voltages that
can exceed the supplies. Other devices use series-protection FETs, for example, the
AMP02 and the AD524, because they act as a low impedance during normal operation,
and a high impedance during overvoltage fault conditions. In any instance, however,
there are always finite safe limits to applied overvoltage (Fig. 2.26, again).
+VS
RLIMIT
+
INPUTS
RLIMIT
IN-AMP
OUTPUT
–
–VS
‹ Always Observe Absolute Maximum Data Sheet Specs!
‹ Schottky Diode Clamps to the Supply Rails Will Limit
Input to Approximately ±VS ±0.3V, TVSs Limit Differential Voltage
‹ External Resistors (or Internal Thin-Film Resistors) Can Limit
Input Current, but will Increase Noise
‹ Some In-Amps Have Series-Protection Input FETs for Lower Noise
and Higher Input Over-Voltages (up to ±60V, Depending on Device)
Figure 2.26: In-Amp Input Overvoltage Considerations
In some instances, an additional Transient Voltage Suppressor (TVS) may be required
across the input pins to limit the maximum differential input voltage. This is especially
applicable to three op amp in-amps operating at high gain with low values of RG.
A more detailed discussion of input overvoltage and EMI/RFI protection can be found in
Chapter 11 of this book.
2.29
BASIC LINEAR DESIGN
Notes:
2.30
OTHER LINEAR CIRCUITS
DIFFERENTIAL AMPLIFIERS
SECTION 2.4: DIFFERENTIAL AMPLIFIERS
Many high performance ADCs are now being designed with differential inputs. A fully
differential ADC design offers the advantages of good CM rejection, reduction in secondorder distortion products, and simplified dc trim algorithms. Although they can be driven
single-ended, a fully differential driver usually optimizes overall performance.
One of the most common ways to drive a differential input ADC is with a transformer.
However, there are many applications where the ADCs cannot be driven with
transformers because the frequency response must extend to dc. In these cases,
differential drivers are required.
A block diagram of the AD813X family of fully differential amplifiers optimized for
ADC driving is shown in Figure 2.27 (see References 3-5). Figure 2.27A shows the
details of the internal circuit, and Figure 2.27B shows the equivalent circuit. The gain is
set by the external RF and RG resistors, and the CM voltage is set by the voltage on the
VOCM pin. The internal CM feedback forces the VOUT+ and VOUT− outputs to be balanced,
i.e., the signals at the two outputs are always equal in amplitude but 180° out of phase per
the equation,
VOCM = ( VOUT+ + VOUT− ) / 2
Eq. 2-9
RF
(A)
V+
+
VIN+
RG
VIN–
RG
VOUT–
–
+
+
–
VOCM
–
–
VOUT+
+
V–
RF
(B)
RF
EQUIVALENT CIRCUIT:
VIN+
GAIN =
RF
RG
~
VIN–
RG
VOUT–
+
RG
VOCM
–
RF
VOUT+
Figure 2.27: AD813x Differential ADC Driver Functional Diagram and
Equivalent Circuit
2.31
BASIC LINEAR DESIGN
The circuit can be used with either a differential or a single-ended input, and the voltage
gain is equal to the ratio of RF to RG.
If a buffered differential voltage output is required from a current output DAC, the
AD813x-series of differential amplifiers can be used as shown in Figure 2.28.
2.49kΩ
0 TO 20mA
0 TO +0.5V
499Ω
+
IOUT
25Ω
CMOS
DAC
AD813X
5V p-p
DIFFERENTIAL
OUTPUT
499Ω
IOUT
–
20 TO 0mA
+0.5 TO 0V
2.49kΩ
25Ω
VOCM
Figure 2.28: Buffering High Speed DACs Using AD813x Differential Amplifier
The DAC output current is first converted into a voltage that is developed across the 25 Ω
resistors. The voltage is amplified by a factor of 5 using the AD813x. This technique is
used in lieu of a direct I/V conversion to prevent fast slewing DAC currents from
overloading the amplifier and introducing distortion. Care must be taken so that the DAC
output voltage is within its compliance rating.
The VOCM input on the AD813x can be used to set a final output CM voltage within the
range of the AD813x. If transmission lines are to be driven at the output, adding a pair of
75 Ω resistors will allow this.
Note also that these amplifiers can be used with single-ended inputs as well. Grounding
one of the inputs turns these amplifiers into single-ended-to-differential converters.
2.32
OTHER LINEAR CIRCUITS
ISOLATION AMPLIFIERS
SECTION 2.5: ISOLATION AMPS
Analog Isolation Techniques
There are many applications where it is desirable, or even essential, for a sensor to have
no direct (“galvanic”) electrical connection with the system to which it is supplying data.
This might be in order to avoid the possibility of dangerous voltages or currents from one
half of the system doing damage in the other, or to break an intractable ground loop. Such
a system is said to be “isolated” and the arrangement that passes a signal without galvanic
connections is known as an isolation barrier.
The protection of an isolation barrier works in both directions, and may be needed in
either, or even in both. The obvious application is where a sensor may encounter high
voltages, such as monitoring the current in an ac induction motor, and the system it is
driving must be protected. Or a sensor may need to be isolated from accidental high
voltages arising downstream, in order to protect its environment: examples include the
need to prevent the ignition of explosive gases by sparks at sensors and the protection
from electric shock of patients whose ECG, EEG, or EMG is being monitored. The ECG
case is interesting, as protection may be required in both directions: the patient must be
protected from accidental electric shock, but if the patient’s heart should stop, the ECG
machine must be protected from the very high voltages (>7.5 kV) applied to the patient
by the defibrillator which will be used to attempt to restart it.
Just as interference, or unwanted information, may be coupled by electric or magnetic
fields, or by electromagnetic radiation, these phenomena may be used for the
transmission of wanted information in the design of isolated systems.
The most common isolation amplifiers use transformers, which exploit magnetic fields,
and another common type uses small high voltage capacitors, exploiting electric fields.
Optoisolators, which consist of an LED and a photocell, provide isolation by using light,
a form of electromagnetic radiation. Different isolators have differing performance: some
are sufficiently linear to pass high accuracy analog signals across an isolation barrier.
With others, the signal may need to be converted to digital form before transmission for
accuracy is to be maintained (note this is a common V/F converter application).
Transformers are capable of analog accuracy of 12 bits to 16 bits and bandwidths up to
several hundred kHz, but their maximum voltage rating rarely exceeds 10 kV, and is
often much lower. Capacitively-coupled isolation amplifiers have lower accuracy,
perhaps 12-bits maximum, lower bandwidth, and lower voltage ratings—but they are low
cost. Optical isolators are fast and cheap, and can be made with very high voltage ratings
(4 kV to 7 kV is one of the more common ratings), but they have poor analog domain
linearity, and are not usually suitable for direct coupling of precision analog signals.
Linearity and isolation voltage are not the only issues to be considered in the choice of
isolation systems. Operating power is, of course, essential. Both the input and the output
circuitry must be powered, and unless there is a battery on the isolated side of the
2.33
BASIC LINEAR DESIGN
isolation barrier (which is possible, but rarely convenient), some form of isolated power
must be provided. Systems using transformer isolation can easily use a transformer
(either the signal transformer or another one) to provide isolated power, but it is
impractical to transmit useful amounts of power by capacitive or optical means. Systems
using these forms of isolation must make other arrangements to obtain isolated power
supplies— this is a powerful consideration in favor of choosing transformer isolated
isolation amplifiers: they almost invariably include an isolated power supply.
The isolation amplifier has an input circuit that is galvanically isolated from the power
supply and the output circuit. In addition, there is minimal capacitance between the input
and the rest of the device. Therefore, there is no possibility for dc current flow, and
minimum ac-coupling. Isolation amplifiers are intended for applications requiring safe,
accurate measurement of low frequency voltage or current (up to about 100 kHz) in the
presence of high common-mode voltage (to thousands of volts) with high common-mode
rejection. They are also useful for line-receiving of signals transmitted at high impedance
in noisy environments, and for safety in general-purpose measurements, where dc and
line-frequency leakage must be maintained at levels well below certain mandated
minimums. Principal applications are in electrical environments of the kind associated
with medical equipment, conventional and nuclear power plants, automatic test
equipment, and industrial process control systems.
AD210 3-Port Isolator
A basic form of isolator is the three-port isolator (input, power, output all isolated) is
shown in Figure 2.29. Note that in this diagram, the input circuits, output circuits, and
power source are all isolated from one another. This figure represents the circuit
architecture of a self-contained isolator, the AD210 (see References 1 and 2).
An isolator of this type requires power from a two-terminal dc power supply (PWR,
PWR COM). An internal oscillator (50 kHz) converts the dc power to ac, which is
transformer-coupled to the shielded input section, then converted to dc for the input stage
and the auxiliary power output. The output current capability of this output is typically
limited to ±15 mA.
The ac carrier is also modulated by the input stage amplifier output, transformer-coupled
to the output stage, demodulated by a phase-sensitive demodulator (using the carrier as
the reference), filtered, and buffered using isolated dc power derived from the carrier.
The AD210 allows the user to select gains from 1 to 100, using external resistors with the
input section op amp. Bandwidth is 20 kHz, and voltage isolation is 2500 V rms
(continuous) and ± 3500 VPEAK (continuous).
2.34
OTHER LINEAR CIRCUITS
ISOLATION AMPLIFIERS
FB
INPUT
–IN
_
+IN
+
OUTPUT
T1
_
DEMOD
FILTER
MOD
VO
+
ICOM
OCOM
+VISS
–VISS
T3
POWER
T2
INPUT
POWER
SUPPLY
+VOSS
OUTPUT
POWER
SUPPLY
–VOSS
POWER
OSCILLATOR
PWR
PWR COM
Figure 2.29: AD210 3-Port Isolation Amplifier
The AD210 is a 3-port isolation amplifier, thus the power circuitry is isolated from both
the input and the output stages and may therefore be connected to either (or to neither),
without change in functionality. It uses transformer isolation to achieve 3500 V isolation
with 12-bit accuracy.
Motor Control Isolation Amplifier
HIGH VOLAGE
AC INPUT < 2500V RMS
+15V
RG
0.01Ω
–IN
+IN
AD620
REF
_
INPUT
FB
+
T1
_
+
OUTPUT
_
DEMOD
FILTER
MOD
+
OUTPUT
VO
OCOM
ICOM
–15V
+VISS
–VISS
M
RG = 499Ω
FOR G = 100
INPUT
POWER
SUPPLY
AD210
POWER
T2
T3
OUTPUT
POWER
SUPPLY
+VOSS
–VOSS
POWER
OSCILLATOR
PWR
PWR COM
+15V
Figure 2.30: Motor Control Current Sensing
A typical isolation amplifier application using the AD210 is shown in Figure 2.30. The
AD210 is used with an AD620 instrumentation amplifier in a current-sensing system for
2.35
BASIC LINEAR DESIGN
motor control. The input of the AD210, being isolated, can be directly connected to a
110 V or 230 V power line without protection being necessary. The input section’s
isolated ±15 V powers the AD620, which senses the voltage drop in a small value current
sensing resistor. The AD210 input stage op amp is simply connected as a unity-gain
follower, which minimizes its error contribution. The 110 V rms or 230 V rms commonmode voltage is ignored by this isolated system
Within this system the AD620 preamp is used as the system scaling control point, and
will produce and output voltage proportional to motor current, as scaled by the sensing
resistor value and gain as set by the AD620’s RG. The AD620 also improves overall
system accuracy, as the AD210 VOS is 15 mV, versus the AD620’s 30 µV (with less drift
also). Note that if higher dc offset and drift are acceptable, the AD620 may be omitted
and the AD210 connected at a gain of 100.
Optional Noise Reduction Post Filter
Due to the nature of this type of carrier-operated isolation system, there will be certain
operating situations where some residual ac carrier component will be superimposed
upon the recovered output dc signal. When this occurs, a low impedance passive RC filter
section following the output stage may be used (if the following stage has a high input
impedance, i.e., nonloading to this filter). Note that will be the case for many high input
impedance sampling ADCs, which appear essentially as small capacitors. A 150 Ω
resistance and 1 nF capacitor will provide a corner frequency of about 1kHz. Note also
that the capacitor should be a film type for low errors, such as polypropylene. As an
option an active filter may be utilized. Since the output of the filter is low impedance (the
output of an op amp) it may be used where the low output is required. Also note that it
may be possible to include the antialiasing requirement of the ADC into this filter.
Two-Port Isolator
A two port isolator differs from a three port isolator in that the power section is not
isolated from the output section. The AD215 is an example of a high speed, two-port
isolation amplifier, designed to isolate and amplify wide bandwidth analog signals (see
Reference 3). The innovative circuit and transformer design of the AD215 ensures wideband dynamic characteristics, while preserving dc performance specifications. An AD215
block diagram is shown in Figure 2.31.
The AD215 provides complete galvanic isolation between the input and output of the
device, which also includes the user-available front-end isolated bipolar power supply.
The functionally complete design, powered by a ±15 V dc supply on the output side,
eliminates the need for a user supplied isolated dc/dc converter. This permits the designer
to minimize circuit overhead and reduce overall system design complexity and
component costs.
2.36
OTHER LINEAR CIRCUITS
ISOLATION AMPLIFIERS
The AD215 has a ±10 V input/output range, a specified gain range of 1 V/V to 10 V/V, a
buffered output with offset trim and a user-available isolated front end power supply
which produces ±15 V dc at ±10 mA.
Figure 2.31: AD215 120 kHz Low Distortion 2-Port Isolation Amplifier
2.37
BASIC LINEAR DESIGN
Notes:
2.38
OTHER LINEAR CIRCUITS
DIGITAL ISOLATION
SECTION 2.6: DIGITAL ISOLATION TECHNIQUES
While not a linear circuit, digital isolation is closely related to isolation amplifiers, so
they will be discussed here.
Analog isolation amplifiers find many applications where a high isolation is required,
such as in medical instrumentation. Digital isolation techniques provide similar galvanic
isolation, and are a reliable method of transmitting digital signals without ground noise.
VDD1 (5V) 425Ω
VIN
CMOS
GATE
GND1
10kΩ
IIN
IOUT
HIGH VOLTAGE
ISOLATION BARRIER
VDD2 (5V)
VOUT
GND2
‹ Uses Light for Transmission Over a High Voltage Barrier
‹ The LED is the Transmitter, and the Phototransistor is the Receiver
‹ High Voltage Isolation: 5000V to 7000V RMS
‹ Non-Linear -- Best for Digital or Frequency Information
‹ Rise and Fall-times can be 10 to 20µs in Slower Devices
‹ Example: Siemens ILQ-1 Quad (http://www.siemens.com)
Figure 2.32: Digital Isolation Using LED / Phototransistor Opto-couplers
Opto-couplers (also called opto-isolators) are useful and available in a wide variety of
styles and packages. A typical opto-coupler based on an LED and a phototransitor is
shown in Figure 2.32. A current of approximately 10 mA drives an LED transmitter, with
light output is received by a phototransistor. The light produced by the LED saturates the
phototransistor. Input/output isolation of 5000 V rms to 7000 V rms is common.
Although fine for digital signals, opto-couplers are too nonlinear for most analog
applications. In addition, the transfer characteristics of the opto-coupler changes with
time. Also, since the phototransistor is often being saturated, response times can range
from 10 µs to 20 µs in slower devices, limiting high speed applications.
A much faster opto-coupler architecture is shown in Figure 2.33 and is based on an LED
and a photodiode. The LED is again driven with a current of approximately 10 mA. This
produces a light output sufficient to generate enough current in the receiving photodiode
to develop a valid high logic level at the output of the transimpedance amplifier. Speed
can vary widely between opto-couplers, and the fastest ones have propagation delays of
20 ns typical, and 40 ns maximum, and can handle data rates up to 25 MBd for NRZ data.
2.39
BASIC LINEAR DESIGN
This corresponds to a maximum square wave operating frequency of 12.5 MHz, and a
minimum allowable passable pulse width of 40 ns.
VDD1
VDD2
(Data In)
VIN
VOUT
(Data Out)
GND1
GND2
‹
‹
‹
‹
‹
‹
‹
‹
+5V Supply Voltage
2500V RMS I/O Withstand Voltage
Logic Signal Frequency: 12.5MHz Maximum
25MBd Maximum Data Rate
40ns Maximum Propagation Delay
9ns Typical Rise/Fall Time
Example: Agilent HCPL-7720
(http://www.semiconductor.agilent.com)
Figure 2.33: Digital Isolation using LED / photodiode opto-couplers
AD260/AD261 High Speed Logic Isolators
The AD260/AD261 family of digital isolators operates on a principle of transformercoupled isolation (see Reference 4). They provide isolation for five digital control signals
to/from high speed DSPs, microcontrollers, or microprocessors. The AD260 also has a
1.5 W transformer for a 3.5 kV rms isolated external dc/dc power supply circuit.
DATA
IN
SCHMITT
TRIGGER
TRI STATE
BUFFER
LATCH
D
XMTR
RCVR
DATA
OUT
E
ENABLE
ENABLE
CONTINUOUS
UPDATE
CIRCUIT
ISOLATED POWER
XFMR (AD260)
37V p-p, 1.5W
NOTE: SINGLE DATA CHANNEL SHOWN
3500V RMS ISOLATION BARRIER
(AD260B/AD261B)
Figure 2.34: AD260/AD261 digital isolators
2.40
OTHER LINEAR CIRCUITS
DIGITAL ISOLATION
Each line of the AD260 can handle digital signals up to 20 MHz (40 MBd) with a
propagation delay of only 14 ns which allows for extremely fast data transmission.
Output waveform symmetry is maintained to within ±1 ns of the input so the AD260 can
be used to accurately isolate time-based pulse width modulator (PWM) signals.
A simplified schematic of one channel of the AD260/AD261 is shown in Figure 2.34.
The data input is passed through a Schmitt trigger circuit, through a latch, and a special
transmitter circuit which differentiates the edges of the digital input signal and drives the
primary winding of a proprietary transformer with a set-high/set-low signal.
The secondary of the isolation transformer drives a receiver with the same set-hi/set-low
data, which regenerates the original logic waveform. An internal circuit operating in the
background interrogates all inputs about every 5 µs, and in the absence of logic
transitions, sends appropriate set-hi/set-low data across the interface. Recovery time from
a fault condition or at power-up is thus between 5 µs and 10 µs.
The power transformer (available on the AD260) is designed to operate between 150 kHz
and 250 kHz and will easily deliver more than 1 W of isolated power when driven pushpull (5 V) on the transmitter side. Different transformer taps, rectifier and regulator
schemes will provide combinations of ±5 V, 15 V, 24 V, or even 30 V or higher.
The transformer output voltage when driven with a low voltage-drop drive will be
37 V p-p across the entire secondary with a 5 V push-pull drive. The availability of low
cost digital isolators such as those previously discussed solves most system isolation
problems in data acquisition systems as shown in Figure 2.35. In the upper example,
digitizing the signal first, then using digital isolation eliminates the problem of analog
isolation amplifiers. While digital isolation can be used with parallel output ADCs
provided the bandwidth of the isolator is sufficient, it is more practical with ADCs that
have serial outputs. This minimizes cost and component count. A 3-wire interface (data,
serial clock, framing clock) is all that is required in these cases.
ISOLATION
BARRIER
SENSOR
ADC
DIGITAL
ISOLATORS
DAC
SENSOR
V/F
CONVERTER
DIGITAL
ISOLATOR
F/V
CONVERTER
GROUND REFERENCE A
GROUND REFERENCE B
Figure 2.35: Practical application of digital isolation in data acquisition systems
2.41
BASIC LINEAR DESIGN
An alternative (lower example) is to use a voltage-to-frequency converter (VFC) as a
transmitter and a frequency-to-voltage converter (FVC) as a receiver. In this case, only
one digital isolator is required.
iCoupler® Technology
In many industrial applications, such as process control systems or data acquisition and
control systems, digital signals must be transmitted from various sensors to a central
controller for processing and analysis. The controller then needs to transmit commands as
a result of the analysis performed, coupled with user inputs to various actuators, to
achieve certain operations. To maintain safety voltage at the user interface and to prevent
transients from being transmitted from the sources, galvanic isolation is required. There
are three commonly known classes of isolation devices: opto-couplers, capacitively
coupled isolators, and transformer-based isolators. Opto-couplers rely on light emitting
diodes to convert the electrical signals to light signals and on photodetectors to convert
the light signals back to electrical signals. The intrinsic low conversion efficiencies for
electrical light conversion and slow response photodetectors lead to opto-coupler
limitations in terms of lifetime, speed, and power assumption. The capacitively coupled
isolators have limitations in their size and ability to reject common-mode voltage
transients, while the traditional transformer assembly based isolators are bulky and
expensive. All these isolators are restricted, moreover, because of integrated circuit
integration limitations and the fact that they often need hybrid packaging.
Recently iCoupler, a new isolation technology based on chip scale transformers, was
developed by Analog Devices. The first product was the ADuM1100 single-channel
digital isolator. iCoupler technology leverages thick-film processing techniques to build
microscale on-chip transformers and achieves thousands of volts of isolation on a chip.
iCoupler isolated transformers can be monolithically integrated with standard silicon ICs
and can be fabricated in single- or multichannel configurations. The bidirectional nature
of inductive coupling further facilitates bidirectional signal transfer. The combination of
high bandwidth for these on-chip transformers and fine scale CMOS circuitry leads to
isolators of unmatched performance characteristics in power, speed, timing accuracy, and
ease of use.
ADuM1100 Architecture: A Single-Channel Digital Isolator
The ADuM1100 is a single-channel 100 Mbps digital isolator. It has two ICs packaged in
an 8-lead SOIC package. A cross-section view of the ADuM1100 is shown in
Figure 2.36. There are two lead frame paddles inside the package, with a gap between
them of about 0.4 mm. The molding compound has breakdown strength over 25 kV/mm,
so the 0.4 mm gap filled with molding compound provides greater than 10 kV insulation
between the substrates of two IC chips.
2.42
OTHER LINEAR CIRCUITS
DIGITAL ISOLATION
The driver chip sitting on the left paddle takes the input digital signal, encodes it, and
drives the encoded differential signal through bond wires to the top coils of the
transformers built on top of the receiver chip sitting on the right paddle. The driver die is
a standard CMOS chip, and the receiver die is a CMOS chip with the additional
structures of two polyimide layers and transformer primary coil fabricated on top of the
passivation. The polyimide between the top and bottom coils is about 20 µm thick. The
breakdown strength of the cured polyimide film is greater than 300 V/m, so 20 µm of
polyimide provides greater than 6 kV of insulation between a given transformer’s coils.
This provides a comfortable margin over the production test voltage of 3 kV rms.
Figure 2.36: Cross-sectional view of ADuM1100 in an 8-lead SOIC package;
Figure b. Cross-sectional view of the top coil and polyimide layers
Because of the structural quality of these wafer processed polyimide films, no partial
discharge over 5 pC can be detected, even at 3 kV rms. The top coil is gold plated, with a
4 µm thick layer, and the coil track width and spacing between the turns are all 4 µm. The
polyimide layers have good mechanical elongation and tensile strength, which also helps
the adhesion between the polyimide layers or between polyimide layer and deposited
metal layer. The minimum interaction between the gold film and the polyimide film,
coupled with high temperature stability of the polyimide film, results in a system that
provides reliable insulation when subjected to various types of environmental stress.
In addition to the fact that thousands of volts of isolation can be achieved on-chip, the
ADuM110 also makes it possible to transmit very high bandwidth signals very
efficiently, accurately, and reliably. Figure 2.37 is a simplified schematic of the
ADuM1100. To guarantee input stability, the front glitch filter filters out pulses narrower
than a pulse width of approximately 2 ns. Upon the receipt of a signal edge, a 1 ns pulse
is sent to either Coil 1 or Coil 2. (For a leading edge signal it is sent to Coil 1, and for a
falling edge signal to Coil 2.) Once the short pulses are transmitted to the secondary coils
(the bottom coils in this case), they are amplified and the input signal is reconstructed
through an SR flip-flop to appear as an isolated output. The wide bandwidth of these
microscale transformers and high speed CMOS make the transmission of these short
nanosecond pulses possible. Since only signal edges are being used, this transmission
scheme is very power efficient. With a very energetic pulse having a current ramping to
100 mA within 1 ns, the average current for a 1 Mbps input signal is only 50 µA. Some
additional power is dissipated by the switching of the surrounded CMOS gates. At 5 V,
2.43
BASIC LINEAR DESIGN
an additional 50 µA/Mbps is needed if the total capacitance of the CMOS gates is 20 pF.
The typical opto-coupler, on the other hand, dissipates over 10 mA, even operating at
1 Mbps. This represents two orders of magnitude (100×) improvement in power
dissipation provided by iCoupler isolators.
If there is no input change for a certain period of time, approximately 1 µs, the
monostable generates a 1 ns pulse and sends it to Coil 1 or Coil 2, depending on the input
logic level. The 1 ns refreshing pulse is sent to Coil 1 if input is high and is sent to Coil 2
if input is low. This helps maintain dc correctness for the isolator because normally
pulses are transmitted only on reception of a signal edge. The receiver includes a
watchdog circuit that will timeout at 2 µs if it is not reset by an incoming pulse. If a
timeout happens, the receiver output will return to a default safe level (logic high in the
ADuM1100). The combination of refresh and watchdog functions provides the additional
advantage of detecting the failure of any field device on the system side. With other
isolators, this would ordinarily require the use of an extra isolated data channel.
The bandwidth of the isolator is dependent on the input filter bandwidth within. For
example, 500 Mbps can be achieved with a 2 ns input filter. For the ADuM1100, we
chose a signal bandwidth of 100 MBd, still 2× faster than the fastest opto-couplers. Very
tight edge symmetry between input and output logic signals is also preserved due to the
instantaneous nature of the inductive coupling between these microscale on-chip coils.
The ADuM1100 has edge symmetry of better than 2 ns for 5 V operation. As the
bandwidth of isolation systems continues to expand, the iCoupler technology will be
capable of meeting the challenge while opto-coupler technology is likely to struggle.
1ns
+
DIFF
SET_HI
SPULE 1
INPUT
1ns
VSS1
VSS2
GLITCH
FILTER
1μs
Q
A STA B LE
+
DIFF
S
REFRESH
R
R
OUTPUT
1ns
1ns
+
DIFF
SET_LO
2μs
WA TCHDOG
Figure 2.37: ADuM1100 Simplified Schematic
2.44
OTHER LINEAR CIRCUITS
DIGITAL ISOLATION
In addition to the improvements in efficiency and bandwidth iCoupler technology
provides, it also offers a more robust and reliable isolation solution than competitive
offerings. Because high voltage transients are present in many data acquisition and
control systems, the ability of the isolator to prevent transients from affecting the logic
controller is very important. High performance opto-couplers have transient immunity of
less than 10 kV/µs, while the ADuM1100 has a transient immunity better than 25 kV/µs.
The induced error voltage at the receiver input induced by an input-output transient is
given by:
V = C*R
dV
dt
Eq. 2-10
where:
C is the capacitance between the input coil and the receiver coil
R is the resistance of the bottom coil
dV/dt is the magnitude of the transient
In the ADuM1100, the capacitance between the top (input) coil and the bottom (receiver)
coil is only 0.2 pF, while the bottom coil has a resistance of 80 Ω. Thus the error signal
induced on the bottom coil by a 25 kV/µs transient on the top coil is only 0.4 V, much
less than the receiver detection threshold. The transient immunity of iCoupler isolators
can be optimized through careful selection of the decoder detection threshold, the
resistance of the receiving coil, and, of course, the capacitance between the top and
bottom coils.
One recurring question about transformer-based isolators involves their magnetic
immunity capability. Since iCouplers use air core technology, no magnetic components
are present and the problem of magnetic saturation for the core material does not exist.
Therefore, iCouplers have essentially infinite dc field immunity. The limitation on the
ADuM1100’s ac magnetic field immunity is set by the condition in which the induced
error voltage in the receiving coil (the bottom coil in this case) is made sufficiently large,
either to falsely set or reset the decoder. The voltage induced across the bottom coil is
given by:
V = - dβ Σπrn2 ; n = 1,2,…N
dt
where:
Eq. 2-11
β = magnetic flux density (Gauss)
N = number of turns in receiving coil
rn = radius of nth turn in receiving coil (cm)
Because of the very small geometry of the receiving coil in the ADuM1100, even a wire
carrying 1000 A at 1 MHz and positioned only 1 cm away from the ADuM1100 would
not induce an error voltage large enough to falsely trigger the decoder. Note that at
combinations of strong magnetic field and high frequency, any loops formed by printed
circuit board traces could induce error voltages sufficiently large to trigger the thresholds
of succeeding circuitry. Typically the PC board design rather than the isolator itself is the
2.45
BASIC LINEAR DESIGN
limiting factor in the presence of such big magnetic transients. In addition to magnetic
immunity, the level of electromagnetic radiation emitted from the iCoupler device is a
concern. Using far-field approximation:
P = 160π2I2Σrn4; n = 1,2,…,N
Eq. 2-12
where:
P = total radiated power
I = coil loop current
Again, given the very small geometry of the coils, the total radiated power is still less
than 50 pW, even if the part is operating at 0.5 GHz.
ADuM130x/ADuM140x: Multichannel Products
In addition to the many performance improvements discussed previously, iCoupler
technology also offers tremendous advantages in terms of integration. The optical
interference makes the realization of multichannel opto-couplers very difficult.
Transformers based on iCoupler technology can be easily integrated onto a single chip.
Furthermore, one data channel can transmit signals in one direction, say from the top coil
to the bottom coil, while the neighboring channel can transmit a signal in the other
direction, from the bottom coil to the top coil. The bidirectional nature of inductive
coupling makes this possible.
Additional products consist of five 3-channel and 4-channel products covering all
possible channel directionality configurations. Besides providing flexible channel
configurations, they support both 3 V and 5 V operation at either side of the isolation
barrier and support the use of these isolators as level translators. One side could be at
2.7 V, for example, while the other side could be at 5.5 V. The edge symmetry of 2 ns is
preserved over all possible supply configurations at all temperatures from –40°C to
+100°C. The ability to mix bidirectional channels of isolation in a single package enables
users to reduce the size and cost of their systems.
For the ADuM1100, two transformers are used to transmit a single channel of data. One
is dedicated to transmit pulses representing the signal’s leading edge or updating input
high, and the other is dedicated to transmit pulses representing the signal’s falling edge or
updating input low. For the ADuM130x/ADuM140x product family, a single transformer
is used for each data channel. The ADuM140x shown in Figure 2.38 has four
transformers in total. The leading edge and falling edge are encoded differently, and the
encoded pulses are combined in the same transformer; as a result, the receiver has
responsibility for decoding the pulses to see whether they are for leading edge or falling
edge. The output signal is then reconstructed correspondingly.
Of course, there is a penalty for using a single transformer per data channel rather than
using two transformers per data channel. The propagation delay is longer for the single
2.46
OTHER LINEAR CIRCUITS
DIGITAL ISOLATION
transformer architecture because of the additional encode and decode time needed. The
penalty for bandwidth is hardly a factor, even at input speed of 100 Mbps.
In contrast to the ADuM1100, the ADuM130x/ADuM140x uses a dedicated transformer
chip, separate from the receiver integrated circuit. This partitioning exemplifies the ease
of integration for iCoupler technology. Besides standalone multichannel isolators, the
iCoupler technology can be embedded with other data acquisition and control ICs to
make the use of isolation even more transparent. Consequently, in the future, system
designers will be able to devote their time to improving system functionality, rather than
worrying about isolation.
Figure 2.38: ADuM140X Die Photograph
2.47
BASIC LINEAR DESIGN
(a)
(b)
(c)
Figure 2.39: Block Diagrams for the ADuM1400 (a), ADuM1401 (b)
and ADuM1402 (c)
2.48
OTHER LINEAR CIRCUITS
ACTIVE FEEDBACK AMPLIFIERS
SECTION 2.7: ACTIVE FEEDBACK AMPLIFIERS
The AD8129/AD8130 differential line receivers, along with their predecessor, the
AD830, utilize a novel amplifier topology called active feedback (see Reference 8). A
simplified block diagram of these devices is shown below in Figure 2.40.
The AD830 and the AD8129/AD8130 have two sets of fully differential inputs, available
at VX1 − VX2 and VY1 − VY2, respectively. Internally, the outputs of the two GM stages are
summed and drive a buffer output stage.
In this device the overall feedback loop forces the internal currents IX and IY to be equal.
This condition forces the differential voltages VX1 – VX2 and VY1 – VY2 to be equal and
opposite in polarity. Feedback is taken from the output back to one input differential pair,
while the other pair is driven directly by an input differential input signal.
VX1
+
VIN
VX2
GM
IX
–
VOUT
A=1
VY1
IY
+
R3
= R1||R2
VY2
GM
–
R2
R1
Feedback Forces IX = IY
∴ VX1 – VX2 = – (VY1 – VY2) = VY2
VOUT =
1+
R2
R1
VOUT =
1+
R2
R1
VY2
VX1 – VX2
Figure 2.40: The AD830/AD8129/AD8130 Active Feedback Amplifier Topology
An important point of this architecture is that high CM rejection is provided by the two
differential input pairs, so CMR is not dependent on resistor bridges and their associated
matching problems. The inherently wideband balanced circuit and the quasi-floating
operation of the driven input provide the high CMR, which is typically 100 dB at dc.
2.49
BASIC LINEAR DESIGN
One way to view this topology is as a standard op amp in a noninverting mode with a pair
of differential inputs in place of the op amps standard inverting and noninverting inputs.
The general expression for the stage’s gain “G” is like a noninverting op amp, or:
G=
VOUT
R2
=1+
.
VIN
R1
Eq. 2-13
As should be noted, this expression is identical to the gain of a noninverting op amp
stage, with R2 and R1 in analogous positions.
The AD8129 is a low noise, high gain (G = 10 or greater) version of this family, intended
for applications with very long cables where signal attenuation is significant. The related
AD8130 device is stable at a gain of one. It is used for those applications where lower
gains are required, such as a gain-of-2, for driving source and load terminated cables.
The AD8129 and AD8130 have a wide power supply range, from single +5 V to ±12 V,
allowing wide common-mode and differential-mode voltage ranges. The wide commonmode range enables the driver/receiver pair to operate without isolation transformers in
many systems where the ground potential difference between driver and receiver
locations is several volts. Both devices include a logic-controlled power-down function.
Both devices have high, balanced input impedances, and achieve 70 dB CMR @
10 MHz, providing excellent rejection of high-frequency common-mode signals.
Figure 2.41 shows AD8130 CMR for various supplies. As can be noted, it can be as high
as 95 dB at 1 MHz, an impressive figure considering that no trimming is required.
COMMON-MODE REJECTION (dB)
–30
–40
–50
–60
–70
–80
–90
V S = ±2.5V
–100
–110
–120
10k
VS = ±5V, ±12V
100k
1M
10M
FREQUENCY (Hz)
100M
Figure 2.41: AD8130 Common-Mode Rejection vs. Frequency
for ±2.5 V, ±5 V, and ±12 V Supplies
2.50
OTHER LINEAR CIRCUITS
ACTIVE FEEDBACK AMPLIFIERS
The typical 3 dB bandwidth for the AD8129 is 200 MHz, while the 0.1 dB bandwidth is
30 MHz in the SOIC package, and 50 MHz in the µSOIC package. The conditions for
these specifications are for VS = ±5 V and G = 10.
The typical 3 dB bandwidth for the AD8130 is 270 MHz, and the 0.1 dB bandwidth is
45 MHz, in either package. The conditions for these specifications are for VS = ±5 V and
G = 1. Typical differential gain and phase specifications for the AD8130 for G = 2,
VS = ±5 V, and RL = 150 Ω are 0.13 % and 0.15°, respectively.
2.51
BASIC LINEAR DESIGN
Notes:
2.52
OTHER LINEAR CIRCUITS
LOGARITHMIC AMPLIFIERS
SECTION 2.8: LOGARITHMIC AMPLIFIERS
The term “logarithmic amplifier” (generally abbreviated “log amp”) is something of a
misnomer, and “Logarithmic Converter” would be a better description. The conversion of
a signal to its equivalent logarithmic value involves a nonlinear operation, the
consequences of which can be confusing if not fully understood. It is important to realize
that many of the familiar concepts of linear circuits are irrelevant to log amps. For
example, the incremental gain of an ideal log amp approaches infinity as the input
approaches zero, and a change of offset at the output of a log amp is equivalent to a
change of amplitude at its input—not a change of input offset.
For the purposes of simplicity in our initial discussions, we shall assume that both the
input and the output of a log amp are voltages, although there is no particular reason why
logarithmic current, transimpedance, or transconductance amplifiers could not also be
designed.
If we consider the equation y = log(x) we find that every time x is multiplied by a
constant A, y increases by another constant A1. Thus if log(K) = K1, then log(AK) = K1
+ A1, log(A2K) = K1 + 2A1, and log(K/A) = K1 – A1. This gives a graph as shown in
Figure 2.42, where y is zero when x is unity, y approaches minus infinity as x approaches
zero, and which has no values for y for which x is negative.
Y
X
X=1
Figure 2.42: Graph of Y = Log(X)
On the whole, log amps do not behave in this way. Apart from the difficulties of
arranging infinite negative output voltages, such a device would not, in fact, be very
useful. A log amp must satisfy a transfer function of the form:
Vout = Vy log(Vin/Vx)
Eq. 2-14
2.53
BASIC LINEAR DESIGN
over some range of input values which may vary from 100:1 (40 dB) to over 1,000,000:1
(120 dB).
With inputs very close to zero, log amps cease to behave logarithmically, and most then
have a linear Vin/Vout law. This behavior is often lost in device noise. Noise often limits
the dynamic range of a log amp. The constant, Vy, has the dimensions of voltage,
because the output is a voltage. The input, Vin, is divided by a voltage, Vx, because the
argument of a logarithm must be a simple dimensionless ratio.
A graph of the transfer characteristic of a log amp is shown in Figure 2.43. The scale of
the horizontal axis (the input) is logarithmic, and the ideal transfer characteristic is a
straight line. When Vin = Vx, the logarithm is zero (log 1 = 0). Vx is therefore known as
the intercept voltage of the log amp because the graph crosses the horizontal axis at this
value of Vin.
VYLOG (VIN /VX )
IDEAL
ACTUAL
2VY
SLOPE = VY
VY
VOUT = VY log10
VIN
VX
+
0
ACTUAL
-
IDEAL
VIN=VX
VIN=10VX
VIN=100VX INPUT ON
LOG SCALE
Figure 2.43: Log Amp Transfer Function
The slope of the line is proportional to Vy. When setting scales, logarithms to the base 10
are most often used because this simplifies the relationship to decibel values: when
Vin = 10 Vx, the logarithm has the value of 1, so the output voltage is Vy. When
Vin = 100 Vx, the output is 2 Vy ,and so forth. Vy can therefore be viewed either as the
“slope voltage” or as the “volts per decade factor.”
The logarithm function is indeterminate for negative values of x. Log amps can respond
to negative inputs in three different ways: (1) They can give a full-scale negative output
as shown in Figure 2.44. (2) They can give an output which is proportional to the log of
the absolute value of the input and disregards its sign as shown in Figure 2.45. This type
of log amp can be considered to be a full-wave detector with a logarithmic characteristic,
2.54
OTHER LINEAR CIRCUITS
LOGARITHMIC AMPLIFIERS
+
OUTPUT
-
+
-
Figure 2.44: Basic Log Amp
(Saturates with Negative Inputs)
+
OUTPUT
-
+
INPUT
-
Figure 2.45: Detecting Log Amp
(Output Polarity Independent of Input Polarity)
+
-
OUTPUT
+
INPUT
-
Figure 2.46: Log Video or "True Log Amp"
(Symmetrical Response to Positive or Negative Signals)
2.55
BASIC LINEAR DESIGN
and is often referred to as a detecting log amp. (3) They can give an output which is
proportional to the log of the absolute value of the input and has the same sign as the
input as shown in Figure 2.46. This type of log amp can be considered to be a video amp
with a logarithmic characteristic, and may be known as a logarithmic video (log video)
amplifier or, sometimes, a true log amp (although this type of log amp is rarely used in
video-display-related applications).
There are three basic architectures which may be used to produce log amps: the basic
diode log amp, the successive detection log amp, and the true log amp which is based on
cascaded semi-limiting amplifiers. The successive detection log amp and the true log amp
are discussed in the RF/IF section.
The voltage across a silicon diode is proportional to the logarithm of the current through
it. If a diode is placed in the feedback path of an inverting op-amp, the output voltage will
be proportional to the log of the input current as shown in Figure 2.47. In practice, the
dynamic range of this configuration is limited to 40 dB to 60 dB because of nonideal
diode characteristic, but if the diode is replaced with a diode-connected transistor as
shown in Figure 2.48, the dynamic range can be extended to 120 dB or more. This type of
log amp has three disadvantages: (1) both the slope and intercept are temperature
dependent; (2) it will only handle unipolar signals; and (3) its bandwidth is both limited
and dependent on signal amplitude.
-VIN
+ V V=
I
RIN
I = I IN
IIN
I
kT
In
IO
q
-
if I >> I O
EO
+
V
EO =
kT
In
q
IIN
IO
0.06log
if IIN >> I O
Figure 2.47: The Diode/Op Amp Log Amp
IC
IE
EO
IIN
+
IIN
kT
EO = q In
IES
Figure 2.48: Transistor/Op Amp Log Amp
2.56
VIN
RINIO
OTHER LINEAR CIRCUITS
LOGARITHMIC AMPLIFIERS
Where several such log amps are used on a single chip to produce an analog computer
which performs both log and antilog operations, the temperature variation in the log
operations is unimportant, since it is compensated by a similar variation in the
antilogging. This makes possible the AD538 (Figure 2.49), a monolithic analog computer
which can multiply, divide, and raise to powers. Where actual logging is required,
however, the AD538 and similar circuits require temperature compensation
(Reference 7). The major disadvantage of this type of log amp for high frequency
applications, though, is its limited frequency response—which cannot be overcome.
However carefully the amplifier is designed, there will always be a residual feedback
capacitance Cc (often known as Miller capacitance), from output to input which limits the
high frequency response.
I Z 11
VZ 2
8 A
LOG
RATIO
25k
17 D
16 IX
B 3
+10V 4
100
100
25k
+2V 5
INTERNAL
VOLTAGE
REFERENCE
+VS 6
–VS 7
AD538
15 VX
14
SIGNAL
GND
13
PWR
GND
12 C
OUTPUT
25k
VO 8
I 9
11 IY
ANTILOG
LOG
25k
10 VY
Figure 2.49: AD538 block diagram
What makes this Miller capacitance particularly troublesome is that the impedance of the
emitter-base junction is inversely proportional to the current flowing in it—so that if the
log amp has a dynamic range of 1,000,000:1, then its bandwidth will also vary by
1,000,000:1. In practice, the variation is less because other considerations limit the large
signal bandwidth, but it is very difficult to make a log amp of this type with a smallsignal bandwidth greater than a few hundred kHz.
We also discuss high speed log amps in the RF/IF section (Section 4.4)
2.57
BASIC LINEAR DESIGN
Notes:
2.58
OTHER LINEAR CIRCUITS
HIGH SPEED CLAMPING AMPLIFIERS
SECTION 2.9 HIGH SPEED CLAMPING AMPLIFIERS
There are many situations where it is desirable to clamp the output of an op amp to
prevent overdriving the circuitry which follows. Specially designed high speed, fast
recovery clamping amplifiers offer an attractive alternative to designing external
clamping/protection circuits. The AD8036/AD8037 low distortion, wide bandwidth
clamp amplifiers represent a significant breakthrough in this technology. These devices
allow the designer to specify a high (VH) and low (VL) clamp voltage. The output of the
device clamps when the input exceeds either of these two levels. The AD8036/AD8037
offer superior clamping performance compared to competing devices that use outputclamping. Recovery time from overdrive is less than 5 ns.
The key to the AD8036 and AD8037’s fast, accurate clamp and amplifier performance is
their proprietary input clamp architecture. This new design reduces clamp errors by more
than 10× over previous output clamp based circuits, as well as substantially increasing the
bandwidth, precision, and versatility of the clamp inputs.
RF
140Ω
-V IN
+VIN
+
A1
-
A
+1
A2
+1
VOUT
S1
VH
+1
VL
+1
B
C
+
S1
V IN > V H
CH
-
V L ≤ V IN ≤ V H
V IN < V L
A B C
0
1
0
1
0
0
0
0
1
+
CL
-
Figure 2.50: AD8036/AD8037 Clamp Amplifier Equivalent Circuit
Figure 2.50 is an idealized block diagram of the AD8036 connected as a unity gain
voltage follower. The primary signal path comprises A1 (a 1200 V/µs, 240 MHz high
voltage gain, differential to single-ended amplifier) and A2 (a G = +1 high current gain
output buffer). The AD8037 differs from the AD8036 only in that A1 is optimized for
closed-loop gains of two or greater.
2.59
BASIC LINEAR DESIGN
The input clamp section is comprised of comparators CH and CL, which drive switch S1
through a decoder. The unity-gain buffers in series with the +VIN, VH, and VL inputs
isolate the input pins from the comparators and S1 without reducing bandwidth or
precision.
The two comparators have about the same bandwidth as A1 (240 MHz), so they can keep
up with signals within the useful bandwidth of the AD8036. To illustrate the operation of
the input clamp circuit, consider the case where VH is referenced to +1 V, VL is open,
and the AD8036 is set for a gain of +1 by connecting its output back to its inverting input
through the recommended 140 Ω feedback resistor. Note that the main signal path always
operates closed loop, since the clamping circuit only affects A1’s noninverting input.
If a 0 V to +2 V voltage ramp is applied to the AD8036’s +VIN for the connection just
described, VOUT should track +VIN perfectly up to +1 V, then should limit at exactly
+1 V as +VIN continues to +2 V.
In practice, the AD8036 comes close to this ideal behavior. As the +VIN input voltage
ramps from zero to 1 V, the output of the high limit comparator CH starts in the off state,
as does the output of CL. When +VIN just exceeds VH (practically, by about 18 mV),
CH changes state, switching S1 from “A” to “B” reference level. Since the + input of A1
is now connected to VH, further increases in +VIN have no effect on the AD8036’s
output voltage. The AD8036 is now operating as a unity-gain buffer for the VH input, as
any variation in VH, for VH > 1 V, will be faithfully produced at VOUT.
Operation of the AD8036 for negative input voltages and negative clamp levels on VL is
similar, with comparator CL controlling S1. Since the comparators see the voltage on the
+VIN pin as their common reference level, the voltage VH and VL are defined as “High”
or “Low” with respect to +VIN. For example, if VIN is set to zero volts, VH is open, and
VL is +1 V, comparator CL will switch S1 to “C,” so the AD8036 will buffer the voltage
on VL and ignore +VIN.
The performance of the AD8036/AD8037 closely matches the ideal just described. The
comparator’s threshold extends from 60 mV inside the clamp window defined by the
voltages on VL and VH to 60 mV beyond the window’s edge. Switch S1 is implemented
with current steering, so that A1’s + input makes a continuous transition from say, VIN to
VH as the input voltage traverses the comparator’s input threshold from 0.9 V to 1.0 V
for VH = 1.0 V.
The practical effect of the nonideal operation is to soften the transition from amplification
to clamping modes, without compromising the absolute clamp limit set by the input
clamping circuit. Figure 2.51 is a graph of VOUT versus VIN for the AD8036 and a
typical output clamp amplifier. Both amplifiers are set for G = +1 and VH = +1 V.
2.60
OTHER LINEAR CIRCUITS
HIGH SPEED CLAMPING AMPLIFIERS
Figure 2.51: Comparison of Input and Output Clamping
The worst-case error between VOUT (ideally clamped) and VOUT (actual) is typically
18 mV times the amplifier closed-loop gain. This occurs when VIN equals VH (or VL).
As VIN goes above and/or below this limit, VOUT will stay within 5 mV of the ideal
value.
In contrast, the output clamp amplifier’s transfer curve typically will show some
compression starting at an input of 0.8 V, and can have an output voltage as far as
200 mV over the clamp limit. In addition, since the output clamp causes the amplifier to
operate open-loop in the clamp mode, the amplifier’s output impedance will increase,
potentially causing additional errors, and the recovery time is significantly longer.
It is important that a clamped amplifier such as the AD8036/AD8037 maintain low levels
of distortion when the input signals approach the clamping voltages. Figure 2.52 shows
the second and third harmonic distortion for the amplifiers as the output approaches the
clamp voltages. The input signal is 20 MHz, the output signal is 2 V peak-to-peak, and
the output load is 100 Ω.
Recovery from step voltage which is two times over the clamping voltage is shown in
Figure 2.53. The input step voltage starts at +2 V and goes to 0 V (left-hand traces on
scope photo). The input clamp voltage (VH) is set at +1 V. The right-hand trace shows
the output waveform.
2.61
BASIC LINEAR DESIGN
-80
-75
-70
AD8037 3RD
HARMONIC
-65
-60
AD8037 2ND
HARMONIC
-55
-50
AD8036 3RD
HARMONIC
AD8036 2ND
HARMONIC
-45
VH
AD8036
+1V
VL
-1V
-0.5V
G
+1
+2
0.8
0.85
-40
-35
-30
0.6
0.65
0.7
0.75
AD8037
+0.5V
0.9
0.95
1.0
ABSOLUTE VALUE OF OUTPUT VOLTAGE - Volts
Figure 2.52: AD8036/AD8037 Distortion Near Clamping Region,
Output = 2 V p-p, Load = 100 Ω, f = 20 MHz
INPUT
OUTPUT
+2V
+1V
0V
1ns
REF
HORIZONTAL
SCALE:
1ns/div
Figure 2.53: AD8036/AD8037 Overdrive (2×) Recovery
Figure 2.54 shows the AD9002 8-bit, 125 MSPS flash converter driven by the AD8037
(240 MHz bandwidth) clamping amplifier. The clamp voltages on the AD8037 are set to
+0.55 V and –0.55 V, referenced to the ±0.5 V input signal, with the external resistive
dividers. The AD8037 also supplies a gain of two, and an offset of –1 V (using the
AD780 voltage reference), to match the 0 V to –2 V input range of the AD9002 flash
converter. The output signal is clamped at +0.1 V and –2.1 V. This multifunction
clamping circuit therefore performs several important functions as well as preventing
damage to the flash converter which occurs if its input exceeds +0.5 V, thereby forward
2.62
OTHER LINEAR CIRCUITS
HIGH SPEED CLAMPING AMPLIFIERS
biasing the substrate diode. The 1N5712 Schottky diode adds further protection during
power-up.
0.1μF
BIPOLAR
SIGNAL
±0.5V
+5V
806Ω
IN5712
+
RT
75Ω
VH = +0.55V
750Ω
AD780
+2.5V
REF
0.1μF
AD9002
FLASH CONVERTER
(8-BITS, 125MSPS)
VIN = -1 ±1V
VL = -0.55V
-
10μF
49.9Ω
AD8037
CLAMPING
AMP
+5V
+
100Ω
806Ω
SUBSTRATE
DIODE
100Ω
R3
-5.2V
0.1μF
-5.2V
R2
301Ω
R1
R1 R3 = R2
499Ω
2.5 R1
R1 + R3
0.1μF
AD8037 OUTPUT
CLAMPS AT +0.1V, -2.1V
= 1 VOLT
Figure 2.54: AD9002 8-Bit, 125 MSPS Flash Converter
Driven by an AD8037 Clamp Amplifier
The feedback resistor, R2 = 301 Ω, is selected for optimum bandwidth per the data sheet
recommendation. For a gain of 2, the parallel combination of R1 and R3 must also equal
R2:
R1 ⋅ R 3
= R 2 = 301Ω
R1 + R 3
Eq. 2-15
(nearest 1% standard resistor value).
In addition, the Thevenin equivalent output voltage of the AD780 +2.5 V reference and
the R3/ R1 divider must be +1 V to provide the –1 V offset at the output of the AD8037.
2.5 ⋅ R1
= 1 volt
R1 + R 3
Eq. 2-16
Solving the equations yields R1 = 499 Ω, R3 = 750 Ω (using the nearest 1% standard
resistor values).
Other input and output voltages ranges can be accommodated by appropriate changes in
the external resistors.
Further examples of applications of these fast clamping op amps are given in
Reference 9.
2.63
BASIC LINEAR DESIGN
Notes:
2.64
OTHER LINEAR CIRCUITS
COMPARATORS
SECTION 2.10: COMPARATORS
A comparator is similar to an op amp. It has two inputs, inverting and noninverting, and
an output. But it is specifically designed to compare the voltages between its two inputs.
Therefore it operates in a nonlinear fashion. The comparator operates open-loop,
providing a two-state logic output voltage. These two states represent the sign of the net
difference between the two inputs (including the effects of the comparator input offset
voltage). Therefore, the comparator’s output will be a Logic 1 if the input signal on the
noninverting input exceeds the signal on the inverting input (plus the offset voltage, Vos)
and a Logic 0 for the opposite case. A comparator is normally used in applications where
some varying signal level is compared to a fixed level (usually a voltage reference). Since
it is, in effect, a 1-bit analog-to-digital converter (ADC), the comparator is a basic
element in all ADCs.
Figure 2.55: Comparator Symbol
Comparator dc specifications are similar to those of op amps: input offset voltage, input
bias current, offset and drift, common-mode input range, gain, CMR, and PSR. Standard
logic-related dc, timing and interface specs are associated with the comparator outputs.
The key comparator ac specification is propagation delay: it is the time required for the
output to reach the 50% point of a transition, after the differential input signal crosses the
offset voltage—when driven by a square wave (typically 100 mV in amplitude) to a
prescribed value of input overdrive (usually 5 mV or 10 mV). See Figure 2.55.
The propagation delay in practical comparators decreases somewhat as the input
overdrive is increased. This variation in propagation delay as a function of overdrive is
called dispersion. See Figure 2.56.
2.65
BASIC LINEAR DESIGN
Figure 2.56: Propagation Delay
The addition of hysteresis, which is application of a small amount of positive feedback, to
a comparator’s transfer function is often useful in a noisy environment, or where it is
undesirable for the comparator to toggle continuously between states when the input
signal is at or near the switching threshold. This is true when a relatively slowly changing
input is compared to a dc level. Noise can cause the output to toggle between the output
levels many times. The transfer function for a comparator with hysteresis is shown in
Figure 2.58.
Figure 2.57: Delay Dispersion
2.66
OTHER LINEAR CIRCUITS
COMPARATORS
If the input voltage approaches the switching threshold (Vos) from the negative direction,
the comparator will switch from a 0 to a 1 when the input crosses Vos + VH/2. The
“new” switching threshold now becomes Vos – VH/2. The comparator output will
remain in a 1 state until the threshold Vos – VH/2 is crossed, coming from the positive
direction. Input noise centered around Vos will not cause the comparator to switch states
unless it exceeds the region bounded by Vos ± VH/2.
Figure 2.58: Effects of Hysteresis
Hysteresis can be accomplished with two resistors, see Figure 2.59, the amount of
hysteresis is proportional to the resistors’ ratio. The signal input to the comparator may
be applied to either the inverting or the noninverting input, but if it is applied to the
inverting input its source impedance must be low enough to have insignificant effect on
R1 (of course if the source impedance is sufficiently predictable it may be used as R1).
OUTPUT SWING: VS
HYSTERESIS =
VS (R1 × R2)
R1
R1
R2
z INPUT SIGNAL MAY BE APPLIED TO EITHER INPUT
BUT ITS SOURCE IMPENDANCE MUST BE LOW IF IT IS APPLIED TO R1
Figure 2.59: Application of Hysteresis
2.67
BASIC LINEAR DESIGN
If the trip voltage is midway between the two comparator output voltages (as is the case
with a symmetrical power supply and a ground reference) then the introduction of
hysteresis will move the positive and negative thresholds equal distances from the trip
point voltage, but if the trip point is nearer to one output than to the other the thresholds
will be asymmetrically placed about the trip point voltage.
To calculate the hysteresis, assume the comparator output voltages are Vp and Vn
respectively. The comparator trip point voltage is VOS . The negative threshold is:
(R1 + R2) VOS – R1Vn
Eq. 2-17
R2
And the positive threshold voltage is:
(R1 + R2) VOS – R1Vp
Eq. 2-18
R2
No external hysteresis
5 mV external hysteresis
Figure 2.60: Hysteresis Helps Clean Up Comparator Response.
A problem encountered with external hysteresis is that output voltage depends on supply
voltage and loading. This means the hysteresis voltage can vary from application to
application; though this affects resolution, it need not be a serious problem, since the
hysteresis is usually a very small fraction of the range and can tolerate a safety margin of
two or three (or more) times what one might calculate. Swapping in a few comparators
can help confidence in the safety margin. Don’t use wirewound resistors for feedback;
their inductance can make matters worse.
Some comparators have hysteresis built in. An example of this is the AD790. See
Figure 2.61. The hysteresis voltage is nominally 500 µV. This, of course, can be
overridden by applying external hysteresis.
The AD790 has an additional advantage. The supplies on the input (analog) side are not
necessarily those on the output. The output swing is from VLOGIC to GND. The input
supplies can be ±15 V down to +5 V and ground.
2.68
OTHER LINEAR CIRCUITS
COMPARATORS
VL OG IC
+
–
+IN
+
–IN
–
A1
Q1
OUT PUT
Av
–
A2
+
GAIN ST AGE
OUT PUT ST AGE
Q2
GND
Figure 2.61: AD790 Block Diagram.
It is quite common for the output of a comparator to be open collector (open drain). This
allows interfacing to whatever logic level is appropriate to the following circuitry. Note
that the maximum allowable output voltage must be observed, but this is usually not too
great an issue.
A window comparator makes use of two comparators with different reference voltages
and a common input voltage. The comparators are connected to logic in such a way that
the final output logic level is asserted when the input signal falls between the two
reference voltages (Figure 2.62).
Figure 2.62: Window Comparator
Many comparators have an internal latch. The latch-enable signal has two states:
compare (track) and latch (hold). When the latch-enable signal is in the compare state,
the comparator output continuously responds to the sign of the net differential input
signal. When the latch-enable signal transitions to the latch state, the comparator output
2.69
BASIC LINEAR DESIGN
goes to either a Logic 1 or a Logic 0, depending on the sign of the differential input
signal at the instant of the transition (at this point, we are neglecting the setup and holdtime, as well as the output propagation delay associated with the latch-enable function).
Even though many comparators have a latch-enable function, they are often operated
only in the compare mode.
The comparator internal latch-enable function is particularly useful in ADC applications
because it allows the comparator decision to be recorded at a known instant of time. Flash
converters make use of this concept and are constructed of many parallel comparators
which share a common latch-enable line. Typical timing associated with the latch-enable
function is shown in Figure 2.63. The delay between the assertion of latch-enable and the
50% point of the output logic swing is referred to as latch-enable to output delay. It may
be different for positive and negative-going outputs. The other key specification
associated with the latch-enable function is the minimum allowable latch-enable pulse
width. This specification determines the maximum frequency at which the comparator
can be strobed.
Figure 2.63: Effects of Output Latch
Fast comparators are somewhat difficult to apply because of their high gain and
bandwidth. Proper application of high speed layout, grounding, decoupling, and signal
routing is mandatory when using comparators. This can not be overemphasized. The
biggest problem is their tendency to oscillate when the input signal is very near to or
equal to the switching threshold. This can also happen when a slow signal is compared to
a dc reference. Hysteresis and the use of a narrow latch-enable pulse will generally help
these conditions. TTL comparators are more likely to oscillate than ECL ones because of
their large output swings and fast edges, often combined with power supply current
spikes as the output changes state. This can lead to feedback to the input in the form of
noise.
2.70
OTHER LINEAR CIRCUITS
COMPARATORS
Using Op Amps as Comparators
Even though op amps and comparators may seem interchangeable at first glance there are
some important differences.
Comparators are designed to work open-loop, they are designed to drive logic from their
outputs, and they are designed to work at high speed with minimal instability. Op-amps
are not designed for use as comparators, they may saturate if over-driven which may
cause it to recover comparatively slowly. Many have input stages which behave in
unexpected ways when used with large differential voltages, in fact, in many cases, the
differential input voltage range of the op amp is limited. And op amp outputs are rarely
compatible with logic.
Yet many people still try to use op amps as comparators. While this may work at low
speeds and low resolutions, many times the results are not satisfactory. Not all of the
issues involved with using an op amp as a comparator can be resolved by reference to the
op amp data sheet, since op amps are not intended for use as comparators.
The most common issues are speed (as we have already mentioned), the effects of input
structures (protection diodes, phase inversion in FET amplifiers, and many others), output
structures which are not intended to drive logic, hysteresis and stability, and commonmode effects.
Speed
Most comparators are quite fast, but so are some op amps. Why must we expect low
speed when using an op amp as a comparator?
A comparator is designed to be used with large differential input voltages, whereas opamps normally operate with their differential input voltage minimized by negative
feedback. When an op amp is over-driven, sometimes by only a few millivolts, some of
its stages may saturate. If this occurs the device will take a comparatively long time to
come out of saturation and will therefore be much slower than if it always remained
unsaturated.
The time to come out of saturation of an overdriven op-amp is likely to be considerably
longer than the normal group delay of the amplifier, and will often depend on the amount
of overdrive. Since few op amps have this desaturation time specified for various
amounts of overdrive it will generally be necessary to determine, by experiment, the
behavior of the amplifier under the conditions of overdrive to be expected in a particular
application.
2.71
BASIC LINEAR DESIGN
The results of such experiments should be regarded with suspicion and the values of
propagation delay through the op-amp comparator which is chosen for worst-case design
calculations should be at least twice the worst value seen in any experiment.
Figure 2.64: Effects of Saturation on Amplifier Speed
When Used as a Comparator
Output Considerations
The output of a comparator will be designed to drive a particular logic family or families,
while the output of an op amp is designed to swing from supply rail to supply rail.
Frequently the logic being driven by the op amp comparator will not share the op amp’s
supplies and the op amp rail-to-rail swing may go outside the logic supply rails—this will
probably destroy the logic circuitry, and the resulting short-circuit may destroy the op
amp as well.
There are three types of logic which we must consider: ECL, TTL and CMOS.
ECL is a very fast current steering logic family. It is unlikely that an op amp would be
used as a comparator in applications where ECL’s highest speed is involved, for reasons
given above, so we shall usually be concerned only to drive ECL logic levels from an op
amp’s signal swing and some additional loss of speed due to stray capacities will be
unimportant. To do this we need only three resistors, as shown in Figure 2.65.
R1, R2, and R3 are chosen so that when the op amp output is positive the level at the gate
is −0.8 V, and when it is low it is −1.6 V. ECL is occasionally used with positive, rather
than negative, supplies (i.e., the other rail is connected to ground), the same basic
interface circuit may be used but the values must be recalculated.
2.72
OTHER LINEAR CIRCUITS
COMPARATORS
R1
R3
R2
ECL GATE
COMPARATOR
–5.2V
ECL SUPPLY
Figure 2.65: Op Amp Comparator Driving ECL Logic
Although CMOS and TTL input structures, logic levels, and current flows are quite
different (although some CMOS is specified to work with TTL input levels) the same
interface circuitry will work perfectly well with both types of logic, since they both work
for Logic 0 near to 0 V and Logic 1 near to 5 V.
+5V
RL
TO CMOS
OR
TTL GATE
+5V
RL
COMPARATOR
TO CMOS
OR
TTL GATE
Figure 2.66: Op Amp Comparator Driving TTL or CMOS Logic
2.73
BASIC LINEAR DESIGN
The simplest interface uses a single N-channel MOS transistor and a pull-up resistor, RL.
A similar circuit may be made with an NPN transistor, RL, and an additional transistor
and diode. These circuits are simple, inexpensive and reliable, and may be connected
with several transistors in parallel and a single RL to give a “wired-or” function, but the
speed of the 0 to 1 transition depends on the value of RL and the stray capacity of the
output node. The lower the value of RL the faster, but the higher the power consumption.
By using two MOS devices, one P-channel and one N-channel, it is possible to make a
CMOS/TTL interface using only two components which has no quiescent power
consumption in either state.
Furthermore, it may be made inverting or noninverting by simple positioning of
components. It does, however, have a large current surge during switching, when both
devices are on at once, and unless MOS devices with high channel resistance are used a
current limiting resistor may be necessary to reduce this effect. It is also important, in this
application and the one in Figure 2.67, to use MOS devices with gate-source breakdown
voltages, Vbgs, greater than the output voltages of the comparator in either direction. A
value of Vbgs > ±25 V is common in MOS devices and is usually adequate, but many
MOS devices contain gate protection diodes which reduce the value—these should not be
used.
Figure 2.67: Op Amp Comparator with CMOS Drive
2.74
OTHER LINEAR CIRCUITS
COMPARATORS
Input Circuitry
There are a number of effects which must be considered regarding the inputs of op amps
used as comparators. The first-level assumption engineers make about all op amps and
comparators is that they have infinite input impedance and can be regarded as open
circuits (except for current feedback (transimpedance) op amps, which have a high
impedance on their noninverting input but a low impedance of a few tens of Ω on their
inverting input).
But many op amps (especially bias-compensated ones such as the OP-07 and its many
descendants) contain protective circuitry to prevent large voltages damaging input
devices.
Others contain more complex input circuitry, which only has high impedance when the
differential voltage applied to it is less than a few tens of mV, or which may actually be
damaged by differential voltages of more than a few volts. It is therefore necessary, when
using an op amp as a comparator, to study the data sheet to determine how the input
circuitry behaves when large differential voltages are applied to it. (It is always necessary
to study the data sheet when using an integrated circuit to ensure that its nonideal
behavior (and every integrated circuit ever made has some nonideal behavior) is
compatible with the proposed application—it is just more important than usual in the
present case.)
Of course some comparator applications never involve large differential voltages—or if
they do the comparator input impedance when large differential voltages are present is
comparatively unimportant. In such cases it may be appropriate to use as a comparator an
op amp whose input circuitry behaves non-linearly—but the issues involved must be
considered, not just ignored.
Figure 2.68: Op Amp Input Structure with Protection
As mentioned elsewhere in this seminar, nearly all BIFET op amps exhibit anomalous
behavior when their inputs are close to one of their supplies (usually the negative supply).
Their inverting and noninverting inputs may become interchanged. If this should occur
when the op amp is being used as a comparator the phase of the system involved will be
2.75
BASIC LINEAR DESIGN
inverted, which could well be inconvenient. The solution is, again, careful reading of the
data sheet to determine just what common-mode range is acceptable.
Also, absence of negative feedback means that, unlike that of op amp circuits, the input
impedance is not multiplied by the loop gain. As a result, the input current varies as the
comparator switches. Therefore the driving impedance, along with parasitic feedbacks,
can play a key role in affecting circuit stability. While negative feedback tends to keep
amplifiers within their linear region, positive feedback forces them into saturation.
2.76
OTHER LINEAR CIRCUITS
ANALOG MULTIPLIERS
SECTION 2.11: ANALOG MULTIPLIERS
A multiplier is a device having two input ports and an output port. The signal at the
output is the product of the two input signals. If both input and output signals are
voltages, the transfer characteristic is the product of the two voltages divided by a scaling
factor, K, which has the dimension of voltage (see Figure 2.69). From a mathematical
point of view, multiplication is a four quadrant operation—that is to say, that both inputs
may be either positive or negative and the output can be positive or negative. Some of the
circuits used to produce electronic multipliers, however, are limited to signals of one
polarity. If both signals must be unipolar, we have a single quadrant multiplier, and the
output will also be unipolar. If one of the signals is unipolar, but the other may have
either polarity, the multiplier is a two quadrant multiplier, and the output may have either
polarity (and is bipolar). The circuitry used to produce one- and two-quadrant multipliers
may be simpler than that required for four quadrant multipliers, and since there are many
applications where full four quadrant multiplication is not required, it is common to find
accurate devices which work only in one or two quadrants. An example is the AD539, a
wideband dual two-quadrant multiplier which has a single unipolar Vy input with a
relatively limited bandwidth of 5 MHz, and two bipolar Vx inputs, one per multiplier,
with bandwidths of 60 MHz. A block diagram of the AD539 is shown in Figure 2.71.
Figure 2.69: Multiplier Block Diagram
Figure 2.70: Multiplier Input/Output Relationships
2.77
BASIC LINEAR DESIGN
Figure 2.71: AD539 Block Diagram
The simplest electronic multipliers use logarithmic amplifiers. The computation relies on
the fact that the antilog of the sum of the logs of two numbers is the product of those
numbers (see Figure 2.72).
Figure 2.72: Log Amps as Multiplier
2.78
OTHER LINEAR CIRCUITS
ANALOG MULTIPLIERS
The disadvantages of this type of multiplication are the very limited bandwidth and single
quadrant operation. A far better type of multiplier uses the Gilbert Cell. This structure
was invented by Barrie Gilbert, now of Analog Devices, in the late 1960s. See
References 1 and 2.
There is a linear relationship between the collector current of a silicon junction transistor
and its transconductance (gain) which is given by
dIc / dVbe = qIc / kT
Eq. 2-19
where
Ic = the collector current
Vbe = the base-emitter voltage
q = the electron charge (1.60219x10-19)
k = Boltzmann's constant (1.38062x10-23)
T = the absolute temperature.
This relationship may be exploited to construct a multiplier with a differential (longtailed) pair of silicon transistors, as shown in Figure 2.73.
This is a rather poor multiplier because (1) the Y input is offset by the Vbe—which
changes nonlinearly with Vy; (2) the X input is non-linear as a result of the exponential
relationship between Ic and Vbe; and (3) the scale factor varies with temperature.
Figure 2.73: Simple Multiplier
2.79
BASIC LINEAR DESIGN
Gilbert realized that this circuit could be linearized and made temperature stable by
working with currents, rather than voltages, and by exploiting the logarithmic Ic/Vbe
properties of transistors (See Figure 2.74). The X input to the Gilbert Cell takes the form
of a differential current, and the Y input is a unipolar current. The differential X currents
flow in two diode-connected transistors, and the logarithmic voltages compensate for the
exponential Vbe/Ic relationship. Furthermore, the q/kT scale factors cancel. This gives
the Gilbert Cell the linear transfer function
ΔIc =
ΔΙx I y
Ix
Eq. 2-20
As it stands, the Gilbert Cell has three inconvenient features: (1) its X input is a
differential current; (2) its output is a differential current; and (3) its Y input is a unipolar
current—so the cell is only a two quadrant multiplier.
By cross-coupling two such cells and using two voltage-to-current converters (as shown
in Figure 2.75), we can convert the basic architecture to a four quadrant device with
voltage inputs, such as the AD534. At low and medium frequencies, a subtractor
amplifier may be used to convert the differential current at the output to a voltage.
Because of its voltage output architecture, the bandwidth of the AD534 is only about
1 MHz, although the AD734, a later version, has a bandwidth of 10 MHz.
Figure 2.74: Four Quadrant Gilbert Cell
In Figure 2.75, Q1A & Q1B, and Q2A & Q2B form the two core long-tailed pairs of the
two Gilbert Cells, while Q3A and Q3B are the linearizing transistors for both cells. In
Figure 2.75 there is an operational amplifier acting as a differential current to single2.80
OTHER LINEAR CIRCUITS
ANALOG MULTIPLIERS
ended voltage converter, but for higher speed applications, the cross-coupled collectors of
Q1 and Q2 form a differential open collector current output (as in the AD834 500 MHz
multiplier).
The translinear multiplier relies on the matching of a number of transistors and currents.
This is easily accomplished on a monolithic chip. Even the best IC processes have some
residual errors, however, and these show up as four dc error terms in such multipliers. In
early Gilbert Cell multipliers, these errors had to be trimmed by means of resistors and
potentiometers external to the chip, which was somewhat inconvenient. With modern
analog processes, which permit the laser trimming of SiCr thin film resistors on the chip
itself, it is possible to trim these errors during manufacture so that the final device has
very high accuracy. Internal trimming has the additional advantage that it does not reduce
the high frequency performance, as may be the case with external trimpots.
Because the internal structure of the translinear multiplier is necessarily differential, the
inputs are usually differential as well (after all, if a single-ended input is required it is not
hard to ground one of the inputs). This is not only convenient in allowing common-mode
signals to be rejected, it also permits more complex computations to be performed. The
AD534 (shown previously in Figure 2.71) is the classic example of a four-quadrant
multiplier based on the Gilbert Cell. It has an accuracy of 0.1% in the multiplier mode,
fully differential inputs, and a voltage output. However, as a result of its voltage output
architecture, its bandwidth is only about 1 MHz.
Figure 2.75: A Multiplier and an Op Amp Configured as a Divider
in Both Inverting and Noninverting Mode.
2.81
BASIC LINEAR DESIGN
Multipliers can be placed in the feedback loop of op amps to form several useful
functions. Figure 2.76 illustrates the basic principle of analog computation that a function
generator in a negative feedback loop computes the inverse function (provided, of course,
that the function is monotonic over the range of operations).
Figure 2.76: Generating an Inverse Function
Figure 2.77: A Divider Circuit
High speed multipliers are also discussed in the RF/IF section (Section 4.3)
2.82
OTHER LINEAR CIRCUITS
RMS TO DC CONVERTERS
SECTION 2.12: RMS TO DC CONVERTERS
The root mean square (rms) is a fundamental measurement of the magnitude of an ac
signal. Defined practically, the rms value assigned to the ac signal is the amount of dc
required to produce an equivalent amount of heat in the same load. Defined
mathematically, the rms value of a voltage is defined as the value obtained by squaring
the signal, taking the average, and then taking the square root. The averaging time must
be sufficiently long to allow filtering at the lowest frequencies of operation desired. A
complete discussion of rms to dc converters can be found in Reference 13, but we will
show a few examples of how efficiently analog circuits can perform this function.
The first method, called the explicit method, is shown in Figure 2.78. The input signal is
first squared by a multiplier. The average value is then taken by using an appropriate
filter, and the square root is taken using an op amp with a second squarer in the feedback
loop. This circuit has limited dynamic range because the stages following the squarer
must try to deal with a signal that varies enormously in amplitude. This restricts this
method to inputs which have a maximum dynamic range of approximately 10:1 (20 dB).
However, excellent bandwidth (greater than 100 MHz) can be achieved with high
accuracy if a multiplier such as the AD834 is used as a building block (see Figure 2.75).
VIN
V
X
VIN
2
AVG
K
VY
R
VO2
K
V =
O
RC >>
VO
C
(VIN2) AVG
VX
VY
1
2 πf
Figure 2.78: Explicit RMS Computation
2.83
BASIC LINEAR DESIGN
75Ω
49.9Ω
1μF
24.9Ω
49.9Ω
INPUT
-6V
8
7
6
5
X2
X1
+VS
W1
100Ω
5μF
AD834
0.1μF
Y1
Y2
-VS
W2
1
2
3
4
5μF
100Ω
33pF
49.9Ω
1μF
1
47.5Ω
24.9Ω
49.9Ω
3
10kΩ
2
49.9Ω
8
15kΩ
15kΩ
7
6
5
X1
+VS
W1
+
-
8
7
2N3904
6
OUTPUT
0.1μF
AD834
Y1
Y2
-VS
W2
1
2
3
4
10Ω
100Ω
100Ω
-5V
Figure 2.79: Wideband RMS Measurement
Figure 2.79 shows the circuit for computing the rms value of a signal using the implicit
method. Here, the output is fed back to the direct-divide input of a multiplier such as the
AD734. In this circuit, the output of the multiplier varies linearly (instead of as the
square) with the rms value of the input. This considerably increases the dynamic range of
the implicit circuit as compared to the explicit circuit. The disadvantage of this approach
is that it generally has less bandwidth than the explicit computation.
VX
VIN
( VIN2 )AVG
VIN 2
VO
VY
VO
R
VO
C
VZ
DIRECT
DIVIDE
INPUT
(AD734)
(DENOMINATOR
INPUT)
V =
O
( VIN2 )AVG
Figure 2.80: Implicit RMS Calculation
2.84
RC >>
1
2π f
OTHER LINEAR CIRCUITS
RMS TO DC CONVERTERS
While it is possible to construct such an rms circuit from an AD734, it is far simpler to
design a dedicated rms circuit. The VIN2/VZ circuit may be current driven and need only
be one quadrant if the input first passes through an absolute value circuit.
Figure 2.81 shows a simplified diagram of a typical monolithic rms/dc converter, the
AD536A. It is subdivided into four major sections: absolute value circuit (active
rectifier), squarer/divider, current mirror, and buffer amplifier. The input voltage VIN,
which can be ac or dc, is converted to a unipolar current, I1, by the absolute value circuit
A1, A2. I1 drives one input of the one-quadrant squarer/divider which has the transfer
function: I4 = I12/I3. The output current, I4, of the squarer/divider drives the current
mirror through a lowpass filter formed by R1 and externally connected capacitor, CAV.
If the R1CAV time constant is much greater than the longest period of the input signal,
then I4 is effectively averaged. The current mirror returns a current, I3, which equals
AVG[I4], back to the squarer/divider to complete the implicit rms computation. Thus:
I4 = AVG [I12/I4] = I1rms
Eq. 2-21
The current mirror also produces the output current, Iout , which equals 2I4. Iout can be
used directly or converted to a voltage with R2 and buffered by A4 to provide a low
impedance voltage output. The transfer function becomes:
Vout = 2R2•Irms = VIN rms
Eq. 2-22
The dB output is derived from the emitter of Q3, since the voltage at this point is
proportional to –logVIN. Emitter follower, Q5, buffers and level shifts this voltage, so
that the dB output voltage is zero when the externally supplied emitter current (IREF) to
Q5 approximates I3. However, the gain of the dB circuit has a TC of approximately
3300 ppm/°C and must be temperature compensated.
There are a number of commercially available rms/dc converters in monolithic form
which make use of these principles. The AD536A is a true rms/dc converter with a
bandwidth of approximately 450 kHz for V rms > 100 mV rms, and 2 MHz bandwidth
for V rms > 1 V rms. The AD636 is designed to provide 1MHz bandwidth for low-level
signals up to 200 mV rms. The AD637 has a 600 kHz bandwidth for 100 mV rms signals,
and an 800 MHz bandwidth for 1 V rms signals. Low cost, general purpose rms/dc
converters such as the AD736 and AD737 (power-down option) are also available.
2.85
BASIC LINEAR DESIGN
CURRENT MIRROR
14
+VS
10
COM
9
RL
5
dB
OUT
R1
25k
R2
I3
ABSOLUTE VALUE/
VOLTAGE - CURRENT
CONVERTER
8
4
CAV
I1
R4
IOUT
IREF
I4
A3
Q1
Q3
VIN
+
1
R3
+
A1
-
BUF
IN
R4
Q5
Q2
A2
+
Q4
7
BUFFER
+
-
ONE - QUADRANT
SQUARER/
DIVIDER
Figure 2.81: The AD536A Monolithic RMS-to-DC Converter
2.86
6
BUF
OUT
3
-VS
OTHER LINEAR CIRCUITS
PROGRAMMABLE GAIN AMPLIFIERS
SECTION 2.13: PROGRAMMABLE GAIN AMPLIFIERS
Most systems with wide dynamic range need some method of adjusting the input signal
level to the analog-to-digital-converter (ADC). The ADC compares the input signal to a
fixed voltage reference (+5 V or +10 V are typical values). To achieve the rated precision
of the converter, the maximum input should be fairly near its full scale voltage. However,
transducers have a wide range of output voltages. High gain is needed for a small sensor
voltage, but with a large transducer output, a high gain will cause the amplifier or ADC to
saturate. So some type of controllable gain device is needed. Such a device has a gain that
is controlled by a dc voltage or, more commonly, a digital input. This device is known as
a programmable gain amplifier, or PGA.
To understand the benefits of variable gain, assume an ideal PGA with two settings, gains
of 1 and 2. The dynamic range of the system is increased by 6 dB. Increasing the gain to
4 results in a 12 dB increase in dynamic range.
If the LSB of an ADC is equivalent to 10 mV of input voltage, the ADC cannot resolve
smaller signals, but when the gain of the PGA is increased to 2, input signals of 5 mV
may be resolved. Thus, the processor can combine PGA gain information with the digital
output of the ADC to increase its resolution by one bit. Essentially, this is the same as
adding additional resolution to the ADC.
GAIN
CONTROL
TRANSDUCER
PGA
ƒ
ƒ
DIGITAL
OUT
ADC
Used to Increase Dynamic Range of Circuit
A PGA With a Gain from 1 to 2 Theoretically
Increases the Dynamic Range by 6dB, A Gain of 1 to
4 Gives 12dB Increase, etc.
Figure 2.82: Programmable Gain Amplifier (PGA)
In practice, PGAs are not ideal, and their error sources must be studied. The most
fundamental problem with PGA design is accurate gain programming. Electromechanical
relays have minimal RON, but are otherwise unsuitable for gain switching. They are
slow, large, and expensive. Silicon switches, as discussed in the section on switches and
2.87
BASIC LINEAR DESIGN
multiplexers (Chapter 7 of this book), have quite large RON, which is both voltage- and
temperature-variable, and stray capacities, which may affect the ac parameters of a PGA
using them.
To understand how RON can affect the performance of a PGA, let us consider a poor
PGA design (Figure 2.83). An op amp is configured in the standard noninverting gain
circuit with four different gain setting resistors, each grounded by a switch. Most silicon
switches have on resistance in the range of 100 Ω to 500 Ω. Even if the on resistance
were as low as 25 Ω, the error for a gain of 16 would be 2.4%, much worse than 8-bits.
Furthermore, RON drifts over temperature, and varies from switch to switch. If the value
of the feedback and gain setting resistors were increased, noise and offset would become
a problem. The only way to achieve accuracy with this circuit is to replace silicon
switches with relays which have virtually no on resistance.
RF
10kΩ
625Ω
1.43kΩ
3.33kΩ
10kΩ
-
G = 16
G=8
G=4
VOUT
G=2
+
VIN
„ Gain Accuracy Limited by Switch’s On Resistance, R and
Ron Modulation
„ Ron Typically 100 - 500Ω for a CMOS Or JFET Switch
„ Even With Ron= 25Ω,There is a 2.4% Gain Error for AV = 16
„ Ron Drift Over Temperature Limits Accuracy
„ Only Solution is to Use Very Low Ron Switches (Relays)
Figure 2.83: How Not to Build a PGA
It is better to use a circuit where RON is unimportant. In Figure 2.84, the switch is placed
in series with the inverting input of an op amp. Since the input impedance of an op amp is
very large, the RON of the switch is irrelevant. The gain is now determined by the
external resistors. The RON may add a small offset error if the op amp bias current is
significant.
2.88
OTHER LINEAR CIRCUITS
PROGRAMMABLE GAIN AMPLIFIERS
VIN
+
VOUT
G=1
500Ω
G=2
1kΩ
1kΩ
„ Ron is Not in Series With Gain Setting Resistors
„ Ron is Very Small Compared to Input Impedance
„ Only a Slight Offset Error Occurs Due to the Bias
Current Flowing Through the Switch
Figure 2.84: Alternative PGA Configuration Negates Effect of Ron
AMPLIFIER
+VS
C1
C2
VIN
OUT
FORCE
N1
N2
-VS
A0
A1
A2
B
CLK
CS
L
A
T
C
H
E
S
C
O
N
T
R
O
L
OUT
SENSE
14k
G=8
3.4k RESISTOR
NETWORK
G=2
L
O
G
I
C
1k
G = 16
1.7k
G=4
DIGITAL
GND
1k
ANALOG
GND2
1.7k
ANALOG
GND1
Figure 2.85: Monolithic Software Programmable PGA Instrumentation Amplifier
(AD526)
2.89
BASIC LINEAR DESIGN
The AD526 amplifier uses this method of building a PGA and integrates it onto a single
chip. The AD526 has five binary gain settings from 1 to 16, and its internal JFET
switches are connected to the inverting input of the amplifier. The gain resistors are laser
trimmed. The maximum gain error is only 0.02%, far better than the 2.4% error in Figure
2.85. The linearity is also very good at 0.001%. The AD526 is controlled by a latched
digital interface.
This same design can be used to build the discrete PGA shown in Figure 2.86. It uses a
single op amp, a quad switch, and precision resistors. The low-noise AD797 replaces the
JFET input op amp of the AD526, but almost any voltage feedback op amp could be used
in this circuit. The ADG412 was picked for its low on resistance of 35 Ω. The resistors
were chosen to give gains of 1, 10, 100, and 1000, but if other gains are required, the
resistor values may easily be altered. Ideally, a trimmed resistor network should be used
both for initial gain accuracy and for low drift over temperature. The 20 pF capacitor
ensures stability and holds the output voltage when the gain is switched. The control
signal to the switches turns one switch off a few nanoseconds before the second switch
turns on. During this break, the op amp is open-loop. If the capacitor was not used, the
output would start slewing. Instead, the capacitor holds the output voltage during the
switching. Since the time that both switches are open is very short, only 20 pF is needed.
For slower switches, a larger capacitor may be necessary.
Figure 2.86: A Very Low Noise PGA
2.90
OTHER LINEAR CIRCUITS
PROGRAMMABLE GAIN AMPLIFIERS
The PGA’s input voltage noise spectral density is only 1.65 nV/√Hz at 1 kHz, only
slightly higher than the noise performance of the AD797 alone. The increase is due to the
noise of the ADG412, and the current noise of the AD797 flowing through the on
resistance. The noise was measured at a gain of 1000 (worst case).
The accuracy of the PGA is important in determining the overall accuracy of a system.
The AD797 has a bias current of 0.9 µA, which, flowing in 35 Ω RON, results in an
additional offset error of 31.5 µV. Combined with the AD797 offset, the total Vos
becomes 71.5 µV (max). Offset temperature drift is affected by the change in bias current
and on resistance. Calculations show that the total temperature coefficient increases from
0.6 µV/ºC to 1.6 µV/ºC. These errors are small, and may not matter, but it is important to
be aware of them. In practice, circuit accuracy and TC will be determined by the external
resistors. Input characteristics such as common-mode range and input bias current are
determined solely by the AD797. The circuit could be converted to single supply simply
by changing the op amp. The switches do not need to be changed.
Another PGA configuration uses a DAC in the feedback loop of an op amp to adjust the
gain under digital control (Figure 2.87). The digital code of the DAC controls its
attenuation. Attenuating the feedback signal increases the closed-loop gain. A
noninverting PGA of this type requires a multiplying DAC with a voltage output (a
multiplying DAC is a DAC with a wide reference voltage range which includes zero). For
most applications of the PGA, the reference input must be capable of handling bipolar
signals. The AD7846 is a 16-bit converter that meets these requirements. In this
application, it is used in standard 2-quadrant multiplying mode. The OP-213 is a low
drift, low noise amplifier, but the choice of amplifier is flexible, and depends on the
application. The input voltage range depends on the output swing of the AD7846, which
is 3 V less than the positive supply, and 4 V above the negative supply. A 1000 pF
capacitor is used in the feedback loop for stability.
.Figure 2.87: Using a Multiplying DAC in a Feedback Loop to Create a Divider
2.91
BASIC LINEAR DESIGN
The gain of the circuit is set by adjusting the digital inputs of the DAC, according to the
equation given in Figure 2.87. D0–15 represents the decimal value of the digital code.
For example, if all the bits were set high, the gain would be 65,536/65,535 = 1.000015. If
the eight least significant bits are set high and the rest low, the gain would be 65,536/255
= 257.
Figure 2.88 shows the small signal response at a gain of 1 with a 100 mV square wave
input. The bandwidth is a fairly high 4 MHz. However, this does reduce with gain, and
for a gain of 256, the bandwidth is only 600 Hz. If the gain-bandwidth product were
constant, the bandwidth in a gain of 256 should be 15.6 kHz; but the internal capacitance
of the DAC reduces the bandwidth to 600 Hz.
Figure 2.88: Performance of the Circuit in Figure 2.87
The gain accuracy of the circuit is determined by the resolution of the DAC and the gain
setting. At a gain of 1, all bits are on, and the accuracy is determined by the DNL
specification of the DAC, which is ±1 LSB maximum. Thus, the gain accuracy is
equivalent to 1 LSB in a 16-bit system, or 0.003%. However, as the gain is increased,
fewer of the bits are on. For a gain of 256, only Bit 8 is turned on. The gain accuracy is
still dependent on the ±1 LSB of DNL, but now that is compared to only the lowest eight
bits. Thus, the gain accuracy is reduced to 1 LSB in an 8-bit system, or 0.4%. If the gain
is increased above 256, the gain accuracy is reduced further. The designer must
determine an acceptable level of accuracy. In this particular circuit, the gain was limited
to 256.
Noninverting PGA circuits using an op amp are easily adaptable to single-supply
operation, but the instrumentation amplifier topology does not lend itself to single-supply
applications. However, the AMP-04 can be used with an external switch to produce the
single supply instrumentation PGA shown in Figure 2.85. This circuit has selectable
gains of 1, 10, 100, and 500, which are controlled by an ADG511. The ADG511 was
2.92
OTHER LINEAR CIRCUITS
PROGRAMMABLE GAIN AMPLIFIERS
chosen as a single-supply switch with a low RON of 45 Ω. The gain of this circuit is
dependent on the RON of the switches. Trimming is required at the higher gains to
achieve accuracy. At a gain of 500, two switches are used in parallel, but their resistance
causes a 10% gain error in the absence of adjustment.
Figure 2.89: Single-Supply Instrumentation Amp PGA
Certain Σ-Δ ADCs (AD7710, AD7711, AD7712 and AD7713, for example) have built-in
PGAs. Circuit design is much easier because an external PGA and its control logic are
not needed. Furthermore, all the errors of the PGA are included in the specifications of
the ADC, making error calculations simple. The PGA gain is controlled over the same
serial interface as the ADC, and the gain setting is factored into the conversion, saving
additional calculations to determine input voltage. This combination of ADC and PGA is
very powerful and enables the realization of a highly accurate system, with a minimum of
circuit design. The PGA function in this case is not a separate block requiring matching
of resistors for accuracy in line with the expectation of the Σ-Δ ADC. It is accomplished
by modulating the duty cycle of the switched capacitors in the modulator, thus changing
the gain.
High speed VGAs are discussed in the RF/IF section (Section 4.6)
2.93
BASIC LINEAR DESIGN
Figure 2.90: PGA built into a Σ-Δ ADC
2.94
OTHER LINEAR CIRCUITS
AUDIO APPLICATIONS
SECTION 2.14: AUDIO APPLICATIONS
Amplifiers
There are no specific audio specifications that apply to amplifiers. Obviously the
amplifier needs to be of appropriate bandwidth and be low distortion. Several types have
found application in the audio field. These include the AD797, OP275 and the AD711/
AD712/ AD713.
One application specific audio IC is the SSM2019 microphone preamplifier (see Figure
2.91). For use as a microphone preamplifier in high fidelity applications, a primary
concern is that the circuit be low noise. The specification for the SSM2019 is 1 nV/√Hz.
The input to the SSM2019 is fully differential to interface with balanced microphones.
Figure 2.91: SSM2019 Microphone Amplifier
There is another application for microphone preamplifiers. Here the emphasis is on voice
intelligibility rather than low noise. The target application is in communications system
and public address systems. The SSM2165/SSM2166/SSM2167 family is a complete and
flexible solution for conditioning microphone inputs. A low noise voltage controlled
amplifier (VCA) provides a gain that is dynamically adjusted by a control loop to
maintain a set compression characteristic. The compression ratio is set by a single resistor
and can be varied from 1:1 to over 15:1 relative to the fixed rotation point. Signals above
the rotation point are limited to prevent overload and to eliminate “popping.” A
downward expander (noise gate) prevents amplification of noise or hum. This results in
2.95
BASIC LINEAR DESIGN
optimized signal levels prior to digitization, thereby eliminating the need for additional
gain or attenuation in the digital domain that could add noise or impair accuracy of
speech recognition algorithms. The block diagram of the SSM2165 is shown in Figure
2.92.
Figure 2.92: Block Diagram of the SSM2165
0
OUT PUT (d Bu )
–10
–20
–30
–40
–50
–60
–80
–70
–60
–50
–40
–30
–20
–10
INPUT (d Bu )
Figure 2.93: Typical Transfer Characteristics for the SSM2165
2.96
OTHER LINEAR CIRCUITS
AUDIO APPLICATIONS
Speaker driver power amplifiers are another application specific audio area. The main
application challenge here is delivering enough audio power in a limited supply voltage
environment such as is typically found in computers and games, while keeping the
package power dissipation down to safe levels. As an example, the SSM2211
(Figure 2.94) is a high performance audio amplifier that delivers 1 W rms of low
distortion audio power into a bridge-connected 8 Ω speaker load, (or 1.5 W rms into 4 Ω
load). The SSM2211 is available in SO-8 and LFCSP (lead frame chip scale package)
surface-mount packages. The SO-8 features the patented Thermal Coastline lead frame.
The thermal coastline package is further discussed in the power section.
Figure 2.94: SSM2211 Typical application
10
10
TA = 25ºC
VDD = 5V
AVD = 2 (BTL)
RL = 8Ω
PL = 1W
CB = 0
1
THD + N (%)
THD + N (%)
1
CB = 0.1μF
0.1
0.01
0.1
CB = 1μF
20
TA = 25°C
VDD = 5V
AVD = 2 (BTL)
RL = 8Ω
FREQUENCY = 20kHz
CB = 0.1μF
100
1k
10k
20k
0.01
20n
FREQUENCY (Hz)
0.1
1
POUTPUT (W)
Figure 2.95: SSM2211 typical performance
2.97
2
BASIC LINEAR DESIGN
VCAs (Voltage Controlled Amplifiers)
Audio signal levels are often controlled by using low distortion VCAs (voltage controlled
amplifiers) in the signal path. By using controlled rate-of-change drive to the VCAs, the
“clicking” associated with switched resistive networks is eliminated. For example, the
SSM2018T is a low noise, low distortion VCA applicable in high performance audio
systems. The “T’ suffix indicates a version that is factory trimmed for distortion and
requires no subsequent user adjustments are required.
Figure 2.96: SSM2018 Block Diagram
Figure 2.97: The Distortion Characteristics of the SSM2018
2.98
OTHER LINEAR CIRCUITS
AUDIO APPLICATIONS
V+
V–
POWER
SUPPLY AND
REFERENCE
GENERATOR
CH1 IN
VC A
VREF
CH1 OUT
5-BIT
CHANNEL
DAC
CH2 IN
VC A
CH2 OUT
5-BIT
CHANNEL
DAC
CH3 IN
VC A
CH3 OUT
5-BIT
CHANNEL
DAC
CH4 IN
CH4 OUT
VC A
5-BIT
CHANNEL
DAC
CH5 IN
CH5 OUT
VC A
5-BIT
CHANNEL
DAC
CH6 IN
VC A
CH6 OUT
5-BIT
CHANNEL
DAC
7-BIT
MASTER
DAC
CLK
DATA
LD
SHIFT REGISTER
AND
ADDRESS
DECODER
CH SET
STEP SIZ E
ADJUST
MSTR SET
MSTR OUT
SSM2160
WRITE
Figure 2.98: SSM2160 block diagram
2.99
BASIC LINEAR DESIGN
Despite the sonic advantages of using analog control of the signal level, it is sometimes
useful to have the control voltage under digital control. In this case a DAC can be added
to the VCA. An example of this configuration is the SSM2160 which allows digital
control of volume of six audio channels, with a master level control and individual
channel controls. Low distortion VCAs (voltage controlled amplifiers) are used in the
signal path. Each channel is controlled by a dedicated 5-bit DAC providing 32 levels of
gain. A master 7-bit DAC feeds every control port giving 128 levels of attenuation. Step
sizes are nominally 1 dB and can be changed by external resistors. Channel balance is
maintained over the entire master control range. Upon power-up, all outputs are
automatically muted. A 3-wire or 4-wire serial data bus enables interfacing with most
popular microcontrollers.
Line Drivers and Receivers
The function of sending/receiving audio signals between various system components has
traditionally involved trade-offs of one form or another. Fully differential or balanced
transmission systems are best at rejecting low frequency and RF noise, so they are used
for highest performance, and are discussed in some detail following.
A typical audio system block diagram using differential or balanced transmission is
shown in Figure 2.99. In concept, a balanced transmission system like this could use
several input/output coupling schemes within the driver and receiver. Some major points
distinguishing coupling method details are discussed briefly below, before addressing
actual circuits.
Figure 2.99: An Audio Balanced Transmission System
A point worth noting is that the ± voltage drive to the line need not be exactly balanced to
reap the benefits of balanced transmission. In fact the drive can be asymmetrical to some
degree, and the signal will still be received at VOUT with correct amplitude, and with good
noise rejection. What does need to be provided is two well-balanced line-driving
impedances, RO1 and RO2. Also, in conjunction with these balanced drive impedances, the
2.100
OTHER LINEAR CIRCUITS
AUDIO APPLICATIONS
associated (+) and (–) receiver input impedances should also be equal. The technical
reasons for this will be apparent shortly.
Audio Line Receivers
An audio line receiver is simply a subtractor amplifier (Figure 2.100). From a dc and ac
trim/balance perspective, the Figure 2.100 topology is most effective with resistors and
amplifiers made simultaneously in a single monolithic IC.
In applying circuits of the Figure 2.100 type (or other topologies which resistively load
the source), a designer must bear in mind that all external resistances added to the four
resistances can potentially degrade CMR, unless kept to proportional value increases. To
place this in perspective, a 2.5 Ω or 0.01% mismatch can easily occur with wiring, and if
not balanced out, this mismatch will degrade the CMR of otherwise perfectly matched
25 kΩ resistors to 86 dB. These circuits are therefore best fed from balanced, low
impedance drive sources, preferably 25 Ω or less.
The SSM2141 and SSM2143 are monolithic IC line receivers which work very much like
the circuit of Figure 2.101 differing only in their individual gains. The SSM2141 operates
as a unity gain device, while the SSM2143 operates either at a nominal gain of 0.5
(−6 dB), or it can optionally be strapped with the input/output of the resistor pairs
reversed, to operate at a gain of 2 (6 dB).
Figure 2.100: A Simple Line Receiver Using a 4-Resistor Differential Amplifier
Both devices operate from supplies up to ±18 V, can drive 600 Ω loads, and they have
low distortion and excellent CMR characteristics. For reference, the op amp used in these
receivers is similar to one half of an OP271. The output appears at Pin 6 and is
uncommitted, with conventional use it gets tied to R4 (Pin 5) for feedback. However, if
2.101
BASIC LINEAR DESIGN
desired, an external in-loop buffer can optionally be added. This step will allow either
line receiver device to drive even lower Z loads if desired.
Perhaps the most outstanding attribute of these devices is their CMR performance, shown
in Figure 2.101(a) (this data are for the SSM2141, but the SSM2143 is similar). For the
SSM2141 the dc-to-1 kHz CMR is typically 100 dB, and even at 10 kHz it is still about
80 dB. The SSM2143 (not shown), using lower resistor values, has a somewhat lower
typical CMR of 90 dB, but maintains this to about 10 kHz. The SSM2141 THD + N
performance also shown in Figure 2.101(b) is also very good for both 600 Ω and 100 kΩ
loads.
Figure 2.101: SSM2143 Common-Mode Rejection and THD
With a companion differential line driver (next section), these two line receivers allow
convenient as well as flexible interfacing between points in audio systems, as well as
other instrumentation up to 100 kHz. However, they both are also more generally useful
as flexible gain blocks within a system, not necessarily requiring the full CM
performance aspects. For example, they are useful as either precise inverting or
noninverting gains blocks, due to the very accurate internal resistor ratios. With the
SSM2141 typical gain accuracy of 0.001%, very precise, single chip unity gain inverters
and summers can be built at low overall cost.
Audio Line Drivers
Unlike the case for the differential line receiver, a standard circuit topology for
differential line drivers is not quite as clear-cut. Two circuit types are discussed in this
section, with their contrasts in performance and complexity.
2.102
OTHER LINEAR CIRCUITS
AUDIO APPLICATIONS
On the other hand, the inherent features of laser trimmed monolithic technology can
make a complex circuit such as the balanced line driver thoroughly practical. Like the
SSM2141 and SSM2143 line receivers, applying these concepts to a driver circuit results
in an efficient and useful IC. This product, the SSM2142 balanced line driver, is shown in
functional form in Figure 2.102.
FUNCTIONAL DIAGRAM
VIN
50Ω
+OUT FORCE
MINI-DIP PACKAGE
+OUT SEN SE
– FORCE 1
8 + FORCE
– SENSE 2
7 + SENSE
GROUND 3
6 +V
VIN 4
5 –V
10kΩ
– OUT SENSE
SSM2142
50Ω
ALL RESISTORS 30kΩ UNLESS
OTHERWISE INDICATED
– OUT FORCE
10k Ω
GND
Figure 2.102 SSM2142 Block Diagram
SSM2142
Figure 2.103: Balanced Audio Transmission System
2.103
BASIC LINEAR DESIGN
The SSM2142 is designed for a single-ended to differential gain of 2 times, and in use
can be simply strapped with the respective FORCE/SENSE pins tied together. In a
system application, the SSM2142 is used with either an SSM2143 or an SSM2141 line
receiver, with the differential mode signal being transmitted via shielded twisted pair
cable. This hookup comprises a complete single-ended to differential and back to singleended transmission system, with noise isolation in the process
With the use of the SSM2143 gain of 0.5, the SSM2142 gain of 2 is complemented, and
the overall system gain is unity. If the SSM2141 is used as the receiver, the gain is 2
overall. The THD + N performance of the unity gain SSM2142/SSM2143 system is
shown in Figure 9.104, for the conditions of a 5 V rms input/output signal, both
with/without a 500' cable.
As should be obvious, these drivers do not offer galvanic isolation, which means that in
all applications there must be a dc current path between the grounds of the driver and the
final receiver. In practice however this isn’t necessarily a problem.
Figure 2.104: Balanced Audio Transmission System Performance
2.104
OTHER LINEAR CIRCUITS
AUDIO APPLICATIONS
Class-D Audio Power Amplifiers
Theory of Operation
A Class-D audio amplifier is basically a switch-mode or PWM (pulse width modulated)
amplifier and is one of a number of different classes of amplifiers. Following is a look at
the definitions for the main classifications:
Class-A – In a Class-A amplifier, the output device(s) are continuously conducting for
the entire cycle, or in other words there is always bias current flowing in the output
devices. This topology has the least distortion and is the most linear, but at the same time
is the least efficient, at around 20%. Therefore, the quiescent dissipation is high. In fact
the dissipation is constant, regardless of how much power is delivered to the load. A
Class-A amplifier output is typically not complementary, with a high and low side output
device(s).
Class-B – In a Class-B amplifier the output device(s) only conduct for half the sinusoidal
cycle (one conducts for the positive half cycle, and one conducts for the negative half
cycle). If there is no signal, then there is no current flow in the output devices. This class
of amplifier is obviously more efficient than Class-A, at about 50%, but has some
distortion at the crossover point due to the time it takes to turn one device off and turn the
other device on. This is referred to as crossover notch distortion. Since it occurs at the
point of minimum signal (the zero crossing), its effect is very obvious.
+VS
+VS
NPN
+VS
NPN
NPN
+
_
+
_
PNP
–VS
(A)
PNP
–VS
(B)
–VS
(C)
Figure 2.105: Example output stages
(A) Class-A, (B) Class-B, and (C) Class-AB
Class-AB – This type of amplifier is a combination of the above two types, and is
probably the most common type of power amplifier in existence. Here both devices are
allowed to conduct at the same time, but just by a small amount near the crossover point.
Hence each device is conducting for more than half a cycle but less than the whole cycle,
2.105
BASIC LINEAR DESIGN
so the inherent nonlinearity (crossover distortion) of Class-B designs is overcome,
without the inefficiencies of a Class-A design. Efficiencies for Class-AB amplifiers can
also be about 50%. There are variations of Class-AB, depending on how much of the
cycle both of the output devices conduct. Obviously, the more they both conduct, the
lower the efficiency, but also the higher the linearity.
Class-D – This class of amplifier is a switching amplifier as mentioned above. In this
type of amplifier, the switches are either fully on or fully off, significantly reducing the
power losses in the output devices. This is very similar to the difference between linear
power regulators and switch-mode regulators.
Efficiencies of 90% to 95% are possible. Audio Class-D amplifiers use a modulator to
convert the input audio signal into a switching waveform used to control the output
switches. Pulse-width modulation (PWM) is the most commonly used modulation
scheme. In PWM, the audio signal is used to modulate a PWM carrier signal which
drives the output devices. The output devices then drive a low-pass filter to remove the
high frequency PWM carrier frequency, while retaining the desired audio content. The
speaker is one of the elements in this filter, and is situated at the filter output.
Class-D amplifiers take on many different forms, some can have digital inputs and some
can have analog inputs.
However, audio quality in PWM amplifiers can be limited: THD is typically 0.1% or
worse, and PSRR is poor (see Reference 1). PSRR can be improved by sensing power
supply variations and adjusting the modulator’s behavior to compensate, as proposed in
(see Reference 1). However, this alone will not suppress the THD produced by inherent
PWM nonlinearity or power-stage non-linearity.
This THD and power supply noise can both be suppressed with feedback from the power
stage outputs (see Reference 2), which incorporates the feedback around an analog PWM
modulator.
PWM is attractive because it allows > 100 dB audio-band SNR at low clock rates near
400 kHz, limiting switching losses. Also, many PWM modulators are stable to near 100%
modulation, allowing high output power before overloading. However, PWM has several
problems. First, the PWM process inherently adds distortion in many modulation
schemes (see Reference 3) and second, harmonics of the PWM switching frequency
produce EMI in the AM band.
Σ∆ modulation does not share these problems, but nonetheless hasn’t traditionally been
used for Class-D (see reference 3), because conventional 1-bit Σ-∆ modulators are only
stable to 50% modulation, and power efficiency is limited because typical output data
rates are > 1 MHz, when ≥ 64x oversampling rate is needed to achieve sufficient audio
band SNR. However, Analog Devices has enhanced the traditional 1-bit Σ-Δ architecture
to overcome these problems, and created Σ-Δ-based Class-D amplifier chips which have
performance advantages over competitor PWM-based products.
2.106
OTHER LINEAR CIRCUITS
AUDIO APPLICATIONS
DEVICE ARCHITECTURE
The AD1990/AD1992/AD1994/AD1996 chips are 2-channel Bridge Tied Load (BTL)
switching audio power amplifiers with integrated Σ-∆ modulator. Hereafter, AD199x will
be used to refer to this product family.
The AD199x modulator accepts a low power analog input signal (of 5 V p-p maximum
amplitude), and generates a switching waveform to drive speakers directly. One of the
two modulators can control both output stages thereby providing twice the current for
single-channel applications. A digital, microcontroller-compatible interface provides
control of reset, mute, and PGA gain as well as output signals for thermal and overcurrent error conditions. The output stage can operate from supply voltages ranging from
8 V to 20 V. The analog modulator and digital logic operate from a 5 V supply.
The power stage of the AD199x is arranged internally as four transistor pairs, which are
used as two H-bridge outputs to provide stereo amplification. The transistor pairs are
driven by the output of the Σ-∆ modulator. A user selectable non-overlap time is provided
between the switching of the high side transistor and low side transistor to ensure that
both transistors are never on at the same time. The AD199x implements turn-on pop
suppression to eliminate any pops or clicks following a reset or un-mute.
Analog Input Section
The analog input section uses an internal amplifier to bias the input signal to the
reference level. A dc blocking capacitor should be connected to remove any external dc
bias contained in the input signal.
The Sigma-Delta Modulator
The modulator uses a 1-bit, seventh-order feedforward architecture. The quantizer output
drives the switching power stage, whose pulses are fed back to a continuous-time (CT)
first integrator. This allows fullest possible integration of the pulse waveform,
maximizing error correction. If the first integrator were discrete-time (DT), its sampling
process would often miss important information about errors in pulse edge timing and
shape, which would reduce the error-correcting effectiveness of the feedback loop.
The CT integrator bandwidth of 100 kHz gives antialias filtering for the subsequent DT,
switched-capacitor (SC) integrators. The SC integrators and quantizer are clocked at
6 MHz, corresponding to 128× oversampling.
For the modulator, seventh order is more than enough to achieve 100 dB SNR with
traditional aggressive noise shaping (see Reference 5). However, this gives instability for
modulation > 50%, limiting the maximum output power with stable operation to just 25%
of theoretical full power. To overcome this limitation, we use less aggressive noiseshaping to maintain stability to 90% modulation. This gives good sound quality at higher
power, but requires high modulator order to get acceptable SNR.
2.107
BASIC LINEAR DESIGN
Fortunately, all integrators after the first can be SC. The high first integrator gain relaxes
noise requirements for the SC ones, allowing small sampling caps (50 fF), and low power
single-stage op amps. Resonator feedbacks to integrators 2, 4, and 6 reduce lowfrequency noise, by placing NTF zeroes at 12 kHz, 22 kHz, and 40 kHz. When PVDD =
12 V, integrated audio-band quantization noise is 25 µV rms and additional thermal noise
yields total integrated audio noise of 50 µV rms. Maximum output is 7.8 V rms, giving
104 dB dynamic range.
The modulator described to this point would become unstable for large inputs > 90% fullscale. Output transients resulting from the instability would bear little relationship to the
desired signal, and would sound bad. To solve this problem, the modulator input is
monitored, and when large signals that could cause instability are detected, Integrators 3
to 7 of the modulator are reset. This effectively converts the modulator to a second order,
unconditionally stable configuration. The loop gain is reduced relative to the “normal”
seventh-order configuration, so that noise-shaping is less effective and more quantization
noise reaches the output. However, this elevated noise is superimposed on a large output
signal that is now closer to the desired waveform, and the composite sound is better than
when the modulator is allowed to become unstable.
Driving the H-Bridge
Each channel of the switching amplifier is controlled by a 4 transistor H-bridge to give a
differential output stage. The outputs of the H-bridges, OUTR+, OUTR-, OUTL+, and
OUTL- will switch between PVDD and PGND as determined by the sigma delta (Σ-∆)
modulator. The power supply that is used to drive the power stage of the AD199x should
be in the range of +8 V to +20 V and be capable of supplying enough current to drive the
load. This power supply is connected across the PVDD and PGND pins. The feedback
pins, NFR+, NFR-, NFL+, and NFL-, are used to supply negative feedback to the
modulator. The pins are connected to the outputs of the H-bridge using a resistor divider
network as shown in Figure 2.106.
External Schottky diodes can be used to reduce power loss during the nonoverlap time
when neither of the high-side or low-side transistors is on. During this time neither
transistor is driving the OUTx pin. The purpose of the inductors is to keep current
flowing.
For example the OUTx pin may approach and pass the PGND level to achieve this. When
the voltage at the OUTx pin is 0.7 V below PGND the parasitic diode associated with the
low side transistor will become forward biased and turn on. When the high side transistor
turns on the voltage at OUTx will rise to PVDD and will reverse bias the parasitic diode.
However, by its nature the parasitic diode has a long reverse recovery time and current
will continue to flow through it to PGND thus causing the entire circuit to draw more
current than necessary. The addition of the Schottky diodes prevents this happening.
When the OUTx pin goes more than 0.3 V below PGND the Schottky diode becomes
forward biased. When the high side transistor turns on the Schottky diode becomes
reverse biased. The reverse recovery time of the Schottky diode is significantly faster
than the parasitic diode so far less current is wasted. A similar effect happens when the
inductor induces a current which drives the OUTx pin above PVDD. Figure 2.106 shows
2.108
OTHER LINEAR CIRCUITS
AUDIO APPLICATIONS
how the external components of a system are connected to the pins of the AD199x to
form the H bridge configuration.
Figure 2.106: H-Bridge Configuration
AMPLIFIER GAIN
Selecting the Modulator Gain
The AD199x modulator can be thought of as a switching analog amplifier with a voltage
gain controlled by two external resistors forming a resistor divider between the OUTxx
pins and PGND. The centre of the resistor divider is connected to the appropriate
feedback pin NFx. Selecting the gain along with the PVDD Voltage will determine how
much power can be delivered to a load for a fixed input signal. The gain of the modulator
is controlled by the values of R1 and R2 (see Figure 2.106) according to the equation:
Gain = (R2 + R1)/R2
(Eq. 2-23)
If the voltage at the NFx pins exceeds 5 V, ESD protection circuitry will turn on, to
protect low voltage circuitry inside the chip that’s connected to NFx. When the protection
circuit is active it introduces nonlinear behavior into the modulator feedback loop, which
degrades audio quality. To avoid this, R1, R2, and the gain should be selected in a
manner that limits max voltage at NFx to < 5 V. For optimal modulator stability and
audio quality, use the formula:
Gain = (R1 + R2)/R2 = PVDD/3.635
(Eq. 2-24)
The ratio of the resistances sets the gain rather than the absolute values. However, the
dividers provide a path from the high voltage supply to ground, so the values should be
large enough to produce negligible loss due to quiescent current.
2.109
BASIC LINEAR DESIGN
The chip contains a calibration circuit to minimize voltage offsets at the speaker, which
helps to minimize clicks and pops when muting or un-muting. Optimal performance is
achieved for the offset calibration circuit when the feedback divider resistors sum to
6 kΩ. (Meaning that (R1+R2) = 6 kΩ).
.
Power-Up Considerations
Careful power-up of the AD100X is necessary to ensure correct operation and avoid
possible latch-up issues. The AD199x should be powered-up with RST/PDN and MUTE
held low until all the power supplies have stabilized. Once the supplies have stabilized
the AD199x can be brought out of reset by bringing RST/PDN high and then MUTE can
be brought high as required.
On/Off/Mute Pop Noise Suppression
The AD199x features pop suppression which is activated when the part is reset or taken
out of mute. The pop suppression is achieved by pulsing the power outputs to bring the
outputs of the LC filter from 0 V to midscale in a controlled fashion. This feature
eliminates unwanted transients on both the outputs and the high voltage power supply.
Thermal Protection
The AD199x features thermal protection. When the die temperature exceeds
approximately 135°C the Thermal Warning Error output (ERR1) is asserted. If the die
temperature exceeds approximately 150°C the Thermal Shutdown Error output (ERR2) is
asserted. If this occurs, the part shuts down to prevent damage. When the die temperature
drops below approximately 120°C both error outputs are negated and the part returns to
normal operation.
Over-Current Protection
The AD199x features over-current or short-circuit protection. If the current through any
power transistors exceeds 4 A the part goes into mute and the over-current error output
(ERR0) is asserted. This is a latched error and does not clear automatically. To clear the
error condition and restore normal operation, the part must be either reset, or MUTE must
be asserted and negated.
Good board layout and decoupling are vital for correct operation of the AD199x. Due to
the fact that the part switches high currents there is the potential for large PVDD bounce
each time a transistor switches. This can cause unpredictable operation of the part. To
avoid this potential problem, close chip decoupling is essential. It is also recommended
that the decoupling capacitors are placed on the same side of the board as the AD199x,
and connected directly to the PVDD and PGND pins. By placing the decoupling
capacitors on the other side of the board and decoupling through vias the effectiveness of
the decoupling is reduced. This is because vias have inductive properties and therefore
2.110
OTHER LINEAR CIRCUITS
AUDIO APPLICATIONS
Figure 2.107: Typical Stereo Mode Application Circuit
prevent very fast discharge of the decoupling capacitors. Best operation is achieved with
at least one decoupling capacitor on each side of the AD199x, or (optionally) two
2.111
BASIC LINEAR DESIGN
capacitors per side can be used to further reduce the series resistance of the capacitor. If
these decoupling recommendations cannot be followed and decoupling through vias is
the only option, the vias should be made as large as possible to increase surface area,
thereby reducing inductance and resistance.
Figure 2.108: AD199x Block diagram
Application Considerations
Audio Fidelity and EMI Reduction
The AD199x amplifiers deliver audiophile sound quality (THD <0.003%; SNR > 103 dB;
PSRR> 65 dB) with 50% lower heat dissipation than traditional linear amplifiers. The
THD performance is 40 dB better than typical open-loop competitors and 10 dB to 20 dB
better than most closed-loop ones. This breakthrough performance is achieved through
Analog Devices’ closed-loop, mixed-signal integration of seventh-order Σ-∆ modulator
technology with high power output drive circuitry and bridge circuitry. Radiated and
conducted out-of-band RF emissions are minimized with Analog Devices’ advanced
modulation techniques and closed-loop, Σ-∆ architecture to enable a significant reduction
in EMI.
Power levels range from stereo 5 W (mono 10 W) to stereo 40 W (mono 80 W). The
AD1994 can be configured in a modulator-only mode. This, coupled with external high
power FETs enables very high power amplification, limited only by the power stage
design. The parts also incorporate critical peripheral functions, including pop/click
suppression circuitry as well as short-circuit, overload, and temperature protection.
2.112
OTHER LINEAR CIRCUITS
AUDIO APPLICATIONS
THD+N for 1 kHz sine wave
Figure 2.109 and 2.110 shows FFTs measured with signals of 1 kHz at 1 µW and 1 W
output power levels. The 1 µW FFT (Figure 2.109) demonstrates that the noise floor is
tone-free for low-power inputs.
Figure 2.109: THD+N for a 1 kHz Sine Wave at 1 μW
The 1 W power level in Figure 2.110 is intended to represent a realistic listening level.
Harmonic distortion is evident, but the 0.00121% THD is an unprecedented level for this
signal condition in a single-chip Class-D amplifier.
Figure 2.110: THD+N for a 1 kHz Sine Wave at 1 W
2.113
BASIC LINEAR DESIGN
Figure 2.111 shows how THD varies with frequency, when the signal condition is a sine
wave of 1W output power.
Figure 2.111: 1 kHz Distortion vs. Frequency
Higher modulator loop gain at low frequencies enables better correction of modulator and
power stage errors, giving better THD than at higher frequencies. THD is actually
0.001% (−100 dB) or better up to hundreds of Hz. The apparent THD improvement at
audio frequencies above 6 kHz is a misleading artifact of the measurement setup, which
was unable to detect harmonics beyond its 20 kHz bandwidth. For 20 kHz fundamentals,
actual THD is near 0.01% (−80 dB), at ultrasonic, inaudible frequencies.
THD + N vs. Output Power, 1 kHz Sine
Figure 14 shows how THD + N varies with output power, for 1 kHz sine waves. There
are two curves in the plot. The first (o) is for a low power application where PVDD = 12
V and the load is 6 Ω (the default measurement setup). The second (x) is for a higher
power application where PVDD = 20 V and the load is 4 Ω.
In these curves, there are three distinct regions of performance. The first is at the lowest
output power levels, where the modulator is seventh order, and THD + N is best. The
second performance region is at significant output power, when the modulator order is
lowered from seventh to second to prevent instability. The second-order configuration
only allows 65 dB THD + N , because quantization noise is now elevated due to the
lower modulator order. However, this higher noise is difficult to hear above the loud,
energetic output. The third performance region is at highest output powers, where
clipping occurs, and distortion associated with clipping causes THD to degrade rapidly.
2.114
OTHER LINEAR CIRCUITS
AUDIO APPLICATIONS
THD+N vs. output power, 1kHz sine, meas. at LC filter output
THD+N (dB, relative to fundamental)
-10
-20
-30
(x): 4 ohm load, PVDD=20V
(o): 6 ohm load, PVDD=12V
-40
-50
-60
-70
-80
-90
-100 -1
10
0
10
Output power (W)
1
10
Figure 2.112: 1 kHz Distortion vs. Output Power
.
IMD
Figure 2.113 shows the intermodulation distortion (IMD) resulting from a 1 W 19 kHz
and 20 kHz twin-tone stimulus. The 1 kHz second-order product is approximately 98 dB
below the tones.
Figure 2.113: Intermodulation distortion (IMD)
2.115
BASIC LINEAR DESIGN
Crosstalk
Crosstalk between channels is a concern in chips with multiple audio channels. To
investigate crosstalk, we drove one channel of the chip with a 1 kHz, 1W (+7.8 dBV) sine
wave, while leaving the other channel idle (0 input). We then measured the idle channel:
results are shown in figure 2.114. The −89 dBV 1 kHz tone in the idle channel is 97 dB
below the driven channel’s signal.
Figure 2.114: Crosstalk
Power efficiency
Figure 2.115 shows power efficiency up to 5 W output power. The 50 mW/channel
modulator power consumption and power stage consumption are both included in this
calculation. (If we included only the power stage consumption but excluded the
modulator, as is sometimes done, the efficiency number would improve).
2.116
OTHER LINEAR CIRCUITS
AUDIO APPLICATIONS
Figure 2.115: Efficiency vs. Output Power
2.117
BASIC LINEAR DESIGN
Notes:
2.118
OTHER LINEAR CIRCUITS
AUTO-ZERO AMPLIFIERS
SECTION 2.15: AUTO-ZERO AMPLIFIERS
Chopper Amplifiers
Chopper type amplifier topologies have existed for decades. Initial chopper designs
actually involved switching the ac coupled input signal and synchronous demodulation of
the ac signal to re-establish the dc signal. See Figure 2.116. While these amplifiers
achieved very low offset, low offset drift, and very high gain, they had limited bandwidth
(it is a sampled system after all) and required filtering to remove the large ripple voltages
generated by the chopping action. In the earliest implementations the chopping switches
were actually relays, commonly switching on the order of 400 Hz.
CHOPPER
SWITCH
DRIVER
VIN
R1
R2
S
Z
C1
C2
S = SAMPLE
Z = AUTO-ZERO
C3
AMP
S
R3
VOUT
Z
C4
RL
Figure 2.116: Classic Chopper Amplifier Simplified Schematic
Virtually all modern IC chopper amplifiers actually use an auto-zero approach utilizing a
two (or more) stage composite amplifier structure similar to the chopper-stabilized
scheme. See Figure 2.117. One stage provides nulling action, while the other provides
wideband response. Together, the two stages provide very high voltage gain as they are
connected in series.
Chopper-stabilized amplifiers solved the bandwidth limitations of the classic
implementation by combining the chopper amplifier (used as a stabilizing amplifier) with
a conventional wideband amplifier that remained in the signal path. Since the main
signal path is not sampled, the bandwidth of the system is determined by the bandwidth
of the signal amplifier. It can exceed the chopping frequency.
2.119
BASIC LINEAR DESIGN
These chopper stabilized designs are capable of inverting operation only since the
stabilizing amplifier is connected to the non-inverting input of the wideband amplifier.
_
–IN
VOUT
A1
NULL
+
+IN
S
Z
C1
S
C2
Z
_
S = SAMPLE
Z = AUTO-ZERO
NULL
A2
+
Figure 2.117: Auto-Zero Amplifier Simplified Schematic
In this approach, the inputs of the nulling stage are shorted together during the first phase
of the operational cycle. During this nulling phase, amplified feedback is used to virtually
eliminate the offset of the nulling stage. The feedback voltage is impressed on a storage
capacitor so that during the second, or “output,” phase the offset remains nulled while the
inputs are now connected to the signal of interest.
In the output phase, the nulled input stage and the wideband stage in series amplify the
signal. The output of the nulled stage is impressed on a storage capacitor so that when the
cycle returns to the nulling phase (inputs shorted together), the output continues to reflect
the last input voltage value. Higher frequency signals bypass the nulling stage through
feed-forward techniques, making wide bandwidth operation possible.
While this technique provides dc accuracy and better frequency response, along with the
flexibility of inverting and noninverting configurations, it is prone to high levels of digital
switching noise that may limit the usefulness of the wider bandwidth.
2.120
OTHER LINEAR CIRCUITS
AUTO-ZERO AMPLIFIERS
Auto-Zero Amplifiers Improves on Choppers
ADI’s auto-zero amplifiers use a similar architecture with some major improvements.
Dual nulling loops, special switching logic and innovative compensation techniques
result in dynamic performance improvements while minimizing total die area. The
result—amplifiers that retain the high gain and dc precision of the auto-zero approach
while minimizing the negative effects of digital switching on the analog signal—at half
the cost. Typical offset voltage is under 1 µV and the offset drift is less than 10 nV/°C.
Voltage gain is more than 10 million, while PSRR and CMRR are well above 120 dB.
Input voltage noise is only 1 µV p-p from dc to 10 Hz.
Many auto-zero amplifiers are plagued by long overload recovery times due to the
complicated settling behavior of the internal nulling loops after saturation of the outputs.
Analog Devices auto-zero amplifiers have been designed so that internal settling occurs
within one or two clock cycles after output saturation occurs. The result is that the
overload recovery time is more than an order of magnitude shorter than previous designs
and is comparable to conventional amplifiers.
The careful design and layout of the AD855x amplifiers reduces digital clock noise and
aliasing effects by as much as 40 dB versus older designs.
In many cases the bandwidth required by the applications is such that the small amount of
digital feedthrough can be eliminated by filtering. Output filtering is also useful in
limiting the broadband noise of the signal amplifier.
The AD857x reduces the effects of digital switching on the analog signal by using a
patented digital spread-spectrum technique. As can be seen from Figures 2.118 and
2.119, the AD857x virtually eliminates the energy spike seen in other auto-zero
amplifiers at the switching frequency. It also reduces aliasing products between the
chopping clock and the input signal to the noise floor. The only penalty for this
breakthrough performance is a slight increase in voltage noise from the industry-best
1 µV p-p from dc to 10 Hz. of the AD855x design.
2.121
BASIC LINEAR DESIGN
AD8551/52/54
FIXED CHOPPING FREQUENCY:
4kHz
AD8571/72/74
PSEUDORANDOM CHOPPING FREQ:
2kHz - 4kHz
VS = +5V
G = 60dB
VS = +5V
G = 60dB
INPUT SIGNAL = 1mV RMS, 200Hz
OUTPUT SIGNAL: 1V RMS, 200Hz
GAIN = 60dB
Figure 2.118: Output Spectrum of Auto-Zero Amplifiers with Fixed Frequency
and Spread Spectrum Chopping
AD8551/52/54
FIXED CHOPPING FREQUENCY:
4kHz
VS = +5V
RS = 0Ω
AD8571/72/74
PSEUDORANDOM CHOPPING FREQUENCY
2kHz - 4kHz
VS = +5V
RS = 0Ω
Figure 2.119: Output Voltage of Auto-Zero Amplifiers with Fixed Frequency and
Spread Spectrum Chopping
2.122
OTHER LINEAR CIRCUITS
AUTO-ZERO AMPLIFIERS
Implementation
The actual circuit implementation of an IC auto-zero amplifier is much more complicated
than the simplified version described above. Multiple nulling loops are combined with
innovative compensation and signal paths are fully differential. Internal voltages are
controlled carefully to prevent saturation of the nulling circuitry. In addition, special logic
designs are utilized and careful layout is required to minimize parasitic effects. These
techniques result in stable, reliable operation and minimize unwanted digital interaction
with the analog signals.
The frequency response of the nulling and wideband amplifiers is carefully tailored so
that low frequency errors (dc circuit offsets and low frequency noise) are nulled while
high frequency signals are amplified as in a conventional op amp. This nulling of low
frequency errors has an important consequence for voltage noise. The very low frequency
1/f noise behavior seen in conventional amplifiers is not present in auto-zero amplifiers.
For applications with long measurement times on slowly varying signals, the noise
performance is better than the best low noise conventional amplifier designs.
In this IC implementation, the size of the on-chip storage capacitors is limited to achieve
a cost-effective die size. The small storage capacitors require careful attention to the
switch design and layout so that charge injection effects do not create large offset errors.
Switch leakage must also be minimized to maintain circuit accuracy, especially at high
temperatures. In the AD855x and AD857x amplifiers, the switches have been optimized
for accurate operation up to +125°C
30
80
Bipolar: OP177
Chopper: AD8571/72/74
25
70
1/F CORNER
FC = 0.7Hz
20
60
15
50
vnw (WHITE)
10
5
40
30
0.1
1
10
FREQUENCY (Hz)
NOISE BW
0.1Hz to 10Hz
0.01Hz to 1Hz
0.001Hz to 0.1Hz
0.0001Hz to 0.01Hz
100
0.01
BIPOLAR (OP177)
0.238µV p-p
0.135µV p-p
0.120µV p-p
0.118µV p-p
0.1
1
FREQUENCY (Hz)
10
CHOPPER (AD8571/72/74)
1.3 µV p-p
0.41µV p-p
0.130µV p-p
0.042µV p-p
Figure 2.120: Noise Comparison between Conventional Precision Amplifiers
and Chopper Stabilized Op Amps
2.123
BASIC LINEAR DESIGN
Operation Description
The simplified circuit (Figure 2.121) consists of the nulling amplifier (AA), the wideband
amplifier (AB), storage capacitors (CM1 and CM2) and switches for the inputs and storage
capacitors. There are two phases (A and B) per clock cycle.
In Phase A, the auto-zero phase, the nulling amplifier auto-zeros itself while the
wideband amplifier amplifies the input signal directly. The inputs of the nulling amp are
shorted together and to the inverting input terminal (common-mode input voltage). The
nulling amplifier nulls its inherent offset voltage through its nulling terminal gain (-BA).
The nulling voltage is also impressed on CM1. The signal at the input terminals is
amplified directly by the wideband amplifier.
Figure 2.121: Auto-Zero Amplifier, Auto-Zero Phase
Figure 2.122: Auto-Zero Amplifier, Output Phase
2.124
OTHER LINEAR CIRCUITS
AUTO-ZERO AMPLIFIERS
In phase B, the output phase, both amplifiers amplify the input signal. The inputs of the
nulling amplifier are connected to the input terminals. The nulling voltage of the nulling
amplifier is now stored on capacitor CM1 and continues to minimize its output offset
voltage. The instantaneous input signal is amplified by the nulling amplifier into the
wideband amplifier through the wideband amplifier nulling terminal gain (BB). The
output voltage of the nulling amplifier is also impressed on storage capacitor CM2. The
total amplifier gain is approximately equal to the product of the nulling amplifier gain
and the wideband amplifier gain. The total offset voltage is approximately equal to the
sum of the nulling amplifier and wideband amplifier offset voltages divided by the gain
of the wideband amplifier nulling terminal. By making this gain very large, the total
amplifier effective offset voltage becomes very small.
Both VOSA and VOSB are high-pass filtered “corner frequency” of high-pass filter set by
chopping frequency.
As the cycle returns to the nulling phase, the stored voltage on CM2 continues to
effectively correct the dc offset of the composite amplifier. The cycle from nulling to
output phase is repeated continuously at a rate set by the internal clock and logic circuits.
This model circuit, while simplified from the actual design, accurately depicts the
essentials of the auto-zero technique.
A more rigorous analysis is available in the data sheets for the AD855x.
2.125
BASIC LINEAR DESIGN
REFERENCES
1.
Daniel H. Sheingold, Editor, Transducer Interfacing Handbook, Analog Devices, Inc., 1981.
2.
C. Kitchin and L. Counts, Instrumentation Amplifier Applications Guide, Analog, Devices,
Inc., 1991.
3.
Amplifier Applications Guide, Analog Devices, Inc., 2002
4.
System Applications Guide, Analog Devices, Inc., 1993.
5.
John Sylvan, Ask The Applications Engineer -5. High-speed comparators provide many useful
circuit functions when used correctly.
6.
Reza Moghimi. Curing Comparator Instability with Hysteresis. Analog Dialogue 34-7 (2000)
7.
George Erdi, "A 300V/µs Monolithic Voltage Follower," IEEE Journal of Solid State Circuits,
Vol. SC-14, No. 6, December, 1979, pp. 1059-1065
8.
Royal A. Gosser, "Wideband Transconductance Generator," US Patent 5,150,074, Filed May 3,
1991, issued September 22, 1992.
9.
Derek F. Bowers, "A 6.8mA Closed-Loop Monolithic Buffer with 120MHz Bandwidth, 4000V/µs
Slew Rate, and ±12V Signal Compatibility," 1994 Bipolar/BiCMOS Circuits and Technology
Meeting 1.3, pp. 23-26.
10.
Barrie Gilbert, ISSCC Digest of Technical Papers 1968, pp. 114-115 February 16, 1968.
11.
Barrie Gilbert, Journal of Solid State Circuits, Vol. SC-3, December 1968, pp. 353-372.
12.
C.L. Ruthroff, Some Broadband Transformers, Proc. I.R.E., Vol.47, August, 1959,
pp.1337-1342.
13.
James M. Bryant, Mixers for High Performance Radio, Wescon 1981: Session 24 (Published by
Electronic Conventions, Inc., Sepulveda Blvd., El Segundo, CA)
14.
P.E. Chadwick, High Performance IC Mixers, IERE Conference on Radio Receivers and
Associated Systems, Leeds, 1981, IERE Conference Publication No. 50.
15.
P.E. Chadwick, Phase Noise, Intermodulation, and Dynamic Range, RF Expo, Anaheim, CA,
January, 1986.
16.
Daniel H. Sheingold, Editor, Nonlinear Circuits Handbook, Analog Devices, Inc., l974.
17.
Richard Smith Hughes, Logarithmic Amplifiers, Artech House, Inc., Dedham, MA., 1986.
18.
William L. Barber and Edmund R. Brown, A True Logarithmic Amplifier for Radar IF
Applications, IEEE Journal of Solid State Circuits, Vol. SC-15, No. 3, June, 1980,
pp. 291-295.
19.
Broadband Amplifier Applications, Plessey Co. Publication P.S. 1938, September, 1984.
20.
M. S. Gay, SL521 Application Note, Plessey Co., 1966.
21.
Amplifier Applications Guide, Analog Devices, Inc., 1992. Section 9.
2.126
OTHER LINEAR CIRCUITS
REFERENCES
22.
Charles Kitchen and Lew Counts, RMS-to-DC Conversion Application Guide, Second Edition,
Analog Devices, Inc., 1986.
23.
Barrie Gilbert, A Low Noise Wideband Variable-Gain Amplifier Using an Interpolated Ladder
Attenuator, IEEE ISSCC Technical Digest, 1991, pp. 280, 281, 330.
24.
Barrie Gilbert, A Monolithic Microsystem for Analog Synthesis of Trigonometric Functions and
their Inverses, IEEE Journal of Solid State Circuits, Vol. SC-17, No. 6, December
25.
Lingli Zhang, et. al, “Real-time Power Supply Compensation for Noise-shaped Class-D
Amplifier”, presented at 117th AES Convention, San Francisco CA, USA, 2004 October 28 – 31.
26.
Marco Berkhout, “Integrated 200-W class-D audio amplifier,” J. Solid State Circuits, vol. 38, pp.
1198-1206, July 2003.
27.
Karsten Nielsen, “A review and comparison of pulse width modulation (PWM) methods for
analog and digital input switching power amplifiers,” presented at 102nd AES Convention,
Munich, Germany, 1997 March 22 – 25.
28.
P. Morrow, et. al, “A 20 Watt Stereo Class-D Audio Output Power Stage in 0.6um BCDMOS
Technology,” J. Solid State Circuits, vol. 39, Nov 2004.
29.
S.R Norsworthy, R. Schreier, G.C Temes editors, Delta-Sigma Data Converters, IEEE press, 1997,
pp. 153-155.
30.
Eric Gaalaas, Bill Yang Liu, Naoaki Nishimura, Robert Adams, Karl Sweetland, and Rajeev
Morajkar, “Integrated stereo Σ-∆ Class-D amplifier”, Presented at the 118th Convention
Barcelona, Spain 2005 May 28–31.
31.
Eric Gaalaas, Bill Yang Liu, Naoaki Nishimura, Robert Adams, and Karl Sweetland
Integrated Stereo Σ-Δ Class-D Amplifier IEEE Journal of Solid-State Circuits, vol. 40, no. 12,
December 2005 pp. 2388-2397
2.127
BASIC LINEAR DESIGN
2.128