Você está na página 1de 47

Analog Integrated Circuits

PDF generated using the open source mwlib toolkit. See http://code.pediapress.com/ for more information.
PDF generated at: Sun, 29 Jan 2012 12:50:46 UTC

Contents
Articles
Bode plot

Current mirror

11

Differential amplifier

19

Operational amplifier

25

References
Article Sources and Contributors

43

Image Sources, Licenses and Contributors

44

Article Licenses
License

45

Bode plot

Bode plot
A Bode plot is a graph of the transfer
function of a linear, time-invariant
system versus frequency, plotted with
a log-frequency axis, to show the
system's frequency response. It is
usually a combination of a Bode
magnitude plot, expressing the
magnitude of the frequency response
gain, and a Bode phase plot,
expressing the frequency response
phase shift.

Overview
Among
his
several
important
contributions to circuit theory and
control theory, engineer Hendrik Wade
Bode (19051982), while working at
Bell Labs in the United States in the
1930s, devised a simple but accurate
method for graphing gain and
phase-shift plots. These bear his name,
Bode gain plot and Bode phase plot
(pronounced Boh-dee in English,
Bow-duh in Dutch).[1]

Figure 1(a): The Bode plot for a first-order (one-pole) highpass filter; the straight-line
approximations are labeled "Bode pole"; phase varies from 90 at low frequencies (due to
the contribution of the numerator, which is 90 at all frequencies) to 0 at high
frequencies (where the phase contribution of the denominator is 90 and cancels the
contribution of the numerator).

The magnitude axis of the Bode plot is


usually expressed as decibels of power,
that is by the 20 log rule: 20 times the
common (base 10) logarithm of the
amplitude gain. With the magnitude
gain being logarithmic, Bode plots
make multiplication of magnitudes a
simple matter of adding distances on
the graph (in decibels), since

A Bode phase plot is a graph of phase


versus frequency, also plotted on a
log-frequency axis, usually used in
conjunction with the magnitude plot, to
evaluate how much a signal will be
phase-shifted. For example a signal

Figure 1(b): The Bode plot for a first-order (one-pole) lowpass filter; the straight-line
approximations are labeled "Bode pole"; phase is 90 lower than for Figure 1(a) because
the phase contribution of the numerator is 0 at all frequencies.

Bode plot
described by: Asin(t) may be attenuated but also phase-shifted. If the system attenuates it by a factor x and phase
shifts it by the signal out of the system will be (A/x)sin(t). The phase shift is generally a function of
frequency.
Phase can also be added directly from the graphical values, a fact that is mathematically clear when phase is seen as
the imaginary part of the complex logarithm of a complex gain.
In Figure 1(a), the Bode plots are shown for the one-pole highpass filter function:

where f is the frequency in Hz, and f1 is the pole position in Hz, f1 = 100Hz in the figure. Using the rules for
complex numbers, the magnitude of this function is

while the phase is:

Care must be taken that the inverse tangent is set up to return degrees, not radians. On the Bode magnitude plot,
decibels are used, and the plotted magnitude is:

In Figure 1(b), the Bode plots are shown for the one-pole lowpass filter function:

Also shown in Figure 1(a) and 1(b) are the straight-line approximations to the Bode plots that are used in hand
analysis, and described later.
The magnitude and phase Bode plots can seldom be changed independently of each other changing the amplitude
response of the system will most likely change the phase characteristics and vice versa. For minimum-phase systems
the phase and amplitude characteristics can be obtained from each other with the use of the Hilbert transform.
If the transfer function is a rational function with real poles and zeros, then the Bode plot can be approximated with
straight lines. These asymptotic approximations are called straight line Bode plots or uncorrected Bode plots and
are useful because they can be drawn by hand following a few simple rules. Simple plots can even be predicted
without drawing them.
The approximation can be taken further by correcting the value at each cutoff frequency. The plot is then called a
corrected Bode plot.

Bode plot

Rules for hand-made Bode plot


The premise of a Bode plot is that one can consider the log of a function in the form:

as a sum of the logs of its poles and zeros:

This idea is used explicitly in the method for drawing phase diagrams. The method for drawing amplitude plots
implicitly uses this idea, but since the log of the amplitude of each pole or zero always starts at zero and only has one
asymptote change (the straight lines), the method can be simplified.

Straight-line amplitude plot


Amplitude decibels is usually done using the

where

and

are constants,

at every value of s where

version. Given a transfer function in the form

, and H is the transfer function:

(a zero), increase the slope of the line by

per decade.

at every value of s where


(a pole), decrease the slope of the line by
per decade.
The initial value of the graph depends on the boundaries. The initial point is found by putting the initial angular
frequency into the function and finding |H(j)|.
The initial slope of the function at the initial value depends on the number and order of zeros and poles that are at
values below the initial value, and are found using the first two rules.
To handle irreducible 2nd order polynomials,

can, in many cases, be approximated as

.
Note that zeros and poles happen when is equal to a certain

or

. This is because the function in question is

the magnitude of H(j), and since it is a complex function,


is

zero

or

pole

involving

the

term

. Thus at any place where there


,

the

magnitude

of

that

term

is

Corrected amplitude plot


To correct a straight-line amplitude plot:
at every zero, put a point
above the line,
at every pole, put a point
below the line,
draw a smooth curve through those points using the straight lines as asymptotes (lines which the curve
approaches).
Note that this correction method does not incorporate how to handle complex values of

or

. In the case of an

irreducible polynomial, the best way to correct the plot is to actually calculate the magnitude of the transfer function
at the pole or zero corresponding to the irreducible polynomial, and put that dot over or under the line at that pole or
zero.

Bode plot

Straight-line phase plot


Given a transfer function in the same form as above:

the idea is to draw separate plots for each pole and zero, then add them up. The actual phase curve is given by
.
To draw the phase plot, for each pole and zero:
if A is positive, start line (with zero slope) at 0 degrees
if A is negative, start line (with zero slope) at 180 degrees
at every

(for stable zeros

beginning one decade before


at every

(E.g.:

(E.g.:

degrees per decade,

), decrease the slope by

degrees per decade,

(for stable poles

beginning one decade before

), increase the slope by

"unstable" (right half plane) poles and zeros (


flatten the slope again when the phase has changed by

) have opposite behavior


degrees (for a zero) or

degrees (for a

pole),
After plotting one line for each pole or zero, add the lines together to obtain the final phase plot; that is, the final
phase plot is the superposition of each earlier phase plot.

Example
A passive (unity pass band gain) lowpass RC filter, for instance has the following transfer function expressed in the
frequency domain:

From the transfer function it can be determined that the cutoff frequency point fc (in hertz) is at the frequency

or (equivalently) at
where

is the angular cutoff frequency in radians per second.

The transfer function in terms of the angular frequencies becomes:

The above equation is the normalized form of the transfer function. The Bode plot is shown in Figure 1(b) above,
and construction of the straight-line approximation is discussed next.

Bode plot

Magnitude plot
The magnitude (in decibels) of the transfer function above, (normalized and converted to angular frequency form),
given by the decibel gain expression
:

when plotted versus input frequency on a logarithmic scale, can be approximated by two lines and it forms the
asymptotic (approximate) magnitude Bode plot of the transfer function:
for angular frequencies below

it is a horizontal line at 0 dB since at low frequencies the

term is small and

can be neglected, making the decibel gain equation above equal to zero,
for angular frequencies above

it is a line with a slope of 20 dB per decade since at high frequencies the

term dominates and the decibel gain expression above simplifies to

which is a straight line with a

slope of 20 dB per decade.


These two lines meet at the corner frequency. From the plot, it can be seen that for frequencies well below the corner
frequency, the circuit has an attenuation of 0 dB, corresponding to a unity pass band gain, i.e. the amplitude of the
filter output equals the amplitude of the input. Frequencies above the corner frequency are attenuated the higher
the frequency, the higher the attenuation.

Phase plot
The phase Bode plot is obtained by plotting the phase angle of the transfer function given by

versus

, where

and

lower than corner, the ratio

are the input and cutoff angular frequencies respectively. For input frequencies much
is small and therefore the phase angle is close to zero. As the ratio increases the

absolute value of the phase increases and becomes 45 degrees when


. As the ratio increases for input
frequencies much greater than the corner frequency, the phase angle asymptotically approaches 90 degrees. The
frequency scale for the phase plot is logarithmic.

Normalized plot
The horizontal frequency axis, in both the magnitude and phase plots, can be replaced by the normalized
(nondimensional) frequency ratio

. In such a case the plot is said to be normalized and units of the frequencies

are no longer used since all input frequencies are now expressed as multiples of the cutoff frequency

An example with pole and zero


Figures 2-5 further illustrate construction of Bode plots. This example with both a pole and a zero shows how to use
superposition. To begin, the components are presented separately.
Figure 2 shows the Bode magnitude plot for a zero and a low-pass pole, and compares the two with the Bode straight
line plots. The straight-line plots are horizontal up to the pole (zero) location and then drop (rise) at 20 dB/decade.
The second Figure 3 does the same for the phase. The phase plots are horizontal up to a frequency factor of ten
below the pole (zero) location and then drop (rise) at 45/decade until the frequency is ten times higher than the pole

Bode plot

(zero) location. The plots then are again horizontal at higher frequencies at a final, total phase change of 90.
Figure 4 and Figure 5 show how superposition (simple addition) of a pole and zero plot is done. The Bode straight
line plots again are compared with the exact plots. The zero has been moved to higher frequency than the pole to
make a more interesting example. Notice in Figure 4 that the 20 dB/decade drop of the pole is arrested by the 20
dB/decade rise of the zero resulting in a horizontal magnitude plot for frequencies above the zero location. Notice in
Figure 5 in the phase plot that the straight-line approximation is pretty approximate in the region where both pole
and zero affect the phase. Notice also in Figure 5 that the range of frequencies where the phase changes in the
straight line plot is limited to frequencies a factor of ten above and below the pole (zero) location. Where the phase
of the pole and the zero both are present, the straight-line phase plot is horizontal because the 45/decade drop of the
pole is arrested by the overlapping 45/decade rise of the zero in the limited range of frequencies where both are
active contributors to the phase.

Example with pole and zero

Figure 2: Bode magnitude plot for zero and low-pass pole; curves
labeled "Bode" are the straight-line Bode plots

Figure 3: Bode phase plot for zero and low-pass pole; curves labeled
"Bode" are the straight-line Bode plots

Figure 4: Bode magnitude plot for pole-zero combination; the location


of the zero is ten times higher than in Figures 2&3; curves labeled
"Bode" are the straight-line Bode plots

Figure 5: Bode phase plot for pole-zero combination; the location of


the zero is ten times higher than in Figures 2&3; curves labeled "Bode"
are the straight-line Bode plots

Gain margin and phase margin


Bode plots are used to assess the stability of negative feedback amplifiers by finding the gain and phase margins of
an amplifier. The notion of gain and phase margin is based upon the gain expression for a negative feedback
amplifier given by

where AFB is the gain of the amplifier with feedback (the closed-loop gain), is the feedback factor and AOL is the
gain without feedback (the open-loop gain). The gain AOL is a complex function of frequency, with both magnitude
and phase.[2] Examination of this relation shows the possibility of infinite gain (interpreted as instability) if the
product AOL = 1. (That is, the magnitude of AOL is unity and its phase is 180, the so-called Barkhausen

Bode plot
stability criterion). Bode plots are used to determine just how close an amplifier comes to satisfying this condition.
Key to this determination are two frequencies. The first, labeled here as f180, is the frequency where the open-loop
gain flips sign. The second, labeled here f0dB, is the frequency where the magnitude of the product | AOL | = 1 (in
dB, magnitude 1 is 0 dB). That is, frequency f180 is determined by the condition:
where vertical bars denote the magnitude of a complex number (for example, |a+jb| = [ a2 + b2]1/2 ), and
frequency f0dB is determined by the condition:
One measure of proximity to instability is the gain margin. The Bode phase plot locates the frequency where the
phase of AOL reaches 180, denoted here as frequency f180. Using this frequency, the Bode magnitude plot finds
the magnitude of AOL. If |AOL|180 = 1, the amplifier is unstable, as mentioned. If |AOL|180 < 1, instability does not
occur, and the separation in dB of the magnitude of |AOL|180 from |AOL| = 1 is called the gain margin. Because a
magnitude of one is 0 dB, the gain margin is simply one of the equivalent forms: 20 log10( |AOL|180) = 20 log10(
|AOL|180) 20 log10( 1 / ).
Another equivalent measure of proximity to instability is the phase margin. The Bode magnitude plot locates the
frequency where the magnitude of |AOL| reaches unity, denoted here as frequency f0dB. Using this frequency, the
Bode phase plot finds the phase of AOL. If the phase of AOL( f0dB) > 180, the instability condition cannot be met
at any frequency (because its magnitude is going to be < 1 when f = f180), and the distance of the phase at f0dB in
degrees above 180 is called the phase margin.
If a simple yes or no on the stability issue is all that is needed, the amplifier is stable if f0dB < f180. This criterion is
sufficient to predict stability only for amplifiers satisfying some restrictions on their pole and zero positions
(minimum phase systems). Although these restrictions usually are met, if they are not another method must be used,
such as the Nyquist plot.[3][4]

Examples using Bode plots


Figures 6 and 7 illustrate the gain behavior and terminology. For a three-pole amplifier, Figure 6 compares the Bode
plot for the gain without feedback (the open-loop gain) AOL with the gain with feedback AFB (the closed-loop gain).
See negative feedback amplifier for more detail.
In this example, AOL = 100 dB at low frequencies, and 1 / = 58 dB. At low frequencies, AFB 58 dB as well.
Because the open-loop gain AOL is plotted and not the product AOL, the condition AOL = 1 / decides f0dB. The
feedback gain at low frequencies and for large AOL is AFB 1 / (look at the formula for the feedback gain at the
beginning of this section for the case of large gain AOL), so an equivalent way to find f0dB is to look where the
feedback gain intersects the open-loop gain. (Frequency f0dB is needed later to find the phase margin.)
Near this crossover of the two gains at f0dB, the Barkhausen criteria are almost satisfied in this example, and the
feedback amplifier exhibits a massive peak in gain (it would be infinity if AOL = 1). Beyond the unity gain
frequency f0dB, the open-loop gain is sufficiently small that AFB AOL (examine the formula at the beginning of this
section for the case of small AOL).
Figure 7 shows the corresponding phase comparison: the phase of the feedback amplifier is nearly zero out to the
frequency f180 where the open-loop gain has a phase of 180. In this vicinity, the phase of the feedback amplifier
plunges abruptly downward to become almost the same as the phase of the open-loop amplifier. (Recall, AFB AOL
for small AOL.)
Comparing the labeled points in Figure 6 and Figure 7, it is seen that the unity gain frequency f0dB and the phase-flip
frequency f180 are very nearly equal in this amplifier, f180 f0dB 3.332kHz, which means the gain margin and
phase margin are nearly zero. The amplifier is borderline stable.

Bode plot

Figures 8 and 9 illustrate the gain margin and phase margin for a different amount of feedback . The feedback
factor is chosen smaller than in Figure 6 or 7, moving the condition | AOL | = 1 to lower frequency. In this example,
1 / = 77 dB, and at low frequencies AFB 77 dB as well.
Figure 8 shows the gain plot. From Figure 8, the intersection of 1 / and AOL occurs at f0dB = 1 kHz. Notice that the
peak in the gain AFB near f0dB is almost gone.[5][6]
Figure 9 is the phase plot. Using the value of f0dB = 1 kHz found above from the magnitude plot of Figure 8, the
open-loop phase at f0dB is 135, which is a phase margin of 45 above 180.
Using Figure 9, for a phase of 180 the value of f180 = 3.332 kHz (the same result as found earlier, of course[7]).
The open-loop gain from Figure 8 at f180 is 58 dB, and 1 / = 77 dB, so the gain margin is 19 dB.
Stability is not the sole criterion for amplifier response, and in many applications a more stringent demand than
stability is good step response. As a rule of thumb, good step response requires a phase margin of at least 45, and
often a margin of over 70 is advocated, particularly where component variation due to manufacturing tolerances is
an issue.[8] See also the discussion of phase margin in the step response article.

Examples

Figure 6: Gain of feedback amplifier AFB in dB and corresponding


open-loop amplifier AOL. Parameter 1/ = 58 dB, and at low
frequencies AFB 58 dB as well. The gain margin in this amplifier is
nearly zero because | AOL| = 1 occurs at almost f = f180.

Figure 7: Phase of feedback amplifier AFB in degrees and


corresponding open-loop amplifier AOL. The phase margin in this
amplifier is nearly zero because the phase-flip occurs at almost the
unity gain frequency f = f0dB where | AOL| = 1.

Figure 8: Gain of feedback amplifier AFB in dB and corresponding


open-loop amplifier AOL. In this example, 1 / = 77 dB. The gain
margin in this amplifier is 19 dB.

Figure 9: Phase of feedback amplifier AFB in degrees and


corresponding open-loop amplifier AOL. The phase margin in this
amplifier is 45.

Bode plot

Bode plotter
The Bode plotter is an electronic
instrument resembling an oscilloscope,
which produces a Bode diagram, or a
graph, of a circuit's voltage gain or
phase shift plotted against frequency in
a feedback control system or a filter.
An example of this is shown in Figure
10. It is extremely useful for analyzing
and testing filters and the stability of
feedback control systems, through the
measurement of corner (cutoff)
frequencies and gain and phase
margins.

Figure 10: Amplitude diagram of a 10th order Chebyshev filter plotted using a Bode
Plotter application. The chebyshev transfer function is defined by poles and zeros which
are added by clicking on a graphical complex diagram.

This is identical to the function


performed by a vector network analyzer, but the network analyzer is typically used at much higher frequencies.
For education/research purposes, plotting Bode diagrams for given transfer functions facilitates better understanding
and getting faster results (see external links).

Related plots
Two related plots that display the same data in different coordinate systems are the Nyquist plot and the Nichols plot.
These are parametric plots, with frequency as the input and magnitude and phase of the frequency response as the
output. The Nyquist plot displays these in polar coordinates, with magnitude mapping to radius and phase to
argument (angle). The Nichols plot displays these in rectangular coordinates, on the log scale.

Related Plots

A Nyquist plot.

A Nichols plot of the same response.

Bode plot

Notes
[1] Van Valkenburg, M. E. University of Illinois at Urbana-Champaign, "In memoriam: Hendrik W. Bode (1905-1982)", IEEE Transactions on
Automatic Control, Vol. AC-29, No 3., March 1984, pp. 193-194. Quote: "Something should be said about his name. To his colleagues at Bell
Laboratories and the generations of engineers that have followed, the pronunciation is boh-dee. The Bode family preferred that the original
Dutch be used as boh-dah."
[2] Ordinarily, as frequency increases the magnitude of the gain drops and the phase becomes more negative, although these are only trends and
may be reversed in particular frequency ranges. Unusual gain behavior can render the concepts of gain and phase margin inapplicable. Then
other methods such as the Nyquist plot have to be used to assess stability.
[3] Thomas H. Lee (2004). The design of CMOS radio-frequency integrated circuits (http:/ / worldcat. org/ isbn/ 0-521-83539-9) (Second Edition
ed.). Cambridge UK: Cambridge University Press. p.14.6 pp. 451453. ISBN0-521-83539-9. .
[4] William S Levine (1996). The control handbook: the electrical engineering handbook series (http:/ / books. google. com/
books?id=2WQP5JGaJOgC& pg=RA1-PA163& lpg=RA1-PA163& dq=stability+ "minimum+ phase"& source=web& ots=P3fFTcyfzM&
sig=ad5DJ7EvVm6In_zhI0MlF_6vHDA) (Second Edition ed.). Boca Raton FL: CRC Press/IEEE Press. p.10.1 p. 163. ISBN0849385709. .
[5] The critical amount of feedback where the peak in the gain just disappears altogether is the maximally flat or Butterworth design.
[6] Willy M C Sansen (2006). Analog design essentials (http:/ / worldcat. org/ isbn/ 0-387-25746-2). Dordrecht, The Netherlands: Springer.
p.0517-0527 pp. 157163. ISBN0-387-25746-2. .
[7] The frequency where the open-loop gain flips sign f180 does not change with a change in feedback factor; it is a property of the open-loop
gain. The value of the gain at f180 also does not change with a change in . Therefore, we could use the previous values from Figures 6 and 7.
However, for clarity the procedure is described using only Figures 8 and 9.
[8] Willy M C Sansen. 0526 p. 162 (http:/ / worldcat. org/ isbn/ 0-387-25746-2). ISBN0-387-25746-2. .

References
External links
Explanation of Bode plots with movies and examples (http://www.facstaff.bucknell.edu/mastascu/
eControlHTML/Freq/Freq5.html)
How to draw piecewise asymptotic Bode plots (http://lpsa.swarthmore.edu/Bode/BodeHow.html)
Summarized drawing rules (http://lims.mech.northwestern.edu/~lynch/courses/ME391/2003/bodesketching.
pdf) (PDF)
Bode plot applet (http://www.uwm.edu/People/msw/BodePlot/) - Accepts transfer function coefficients as
input, and calculates magnitude and phase response
Circuit analysis in electrochemistry (http://www.abc.chemistry.bsu.by/vi/fit.htm)
Tim Green: Operational amplifier stability (http://www.en-genius.net/includes/files/acqt_013105.pdf)
Includes some Bode plot introduction
Gnuplot code for generating Bode plot: DIN-A4 printing template (pdf)

10

Current mirror

Current mirror
A current mirror is a circuit designed to copy a current through one active device by controlling the current in
another active device of a circuit, keeping the output current constant regardless of loading. The current being
'copied' can be, and sometimes is, a varying signal current. Conceptually, an ideal current mirror is simply an ideal
inverting current amplifier that reverses the current direction as well or it is a current-controlled current source
(CCCS). The current mirror is used to provide bias currents and active loads to circuits

Mirror characteristics
There are three main specifications that characterize a current mirror. The first is the transfer ratio (in the case of a
current amplifier) or the output current magnitude (in the case of a constant current source CCS). The second is its
AC output resistance, which determines how much the output current varies with the voltage applied to the mirror.
The third specification is the minimum voltage drop across the output part of the mirror necessary to make it work
properly. This minimum voltage is dictated by the need to keep the output transistor of the mirror in active mode.
The range of voltages where the mirror works is called the compliance range and the voltage marking the boundary
between good and bad behavior is called the compliance voltage. There are also a number of secondary performance
issues with mirrors, for example, temperature stability.

Practical approximations
For small-signal analysis the current mirror can be approximated by its equivalent Norton impedance .
In large-signal hand analysis, a current mirror is usually and simply approximated by an ideal current source.
However, an ideal current source is unrealistic in several respects:
it has infinite AC impedance, while a practical mirror has finite impedance
it provides the same current regardless of voltage, that is, there are no compliance range requirements
it has no frequency limitations, while a real mirror has limitations due to the parasitic capacitances of the
transistors
the ideal source has no sensitivity to real-world effects like noise, power-supply voltage variations and component
tolerances.

Circuit realizations of current mirrors


Basic idea
A bipolar transistor can be used as the
simplest current-to-current converter
but its transfer ratio would highly
depend on temperature variations,
tolerances, etc. To eliminate these
undesired disturbances, a current
mirror is composed of two cascaded
A current mirror consists of two cascaded inverse converters with mirrored transfer
current-to-voltage
and
characteristic.
voltage-to-current converters placed at
the same conditions and having reverse characteristics. They have not to be obligatory linear; the only requirement is
their characteristics to be mirrorlike (for example, in the BJT current mirror below, they are logarithmic and

11

Current mirror

12

exponential). Usually, two identical converters are used but the characteristic of the first one is reversed by applying
a negative feedback. Thus a current mirror consists of two cascaded equal converters (the first - reversed and the
second - direct).

Basic BJT current mirror


If a voltage is applied to the BJT base-emitter junction as an input
quantity and the collector current is taken as an output quantity,
the transistor will act as an exponential voltage-to-current
converter. By applying a negative feedback (simply joining the
base and collector) the transistor can be "reversed" and it will
begin acting as the opposite logarithmic current-to-voltage
converter; now it will adjust the "output" base-emitter voltage so
as to pass the applied "input" collector current.
The simplest bipolar current mirror (shown in Figure 1)
implements this idea. It consists of two cascaded transistor stages
acting accordingly as a reversed and direct voltage-to-current
Figure 1: A current mirror implemented with npn
converters. Transistor Q1 is connected to ground. Its collector-base
bipolar transistors using a resistor to set the reference
voltage is zero as shown. Consequently, the voltage drop across Q1
current IREF; VCC = supply voltage
is VBE, that is, this voltage is set by the diode law and Q1 is said to
be diode connected. (See also Ebers-Moll model.) It is important to have Q1 in the circuit instead of a simple diode,
because Q1 sets VBE for transistor Q2. If Q1 and Q2 are matched, that is, have substantially the same device
properties, and if the mirror output voltage is chosen so the collector-base voltage of Q2 is also zero, then the
VBE-value set by Q1 results in an emitter current in the matched Q2 that is the same as the emitter current in Q1.
Because Q1 and Q2 are matched, their 0-values also agree, making the mirror output current the same as the
collector current of Q1. The current delivered by the mirror for arbitrary collector-base reverse bias VCB of the output
transistor is given by (see bipolar transistor):
,
where IS = reverse saturation current or scale current, VT = thermal voltage and VA = Early voltage. This current is
related to the reference current IREF when the output transistor VCB = 0 V by:

as found using Kirchhoff's current law at the collector node of Q1:


The reference current supplies the collector current to Q1 and the base currents to both transistors when both
transistors have zero base-collector bias, the two base currents are equal, IB1=IB2=IB.

Parameter 0 is the transistor -value for VCB = 0 V.

Current mirror

13

Output resistance
If VCB is greater than zero in output transistor Q2, the collector current in Q2 will be somewhat larger than for Q1 due
to the Early effect. In other words, the mirror has a finite output (or Norton) resistance given by the rO of the output
transistor, namely (see Early effect):
,
where VA = Early voltage and VCB = collector-to-base bias.
Compliance voltage
To keep the output transistor active, VCB 0 V. That means the lowest output voltage that results in correct mirror
behavior, the compliance voltage, is VOUT = VCV = VBE under bias conditions with the output transistor at the output
current level IC and with VCB = 0 V or, inverting the I-V relation above:

where VT = thermal voltage and IS = reverse saturation current or scale current.


Extensions and complications
When Q2 has VCB > 0 V, the transistors no longer are matched. In particular, their -values differ due to the Early
effect, with

where VA is the Early voltage and 0 = transistor for VCB = 0 V. Besides the difference due to the Early effect, the
transistor -values will differ because the 0-values depend on current, and the two transistors now carry different
currents (see Gummel-Poon model).
Further, Q2 may get substantially hotter than Q1 due to the associated higher power dissipation. To maintain
matching, the temperature of the transistors must be nearly the same. In integrated circuits and transistor arrays
where both transistors are on the same die, this is easy to achieve. But if the two transistors are widely separated, the
precision of the current mirror is compromised.
Additional matched transistors can be connected to the same base and will supply the same collector current. In other
words, the right half of the circuit can be duplicated several times with various resistor values replacing R2 on each.
Note, however, that each additional right-half transistor "steals" a bit of collector current from Q1 due to the non-zero
base currents of the right-half transistors. This will result in a small reduction in the programmed current.
An example of a mirror with emitter degeneration to increase mirror resistance is found in two-port networks.
For the simple mirror shown in the diagram, typical values of

will yield a current match of 1% or better.

Current mirror

14

Basic MOSFET current mirror


The basic current mirror can also be implemented using MOSFET
transistors, as shown in Figure 2. Transistor M1 is operating in the
saturation or active mode, and so is M2. In this setup, the output
current IOUT is directly related to IREF, as discussed next.
The drain current of a MOSFET ID is a function of both the
gate-source voltage and the drain-to-gate voltage of the MOSFET
given by ID = f (VGS, VDG), a relationship derived from the
functionality of the MOSFET device. In the case of transistor M1
of the mirror, ID = IREF. Reference current IREF is a known
current, and can be provided by a resistor as shown, or by a
"threshold-referenced" or "self-biased" current source to ensure
that it is constant, independent of voltage supply variations.[1]

Figure 2: An n-channel MOSFET current mirror with a


resistor to set the reference current IREF; VDD is the
supply voltage

Using VDG=0 for transistor M1, the drain current in M1 is ID = f


(VGS,VDG=0), so we find: f (VGS, 0) = IREF, implicitly determining
the value of VGS. Thus IREF sets the value of VGS. The circuit in the diagram forces the same VGS to apply to
transistor M2. If M2 is also biased with zero VDG and provided transistors M1 and M2 have good matching of their
properties, such as channel length, width, threshold voltage etc., the relationship IOUT = f (VGS,VDG=0 ) applies, thus
setting IOUT = IREF; that is, the output current is the same as the reference current when VDG=0 for the output
transistor, and both transistors are matched.
The drain-to-source voltage can be expressed as VDS=VDG +VGS. With this substitution, the Shichman-Hodges
model provides an approximate form for function f (VGS,VDG):[1]

where,

is a technology related constant associated with the transistor, W/L is the width to length ratio of the

transistor, VGS is the gate-source voltage, Vth is the threshold voltage, is the channel length modulation constant,
and VDS is the drain source voltage.
Output resistance
Because of channel-length modulation, the mirror has a finite output (or Norton) resistance given by the ro of the
output transistor, namely (see channel length modulation):
,
where = channel-length modulation parameter and VDS = drain-to-source bias.
Compliance voltage
To keep the output transistor resistance high, VDG 0 V.[2] (see Baker).[3] That means the lowest output voltage that
results in correct mirror behavior, the compliance voltage, is VOUT = VCV = VGS for the output transistor at the output
current level with VDG = 0 V, or using the inverse of the f-function, f 1:
.
For Shichman-Hodges model, f

-1

is approximately a square-root function.

Current mirror

15

Extensions and reservations


A useful feature of this mirror is the linear dependence of f upon device width W, a proportionality approximately
satisfied even for models more accurate than the Shichman-Hodges model. Thus, by adjusting the ratio of widths of
the two transistors, multiples of the reference current can be generated.
It must be recognized that the Shichman-Hodges model[4] is accurate only for rather dated technology, although it
often is used simply for convenience even today. Any quantitative design based upon new technology uses computer
models for the devices that account for the changed current-voltage characteristics. Among the differences that must
be accounted for in an accurate design is the failure of the square law in Vgs for voltage dependence and the very
poor modeling of Vds drain voltage dependence provided by Vds. Another failure of the equations that proves very
significant is the inaccurate dependence upon the channel length L. A significant source of L-dependence stems from
, as noted by Gray and Meyer, who also note that usually must be taken from experimental data.[1]

Feedback assisted current mirror


Figure 3 shows a mirror using negative
feedback to increase output resistance.
Because of the op amp, these circuits are
sometimes called gain-boosted current
mirrors. Because they have relatively low
compliance voltages, they also are called
wide-swing current mirrors. A variety of
circuits based upon this idea are in
use,[5][6][7] particularly for MOSFET
mirrors because MOSFETs have rather low
intrinsic output resistance values. A
MOSFET version of Figure 3 is shown in
Figure 4 where MOSFETs M3 and M4
operate in Ohmic mode to play the same
role as emitter resistors RE in Figure 3, and
MOSFETs M1 and M2 operate in active
mode in the same roles as mirror transistors
Q1 and Q2 in Figure 3. An explanation
follows of how the circuit in Figure 3 works.

Figure 3: Gain-boosted current mirror with op amp feedback to increase output


resistance

The operational amplifier is fed the


difference in voltages V1 - V2 at the top of
the two emitter-leg resistors of value RE. This difference is amplified by the op amp and fed to the base of output
transistor Q2. If the collector base reverse bias on Q2 is increased by increasing the applied voltage VA, the current in
Q2 increases, increasing V2 and decreasing the difference V1 - V2 entering the op amp. Consequently, the base
voltage of Q2 is decreased, and VBE of Q2 decreases, counteracting the increase in output current.

Current mirror

16

If the op amp gain Av is large, only a very


small difference V1 - V2 is sufficient to
generate the needed base voltage VB for Q2,
namely

Consequently, the currents in the two leg


resistors are held nearly the same, and the
output current of the mirror is very nearly
the same as the collector current IC1 in Q1,
which in turn is set by the reference current
as

where 1 for transistor Q1 and 2 for Q2


differ due to the Early effect if the reverse
bias across the collector-base of Q2 is
non-zero.
Figure 4: MOSFET version of wide-swing current mirror; M1 and M2 are in active
mode, while M3 and M4 are in Ohmic mode and act like resistors

Figure 5: Small-signal circuit to determine output resistance of mirror; transistor Q2


is replaced with its hybrid-pi model; a test current IX at the output generates a
voltage VX, and the output resistance is Rout = VX / IX.

Output resistance
An idealized treatment of output resistance is given in the footnote.[8] A small-signal analysis for an op amp with
finite gain Av but otherwise ideal is based upon Figure 5 (, rO and r refer to Q2). To arrive at Figure 5, notice that
the positive input of the op amp in Figure 3 is at AC ground, so the voltage input to the op amp is simply the AC
emitter voltage Ve applied to its negative input, resulting in a voltage output of Av Ve. Using Ohm's law across the
input resistance r determines the small-signal base current Ib as:

Combining this result with Ohm's law for RE, Ve can be eliminated, to find:[9]

Current mirror

Kirchhoff's voltage law from the test source IX to the ground of RE provides:
Substituting for Ib and collecting terms the output resistance Rout is found to be:

For a large gain Av >> r / RE the maximum output resistance obtained with this circuit is
a substantial improvement over the basic mirror where Rout = rO.
The small-signal analysis of the MOSFET circuit of Figure 4 is obtained from the bipolar analysis by setting = gm
r in the formula for Rout and then letting r . The result is
This time, RE is the resistance of the source-leg MOSFETs M3, M4. Unlike Figure 3, however, as Av is increased
(holding RE fixed in value), Rout continues to increase, and does not approach a limiting value at large Av.
Compliance voltage
For Figure 3, a large op amp gain achieves the maximum Rout with only a small RE. A low value for RE means V2
also is small, allowing a low compliance voltage for this mirror, only a voltage V2 larger than the compliance voltage
of the simple bipolar mirror. For this reason this type of mirror also is called a wide-swing current mirror, because it
allows the output voltage to swing low compared to other types of mirror that achieve a large Rout only at the
expense of large compliance voltages.
With the MOSFET circuit of Figure 4, like the circuit in Figure 3, the larger the op amp gain Av, the smaller RE can
be made at a given Rout, and the lower the compliance voltage of the mirror.

Other current mirrors


There are many sophisticated current mirrors that have higher output resistances than the basic mirror (more closely
approach an ideal mirror with current output independent of output voltage) and produce currents less sensitive to
temperature and device parameter variations and to circuit voltage fluctuations. These multi-transistor mirror circuits
are used both with bipolar and MOS transistors. These circuits include:
the Widlar current source
the Wilson current source
the Cascoded current sources

17

Current mirror

Notes
[1] Paul R. Gray, Paul J. Hurst, Stephen H. Lewis, Robert G. Meyer (2001). Analysis and Design of Analog Integrated Circuits (Fourth Edition
ed.). New York: Wiley. p.308309. ISBN0471321680.
[2] Keeping the output resistance high means more than keeping the MOSFET in active mode, because the output resistance of real MOSFETs
only begins to increase on entry into the active region, then rising to become close to maximum value only when VDG 0 V.
[3] R. Jacob Baker (2010). CMOS Circuit Design, Layout and Simulation (Third ed.). New York: Wiley-IEEE. pp.297, 9.2.1 and Figure 20.28,
p. 636. ISBN978-0-470-88132-3.
[4] NanoDotTek Report NDT14-08-2007, 12 August 2007 (http:/ / www. nanodottek. com/ NDT14_08_2007. pdf)
[5] R. Jacob Baker. 20.2.4 pp. 645646. ISBN978-0-470-88132-3.
[6] Ivanov VI and Filanovksy IM (2004). Operational amplifier speed and accuracy improvement: analog circuit design with structural
methodology (http:/ / books. google. com/ books?id=IuLsny9wKIIC& pg=PA110& dq=gain+ boost+ wide+ + "current+ mirror"#PPA107,M1)
(The Kluwer international series in engineering and computer science, v. 763 ed.). Boston, Mass.: Kluwer Academic. p.6.1, p. 105108.
ISBN1-4020-7772-6. .
[7] W. M. C. Sansen (2006). Analog design essentials. New York ; Berlin: Springer. p.0310, p. 93. ISBN0-387-25746-2.
[8] An idealized version of the argument in the text, valid for infinite op amp gain, is as follows. If the op amp is replaced by a nullor, voltage V2
= V1, so the currents in the leg resistors are held at the same value. That means the emitter currents of the transistors are the same. If the VCB of
Q2 increases, so does the output transistor because of the Early effect: = 0 ( 1 + VCB / VA ). Consequently the base current to Q2 given by
IB = IE / ( + 1) decreases and the output current Iout = IE / (1 + 1 / ) increases slightly because increases slightly. Doing the math,

where the transistor output resistance is given by rO = ( VA + VCB ) / Iout. That is, the ideal mirror resistance for the
circuit using an ideal op amp nullor is Rout = ( + 1 ) rO, in agreement with the value given later in the text when the
gain .
[9] Notice that as Av , Ve 0 and Ib IX.

References
External links
Patent (1968) by RJ Widlar: Biasing scheme especially suited for integrated circuits (http://www.google.com/
patents?hl=en&id=IhlMAAAAEBAJ&dq=+"Biasing+scheme+especially+suited+for+integrated+circuits"&
printsec=frontcover&source=web&ots=Oq1T4owqMu&sig=oYenqTCaZip9QTcLwVjIcOgIc-U&sa=X&
oi=book_result&resnum=4&ct=result#PPP3,M1)
Current mirrors (http://www.google.bg/search?hl=en&source=hp&biw=1266&bih=641&q=current+
mirror&oq=current+mirror&aq=f&aqi=g10&aql=&gs_sm=e&
gs_upl=1781l6281l0l14l14l0l3l3l0l328l2455l0.4.6.1)
4QD tec - Current sources and mirrors (http://www.4qdtec.com/csm.html) Compendium of circuits and
descriptions

18

Differential amplifier

19

Differential amplifier
A differential amplifier is a type of electronic amplifier that amplifies
the difference between two voltages but does not amplify the particular
voltages.

Theory
Many electronic devices use differential amplifiers internally. The
output of an ideal differential amplifier is given by:

Where

and

are the input voltages and

is the differential

gain.
In practice, however, the gain is not quite equal for the two inputs. This
means, for instance, that if
and
are equal, the output will not
be zero, as it would be in the ideal case. A more realistic expression for
the output of a differential amplifier thus includes a second term.

Differential amplifier symbolThe inverting and


non-inverting inputs are distinguished by "" and
"+" symbols (respectively) placed in the amplifier
triangle. Vs+ and Vs are the power supply
voltages; they are often omitted from the diagram
for simplicity, but of course must be present in
the actual circuit.

is called the common-mode gain of the amplifier.


As differential amplifiers are often used to null out noise or bias-voltages that appear at both inputs, a low
common-mode gain is usually desired.
The common-mode rejection ratio (CMRR), usually defined as the ratio between differential-mode gain and
common-mode gain, indicates the ability of the amplifier to accurately cancel voltages that are common to both
inputs. The common-mode rejection ratio is defined as:

In a perfectly symmetrical differential amplifier,

is zero and the CMRR is infinite. Note that a differential

amplifier is a more general form of amplifier than one with a single input; by grounding one input of a differential
amplifier, a single-ended amplifier results.

Long-tailed pair
Historical background
The long-tailed pair was originally implemented using a pair of vacuum tubes. The circuit works the same way for
all three-terminal devices with current gain. Today, its main feature is mostly vestigial, by virtue of the fact that
long-tail resistor circuit bias points are largely determined by Ohm's Law and less so by active component
characteristics.
The long-tailed pair was developed from earlier knowledge of push-pull circuit techniques and measurement
bridges.[1] The earliest circuit that is truly recognizable as a long-tailed pair in its conventional form is given by
Matthews (1934)[2] and the same circuit form appears in a patent submitted by Alan Blumlein in 1936.[3] By the end
of the 1930s the topology was well established and had been described by various authors including Offner (1937),[4]
Schmitt (1937)[5] and Toennies (1938) and It was particularly used for detection and measurement of physiological
impulses.[6]

Differential amplifier

20

The long-tailed pair was very successfully used in early British computing, most notably the Pilot ACE Model and
descendants,[7] Wilkes' EDSAC, and probably others designed by people who worked with Blumlein or his peers.
The long-tailed pair has many attributes as a switch: largely immune to tube (transistor) variations (of great
importance when machines contained 1,000 or more tubes), high gain, gain stability, high input impedance,
medium/low output impedance, good clipper (with not-too-long tail), non-inverting (EDSAC contained no
inverters!) and large output voltage swings. One disadvantage is that the output voltage swing (typically 1020V)
was imposed upon a high DC voltage (200V or so), requiring care in signal coupling, usually some form of
wide-band DC coupling. Many computers of this time tried to avoid this problem by using only AC-coupled pulse
logic, which made them very large and overly complex (ENIAC: 18,000 tubes for a 20 digit calculator) or unreliable.
DC-coupled circuitry became the norm after the first generation of vacuum tube computers.

Configurations
A differential (long-tailed,[8] emitter-coupled) pair amplifier consists of two amplifying stages with common
(emitter, source or cathode) degeneration.
Differential output
With two inputs and two outputs, this forms a differential amplifier
stage (Fig. 2). The two bases (or grids or gates) are inputs which are
differentially amplified (subtracted and multiplied) by the pair; they
can be fed with a differential (balanced) input signal, or one input
could be grounded to form a phase splitter circuit. An amplifier with
differential output can drive floating load or another stage with
differential input.
Single-ended output
If the differential output is not desired, then only one output can be
used (taken from just one of the collectors (or anodes or drains),
disregarding the other output without a collector resistor; this
configuration is referred to as single-ended output. The gain is half that
of the stage with differential output. To avoid sacrificing gain, a
differential to single-ended converter can be utilized. This is often
implemented as a current mirror (Fig. 3).

Figure 2: A classic long-tailed pair

Operation
To explain the circuit operation, four particular modes are isolated below although, in practice, some of them act
simultaneously and their effects are superimposed.
Biasing
In contrast with classic amplifying stages that are biased from the side of the base (and so they are highly
-dependent), the differential pair is directly biased from the side of the emitters by sinking/injecting the total
quiescent current. The series negative feedback (the emitter degeneration) makes the transistors act as voltage
stabilizers; it forces them to adjust their VBE voltages (base currents) so that to pass the quiescent current through
their collector-emitter junctions.[9] So, due to the negative feedback, the quiescent current depends slightly on the
transistor's .

Differential amplifier
The biasing base currents needed to evoke the quiescent collector currents usually come from the ground, pass
through the input sources and enter the bases. So, the sources have to be galvanic (DC) to ensure paths for the
biasing currents and low resistive enough to not create significant voltage drops across them. Otherwise, additional
DC elements should be connected between the bases and the ground (or the positive power supply).
Common mode
At common mode (the two input voltages change in the same directions), the two voltage (emitter) followers
cooperate with each other working together on the common high-resistive emitter load (the "long tail"). They all
together increase or decrease the voltage of the common emitter point (figuratively speaking, they together "pull up"
or "loose" it so that it moves). In addition, the dynamic load "helps" them by changing its instant ohmic resistance in
the same direction as the input voltages (it increases when the voltage increases and vice versa.) thus keeping up
constant total resistance between the two supply rails. There is a full (100%) negative feedback; the two input base
voltages and the emitter voltage change simultaneously while the collector currents and the total current do not
change. As a result, the output collector voltages do not change as well.
Differential mode
Normal. At differential mode (the two input voltages change in opposite directions), the two voltage (emitter)
followers oppose each other - while one of them tries to increase the voltage of the common emitter point, the other
tries to decrease it (figuratively speaking, one of them "pulls up" the common point while the other "looses" it so that
it stays immovable) and v.v. So, the common point does not change its voltage; it behaves like a virtual ground with
a magnitude determined by the common-mode input voltages. The high-resistive emitter element does not play any
role since it is shunted by the other low-resistive emitter follower. There is no negative feedback since the emitter
voltage does not change at all when the input base voltages change. he common quiescent current vigorously steers
between the two transistors and the output collector voltages vigorously change. The two transistors mutually ground
their emitters; so, although they are common-collector stages, they actually act as common-emitter stages with
maximum gain. Bias stability and independence from variations in device parameters can be improved by negative
feedback introduced via cathode/emitter resistors with relatively small resistances.
Overdriven. If the input differential voltage changes significantly (more than about a hundred millivolts), the
base-emitter junction of the transistor driven by the lower input voltage becomes backward biased and its collector
voltage reaches the positive supply rail. The other transistor (driven by the higher input voltage) saturates and its
collector voltage begins following the input one. This mode is used in differential switches and ECL gates.
Breakdown. If the input voltage continues increasing and exceeds the base-emitter breakdown voltage, the
base-emitter junction of the transistor driven by the lower input voltage breaks down. If the input sources are low
resistive, an unlimited current will flow directly through the "diode bridge" between the two input sources and will
damage them.
At common mode, the emitter voltage follows the input voltage variations; there is a full negative feedback and the
gain is minimum. At differential mode, the emitter voltage is fixed (equal to the instant common input voltage); there
is no negative feedback and the gain is maximum.

21

Differential amplifier
Single-ended input
The differential pair can be used as an amplifier with a single-ended input if one of the inputs is grounded or fixed to
a reference voltage (usually, the other collector is used as a single-ended output) This arrangement can be thought as
of cascaded common-collector and common-base stages or as a buffered common-base stage.[10]
The emitter-coupled amplifier is compensated for temperature drifts, VBE is cancelled, Miller effect and transistor
saturation are beated. That is why it is used to form emitter-coupled amplifiers (avoiding Miller effect), phase splitter
circuits (obtaining two inverse voltages), ECL gates and switches (avoiding transistor saturation), etc.

Improvements
Emitter constant current source
The quiescent current has to be constant to ensure constant collector
voltages at common mode. This requirement is not so important in the
case of a differential output since the two collector voltages will vary
simultaneously but their difference (the output voltage) will not vary;
only the output range will decrease. But in the case of a single-ended
output, it is extremely important to keep a constant current since the
output collector voltage will vary. Thus the higher the resistance of the
current source
, the lower
is, and the better the CMRR. The
constant current needed can be produced by connecting an element
(resistor) with very high resistance between the shared emitter node
Figure 3: An improved long-tailed pair with
and the supply rail (negative for NPN and positive for PNP transistors)
current-mirror load and constant-current biasing
but this will require high supply voltage. That is why, in more
sophisticated designs, an element with high differential (dynamic)
resistance approximating a constant current source/sink is substituted for the long tail (Fig. 3). It is usually
implemented by a current mirror because of its high compliance voltage (small voltage drop across the output
transistor).
The same arrangement is widely used in cascode circuits as well. It can be generalized by an equivalent circuit
consisting of a constant current source loaded by two connected in parallel voltage sources with equal voltages. The
current source determines the common current flowing through the voltage sources while the voltage sources fix the
voltage across the current source. The emitter current source is usually implemented as a common-emitter transistor
stage with constant base voltage driving with current the two common-base transistor stages. So, this arrangement
can be considered as a cascode consisting of cascaded common-emitter and common-base stages.
Collector current mirror
The collector resistors can be replaced by a current mirror, whose output part acts as an active load (Fig. 3). Thus the
differential collector current signal is converted to a single ended voltage signal without the intrinsic 50% losses and
the gain is extremely increased. This is achieved by copying the input collector current from the right to the left side
where the magnitudes of the two input signals add. For this purpose, the input of the current mirror is connected to
the right output and the output of the current mirror is connected to the left output of the differential amplifier.
The current mirror inverts the right collector current and tries to pass it through the left transistor that produces the
left collector current. In the middle point between the two left transistors, the two signal currents (current changes)
are subtracted. In this case (differential input signal), they are equal and opposite. Thus, the difference is twice the
individual signal currents (I - (-I) = 2I) and the differential to single ended conversion is completed without gain
losses.

22

Differential amplifier

23

Interfacing considerations
Floating input source
It is possible to connect a floating source between the two bases, but it is necessary to ensure paths for the biasing
base currents. In the case of galvanic source, only one resistor has to be connected between one of the bases and the
ground. The biasing current will enter directly this base and indirectly (through the input source) the other one. If the
source is capacitive, two resistors have to be connected between the two bases and the ground to ensure different
paths for the base currents.
Input/output impedance
The input impedance of the differential pair highly depends on the input mode. At common mode, the two parts
behave as common-collector stages with high emitter loads; so, the input impedances are extremely high. At
differential mode, they behave as common-emitter stages with grounded emitters; so, the input impedances are low.
The output impedance of the differential pair is high (especially for the improved differential pair from Fig. 3).
Input/output range
The common-mode input voltage can vary between the two supply rails but cannot closely reach them since some
voltage drops (minimum 1 volt) have to remain across the output transistors of the two current mirrors.

Other differential amplifiers


An operational amplifier, or op-amp, is a
differential
amplifier
with
very
high
differential-mode gain, very high input
impedances, and a low output impedance. By
applying negative feedback an op-amp
differential amplifier (Fig. 4) with predictable
and stable gain can be built.[11] Some kinds of
differential amplifier usually include several
simpler differential amplifiers. For example, an
instrumentation amplifier, a fully differential
amplifier, an instrument amplifier, or an
isolation amplifier are often built from several
op-amps.

Figure 4: Op-amp differential amplifier

Applications
Differential amplifiers are found in many circuits that utilise series negative feedback (op-amp follower,
non-inverting amplifier, etc.), where one input is used for the input signal, the other for the feedback signal (usually
implemented by operational amplifiers). For comparison, the old-fashioned inverting single-ended op-amps from the
early 40's could realize only parallel negative feedback by connecting additional resistor networks (an op-amp
inverting amplifier is the most popular example). A common application is for the control of motors or servos, as
well as for signal amplification applications. In discrete electronics, a common arrangement for implementing a
differential amplifier is the long-tailed pair, which is also usually found as the differential element in most op-amp
integrated circuits. A long-tailed pair can be used as an analog multiplier with the differential voltage as one input
and the biasing current as another.

Differential amplifier
A differential amplifier is used as the input stage emitter coupled logic gates and as switch. When used as a switch,
the "left" base/grid is used as signal input and the "right" base/grid is grounded; output is taken from the right
collector/plate. When the input is zero or negative, the output is close to zero (but can be not saturated); when the
input is positive, the output is most-positive, dynamic operation being the same as the amplifier use described above.

Footnotes
[1] Eglin, J. M. A Direct Current Amplifier for Measuring Small Currents. Journal of the Optical Society of America and Review of Scientific
Instruments, May 1929, p.393-402.
[2] Matthews, B. H. C. A Special Purpose Amplifier. The Journal of Physiology, March 17, 1934, 81, 28P-29P.
[3] Blumlein, A. D. UK Patent No: 482740; July 4, 1936. US Patent 2185367; filed June 24, 1937
[4] Offner, F. Push-Pull Resistance Coupled Amplifiers. Review of Scientific Instruments, January 1937, p.20-21.
[5] Schmitt, O. H. Cathode Phase Inversion. Review of Scientific Instruments, 1937, p.100-101.
[6] Geddes, L. A. Who Invented the Differential Amplifier?. IEEE Engineering in Medicine and Biology, May/June 1996, p.116-117.
[7] Details of the long-tailed pair circuitry used in early computing can be found in "Alan Turing's Automatic Computing Engine" (Oxford
University Press, 2005, ISBN 0-19-856593-3) in Part IV, 'ELECTRONICS'
[8] Long-tail is a figurative name of high resistance that represents the high emitter resistance at common mode with a common long tail with a
proportional length (at differential mode this tail shortens up to zero). If additional emitter resistors with small resistances are included
between the emitters and the common node (to introduce a small negative feedback at differential mode), they can be figuratively represented
by short tails.
[9] It is interesting fact that the negative feedback as though has reversed the transistor behavior - the collector current has become an input
quantity while the base current serves as an output one.
[10] More generally, this arrangement can be considered as two interacting voltage followers with negative feedback: the output part of the
differential pair acts as a voltage follower with constant input voltage (a voltage stabilizer) producing constant output voltage; the input part
acts as a voltage follower with varying input voltage trying to change the steady output voltage of the stabilizer. The stabilizer reacts to this
intervention by changing its output quantity (current, respectively voltage) that serves as a circuit output.
[11] It seems strange that, in this arrangement, a high-gain differential amplifier (op-amp) is used as a component to build a low-gain differential
amplifier like a high-gain inverting amplifier (op-amp) serves as a component in a low-gain inverting amplifier; but just this paradox of
negative feedback amplifiers has impeded Harold Black to obtain his patent.

References
External links
BJT Differential Amplifier (http://www.ecircuitcenter.com/Circuits/BJT_Diffamp1/BJT_Diffamp1.htm)
Circuit and explanation
A testbench for differential circuits (http://www.designers-guide.org/Analysis/diff.pdf)
A discrete OpAmp with complimentary differential amplifier (http://gaedtke.name/SubMenu_Verstaerker/
KompDiffAmp.html) (in German)
Application Note: Terminating a Differential Amplifier in Single-Ended Input Applications (http://www.analog.
com/static/imported-files/application_notes/AN-0990.pdf)

24

Operational amplifier

25

Operational amplifier
An operational amplifier ("op-amp") is a
DC-coupled high-gain electronic voltage
amplifier with a differential input and,
usually, a single-ended output.[1] An op-amp
produces an output voltage that is typically
hundreds of thousands times larger than the
voltage difference between its input
terminals.[2]
Operational amplifiers are important
building blocks for a wide range of
electronic circuits. They had their origins in
analog computers where they were used in
many
linear,
non-linear
and
A Signetics a741 operational amplifier, one of the most successful op-amps.
frequency-dependent
circuits.
Their
popularity in circuit design largely stems from the fact that characteristics of the final op-amp circuits with negative
feedback (such as their gain) are set by external components with little dependence on temperature changes and
manufacturing variations in the op-amp itself.
Op-amps are among the most widely used electronic devices today, being used in a vast array of consumer,
industrial, and scientific devices. Many standard IC op-amps cost only a few cents in moderate production volume;
however some integrated or hybrid operational amplifiers with special performance specifications may cost over
$100 US in small quantities. Op-amps may be packaged as components, or used as elements of more complex
integrated circuits.
The op-amp is one type of differential amplifier. Other types of differential amplifier include the fully differential
amplifier (similar to the op-amp, but with two outputs), the instrumentation amplifier (usually built from three
op-amps), the isolation amplifier (similar to the instrumentation amplifier, but with tolerance to common-mode
voltages that would destroy an ordinary op-amp), and negative feedback amplifier (usually built from one or more
op-amps and a resistive feedback network).

Circuit notation
The circuit symbol for an op-amp is shown to the right, where:
V+: non-inverting input
V: inverting input
Vout: output
VS+: positive power supply
VS: negative power supply
The power supply pins (VS+ and VS) can be labeled in different ways
(See IC power supply pins). Despite different labeling, the function
remains the same to provide additional power for amplification of
the signal. Often these pins are left out of the diagram for clarity, and
the power configuration is described or assumed from the circuit.

Circuit diagram symbol for an op-amp

Operational amplifier

26

Operation
The amplifier's differential inputs consist of a V+ input and a V input,
and ideally the op-amp amplifies only the difference in voltage
between the two, which is called the differential input voltage. The
output voltage of the op-amp is given by the equation,

where V+ is the voltage at the non-inverting terminal, V is the voltage


at the inverting terminal and AOL is the open-loop gain of the amplifier
(the term "open-loop" refers to the absence of a feedback loop from the
output to the input).
The magnitude of AOL is typically very large10,000 or more for
An op-amp without negative feedback (a
integrated circuit op-ampsand therefore even a quite small difference
comparator)
between V+ and V drives the amplifier output nearly to the supply
voltage. This is called saturation of the amplifier. The magnitude of
AOL is not well controlled by the manufacturing process, and so it is impractical to use an operational amplifier as a
stand-alone differential amplifier. Without negative feedback, and perhaps with positive feedback for regeneration,
an op-amp acts as a comparator. If the inverting input is held at ground (0 V) directly or by a resistor, and the input
voltage Vin applied to the non-inverting input is positive, the output will be maximum positive; if Vin is negative, the
output will be maximum negative. Since there is no feedback from the output to either input, this is an open loop
circuit acting as a comparator. The circuit's gain is just the AOL< of the op-amp.
If predictable operation is desired, negative feedback is used, by
applying a portion of the output voltage to the inverting input. The
closed loop feedback greatly reduces the gain of the amplifier. If
negative feedback is used, the circuit's overall gain and other
parameters become determined more by the feedback network than by
the op-amp itself. If the feedback network is made of components with
relatively constant, stable values, the unpredictability and inconstancy
of the op-amp's parameters do not seriously affect the circuit's
performance. Typically the op-amp's very large gain is controlled by
negative feedback, which largely determines the magnitude of its
output ("closed-loop") voltage gain in amplifier applications, or the
transfer function required (in analog computers). High input impedance
at the input terminals and low output impedance at the output
terminal(s) are important typical characteristics.

An op-amp with negative feedback (a


non-inverting amplifier)

For example, in a non-inverting amplifier (see the figure on the right) adding a negative feedback via the voltage
divider Rf,Rg reduces the gain. Equilibrium will be established when Vout is just sufficient to reach around and "pull"
the inverting input to the same voltage as Vin. The voltage gain of the entire circuit is determined by 1 + Rf/Rg. As a
simple example, if Vin = 1 V and Rf = Rg, Vout will be 2 V, the amount required to keep V at 1 V. Because of the
feedback provided by Rf,Rg this is a closed loop circuit. Its overall gain Vout/Vin is called the closed-loop gain ACL.
Because the feedback is negative, in this case ACL is less than the AOL of the op-amp.
Another way of looking at it is to make two relatively valid assumptions: One, that when an op-amp is being
operated in linear mode, the difference in voltage between the non-inverting (+) pin and the inverting (-) pin is so
small as to be considered negligible.[3] The second assumption is that the input impedance at both + and - pins is
extremely high (at least several megaohms with modern op-amps). Thus, when the circuit to the right is operated as a

Operational amplifier

27

non-inverting linear amplifier, Vin will appear at the + and - pins and create a current i through Rg equal to Vin/Rg.
Since Kirchoff's current law states that the same current must leave a node as enter it, and since the impedance into
the - pin is near infinity, we can assume the overwhelming majority of the same current i travels through Rf, creating
an output voltage equal to Vin + i*Rf. By combining terms, we can easily determine the gain of this particular type
of circuit.
i = Vin / Rg
Vout = Vin + i*Rf
Vout = Vin + (Vin / Rg) * Rf
Vout = Vin + (Vin * Rf) / Rg
G = Vout/Vin = (Vin + (Vin * Rf) / Rg) / Vin
G = (Vin/Vin) + (Vin * Rf) / Rg) / Vin
G = 1 + (Vin * Rf) / (Rg * Vin)
G = 1 + Rf/Rg

Op-amp characteristics
Ideal op-amps
An ideal op-amp is usually considered to have the
following properties, and they are considered to hold
for all input voltages:
Infinite open-loop gain (when doing theoretical
analysis, a limit may be taken as open loop gain AOL
goes to infinity).
Infinite voltage range available at the output (
)
(in practice the voltages available from the output
are limited by the supply voltages
and
).
The power supply sources are called rails.
Infinite bandwidth (i.e., the frequency magnitude
response is considered to be flat everywhere with
zero phase shift).
Infinite input impedance (so, in the diagram,
, and zero current flows from
to

An equivalent circuit of an operational amplifier that models some


resistive non-ideal parameters.

).

Zero input current (i.e., there is assumed to be no leakage or bias current into the device).
Zero input offset voltage (i.e., when the input terminals are shorted so that

, the output is a virtual

ground or
).
Infinite slew rate (i.e., the rate of change of the output voltage is unbounded) and power bandwidth (full output
voltage and current available at all frequencies).

Zero output impedance (i.e.,


, so that output voltage does not vary with output current).
Zero noise.
Infinite Common-mode rejection ratio (CMRR).
Infinite Power supply rejection ratio for both power supply rails.

These ideals can be summarized by the two "golden rules":


I. The output attempts to do whatever is necessary to make the voltage difference between the inputs zero.
II. The inputs draw no current.[4]:177

Operational amplifier
The first rule only applies in the usual case where the op-amp is used in a closed-loop design (negative feedback,
where there is a signal path of some sort feeding back from the output to the inverting input). These rules are
commonly used as a good first approximation for analyzing or designing op-amp circuits.[4]:177
In practice, none of these ideals can be perfectly realized, and various shortcomings and compromises have to be
accepted. Depending on the parameters of interest, a real op-amp may be modeled to take account of some of the
non-infinite or non-zero parameters using equivalent resistors and capacitors in the op-amp model. The designer can
then include the effects of these undesirable, but real, effects into the overall performance of the final circuit. Some
parameters may turn out to have negligible effect on the final design while others represent actual limitations of the
final performance, that must be evaluated.

Real op-amps
Real op-amps differ from the ideal model in various respects.
DC imperfections
Real operational amplifiers suffer from several non-ideal effects:
Finite gain
Open-loop gain is infinite in the ideal operational amplifier but finite in real operational amplifiers. Typical
devices exhibit open-loop DC gain ranging from 100,000 to over 1 million. So long as the loop gain (i.e., the
product of open-loop and feedback gains) is very large, the circuit gain will be determined entirely by the
amount of negative feedback (i.e., it will be independent of open-loop gain). In cases where closed-loop gain
must be very high, the feedback gain will be very low, and the low feedback gain causes low loop gain; in
these cases, the operational amplifier will cease to behave ideally.
Finite input impedances
The differential input impedance of the operational amplifier is defined as the impedance between its two
inputs; the common-mode input impedance is the impedance from each input to ground. MOSFET-input
operational amplifiers often have protection circuits that effectively short circuit any input differences greater
than a small threshold, so the input impedance can appear to be very low in some tests. However, as long as
these operational amplifiers are used in a typical high-gain negative feedback application, these protection
circuits will be inactive. The input bias and leakage currents described below are a more important design
parameter for typical operational amplifier applications.
Non-zero output impedance
Low output impedance is important for low-impedance loads; for these loads, the voltage drop across the
output impedance of the amplifier will be significant. Hence, the output impedance of the amplifier limits the
maximum power that can be provided. In configurations with a voltage-sensing negative feedback, the output
impedance of the amplifier is effectively lowered; thus, in linear applications, op-amps usually exhibit a very
low output impedance indeed. Negative feedback can not, however, reduce the limitations that Rload in
conjunction with Rout place on the maximum and minimum possible output voltages; it can only reduce output
errors within that range.
Low-impedance outputs typically require high quiescent (i.e., idle) current in the output stage and will
dissipate more power, so low-power designs may purposely sacrifice low output impedance.
Input current
Due to biasing requirements or leakage, a small amount of current (typically ~10 nanoamperes for bipolar
op-amps, tens of picoamperes for JFET input stages, and only a few pA for MOSFET input stages) flows into
the inputs. When large resistors or sources with high output impedances are used in the circuit, these small
currents can produce large unmodeled voltage drops. If the input currents are matched, and the impedance

28

Operational amplifier
looking out of both inputs are matched, then the voltages produced at each input will be equal. Because the
operational amplifier operates on the difference between its inputs, these matched voltages will have no effect
(unless the operational amplifier has poor CMRR, which is described below). It is more common for the input
currents (or the impedances looking out of each input) to be slightly mismatched, and so a small offset voltage
(different from the input offset voltage below) can be produced. This offset voltage can create offsets or
drifting in the operational amplifier. It can often be nulled externally; however, many operational amplifiers
include offset null or balance pins and some procedure for using them to remove this offset. Some operational
amplifiers attempt to nullify this offset automatically.
Input offset voltage
This voltage, which is what is required across the op-amp's input terminals to drive the output voltage to
zero,[5][6] is related to the mismatches in input bias current. In the perfect amplifier, there would be no input
offset voltage. However, it exists in actual op-amps because of imperfections in the differential amplifier that
constitutes the input stage of the vast majority of these devices. Input offset voltage creates two problems:
First, due to the amplifier's high voltage gain, it virtually assures that the amplifier output will go into
saturation if it is operated without negative feedback, even when the input terminals are wired together.
Second, in a closed loop, negative feedback configuration, the input offset voltage is amplified along with the
signal and this may pose a problem if high precision DC amplification is required or if the input signal is very
small.[7]
Common-mode gain
A perfect operational amplifier amplifies only the voltage difference between its two inputs, completely
rejecting all voltages that are common to both. However, the differential input stage of an operational
amplifier is never perfect, leading to the amplification of these identical voltages to some degree. The standard
measure of this defect is called the common-mode rejection ratio (denoted CMRR). Minimization of common
mode gain is usually important in non-inverting amplifiers (described below) that operate at high
amplification.
Output sink current
The output sink current is maximum current allowed to sink into the output stage. Some manufacturers show
the output voltage vs. the output sink current plot, which gives an idea of the output voltage when it is sinking
current from another source into the output pin.
Temperature effects
All parameters change with temperature. Temperature drift of the input offset voltage is especially important.
Power-supply rejection
The output of a perfect operational amplifier will be completely independent from ripples that arrive on its
power supply inputs. Every real operational amplifier has a specified power supply rejection ratio (PSRR) that
reflects how well the op-amp can reject changes in its supply voltage. Copious use of bypass capacitors can
improve the PSRR of many devices, including the operational amplifier.
Drift
Real op-amp parameters are subject to slow change over time and with changes in temperature, input
conditions, etc.
Noise
Amplifiers generate random voltage at the output even when there is no signal applied. This can be due to
thermal noise and flicker noise of the devices. For applications with high gain or high bandwidth, noise
becomes a very important consideration.

29

Operational amplifier
AC imperfections
The op-amp gain calculated at DC does not apply at higher frequencies. Thus, for high-speed operation, more
sophisticated considerations must be used in an op-amp circuit design.
Finite bandwidth
All amplifiers have finite bandwidth. To a first approximation, the op-amp has the frequency response of an
integrator with gain. That is, the gain of a typical op-amp is inversely proportional to frequency and is
characterized by its gainbandwidth product(GBWP). For example, an op-amp with a GBWP of 1MHz
would have a gain of 5 at 200kHz, and a gain of 1 at 1MHz. This dynamic response coupled with the very
high DC gain of the op-amp gives it the characteristics of a first-order low-pass filter with very high DC gain
and low cutoff frequency given by the GBWP divided by the DC gain.
The finite bandwidth of an op-amp can be the source of several problems, including:
Stability. Associated with the bandwidth limitation is a phase difference between the input signal and the
amplifier output that can lead to oscillation in some feedback circuits. For example, a sinusoidal output signal
meant to interfere destructively with an input signal of the same frequency will interfere constructively if
delayed by 180 degrees. In these cases, the feedback circuit can be stabilized by means of frequency
compensation, which increases the gain or phase margin of the open-loop circuit. The circuit designer can
implement this compensation externally with a separate circuit component. Alternatively, the compensation
can be implemented within the operational amplifier with the addition of a dominant pole that sufficiently
attenuates the high-frequency gain of the operational amplifier. The location of this pole may be fixed
internally by the manufacturer or configured by the circuit designer using methods specific to the op-amp. In
general, dominant-pole frequency compensation reduces the bandwidth of the op-amp even further. When the
desired closed-loop gain is high, op-amp frequency compensation is often not needed because the requisite
open-loop gain is sufficiently low; consequently, applications with high closed-loop gain can make use of
op-amps with higher bandwidths.
Noise, Distortion, and Other Effects. Reduced bandwidth also results in lower amounts of feedback at higher
frequencies, producing higher distortion, noise, and output impedance and also reduced output phase linearity
as the frequency increases.
Typical low-cost, general-purpose op-amps exhibit a GBWP of a few megahertz. Specialty and high-speed
op-amps exist that can achieve a GBWP of hundreds of megahertz. For very high-frequency circuits, a
current-feedback operational amplifier is often used.
Input capacitance
Most important for high frequency operation because it further reduces the open-loop bandwidth of the
amplifier.
Common-mode gain
See DC imperfections, above.
Non-linear imperfections
Saturation
output voltage is limited to a minimum and maximum value close to the power supply voltages.[8] Saturation
occurs when the output of the amplifier reaches this value and is usually due to:
In the case of an op-amp using a bipolar power supply, a voltage gain that produces an output that is more
positive or more negative than that maximum or minimum; or
In the case of an op-amp using a single supply voltage, either a voltage gain that produces an output that is
more positive than that maximum, or a signal so close to ground that the amplifier's gain is not sufficient to
raise it above the lower threshold.[9]

30

Operational amplifier

31

Slewing
the amplifier's output voltage reaches its maximum rate of change. Measured as the slew rate, it is usually
specified in volts per microsecond. When slewing occurs, further increases in the input signal have no effect
on the rate of change of the output. Slewing is usually caused by internal capacitances in the amplifier,
especially those used to implement its frequency compensation.
Non-linear input-output relationship
The output voltage may not be accurately proportional to the difference between the input voltages. It is
commonly called distortion when the input signal is a waveform. This effect will be very small in a practical
circuit if substantial negative feedback is used.
Power considerations
Limited output current
The output current must be finite. In practice, most op-amps are designed to limit the output current so as not
to exceed a specified level around 25mA for a type 741 IC op-amp thus protecting the op-amp and
associated circuitry from damage. Modern designs are electronically more rugged than earlier implementations
and some can sustain direct short circuits on their outputs without damage.
Limited dissipated power
The output current flows through the op-amp's internal output impedance, dissipating heat. If the op-amp
dissipates too much power, then its temperature will increase above some safe limit. The op-amp may enter
thermal shutdown, or it may be destroyed.
Modern integrated FET or MOSFET op-amps approximate more closely the ideal op-amp than bipolar ICs when it
comes to input impedance and input bias and offset currents. Bipolars are generally better when it comes to input
voltage offset, and often have lower noise. Generally, at room temperature, with a fairly large signal, and limited
bandwidth, FET and MOSFET op-amps now offer better performance.

Internal circuitry of 741 type op-amp


Though designs vary between products
and manufacturers, all op-amps have
basically the same internal structure,
which consists of three stages:
1. Differential amplifier provides
low noise amplification, high input
impedance, usually a differential
output.
2. Voltage amplifier provides high
voltage gain, a single-pole
frequency roll-off, usually
single-ended output.
3. Output amplifier provides high
current driving capability, low
output impedance, current limiting
and short circuit protection
circuitry.

A component level diagram of the common 741 op-amp. Dotted lines outline: current
mirrors (red); differential amplifier (blue); class A gain stage (magenta); voltage level
shifter (green); output stage (cyan).

Operational amplifier
IC op-amps as implemented in practice are moderately complex integrated circuits. A typical example is the
ubiquitous 741 op-amp designed by Dave Fullagar in Fairchild Semiconductor after the remarkable Widlar LM301.
Thus the basic architecture of the 741 is identical to that of the 301.[10]

Input stage
The input stage is a composed differential amplifier with a complex biasing circuit and a current mirror active load.
Differential amplifier
It is implemented by two cascaded stages satisfying the conflicting requirements. The first stage consists of the
NPN-based input emitter followers Q1 and Q2 that provide high input impedance. The next is the PNP-based
common base pair Q3 and Q4 that eliminates the undesired Miller effect, shifts the voltage level downwards and
provides a sufficient voltage gain to drive the next class A amplifier. The PNP transistors also help to increase the
reverse Vbe rating (the base-emitter junctions of the NPN transistors Q1 and Q2 break down at around 7 V but the
PNP transistors Q3 and Q4 have breakdown voltages around 50 V).[11]
Biasing circuit
The classical emitter-coupled differential stage is biased from the side of the emitters by connecting a constant
current source to them. The series negative feedback (the emitter degeneration) makes the transistors act as voltage
stabilizers; it forces them to adjust their VBE voltages so that to pass the current through their collector-emitter
junctions. As a result, the quiescent current is -independent.
Here, the Q3/Q4 emitters are already used as inputs. Their collectors are separated and cannot be used as inputs for
the quiescent current source since they behave as current sources. So, the quiescent current can be set only from the
side of the bases by connecting a constant current source to them. To make it not depend on as above, a negative
but parallel feedback is used. For this purpose, the total quiescent current is mirrored by Q8-Q9 current mirror and
the negative feedback is taken from the Q9 collector. Now it makes the transistors Q1-Q4 adjust their VBE voltages
so that to pass the desired quiescent current. The effect is the same as at the classical emitter-coupled pair - the
quiescent current is -independent. It is interesting fact that "to the extent that all PNP s match, this clever circuit
generates just the right -dependent base current to produce a -independent collector current".[10] The biasing base
currents are usually provided only by the negative power supply; they should come from the ground and enter the
bases. But to ensure maximum high input impedances, the biasing loops are not internally closed between the base
and ground; it is expected they will be closed externally by the input sources. So, the sources have to be galvanic
(DC) to ensure paths for the biasing currents and low resistive enough (tens or hundreds kilohms) to not create
significant voltage drops across them. Otherwise, additional DC elements should be connected between the bases
and the ground (or the positive power supply).
The quiescent current is set by the 39 k resistor that is common for the two current mirrors Q12-Q13 and Q10-Q11.
The current determined by this resistor acts also as a reference for the other bias currents used in the chip. The
Widlar current mirror built by Q10, Q11, and the 5 k resistor produces a very small fraction of
at the Q10
collector. This small constant current through Q10's collector supplies the base currents for Q3 and Q4 as well as the
Q9 collector current. The Q8/Q9 current mirror tries to make Q9 collector current the same as the Q3 and Q4
collector currents and succeeds with the help of the negative feedback. The Q9 collector voltage changes until the
ratio between the Q3/Q4 base and collector currents becomes equal to . Thus Q3 and Q4's combined base currents
(which are of the same order as the overall chip's input currents) are a small fraction of the already small Q10
current.
Thus the quiescent current is set by Q10-Q11 current mirror without using a current-sensing negative feedback. The
voltage-sensing negative feedback only helps this process by stabilizing Q9 collector (Q3/Q4 base) voltage.[12] The
feedback loop also isolates the rest of the circuit from common-mode signals by making the base voltage of Q3/Q4

32

Operational amplifier
follow tightly

33
below the higher of the two input voltages.

Current mirror active load


The differential amplifier formed by Q1Q4 drives an active load implemented as an improved current mirror
(Q5Q7) whose role is to convert the differential current input signal to a single ended voltage signal without the
intrinsic 50% losses and to increase extremely the gain. This is achieved by copying the input signal from the left to
the right side where the magnitudes of the two input signals add (Widlar used the same trick in A702 and A709).
For this purpose, the input of the current mirror (Q5 collector) is connected to the left output (Q3 collector) and the
output of the current mirror (Q6 collector) is connected to the right output of the differential amplifier (Q4 collector).
Q7 increases the accuracy of the current mirror by decreasing the amount of signal current required from Q3 to drive
the bases of Q5 and Q6.
Operation
Differential mode
The input voltage sources are connected through two "diode" strings, each of them consisting of two connected in
series base-emitter junctions (Q1-Q3 and Q2-Q4), to the common point of Q3/Q4 bases. So, if the input voltages
change slightly in opposite directions, Q3/Q4 bases stay at relatively constant voltage and the common base current
does not change as well; it only vigorously steers between Q3/Q4 bases and makes the common quiescent current
distribute between Q3/Q4 collectors in the same proportion.[13] The current mirror inverts Q3 collector current and
tries to pass it through Q4. In the middle point between Q4 and Q6, the signal currents (current changes) of Q3 and
Q4 are subtracted. In this case (differential input signal), they are equal and opposite. Thus, the difference is twice
the individual signal currents (I - (-I) = 2I) and the differential to single ended conversion is completed without
gain losses. The open circuit signal voltage appearing at this point is given by the product of the subtracted signal
currents and the total circuit impedance (the paralleled collector resistances of Q4 and Q6). Since the collectors of
Q4 and Q6 appear as high differential resistances to the signal current (Q4 and Q6 behave as current sources), the
open circuit voltage gain of this stage is very high.[14]
More intuitively, the transistor Q6 can be considered as a duplicate of Q3 and the combination of Q4 and Q6 can be
thought as of a varying voltage divider composed of two voltage-controlled resistors. For differential input signals,
they vigorously change their instant resistances in opposite directions but the total resistance stays constant (like a
potentiometer with quickly moving slider). As a result, the current stays constant as well but the voltage at the
middle point changes vigorously. As the two resistance changes are equal and opposite, the effective voltage change
is twice the individual change.
The base current at the inputs is not zero and the effective differential input impedance of a 741 is about 2 M. The
"offset null" pins may be used to place external resistors in parallel with the two 1 k resistors (typically in the form
of the two ends of a potentiometer) to adjust the balancing of the Q5/Q6 current mirror and thus indirectly control
the output of the op-amp when zero signal is applied between the inputs.

Operational amplifier
Common mode
If the input voltages change in the same direction, the negative feedback makes Q3/Q4 base voltage follow (with
2VBE below) the input voltage variations. Now the output part (Q10) of Q10-Q11 current mirror keeps up the
common current through Q9/Q8 constant in spite of varying voltage. Q3/Q4 collector currents and accordingly, the
output voltage in the middle point between Q4 and Q6, remain unchanged.
The following negative feedback (bootstrapping) increases virtually the effective op-amp common-mode input
impedance.

Class A gain stage


The section outlined in magenta is the class A gain stage. The top-right current mirror Q12/Q13 supplies this stage
by a constant current load, via the collector of Q13, that is largely independent of the output voltage. The stage
consists of the two NPN transistors Q15/Q19 connected in a Darlington configuration and uses the output side of a
current mirror as its collector (dynamic) load to achieve high gain. The transistor Q22 prevents this stage from
saturating by diverting the excessive Q15 base current (it acts as a Baker clamp).
The 30 pF capacitor provides frequency selective negative feedback around the class A gain stage as a means of
frequency compensation to stabilise the amplifier in feedback configurations. This technique is called Miller
compensation and functions in a similar manner to an op-amp integrator circuit. It is also known as 'dominant pole
compensation' because it introduces a dominant pole (one which masks the effects of other poles) into the open loop
frequency response. This pole can be as low as 10Hz in a 741 amplifier and it introduces a 3 dB loss into the open
loop response at this frequency. This internal compensation is provided to achieve unconditional stability of the
amplifier in negative feedback configurations where the feedback network is non-reactive and the closed loop gain is
unity or higher. Hence, the use of the operational amplifier is simplified because no external compensation is
required for unity gain stability; amplifiers without this internal compensation such as the 748 may require external
compensation or closed-loop gains significantly higher than unity.

Output bias circuitry


The green outlined section (based on Q16) is a voltage
level shifter named rubber diode, transistor Zener or
VBE multiplier. In the circuit as shown, Q16 provides a
constant voltage drop across its collector-emitter
junction regardless of the current through it (it acts as a
voltage stabilizer). This is achieved by introducing a
negative feedback between Q16 collector and its base,
i.e. by connecting a voltage divider with ratio = 7.5
The circuit presented as a negative feedback amplifier with constant
k / (4.5 k + 7.5 k) = 0.625 composed by the two
input voltage
resistors. If the base current to the transistor is assumed
to be zero, the negative feedback forces the transistor to increase its collector-emitter voltage up to 1 V until its
base-emitter voltage reaches 0.625 V (a typical value for a BJT in the active region). This serves to bias the two
output transistors slightly into conduction reducing crossover distortion (in some discrete component amplifiers, this
function is usually achieved with a string of two silicon diodes).
The circuit can be presented as a negative feedback voltage amplifier with constant input voltage of 0.625 V and a
feedback ratio of = 0.625 (a gain of 1/ = 1.6). The same circuit but with = 1 is used in the input current-setting
part of the classical BJT current mirror.

34

Operational amplifier

Output stage
The output stage (outlined in cyan) is a Class AB push-pull emitter follower (Q14, Q20) amplifier with the bias set
by the
multiplier voltage source Q16 and its base resistors. This stage is effectively driven by the collectors of
Q13 and Q19. Variations in the bias with temperature, or between parts with the same type number, are common so
crossover distortion and quiescent current may be subject to significant variation. The output range of the amplifier
is about one volt less than the supply voltage, owing in part to
of the output transistors Q14 and Q20.
The 25 resistor in the output stage acts as a current sense to provide the output current-limiting function which
limits the current in the emitter follower Q14 to about 25 mA for the 741. Current limiting for the negative output is
done by sensing the voltage across Q19's emitter resistor and using this to reduce the drive into Q15's base. Later
versions of this amplifier schematic may show a slightly different method of output current limiting. The output
resistance is not zero, as it would be in an ideal op-amp, but with negative feedback it approaches zero at low
frequencies.

Some considerations
Note: while the 741 was historically used in audio and other sensitive equipment, such use is now rare because of
the improved noise performance of more modern op-amps. Apart from generating noticeable hiss, 741s and other
older op-amps may have poor common-mode rejection ratios and so will often introduce cable-borne mains hum and
other common-mode interference, such as switch 'clicks', into sensitive equipment.
The "741" has come to often mean a generic op-amp IC (such as A741, LM301, 558, LM324, TBA221 - or a more
modern replacement such as the TL071). The description of the 741 output stage is qualitatively similar for many
other designs (that may have quite different input stages), except:
Some devices (A748, LM301, LM308) are not internally compensated (require an external capacitor from output
to some point within the operational amplifier, if used in low closed-loop gain applications).
Some modern devices have rail-to-rail output capability (output can be taken to positive or negative power supply
rail within a few millivolts).

Classification
Op-amps may be classified by their construction:
discrete (built from individual transistors or tubes/valves)
IC (fabricated in an Integrated circuit) - most common
hybrid
IC op-amps may be classified in many ways, including:
Military, Industrial, or Commercial grade (for example: the LM301 is the commercial grade version of the
LM101, the LM201 is the industrial version). This may define operating temperature ranges and other
environmental or quality factors.
Classification by package type may also affect environmental hardiness, as well as manufacturing options; DIP,
and other through-hole packages are tending to be replaced by surface-mount devices.
Classification by internal compensation: op-amps may suffer from high frequency instability in some negative
feedback circuits unless a small compensation capacitor modifies the phase and frequency responses. Op-amps
with a built-in capacitor are termed "compensated", or perhaps compensated for closed-loop gains down to (say)
5. All others are considered uncompensated.
Single, dual and quad versions of many commercial op-amp IC are available, meaning 1, 2 or 4 operational
amplifiers are included in the same package.
Rail-to-rail input (and/or output) op-amps can work with input (and/or output) signals very close to the power
supply rails.

35

Operational amplifier

36

CMOS op-amps (such as the CA3140E) provide extremely high input resistances, higher than JFET-input
op-amps, which are normally higher than bipolar-input op-amps.
other varieties of op-amp include programmable op-amps (simply meaning the quiescent current, gain, bandwidth
and so on can be adjusted slightly by an external resistor).
manufacturers often tabulate their op-amps according to purpose, such as low-noise pre-amplifiers, wide
bandwidth amplifiers, and so on.

Applications
Use in electronics system design
The use of op-amps as circuit blocks is much easier and clearer than
specifying all their individual circuit elements (transistors, resistors,
etc.), whether the amplifiers used are integrated or discrete. In the first
approximation op-amps can be used as if they were ideal differential
gain blocks; at a later stage limits can be placed on the acceptable
range of parameters for each op-amp.

DIP pinout for 741-type operational amplifier

Circuit design follows the same lines for all electronic circuits. A specification is drawn up governing what the
circuit is required to do, with allowable limits. For example, the gain may be required to be 100 times, with a
tolerance of 5% but drift of less than 1% in a specified temperature range; the input impedance not less than one
megohm; etc.
A basic circuit is designed, often with the help of circuit modeling (on a computer). Specific commercially available
op-amps and other components are then chosen that meet the design criteria within the specified tolerances at
acceptable cost. If not all criteria can be met, the specification may need to be modified.
A prototype is then built and tested; changes to meet or improve the specification, alter functionality, or reduce the
cost, may be made.

Applications without using any feedback


That is, the op-amp is being used as a voltage comparator. Note that a device designed primarily as a comparator
may be better if, for instance, speed is important or a wide range of input voltages may be found, since such devices
can quickly recover from full on or full off ("saturated") states.
A voltage level detector can be obtained if a reference voltage Vref is applied to one of the op-amp's inputs. This
means that the op-amp is set up as a comparator to detect a positive voltage. If the voltage to be sensed, Ei, is applied
to op amp's (+) input, the result is a noninverting positive-level detector: when Ei is above Vref, VO equals +Vsat;
when Ei is below Vref, VO equals -Vsat. If Ei is applied to the inverting input, the circuit is an inverting positive-level
detector: When Ei is above Vref, VO equals -Vsat.
A zero voltage level detector (Ei = 0) can convert, for example, the output of a sine-wave from a function generator
into a variable-frequency square wave. If Ei is a sine wave, triangular wave, or wave of any other shape that is
symmetrical around zero, the zero-crossing detector's output will be square. Zero-crossing detection may also be
useful in triggering TRIACs at the best time to reduce mains interference and current spikes.

Operational amplifier

37

Positive feedback applications


Another typical configuration of op-amps is with positive feedback, which takes a fraction of the output signal back
to the non-inverting input. An important application of it is the comparator with hysteresis, the Schmitt trigger. Some
circuits may use Positive feedback and Negative feedback around the same amplifier, for example Triangle wave
oscillators and active filters.
Because of the wide slew-range and lack of positive feedback, the response of all the open-loop level detectors
described above will be relatively slow. External overall positive feedback may be applied but (unlike internal
positive feedback that may be applied within the latter stages of a purpose-designed comparator) this markedly
affects the accuracy of the zero-crossing detection point. Using a general-purpose op-amp, for example, the
frequency of Ei for the sine to square wave converter should probably be below 100Hz.

Negative feedback applications


Non-inverting amplifier
In a non-inverting amplifier, the output voltage changes in the same
direction as the input voltage.
The gain equation for the op-amp is:

However, in this circuit

is a function of

negative feedback through the


voltage divider, and as

because of the

network.

and

form a

An op-amp connected in the non-inverting


amplifier configuration

is a high-impedance input, it does not load

it appreciably. Consequently:
where

Substituting this into the gain equation, we obtain:

Solving for

If

is very large, this simplifies to


.

Note that the non-inverting input of the operational amplifier will need a path for DC to ground; if the signal source
might not give this, or if that source requires a given load impedance, the circuit will require another resistor - from
input to ground. In either case, the ideal value for the feedback resistors (to give minimum offset voltage) will be
such that the two resistances in parallel roughly equal the resistance to ground at the non-inverting input pin.

Operational amplifier

38

Inverting amplifier
In an inverting amplifier, the output voltage changes in an opposite
direction to the input voltage.
As with the non-inverting amplifier, we start with the gain equation of
the op-amp:

This time,

is a function of both

divider formed by

and

and

due to the voltage

. Again, the op-amp input does not

apply an appreciable load, so:

Substituting this into the gain equation and solving for

If

An op-amp connected in the inverting amplifier


configuration

is very large, this simplifies to


.

A resistor is often inserted between the non-inverting input and ground (so both inputs "see" similar resistances),
reducing the input offset voltage due to different voltage drops due to bias current, and may reduce distortion in
some op-amps.
A DC-blocking capacitor may be inserted in series with the input resistor when a frequency response down to DC is
not needed and any DC voltage on the input is unwanted. That is, the capacitive component of the input impedance
inserts a DC zero and a low-frequency pole that gives the circuit a bandpass or high-pass characteristic.
The potentials at the operational amplifier inputs remain virtually constant (near ground) in the inverting
configuration. The constant operating potential typically results in distortion levels that are lower than those
attainable with the non-inverting topology.

Other applications

audio- and video-frequency pre-amplifiers and buffers


differential amplifiers
differentiators and integrators
filters
precision rectifiers
precision peak detectors
voltage and current regulators
analog calculators
analog-to-digital converters
digital-to-analog converters
voltage clamps
oscillators and waveform generators

Most single, dual and quad op-amps available have a standardized pin-out which permits one type to be substituted
for another without wiring changes. A specific op-amp may be chosen for its open loop gain, bandwidth, noise
performance, input impedance, power consumption, or a compromise between any of these factors.

Operational amplifier

39

Historical timeline
1941: A vacuum tube op-amp. An op-amp, defined as a general-purpose, DC-coupled, high gain, inverting
feedback amplifier, is first found in U.S. Patent 2,401,779 [15] "Summing Amplifier" filed by Karl D. Swartzel Jr. of
Bell Labs in 1941. This design used three vacuum tubes to achieve a gain of 90 dB and operated on voltage rails of
350 V. It had a single inverting input rather than differential inverting and non-inverting inputs, as are common in
today's op-amps. Throughout World War II, Swartzel's design proved its value by being liberally used in the M9
artillery director designed at Bell Labs. This artillery director worked with the SCR584 radar system to achieve
extraordinary hit rates (near 90%) that would not have been possible otherwise.[16]
1947: An op-amp with an explicit non-inverting input. In 1947, the operational
amplifier was first formally defined and named in a paper by Professor John R.
Ragazzini of Columbia University. In this same paper a footnote mentioned an
op-amp design by a student that would turn out to be quite significant. This op-amp,
designed by Loebe Julie, was superior in a variety of ways. It had two major
innovations. Its input stage used a long-tailed triode pair with loads matched to
reduce drift in the output and, far more importantly, it was the first op-amp design to
have two inputs (one inverting, the other non-inverting). The differential input made
a whole range of new functionality possible, but it would not be used for a long time
due to the rise of the chopper-stabilized amplifier.[17]
1949: A chopper-stabilized op-amp. In 1949, Edwin A. Goldberg designed a
chopper-stabilized op-amp.[18] This set-up uses a normal op-amp with an additional
AC amplifier that goes alongside the op-amp. The chopper gets an AC signal from
DC by switching between the DC voltage and ground at a fast rate (60Hz or
400Hz). This signal is then amplified, rectified, filtered and fed into the op-amp's
GAP/R's K2-W: a vacuum-tube
non-inverting input. This vastly improved the gain of the op-amp while significantly
op-amp (1953)
reducing the output drift and DC offset. Unfortunately, any design that used a
chopper couldn't use their non-inverting input for any other purpose. Nevertheless, the much improved
characteristics of the chopper-stabilized op-amp made it the dominant way to use op-amps. Techniques that used the
non-inverting input regularly would not be very popular until the 1960s when op-amp ICs started to show up in the
field.
In 1953, vacuum tube op-amps became commercially available with the release of the model K2-W from George A.
Philbrick Researches, Incorporated. The designation on the devices shown, GAP/R, is an acronym for the complete
company name. Two nine-pin 12AX7 vacuum tubes were mounted in an octal package and had a model K2-P
chopper add-on available that would effectively "use up" the non-inverting input. This op-amp was based on a
descendant of Loebe Julie's 1947 design and, along with its successors, would start the widespread use of op-amps in
industry.
1961: A discrete IC op-amps. With the birth of the transistor in 1947, and the
silicon transistor in 1954, the concept of ICs became a reality. The introduction of
the planar process in 1959 made transistors and ICs stable enough to be
commercially useful. By 1961, solid-state, discrete op-amps were being produced.
These op-amps were effectively small circuit boards with packages such as edge
connectors. They usually had hand-selected resistors in order to improve things such
as voltage offset and drift. The P45 (1961) had a gain of 94dB and ran on 15V
rails. It was intended to deal with signals in the range of 10 V.

GAP/R's model P45: a


solid-state, discrete op-amp
(1961).

Operational amplifier

40

1961: A varactor bridge op-amps. There have been many different directions taken in op-amp design. Varactor
bridge op-amps started to be produced in the early 1960s.[19][20] They were designed to have extremely small input
current and are still amongst the best op-amps available in terms of common-mode rejection with the ability to
correctly deal with hundreds of volts at their inputs.
1962: An op-amps in potted modules. By 1962, several companies were producing
modular potted packages that could be plugged into printed circuit boards. These
packages were crucially important as they made the operational amplifier into a
single black box which could be easily treated as a component in a larger circuit.
1963: A monolithic IC op-amp. In 1963, the first monolithic IC op-amp, the
A702 designed by Bob Widlar at Fairchild Semiconductor, was released.
Monolithic ICs consist of a single chip as opposed to a chip and discrete parts (a
discrete IC) or multiple chips bonded and connected on a circuit board (a hybrid IC).
Almost all modern op-amps are monolithic ICs; however, this first IC did not meet
with much success. Issues such as an uneven supply voltage, low gain and a small
dynamic range held off the dominance of monolithic op-amps until 1965 when the
A709[21] (also designed by Bob Widlar) was released.

GAP/R's model PP65: a


solid-state op-amp in a potted
module (1962)

1968: Release of the A741. The popularity of monolithic op-amps was further improved upon the release of the
LM101 in 1967, which solved a variety of issues, and the subsequent release of the A741 in 1968. The A741 was
extremely similar to the LM101 except that Fairchild's facilities allowed them to include a 30pF compensation
capacitor inside the chip instead of requiring external compensation. This simple difference has made the 741 the
canonical op-amp and many modern amps base their pinout on the 741s. The A741 is still in production, and has
become ubiquitous in electronicsmany manufacturers produce a version of this classic chip, recognizable by part
numbers containing 741. The same part is manufactured by several companies.
1970: First high-speed, low-input current FET design. In the 1970s high speed, low-input current designs started
to be made by using FETs. These would be largely replaced by op-amps made with MOSFETs in the 1980s. During
the 1970s single sided supply op-amps also became available.
1972: Single sided supply op-amps being produced. A single sided supply op-amp
is one where the input and output voltages can be as low as the negative power
supply voltage instead of needing to be at least two volts above it. The result is that
it can operate in many applications with the negative supply pin on the op-amp
being connected to the signal ground, thus eliminating the need for a separate
negative power supply.
The LM324 (released in 1972) was one such op-amp that came in a quad package
(four separate op-amps in one package) and became an industry standard. In
ADI's HOS-050: a high speed
hybrid IC op-amp (1979)
addition to packaging multiple op-amps in a single package, the 1970s also saw the
birth of op-amps in hybrid packages. These op-amps were generally improved
versions of existing monolithic op-amps. As the properties of monolithic op-amps improved, the more complex
hybrid ICs were quickly relegated to systems that are required to have extremely long service lives or other specialty
systems.

Operational amplifier

41

Recent trends. Recently supply voltages in analog circuits have decreased (as they
have in digital logic) and low-voltage op-amps have been introduced reflecting this.
Supplies of 5V and increasingly 3.3V (sometimes as low as 1.8V) are common.
To maximize the signal range modern op-amps commonly have rail-to-rail output
(the output signal can range from the lowest supply voltage to the highest) and
sometimes rail-to-rail inputs.
An op-amp in a modern mini
DIP

Notes
[1] MAXIM Application Note 1108: Understanding Single-Ended, Pseudo-Differential and Fully-Differential ADC Inputs (http:/ / www.
maxim-ic. com/ appnotes. cfm/ an_pk/ 1108) Retrieved November 10, 2007
[2] Analog devices MT-044 TUTORIAL (http:/ / www. analog. com/ static/ imported-files/ tutorials/ MT-044. pdf)
[3] Jacob Millman, Microelectronics: Digital and Analog Circuits and Systems, McGraw-Hill, 1979, ISBN 0-07-042327-x, pp. 523-527
[4] Horowitz, Paul; Hill, Winfield (1989). [[The Art of Electronics (http:/ / books. google. com/ books?id=bkOMDgwFA28C& pg=PA177&
lpg=PA177#v=onepage& q& f=false)]]. Cambridge, UK: Cambridge University Press. ISBN0521370957. .
[5] D.F. Stout Handbook of Operational Amplifier Circuit Design (McGraw-Hill, 1976, ISBN 007061797X ) pp.111.
[6] This definition hews to the convention of measuring op-amp parameters with respect to the zero voltage point in the circuit, which is usually
half the total voltage between the amplifier's positive and negative power rails.
[7] Many older designs of operational amplifiers have offset null inputs to allow the offset to be manually adjusted away. Modern precision
op-amps can have internal circuits that automatically cancel this offset using choppers or other circuits that measure the offset voltage
periodically and subtract it from the input voltage.
[8] That the output cannot reach the power supply voltages is usually the result of limitations of the amplifier's output stage transistors. See
Output stage.
[9] The output of older op-amps can reach to within one or two volts of the supply rails. The output of newer so-called "rail to rail" op-amps can
reach to within millivolts of the supply rails when providing low output currents.
[10] Lee, Thomas H. (November 18, 2002), IC Op-Amps Through the Ages (http:/ / www. stanford. edu/ class/ archive/ ee/ ee214/ ee214. 1032/
Handouts/ ho18opamp. pdf), Stanford University, ; Handout #18: EE214 Fall 2002
[11] The uA741 Operational Amplifier (http:/ / ecow. engr. wisc. edu/ cgi-bin/ get/ ece/ 342/ schowalter/ notes/ chapter10/
theua741operationalamplifier. ppt)
[12] This arrangement can be generalized by an equivalent circuit consisting of a constant current source loaded by a voltage source; the voltage
source fixes the voltage across the current source while the current source sets the current through the voltage source. As the two
heterogeneous sources provide ideal load conditions for each other, this circuit solution is widely used in cascode circuits, Wilson current
mirror, the input part of the simple current mirror, emitter-coupled and other exotic circuits.
[13] If the input differential voltage changes significantly (with more than about a hundred millivolts), the base-emitter junctions of the
transistors driven by the lower input voltage (e.g., Q1 and Q3) become backward biased and the total common base current flows through the
other (Q2 and Q4) base-emitter junctions. However, the high breakdown voltage of the PNP transistors Q3/Q4 prevents Q1/Q2 base-emitter
junctions from damaging when the input difference voltage increases up to 50 V because of the unlimited current that may flow directly
through the "diode bridge" between the two input sources.
[14] This circuit (and geometrical) phenomenon can be illustrated graphically by superimposing the Q4 and Q6 output characteristics (almost
parallel horizontal lines) on the same coordinate system. When the input voltages vary slightly in opposite directions, the two curves move
slightly toward each other in the vertical direction but the operating (cross) point moves vigorously in the horizontal direction. The ratio
between the two movements represents the high amplification.
[15] http:/ / www. google. com/ patents?vid=2,401,779
[16] Jung, Walter G. (2004). "Chapter 8: Op Amp History" (http:/ / books. google. com/ books?id=dunqt1rt4sAC). Op Amp Applications
Handbook. Newnes. p.777. ISBN9780750678445. . Retrieved 2008-11-15.
[17] Jung, Walter G. (2004). "Chapter 8: Op Amp History" (http:/ / books. google. com/ books?id=dunqt1rt4sAC). Op Amp Applications
Handbook. Newnes. p.779. ISBN9780750678445. . Retrieved 2008-11-15.
[18] http:/ / www. analog. com/ library/ analogDialogue/ archives/ 39-05/ Web_ChH_final. pdf
[19] http:/ / www. philbrickarchive. org/
[20] June 1961 advertisement for Philbrick P2, http:/ / www. philbrickarchive. org/
p2%20and%206033%20ad%20rsi%20vol32%20no6%20june1961. pdf
[21] A.P. Malvino, Electronic Principles (2nd Ed. 1979. ISBN 0-07-039867-4) p.476.

Operational amplifier

References
Further reading
Basic Operational Amplifiers and Linear Integrated Circuits; 2nd Ed; Thomas L Floyd; 593 pages; 1998; ISBN
978-0130829870.
Design with Operational Amplifiers and Analog Integrated Circuits; 3rd Ed; Sergio Franco; 672 pages; 2001;
ISBN 978-0072320848.
Operational Amplifiers and Linear Integrated Circuits; 6th Ed; Robert F Coughlin; 529 pages; 2000; ISBN
978-0130149916.
Op-Amps and Linear Integrated Circuits; 4th Ed; Ram Gayakwad; 543 pages; 1999; ISBN 978-0132808682.

External links
Simple Op Amp Measurements (http://www.analog.com/library/analogDialogue/archives/45-04/
op_amp_measurements.html) How to measure offset voltage, offset and bias current, gain, CMRR, and PSRR.
Introduction to op-amp circuit stages, second order filters, single op-amp bandpass filters, and a simple intercom
(http://www.bowdenshobbycircuits.info/opamp.htm)
Hyperphysics descriptions of common applications (http://hyperphysics.phy-astr.gsu.edu/hbase/electronic/
opampvar.html)
Single supply op-amp circuit collection (http://instruct1.cit.cornell.edu/courses/bionb440/datasheets/
SingleSupply.pdf)
Op-amp circuit collection (http://www.national.com/an/AN/AN-31.pdf)
Opamps for everyone (http://focus.ti.com/lit/an/slod006b/slod006b.pdf) Downloadable book.
MOS op amp design: A tutorial overview (http://www.ee.unb.ca/Courses/EE3122/DFL/AdditionalMaterial/
OpAmps/MOS_OpAmpTutorial.pdf)
High Speed OpAmp Techniques (http://www.linear.com/docs/4138) very practical and readable with photos
and real waveforms
Op Amp Applications (http://www.analog.com/library/analogDialogue/archives/39-05/
op_amp_applications_handbook.html) Downloadable book. Can also be bought
Operational Amplifier Noise Prediction (All Op Amps) (http://www.intersil.com/data/an/an519.pdf) using
spot noise
Operational Amplifier Basics (http://www.williamson-labs.com/480_opam.htm)
History of the Op-amp (http://www.analog.com/library/analogDialogue/archives/39-05/Web_ChH_final.
pdf) from vacuum tubes to about 2002. Lots of detail, with schematics. IC part is somewhat ADI-centric.
ECE 209: Operational amplifier basics (http://www.tedpavlic.com/teaching/osu/ece209/support/
opamp_basics.pdf) Brief document explaining zero error by naive high-gain negative feedback. Gives single
OpAmp example that generalizes typical configurations.
Loebe Julie historical OpAmp interview by Bob Pease (http://electronicdesign.com/article/
analog-and-mixed-signal/what-s-all-this-julie-stuff-anyhow-6071.aspx)
www.PhilbrickArchive.org (http://www.PhilbrickArchive.org) A free repository of materials from George A
Philbrick / Researches - Operational Amplifier Pioneer

42

Article Sources and Contributors

Article Sources and Contributors


Bode plot Source: http://en.wikipedia.org/w/index.php?oldid=470369397 Contributors: A2Kafir, Andy.ackland, Anna Lincoln, Arcturus4669, Avicennasis, Avoided, BenFrantzDale,
Betacommand, Bobblehead, Brews ohare, Buffetrand, Chaosgate, Clam0p, Clemens vi, Dbtfz, Derdeib, Dicklyon, Dr.K., Drew335, Ec5618, ExportRadical, Falcorian, Fresheneesz, Gaius
Cornelius, Ganzor, Giftlite, Gmoose1, Graibeard, Gravix, Heron, Hooperbloob, Huntthetroll, Jamelan, Jeff P Harris, Jiuguang Wang, Jive Dadson, Joy, Jshen6, KPH2293, Kenyon, Kmellem,
Lambertch, Mdd, Mehran, Mellery, Michael Devore, Michael Hardy, Miguelgoldstein, Mivey6, Mr. PIM, Nabla, Nbarth, Niketmjoshi, Nitin.mehta, Nodlit, Omegatron, Patrick, Peter.scottie,
Phosphoricx, Poulpy, Private Pilot, Redheylin, Rehnn83, Requestion, Shadowjams, SimonP, Sommacal alfonso, Spradlig, Stw, Tim Starling, Toffile, Tomer shalev, User A1, Vukg, WakingLili,
Wtshymanski, Zoomzoom1, 114 anonymous edits
Current mirror Source: http://en.wikipedia.org/w/index.php?oldid=472931941 Contributors: Albert Castillo, Albugwiki, Alfred Centauri, Azaghal of Belegost, Bjf, BorgQueen, Brews ohare,
Chowbok, Circuit dreamer, DabMachine, Dr.K., Editor at Large, Everyking, Hooperbloob, ICE77, Iandiver, Inductiveload, Iridescent, Kungfuadam, Leonard G., Light current, Lindenh248,
Lovibond, Mako098765, Matt B., Mboverload, Msiddalingaiah, Omegatron, Pdn, Petrb, Pgadfor, Rich Farmbrough, Rogerbrent, Rohitbd, Ronz, Royboycrashfan, Rpyle731, Rsashwinkumar,
Searchme, Slasher-fun, Soap, Zangar, 91 anonymous edits
Differential amplifier Source: http://en.wikipedia.org/w/index.php?oldid=470589040 Contributors: Ahoerstemeier, AndreAdrian, Audriusa, AugustWest, Avian, Binksternet, Brews ohare,
Circuit dreamer, Colorbow, Dicklyon, Docu, Estarapapax, Giraultjones, Glenn, GoingBatty, Harumphy, Hooperbloob, ICE77, J.delanoy, John of Reading, JulesH, Jwortzel, Lv131, Maitchy,
Materialscientist, Omegatron, Ong saluri, Prari, RTC, Radagast83, RedWolf, Rogerbrent, Rohitbd, RyanCross, Shadowriver, Siva6085, Spike Wilbury, TedPavlic, Tom.jennings, TuringBirds,
Vivio Testarossa, Woohookitty, Xavier016, Yuckhil, Yves-Laurent, Zen-in, 64 anonymous edits
Operational amplifier Source: http://en.wikipedia.org/w/index.php?oldid=473659752 Contributors: .:Ajvol:., .K, 213.20.48.xxx, 48v, Abhishekchavan79, Acu7, Adashiel, Aitias, Alai,
Alejo2083, AlexPlank, Alexius08, Alfred Centauri, Analogkidr, Andy Dingley, Anitauky, AnnaFrance, Annomination, AnonMoos, Anoneditor, Antireconciler, Arabani, Archimerged,
ArnoldReinhold, Atlant, Audriusa, Aulis Eskola, Awbliven, Bakkster Man, Barticus88, Bdieseldorff, Bemoeial, Berrinkursun, Binksternet, Bizarro Bull, Bobo192, Borgx, Breezeboy, Brews
ohare, Bruno gouveia rodrigues, Bvankuik, Caltas, Can't sleep, clown will eat me, CanDo, CanisRufus, Cataclysm, Cbdorsett, Charles.small, Chris Pressey, Christian75, Circuit dreamer,
Clankypup, Clemwang, Closedmouth, Cocytus, Conversion script, CosineKitty, Crunchy Numbers, CyborgTosser, CyrilB, Czar44, DARTH SIDIOUS 2, DRE, DSatz, Danielop-NJITWILL,
DavidCary, DeadEyeArrow, Dicklyon, Donreed, Dpotter, Dprabhu, Draurbilla, Ds13, Dysprosia, ESkog, East of Borschov, Editor at Large, Edward, Efficacy, Electron9, Elkman, EncMstr,
Endothermic, Eng general, Enon, Eranb, Evaluist, Ezhuttukari, Faradayplank, Favonian, First Harmonic, Foobar, Foobaz, FredStrauss, Frencheigh, Fresheneesz, GRAHAMUK, Gaius Cornelius,
Gdje je nestala dua svijeta, Gene Nygaard, Gggh, Ghewgill, Giftlite, Gimboid13, Glenn, Glrx, Grafen, Ground Zero, Hankwang, Henning Makholm, HermanFinster, Heron, Homer Landskirty,
Hoo man, Hooperbloob, I dream of horses, ICE77, Ianr44, Iinvnt, Inductiveload, Inkling, J.delanoy, JForget, Jaanus.kalde, Jascii, Jcurie, Jerome Baum, Jim1138, Jjwilkerson, John of Reading,
JohnWittle, Johnuniq, Jonnie5, Josemiotto, Jp314159, Julesd, Jwestbrook, KD5TVI, Kasper Meerts, Keenan Pepper, KeithTyler, Khattab01, L Kensington, La Pianista, Lain.cai, Leonard G.,
Light current, Lindosland, LiuyuanChen, LouScheffer, Lovibond, Luk, Luna Santin, M Puddick, M1ss1ontomars2k4, Mahjongg, Maitchy, Malcolm Farmer, Man with two legs, Manavbhardwaj,
Mav, Mboverload, Mebden, Meht7860, Michael Hardy, Mike R, Mike1024, MikeLynch, Mikespedia, Mild Bill Hiccup, Mindmatrix, Mintleaf, Missamo80, Mortense, Msiddalingaiah, Mrten
Berglund, Naspilot, Neonumbers, Nick Number, Nigelj, Novangelis, Oddbodz, Ohconfucius, Oli Filth, Omegatron, Ong saluri, Opamp, OsamaBinLogin, OscarJuan, Overjive, PHermans, Papa
November, Pengo, Peranders, Pgadfor, Phil Christs, Piano non troppo, Pietrow, PleaseStand, Plugwash, Pmoseley, Pol098, Ppj4, Punkguitar, RTC, Raidfibre, Rcsprinter123, Red Thrush,
RedWolf, RexNL, RobertTanzi, Rogerbrent, Rohitbd, Roman12345, Royboycrashfan, Rpyle731, Rumpelstiltskin223, SJP, Saintrain, Sakthivel30, Salam32, Sbmeirow, Searchme, Secret Squrrel,
Sellyme, Serych, Shanes, Sigmundg, Sjakkalle, Skonieczny, Sldghmmr, Smart Viral, Smither, Sn0wflake, Snafflekid, Some jerk on the Internet, Sparkinark, Speight, Spinningspark, Srinivasbt,
Srleffler, Sudeepa123, Sudheerp99, Sunny sin2005, Supav1nnie, Supten, Synaptidude, Szzuk, TDogg310, TWINE006, Tawsifkhan, TedPavlic, Teravolt, The Anome, Theleftorium, Torturella,
Unixxx, Vasurak, Vhann, VictorianMutant, Viscious81, Voidxor, Wbm1058, Wernher, Weyes, Wikigayburgers, WillWare, Woohookitty, Wstorr, Wtshymanski, Yves-Laurent, Zen-in, ZooFari,
neas, , 508 anonymous edits

43

Image Sources, Licenses and Contributors

Image Sources, Licenses and Contributors


Image:Bode High-Pass.PNG Source: http://en.wikipedia.org/w/index.php?title=File:Bode_High-Pass.PNG License: GNU Free Documentation License Contributors: Brews ohare
Image:Bode Low-Pass.PNG Source: http://en.wikipedia.org/w/index.php?title=File:Bode_Low-Pass.PNG License: Creative Commons Attribution-Share Alike Contributors: Brews ohare
Image:Bode Low Pass Magnitude Plot.PNG Source: http://en.wikipedia.org/w/index.php?title=File:Bode_Low_Pass_Magnitude_Plot.PNG License: GNU Free Documentation License
Contributors: Brews ohare
Image:Bode Low Pass Phase Plot.PNG Source: http://en.wikipedia.org/w/index.php?title=File:Bode_Low_Pass_Phase_Plot.PNG License: GNU Free Documentation License Contributors:
Brews ohare
Image:Bode Pole-Zero Magnitude Plot.PNG Source: http://en.wikipedia.org/w/index.php?title=File:Bode_Pole-Zero_Magnitude_Plot.PNG License: GNU Free Documentation License
Contributors: Brews ohare
Image:Bode Pole-Zero Phase Plot.PNG Source: http://en.wikipedia.org/w/index.php?title=File:Bode_Pole-Zero_Phase_Plot.PNG License: GNU Free Documentation License Contributors:
Brews ohare
Image:Magnitude of feedback amplifier.PNG Source: http://en.wikipedia.org/w/index.php?title=File:Magnitude_of_feedback_amplifier.PNG License: Creative Commons Attribution-Share
Alike Contributors: Brews ohare
Image:Phase of feedback amplifier.PNG Source: http://en.wikipedia.org/w/index.php?title=File:Phase_of_feedback_amplifier.PNG License: GNU Free Documentation License Contributors:
Brews ohare
Image:Gain Margin.PNG Source: http://en.wikipedia.org/w/index.php?title=File:Gain_Margin.PNG License: GNU Free Documentation License Contributors: Brews ohare
Image:Phase Margin.PNG Source: http://en.wikipedia.org/w/index.php?title=File:Phase_Margin.PNG License: GNU Free Documentation License Contributors: Brews ohare
Image:Bodeplot.png Source: http://en.wikipedia.org/w/index.php?title=File:Bodeplot.png License: Public Domain Contributors: derivative work: Stw (talk) Bodeplot.JPG: Vukg
Image:Nyquist.svg Source: http://en.wikipedia.org/w/index.php?title=File:Nyquist.svg License: Public Domain Contributors: Engelec
Image:Nichols.svg Source: http://en.wikipedia.org/w/index.php?title=File:Nichols.svg License: Public domain Contributors: Engelec (talk)
Image:Reverse function 450.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Reverse_function_450.jpg License: GNU Free Documentation License Contributors: Cyril Mechkov -Circuit-fantasist
Image:Simple bipolar mirror.svg Source: http://en.wikipedia.org/w/index.php?title=File:Simple_bipolar_mirror.svg License: Creative Commons Attribution-Sharealike 3.0,2.5,2.0,1.0
Contributors: Simple_bipolar_mirror.PNG: Brews ohare derivative work: Azaghal of Belegost (talk)
Image:Simple MOSFET mirror.PNG Source: http://en.wikipedia.org/w/index.php?title=File:Simple_MOSFET_mirror.PNG License: Creative Commons Attribution-Share Alike
Contributors: Brews ohare
Image:Gain-assisted current mirror.svg Source: http://en.wikipedia.org/w/index.php?title=File:Gain-assisted_current_mirror.svg License: Creative Commons Attribution 2.5 Contributors:
Albedo-ukr, Ea91b3dd
Image:Wide-swing MOSFET mirror.svg Source: http://en.wikipedia.org/w/index.php?title=File:Wide-swing_MOSFET_mirror.svg License: Creative Commons Attribution 2.5 Contributors:
Ea91b3dd
Image:Mirror output resistance.PNG Source: http://en.wikipedia.org/w/index.php?title=File:Mirror_output_resistance.PNG License: Creative Commons Attribution-Share Alike
Contributors: Brews ohare
Image:Op-amp symbol.svg Source: http://en.wikipedia.org/w/index.php?title=File:Op-amp_symbol.svg License: GNU Free Documentation License Contributors: User:Omegatron
Image:Difference amplifier.png Source: http://en.wikipedia.org/w/index.php?title=File:Difference_amplifier.png License: GNU Free Documentation License Contributors: User:Rohitbd
Image:Long-tailed-pair.gif Source: http://en.wikipedia.org/w/index.php?title=File:Long-tailed-pair.gif License: GNU Free Documentation License Contributors: User:Rohitbd
Image:Op-Amp Differential Amplifier.svg Source: http://en.wikipedia.org/w/index.php?title=File:Op-Amp_Differential_Amplifier.svg License: Public Domain Contributors: Inductiveload
Image:Ua741 opamp.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Ua741_opamp.jpg License: Creative Commons Attribution 3.0 Contributors: Teravolt (talk). Original uploader
was Teravolt at en.wikipedia
Image:Op-amp open-loop 1.svg Source: http://en.wikipedia.org/w/index.php?title=File:Op-amp_open-loop_1.svg License: GNU Free Documentation License Contributors: Ong saluri
Image:Operational amplifier noninverting.svg Source: http://en.wikipedia.org/w/index.php?title=File:Operational_amplifier_noninverting.svg License: GNU Free Documentation License
Contributors: Ong saluri
Image:Op-Amp Internal.svg Source: http://en.wikipedia.org/w/index.php?title=File:Op-Amp_Internal.svg License: Public Domain Contributors: Inductiveload
Image:OpAmpTransistorLevel Colored Labeled.svg Source: http://en.wikipedia.org/w/index.php?title=File:OpAmpTransistorLevel_Colored_Labeled.svg License: Creative Commons
Attribution-ShareAlike 3.0 Unported Contributors: Daniel Braun
Image:Block Diagram for Feedback.svg Source: http://en.wikipedia.org/w/index.php?title=File:Block_Diagram_for_Feedback.svg License: Creative Commons Attribution-Sharealike
3.0,2.5,2.0,1.0 Contributors: Block_diagram_for_feedback.PNG: Brews ohare derivative work: Serenthia
Image:Generic 741 pinout top.png Source: http://en.wikipedia.org/w/index.php?title=File:Generic_741_pinout_top.png License: Creative Commons Attribution-Sharealike 3.0 Contributors:
TedPavlic
Image:Op-Amp Non-Inverting Amplifier.svg Source: http://en.wikipedia.org/w/index.php?title=File:Op-Amp_Non-Inverting_Amplifier.svg License: Public Domain Contributors:
Inductiveload
Image:Op-Amp Inverting Amplifier.svg Source: http://en.wikipedia.org/w/index.php?title=File:Op-Amp_Inverting_Amplifier.svg License: Public Domain Contributors: Inductiveload
Image:K2-w vaccuum tube op-amp.jpg Source: http://en.wikipedia.org/w/index.php?title=File:K2-w_vaccuum_tube_op-amp.jpg License: Creative Commons Attribution-Sharealike 3.0
Contributors: Bdieseldorff
Image:Discrete opamp.png Source: http://en.wikipedia.org/w/index.php?title=File:Discrete_opamp.png License: Creative Commons Attribution-Sharealike 3.0 Contributors: Analog Devices
Image:Modular opamp.png Source: http://en.wikipedia.org/w/index.php?title=File:Modular_opamp.png License: Creative Commons Attribution-Sharealike 3.0 Contributors: Analog Devices
Image:Hybrid opamp.png Source: http://en.wikipedia.org/w/index.php?title=File:Hybrid_opamp.png License: Creative Commons Attribution-Sharealike 3.0 Contributors: Analog Devices
Image:Lm356.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Lm356.jpg License: Creative Commons Attribution-Sharealike 3.0 Contributors: Bdieseldorff

44

License

License
Creative Commons Attribution-Share Alike 3.0 Unported
//creativecommons.org/licenses/by-sa/3.0/

45

Você também pode gostar