Pole splitting

Pole splitting

Pole splitting is a type of frequency compensation used in an electronic amplifier in which a capacitor is introduced between the input and output sides of the amplifier with the intention of moving the pole lowest in frequency (usually an input pole) to lower frequencies and the pole next in frequency (usually an output pole) to higher frequency. This pole movement increases the stability of the amplifier and improves its step response at the cost of decreased speed. [ That is, the rise time is selected to be the fastest possible consistent with low overshoot and ringing.] cite book
author=C. Toumazu, Moschytz GS & Gilbert B (Editors)
title=Trade-offs in analog circuit design: the designer's companion
year= 2007
pages=pp. 272-275
publisher=Springer
location=New York/Berlin/Dordrecht
isbn= 1402070373 |url=http://books.google.com/books?id=VoBIOvirkiMC&pg=PA272&lpg=PA272&dq=%22pole+splitting%22&source=web&ots=MC083mOWhv&sig=duZQKaGECaAH80qDj-YNMdRd8nA
] cite book
author=Marc T. Thompson
title=Intuitive analog circuit design: a problem-solving approach using design case studies
year= 2006
pages=p. 200
publisher=Elsevier Newnes
location=Amsterdam
isbn= 0750677864 |url=http://books.google.com/books?id=1Tyzjmf0DI8C&pg=PA200&dq=pole+splitting+analog+amplifier&lr=&as_brr=0&sig=gmvG9dtlK48hcqpvf3NwwqcF2Hk
] cite book
author=Wally M. C. Sansen
title=Analog design essentials
year= 2006
pages=§097, p. 266 "et seq"
publisher=Springer
location=New York; Berlin
isbn=0-387-25746-2
url=http://worldcat.org/isbn/0-387-25746-2
]

Example of pole splitting

This example shows that introduction of the capacitor "CC" in the amplifier of Figure 1 has two results: first it causes the lowest frequency pole of the amplifier to move still lower in frequency and second, it causes the higher pole to move higher in frequency. [Although this example appears very specific, the associated mathematical analysis is very much used in circuit design.] The amplifier of Figure 1 has a low frequency pole due to the added input resistance "Ri" and capacitance "Ci", with the time constant "Ci" ( "RA // Ri" ). This pole is moved down in frequency by the Miller effect. The amplifier is given a high frequency output pole by addition of the load resistance "RL" and capacitance "CL", with the time constant "CL" (" Ro // RL" ). The upward movement of the high-frequency pole occurs because the Miller-amplified compensation capacitor "CC" alters the frequency dependence of the output voltage divider.

The first objective, to show the lowest pole moves down in frequency, is established using the same approach as the Miller's theorem article. Following the procedure described in the article on Miller's theorem, the circuit of Figure 1 is transformed to that of Figure 2, which is electrically equivalent to Figure 1. Application of Kirchhoff's current law to the input side of Figure 2 determines the input voltage v_i to the ideal op amp as a function of the applied signal voltage v_a, namely,

::

frac {v_i} {v_a} = frac {R_i} {R_i+R_A} frac {1} {1+j omega (C_M+C_i) (R_A//R_i)} ,

which exhibits a roll-off with frequency beginning at "f1" (say) where

::

egin{align} f_{1} & = frac {1} {2 pi (C_M+C_i)(R_A//R_i) } \ & = frac {1} {2 pi au_1} , \end{align}

which introduces notation au_1 for the time constant of the lowest pole. Clearly, this frequency is lower than the initial low frequency of the amplifier, which for "CC" = 0 F is just 1/ [ 2 π "Ci" ( "RA // Ri" )] .

Turning to the second objective, showing the higher pole moves still higher in frequency, it is necessary to look at the output side of the circuit, which contributes a second factor to the overall gain, and additional frequency dependence. The voltage v_o is determined by the gain of the ideal op amp inside the amplifier as

:: v_o = A_v v_i .

Using this relation and applying Kirchhoff's current law to the output side of the circuit determines the load voltage v_{ell} as a function of the voltage v_{i} at the input to the ideal op amp as:

:: frac {v_{ell {v_i} = A_v frac {R_L} {R_L+R_o},!sdot frac {1+j omega C_C R_o/A_v } {1+j omega (C_L + C_C ) (R_o//R_L) } .

This expression is combined with the gain factor found earlier for the input side of the circuit to obtain the overall gain as

::

frac {v_{ell {v_a} = frac {v_{ell{v_i} frac {v_i} {v_a}

:::= A_v frac {R_i} {R_i+R_A}sdot frac {R_L} {R_L+R_o},! sdot frac {1} {1+j omega (C_M+C_i) (R_A//R_i)} ,! sdot frac {1+j omega C_C R_o/A_v } {1+j omega (C_L + C_C ) (R_o//R_L) } .

This gain formula appears to show a simple two-pole response with two time constants. (It also exhibits a zero in the numerator but, assuming the amplifier gain "Av" is large, this zero is important only at frequencies too high to matter in this discussion , so the numerator can be approximated as unity.) However, although the amplifier does have a two-pole behavior, the two time-constants are more complicated than the above expression suggests because the Miller capacitance contains a buried frequency dependence that has no importance at low frequencies, but has considerable effect at high frequencies. That is, assuming the output "R-C" product, "CL" ( "Ro // RL" ), corresponds to a frequency well above the low frequency pole, the accurate form of the Miller capacitance must be used, rather than the Miller approximation. According to the article on Miller effect, the Miller capacitance is given by

::egin{align}C_M & = C_C left( 1 - frac {v_{ell {v_i} ight) \ & = C_C left( 1 - A_v frac {R_L} {R_L+R_o} frac {1+j omega C_C R_o/A_v } {1+j omega (C_L + C_C ) (R_o//R_L) } ight ) . \end{align}

(For a positive Miller capacitance, "Av" is negative.) Upon substitution of this result into the gain expression and collecting terms, the gain is rewritten as:

:: frac {v_{ell {v_a} = A_v frac {R_i} {R_i+R_A} frac {R_L} {R_L+R_o} frac {1+j omega C_C R_o/A_v } {D_{ omega ,

with "Dω" given by a quadratic in ω, namely:

::D_{ omega },! = [1+j omega (C_L+C_C) (R_o//R_L)] ,! sdot [ 1+j omega C_i (R_A//R_i)] ,! +j omega C_C (R_A//R_i),! sdot left( 1-A_v frac {R_L} {R_L+R_O} ight) ,! +(j omega) ^2 C_C C_L (R_A//R_i) (R_O//R_L) .

Every quadratic has two factors, and this expression looks simpler if it is rewritten as

:: D_{ omega } =(1+j omega { au}_1 )(1+j omega { au}_2 )

::: = 1 + j omega ( { au}_1+{ au}_2) ) +(j omega )^2 au_1 au_2 ,

where au_1 and au_2 are combinations of the capacitances and resistances in the formula for "Dω". [The sum of the time constants is the coefficient of the term linear in jω and the product of the time constants is the coefficient of the quadratic term in (jω)2.] They correspond to the time constants of the two poles of the amplifier. One or the other time constant is the longest; suppose au_1 is the longest time constant, corresponding to the lowest pole, and suppose au_1 >> au_2. (Good step response requires au_1 >> au_2. See Selection of CC below.)

At low frequencies near the lowest pole of this amplifier, ordinarily the linear term in ω is more important than the quadratic term, so the low frequency behavior of "Dω" is:

::egin{align} D_{ omega } & = 1+ j omega [(C_M+C_i) (R_A//R_i) +(C_L+C_C) (R_o//R_L)] \ & = 1+j omega ( au_1 + au_2) approx 1 + j omega au_1 , \end{align}

where now "CM" is redefined as

:: C_M= C_C left( 1 - A_v frac {R_L}{R_L+R_o} ight) ,

which is simply the previous Miller capacitance evaluated at low frequencies, in other words, using the Miller approximation. On this basis au_1 is determined, provided au_1 >> au_2. Because "CM" is large, the time constant { au}_1 is much larger than its original value of "Ci" ( "RA // Ri" ). [ The expression for au_1 differs a little from ( "CM+Ci" ) ( "RA" // "Ri" ) as found initially for "f1", but the difference is minor assuming the load capacitance is not so large that it controls the low frequency response instead of the Miller capacitance.]

At high frequencies the quadratic term becomes important. Assuming the above result for au_1 is valid, the second time constant, the position of the high frequency pole, is found from the quadratic term in "Dω" as

:: au_2 = frac { au_1 au_2} { au_1} approx frac { au_1 au_2} { au_1 + au_2} .

Substituting in this expression the quadratic coefficient corresponding to the product au_1 au_2 along with the estimate for au_1, an estimate for the position of the second pole is found:

::egin{align} au_2 & = frac {(C_C C_L +C_L C_i+C_i C_C)(R_A//R_i) (R_O//R_L) } {(C_M+C_i) (R_A//R_i) +(C_L+C_C) (R_o//R_L)} \ & approx frac {C_C C_L +C_L C_i+C_i C_C} {C_M} (R_O//R_L) , \end{align}

and because "CM" is large, it seems au_2 is reduced in size from its original value "CL" ( "Ro" // "RL" ); that is, the higher pole has moved still higher in frequency because of "CC". [As an aside, the higher the high-frequency pole is made in frequency, the more likely it becomes for a real amplifier that other poles (not considered in this analysis) play a part.]

In short, introduction of capacitor "CC" moved the low pole lower and the high pole higher, so the term pole splitting seems a good description.

Selection of CC

What value is a good choice for "CC"? For general purpose use, traditional design (often called "dominant-pole" or "single-pole compensation") requires the amplifier gain to drop at 20 dB/decade from the corner frequency down to 0 dB gain, or even lower. cite book
author=A.S. Sedra and K.C. Smith
title=Microelectronic circuits
year= 2004
pages=p. 849 and Example 8.6, p. 853
publisher=Oxford University Press
edition=Fifth Edition
location=New York
isbn= 0-19-514251-9
url=http://worldcat.org/isbn/0-19-514251-9
] cite book
author=Huijsing, Johan H.
title=Operational amplifiers: theory and design
year= 2001
pages=§6.2, pp.205-206 and Figure 6.2.1
publisher=Kluwer Academic
location=Boston, MA
isbn= 0-7923-7284-0
url=http://books.google.com/books?id=tiuV_agzk_EC&pg=PA102&dq=isbn:0792372840&sig=d-oEw_n992coA6bU0h6gkoJzoUo#PPA206,M1
] With this design the amplifier is stable and has near-optimal step response even as a unity gain voltage buffer. A more aggressive technique is two-pole compensation. [ Feucht, Dennis: [http://www.analogzone.com/col_0719.pdf "Two-pole compensation"] ] cite book
author=Self, Douglas
title=Audio power amplifier design handbook
year= 2006
pages=pp.191-193
publisher=Newnes
location=Oxford
isbn= 0750680725
url=http://books.google.com/books?id=BRQZppvawWwC&pg=PA191&lpg=PA191&dq=%22two+pole+compensation%22&source=web&ots=qsxRG-z1Xl&sig=41uVzeYZW3vi3BndJORUNHNZqPY#PPA191,M1
]

The way to position "f"2 to obtain the design is shown in Figure 3. At the lowest pole "f"1, the Bode gain plot breaks slope to fall at 20 dB/decade. The aim is to maintain the 20 dB/decade slope all the way down to zero dB, and taking the ratio of the desired drop in gain (in dB) of 20 log10 "Av" to the required change in frequency (on a log frequency scale [That is, the frequency is plotted in powers of ten, as 1, 10, 102 "etc".] ) of ( log10 "f"2 − log10 "f"1 ) = log10 ( "f"2 / "f"1 ) the slope of the segment between "f"1 and "f"2 is:

::Slope per decade of frequency =20 frac {mathrm{log_{10 ( A_v )} {mathrm{log_{10 (f_2 / f_1 ) } ,

which is 20 dB/decade provided "f2 = Av f1" . If "f2" is not this large, the second break in the Bode plot that occurs at the second pole interrupts the plot before the gain drops to 0 dB with consequent lower stability and degraded step response.

Figure 3 shows that to obtain the correct gain dependence on frequency, the second pole is at least a factor "Av" higher in frequency than the first pole. The gain is reduced a bit by the voltage dividers at the input and output of the amplifier, so with corrections to "Av" for the voltage dividers at input and output the pole-ratio condition for good step response becomes:

:: frac { au_1} { au_2} approx A_v frac {R_i} {R_i+R_A}sdot frac {R_L} {R_L+R_o} ,

Using the approximations for the time constants developed above,

:: frac { au_1} { au_2} approx frac {( au_1 + au_2 ) ^2} { au_1 au_2} approx A_v frac {R_i} {R_i+R_A}sdot frac {R_L} {R_L+R_o} ,

or

:: frac { [(C_M+C_i) (R_A//R_i) +(C_L+C_C) (R_o//R_L)] ^2} {(C_C C_L +C_L C_i+C_i C_C)(R_A//R_i) (R_O//R_L) } ,! {color{White}sdot} = A_v frac {R_i} {R_i+R_A}sdot frac {R_L} {R_L+R_o} ,

which provides a quadratic equation to determine an appropriate value for "CC". Figure 4 shows an example using this equation. At low values of gain this example amplifier satisfies the pole-ratio condition without compensation, but as gain increases, a compensation capacitance rapidly becomes necessary because the necessary pole ratio increases with gain. For still larger gain, the necessary "CC" drops with increasing gain because the Miller amplification of "CC", which increases with gain, allows a smaller value for "CC".

To provide more safety margin for design uncertainties, often "Av" is increased to two or three times "Av" on the right side of this equation. [A factor of two results in the "maximally flat" or Butterworth design for a two-pole amplifier. However, real amplifiers have more than two poles, and a factor greater than two often is necessary.] See Sansen or Huijsing and article on step response.

lew rate

The above is a small-signal analysis. However, when large signals are used, the need to charge and discharge the compensation capacitor adversely affects the amplifier slew rate; in particular, the response to an input ramp signal is limited by the need to charge "CC".

See also

* Frequency compensation
* Step response
* Miller effect
* Common source
* Bode plot
* Step response

References and notes

External links

* [http://en.wikibooks.org/wiki/Circuit_Theory/Bode_Plots Bode Plots] in the Circuit Theory Wikibook
* [http://en.wikibooks.org/wiki/Control_Systems/Bode_Plots Bode Plots] in the Control Systems Wikibook


Wikimedia Foundation. 2010.

Игры ⚽ Нужно решить контрольную?

Look at other dictionaries:

  • Frequency compensation — In electrical engineering, frequency compensation is a technique used in amplifiers, and especially in amplifiers employing negative feedback. It usually has two primary goals: To avoid the unintentional creation of positive feedback, which will… …   Wikipedia

  • Step response — The step response of a system in a given initial state consists of the time evolution of its outputs when its control inputs are Heaviside step functions. In electronic engineering and control theory, step response is the time behaviour of the… …   Wikipedia

  • Negative feedback amplifier — Figure 1: Ideal negative feedback model A negative feedback amplifier (or more commonly simply a feedback amplifier) is an amplifier which combines a fraction of the output with the input so that a negative feedback opposes the original signal.… …   Wikipedia

  • Open-circuit time constant method — The open circuit time constant method is an approximate analysis technique used in electronic circuit design to determine the corner frequency of complex circuits. It also is known as the zero value time constant technique. The method provides a… …   Wikipedia

  • Common source — Figure 1: Basic N channel JFET common source circuit (neglecting biasing details). Figure 2: Basic N channel JFET common source circuit with source degeneration …   Wikipedia

  • Miller effect — In electronics, the Miller effect accounts for the increase in the equivalent input capacitance of an inverting voltage amplifier due to amplification of the effect of capacitance between the input and output terminals. The virtually increased… …   Wikipedia

  • Full state feedback — (FSF), or pole placement, is a method employed in feedback control system theory to place the closed loop poles of a plant in pre determined locations in the s plane. Placing poles is desirable because the location of the poles corresponds… …   Wikipedia

  • Feedback — For other uses, see Feedback (disambiguation). Feedback describes the situation when output from (or information about the result of) an event or phenomenon in the past will influence an occurrence or occurrences of the same (i.e. same defined)… …   Wikipedia

  • Fully differential amplifier — A fully differential amplifier, usually referred to as an FDA for brevity, is a DC coupled high gain electronic voltage amplifier with differential inputs and differential outputs. In its ordinary usage, the output of the FDA is controlled by two …   Wikipedia

  • Phase margin — In electronic amplifiers, phase margin is the difference, measured in degrees, between the phase of the amplifier s output signal and 360°. In feedback amplifiers, the phase margin is measured at the frequency at which the open loop voltage gain… …   Wikipedia

Share the article and excerpts

Direct link
Do a right-click on the link above
and select “Copy Link”