User:John R. Brews/WP Import: Difference between revisions

From Citizendium
Jump to navigation Jump to search
imported>John R. Brews
imported>John R. Brews
No edit summary
Line 1: Line 1:
{{TOC|right}}
{{TOC|right}}
In [[electronics]], the '''Miller effect''' accounts for the increase in the equivalent input [[capacitance]] of an inverting voltage [[amplifier]] due to amplification of the capacitance between the input and output terminals. The additional input capacitance due to the Miller effect is given by
:<math>C_{M}=C (1+A_v)\,</math>
where <math>-A_v</math> is the gain of the amplifier and C is the feedback capacitance.
Although the term ''Miller effect'' normally refers to capacitance, any impedance connected between the input and another node exhibiting gain can modify the amplifier input impedance via this effect. These properties of Miller effect are generalized by [[Miller theorem]].


The '''hybrid-pi model''' is a popular [[circuit]] model used for analyzing the [[small signal]] behavior of [[transistors]]. The model can be quite accurate for low-frequency circuits and can easily be adapted for higher frequency circuits with the addition of appropriate inter-electrode [[capacitance]]s and other parasitic elements.


==Bipolar transistor==
== History ==
{{Image|Bipolar hybrid-pi model.PNG|right|200px| Simplified, low-frequency hybrid-pi [[BJT]] model.}}
The Miller effect was named after [[John Milton Miller]].<ref name=Miller/> When Miller published his work in 1920, he was working on [[vacuum tube]] triodes, however the same theory applies to more modern devices such as bipolar and MOS [[transistors]].
{{Image|Bipolar hybrid-pi capacitances.PNG|right|300px|Bipolar hybrid-pi model with parasitic capacitances.}}
The hybrid-pi model is a linearized, ''y''-parameter [[Two-port_network#Admittance_parameters_.28y-parameters.29|two-port network]] approximation to the transistor using the small-signal base-emitter voltage ''v<sub>&pi;</sub>'' and collector-emitter voltage ''v<sub>ce</sub>'' as independent variables, and the small-signal base current ''i<sub>b</sub>'' and collector current ''i<sub>c</sub>'' as dependent variables. (See Jaeger and Blalock.<ref name=Jaeger1/>)


A basic, low-frequency hybrid-pi model for the [[bipolar transistor]] is shown in the figure. The three transistor terminals are ''E'' = emitter, ''B'' = base, and ''C'' = collector. The base-emitter connection is through a resistor ''r<sub>&pi;</sub>'', and the base current causes a small-signal voltage drop across it, ''v<sub>&pi;</sub>'' (the &pi; notation is standard). Voltage ''v<sub>&pi;</sub>'' induces a small-signal collector current ''via'' the voltage-controlled current source with current ''g<sub>m</sub>v<sub>&pi;</sub>'', ''g<sub>m</sub>'' is the transistor ''transconductance''.
== Derivation ==
[[Image:Miller cir.png|right|frame|Figure 1: Circuit for deriving the Miller effect, with ideal inverting voltage amplifier]]
Consider an ideal inverting voltage [[amplifier]] of gain <math>-A_v</math> with an [[Electrical impedance|impedance]] <math>Z</math> connected between its input and output nodes. The output voltage is therefore <math>V_o =- A_v V_i</math>.  Assuming that the amplifier input draws no current, all of the input current flows through <math>Z</math>, and is therefore given by


The various parameters are as follows.
:<math>I_i = \frac{V_i - V_o}{Z} = \frac{V_i (1 + A_v)}{Z}</math>.


*<math>g_m = \frac{i_{c}}{v_{be}}\Bigg |_{v_{ce}=0} = \begin{matrix}\frac {I_\mathrm{C}}{ V_\mathrm{T} }\end{matrix} </math> is the [[transconductance]] in [[Siemens (unit)|siemens]], evaluated in a simple model (see Jaeger and Blalock<ref name=Jaeger/>)
The input impedance of the circuit is  
:where:
:* <math>I_\mathrm{C} \,</math> is the [[quiescent]] collector current (also called the collector bias or DC collector current)
:* <math>V_\mathrm{T} = \begin{matrix}\frac {kT}{ q}\end{matrix}</math> is the ''[[Boltzmann constant#Role in semiconductor physics: the thermal voltage|thermal voltage]]'', calculated from [[Boltzmann's constant]], the [[elementary charge|charge on an electron]], and the transistor temperature in [[kelvin]]s. At 290 K (an approximation to room temperature: 70°F ≈ 294 K) ''V<sub>T</sub>'' is very nearly 25 mV ([http://www.google.com/search?hl=en&q=290+kelvin+*+k+%2F+elementary+charge+in+millivolts+%3D Google calculator]).
* <math>r_{\pi} = \frac{v_{be}}{i_{b}}\Bigg |_{v_{ce}=0} = \frac{\beta_0}{g_m} = \frac{V_\mathrm{T}}{I_\mathrm{B}} \,</math> in [[Ohm (unit)|ohm]]s
:where:
:* <math>\beta_0 = \frac{I_\mathrm{C}}{I_\mathrm{B}} \,</math> is the current gain at low frequencies (commonly called h<sub>FE</sub>). Here <math>I_B</math> is the Q-point base current.  This is a parameter specific to each transistor, and can be found on a datasheet; <math>\beta</math> is a function of the choice of collector current.
*<math> r_O = \frac{v_{ce}}{i_{c}}\Bigg |_{v_{be}=0} = \begin{matrix}\frac {V_A+V_{CE}}{I_C}\end{matrix} \approx \begin{matrix} \frac {V_A}{I_C}\end{matrix}</math> is the output resistance due to the [[Early effect]], and ''V<sub>A</sub>'' is called the ''Early voltage''.


===Higher frequencies===
:<math>Z_{in} = \frac{V_i}{I_i} = \frac{Z}{1+A_v}.</math>
As frequencies are increased, amplifier gain drops off with increasing frequency because the bipolar transistor is degraded by parasitic capacitances. These effects must be modeled to assess the limitations of bipolar circuits. Three parasitic elements are commonly added to the hybrid-pi model to make frequency predictions. One is the base-emitter capacitance ''C<sub>&pi;</sub>'', which includes the base-emitter [[diffusion capacitance]] and the base-emitter junction capacitance. At high enough frequencies, ''C<sub>&pi;</sub>'' shorts out the transistor and it fails to function. A second parasitic capacitance ''C<sub>&mu;</sub>'' connects the base to the collector. Although this capacitor is small, its influence is multiplied by the transistor gain which amplifies the voltage variation at the base connection and applies it to the collector connection, exaggerating the influence of this capacitor through the [[Miller effect]]. The small series base resistance ''r<sub>x</sub>'' plays a role only when there is no other resistance introduced by the circuit into the base lead.


===Related terms===
If Z represents a capacitor with impedance <math>Z = \frac{1}{s C}</math>, the resulting input impedance is
The reciprocal of the output resistance is named the '''output conductance'''
:*<math>g_{ce} = \frac {1} {r_O} </math>.


The reciprocal of g<sub>m</sub> is called the '''intrinsic resistance'''
:<math>Z_{in} = \frac{1}{s C_{M}} \quad \mathrm{where} \quad C_{M}=C (1+A_v).</math>
:*<math>r_{E} = \frac {1} {g_m} </math>.


==MOSFET parameters==
Thus the effective or '''Miller capacitance''' ''C<sub>M</sub>'' is the physical ''C'' multiplied by the factor <math>(1+A_v)</math><ref name=Spencer/>.
{{Image|MOSFET hybrid-pi with capactances.PNG|right|300px|Simplified, three-terminal MOSFET hybrid-pi model.}}
{{Image|MOSFET four-terminal hybrid-pi circuit.PNG|right|350px|Four-terminal small-signal MOSFET circuit.}}
Unlike the bipolar transistor, the [[MOSFET]] fundamentally is a ''four''-terminal, not a three-terminal, device, the terminals being ''S'' = source, ''D'' = drain, ''G'' = gate, and ''B'' = body or substrate. The body is not normally used deliberately as a signal node, and the analogy to the bipolar device is stretched a bit to minimize differences between these devices.


The upper figure shows the three-terminal small-signal equivalent circuit for the MOSFET, which ignores the body contact altogether. At low frequencies,the capacitances all are replaced with open circuits to obtain a frequency-independent circuit depending upon resistors only. The various parameters are as follows.
== Effects ==


*<math>g_m = \frac{i_{d}}{v_{gs}}\Bigg |_{v_{ds}=0}</math>  
As most amplifiers are inverting (i.e. <math>A_v < 0</math>), the effective capacitance at their inputs is increased due to the Miller effect. This can lower the bandwidth of the amplifier, reducing its range of operation to lower frequencies. The tiny junction and stray capacitances between the base and collector terminals of a [[Darlington transistor]], for example, may be drastically increased by the Miller effects due to its high gain, lowering the high frequency response of the device.


is the [[transconductance]] in [[Siemens (unit)|siemens]], evaluated in the Shichman-Hodges model in terms of the [[Q-point]] drain current <math> I_D</math> by (see Jaeger and Blalock<ref name=Jaeger2/>):
It is also important to note that the Miller capacitance is the capacitance seen looking into the input.  If looking for all of the [[RC time constant]]s (poles) it is important to include as well the capacitance seen by the output.  The capacitance on the output is often neglected since it sees <math>{C}({1-1/A_v})</math> and amplifier outputs are typically low impedance.  However if the amplifier has a high impedance output, such as if a gain stage is also the output stage, then this RC can have a [[open-circuit time constant method|significant impact]] on the performance of the amplifier.  This is when [[pole splitting]] techniques are used. 


:::<math>\ g_m = \begin{matrix}\frac {2I_\mathrm{D}}{ V_{\mathrm{GS}}-V_\mathrm{th} }\end{matrix}</math>,
The Miller effect may also be exploited to synthesize larger capacitors from smaller ones. One such example is in the stabilization of [[negative feedback amplifier|feedback amplifiers]], where the required capacitance may be too large to practically include in the circuit. This may be particularly important in the design of [[integrated circuit]], where capacitors can consume significant area, increasing costs.
:where:
::<math>I_\mathrm{D} </math> is the [[quiescent]] drain current (also called the drain bias or DC drain current)
::<math>V_{th}</math> = [[threshold voltage]] and <math>V_{GS}</math> = gate-to-source voltage.


The combination:
===Mitigation===
The Miller effect may be undesired in many cases, and approaches may be sought to lower its impact. Several such techniques are used in the design of amplifiers.


:: <math>\ V_{ov}=( V_{GS}-V_{th})</math>  
A current buffer stage may be added at the output to lower the gain <math>A_v</math> between the input and output terminals of the amplifier (though not necessarily the overall gain). For example, a [[common base]] may be used as a current buffer at the output of a [[common emitter]] stage, forming a [[cascode]]. This will typically reduce the Miller effect and increase the bandwidth of the amplifier.


often is called the ''overdrive voltage''.
Alternatively, a voltage buffer may be used before the amplifier input, reducing the effective source impedance seen by the input terminals. This lowers the <math>RC</math> time constant of the circuit and typically increases the bandwidth.


*<math> r_O = \frac{v_{ds}}{i_{d}}\Bigg |_{v_{gs}=0}</math> is the output resistance due to [[channel length modulation]], calculated using the Shichman-Hodges model as
== Impact on frequency response ==
[[Image:Miller before transform.PNG|thumbnail|250px|Figure 2: Operational amplifier with feedback capacitor ''C<sub>C</sub>''.]]
:::<math>r_O = \begin{matrix}\frac {1/\lambda+V_{DS}}{I_D}\end{matrix} \approx \begin{matrix} \frac {V_E L}{I_D}\end{matrix} </math>,
[[Image:Miller after transform.PNG|thumbnail|250px|Figure 3: Circuit of Figure 2 transformed using Miller's theorem, introducing the '''Miller capacitance''' on the input side of the circuit.]]
using the approximation for the '''channel length modulation''' parameter &lambda;<ref name=Sansen/>
Figure 2 shows an example of Figure 1 where the impedance coupling the input to the output is the coupling capacitor ''C<sub>C</sub>''. A [[Thévenin's theorem|Thévenin voltage]] source ''V<sub>A</sub>'' drives the circuit with Thévenin resistance ''R<sub>A</sub>''. At the output a parallel ''RC''-circuit serves as load. (The load is irrelevant to this discussion: it just provides a path for the current to leave the circuit.) In Figure 2, the coupling capacitor delivers a current jω''C<sub>C</sub>( v<sub>i</sub> - v<sub>o</sub> )'' to the output circuit.
:::<math> \lambda =\begin{matrix} \frac {1}{V_E L} \end{matrix} </math>.
Here ''V<sub>E</sub>'' is a technology related parameter (about 4 V / μm for the [[65 nanometer|65 nm]] technology node<ref name = Sansen/>)  and ''L'' is the length of the source-to-drain separation.


The reciprocal of the output resistance is named the '''drain conductance'''
Figure 3 shows a circuit electrically identical to Figure 2 using Miller's theorem. The coupling capacitor is replaced on the input side of the circuit by the Miller capacitance ''C<sub>M</sub>'', which draws the same current from the driver as the coupling capacitor in Figure 2. Therefore, the driver sees exactly the same loading in both circuits. On the output side, a dependent current source in Figure 3 delivers the same current to the output as does the coupling capacitor in Figure 2. That is, the ''R-C''-load sees the same current in Figure 3 that it does in Figure 2.
*<math>g_{ds} = \frac {1} {r_O} </math>.


It must be pointed out that the Shichman-Hodges model has absolutely no claim to accuracy, especially in today's technology. However, it still is widely used in textbooks to get the flavor of MOSFET circuit design. Nobody would actually implement a circuit expecting this model to be quantitative, however, and real design work always involves extensive computer modeling, for example, using the BSIM computer model.<ref name=BSIM/>
In order that the Miller capacitance draw the same current in Figure 3 as the coupling capacitor in Figure 2, the Miller transformation is used to relate ''C<sub>M</sub>'' to ''C<sub>C</sub>''. In this example, this transformation is equivalent to setting the currents equal, that is
::<math>\  j\omega C_C ( v_i - v_O ) = j \omega C_M v_i, </math>
or, rearranging this equation
:: <math> C_M = C_C \left( 1 + \frac {v_o} {v_i} \right )  = C_C (1 + A_v). </math>
This result is the same as ''C<sub>M</sub>'' of the ''Derivation Section''.


The lower figure explicitly introduces the body contact and the capacitances to the body. This circuit is a great deal more complicated than the three-terminal circuit, and is avoided where possible. However, in some circuits the body is not held at signal ground, and the body must be explicitly handled.
The present example with ''A<sub>v</sub>'' frequency independent shows the implications of the Miller effect, and therefore of  ''C<sub>C</sub>'', upon the frequency response of this circuit, and is typical of the impact of the Miller effect (see, for example, [[common source]]). If ''C<sub>C</sub>'' = 0 F, the output voltage of the circuit is simply ''A<sub>v</sub> v<sub>A</sub>'', independent of frequency. However, when ''C<sub>C</sub>'' is not zero, Figure 3 shows the large Miller capacitance appears at the input of the circuit. The voltage output of the circuit now becomes


==References and notes==
::<math> v_o =- A_v v_i = A_v \frac {v_A} {1+j \omega C_M R_A}, </math>


{{reflist |refs=
and rolls off with frequency once frequency is high enough that ω ''C<sub>M</sub>R<sub>A</sub>'' ≥ 1. It is a [[low-pass filter]]. In analog amplifiers this curtailment of frequency response is a major implication of the Miller effect. In this example, the frequency ω''<sub>3dB</sub>'' such that ω''<sub>3dB</sub>'' ''C<sub>M</sub>R<sub>A</sub>'' = 1 marks the end of the low-frequency response region and sets the [[Bandwidth (signal processing)|bandwidth]] or [[cutoff frequency]] of the amplifier.
<ref name=Jaeger1>  
{{cite book
|author=R.C. Jaeger and T.N. Blalock
|title=Microelectronic Circuit Design
|year= 2004
|edition=Second Edition
|publisher=McGraw-Hill
|location=New York
|isbn=0-07-232099-0
|pages=Section 13.5, esp. Eqs. 13.19
|url=http://worldcat.org/isbn/0072320990}}
</ref>)


<ref name=Jaeger>
It is important to notice that the effect of ''C''<sub>M</sub> upon the amplifier bandwidth is greatly reduced for low impedance drivers (''C''<sub>M</sub> ''R''<sub>A</sub> is small if ''R''<sub>A</sub> is small). Consequently, one way to minimize the Miller effect upon bandwidth is to use a low-impedance driver, for example, by interposing a [[voltage follower]] stage between the driver and the amplifier, which reduces the apparent driver impedance seen by the amplifier.
{{cite book
|author=R.C. Jaeger and T.N. Blalock
|title=Eq. 5.45 pp. 242 and Eq. 13.25 p. 682
|isbn=0-07-232099-0
|url=http://worldcat.org/isbn/0072320990}}
</ref>)


<ref name=Jaeger2>
The output voltage of this simple circuit is always ''A<sub>v</sub> v<sub>i</sub>''. However, real amplifiers have output resistance. If the amplifier output resistance is included in the analysis, the output voltage exhibits a more complex frequency response and the impact of the frequency-dependent current source on the output side must be taken into account.<ref name = PoleSplitting/> Ordinarily these effects show up only at frequencies much higher than the [[roll-off]] due to the Miller capacitance, so the analysis presented here is adequate to determine the useful frequency range of an amplifier dominated by the Miller effect.
{{cite book
|author=R.C. Jaeger and T.N. Blalock
|title=Eq. 4.20 pp. 155 and Eq. 13.74 p. 702
|isbn=0-07-232099-0
|url=http://worldcat.org/isbn/0072320990}}
</ref>


<ref name=Sansen>
===Miller approximation===
{{cite book
This example also assumes ''A<sub>v</sub>'' is frequency independent, but more generally there is  frequency dependence of the amplifier contained implicitly in ''A<sub>v</sub>''. Such frequency dependence of ''A<sub>v</sub>'' also makes the Miller capacitance frequency dependent, so interpretation of ''C<sub>M</sub>'' as a capacitance becomes more difficult. However, ordinarily any frequency dependence of ''A<sub>v</sub>'' arises only at frequencies much higher than the roll-off with frequency caused by the Miller effect, so for frequencies up to the Miller-effect roll-off of the gain, ''A<sub>v</sub>'' is accurately approximated by its low-frequency value. Determination of ''C<sub>M</sub>'' using ''A<sub>v</sub>'' at low frequencies is the so-called '''Miller approximation'''.<ref name=Spencer/> With the Miller approximation, ''C<sub>M</sub>'' becomes frequency independent, and its interpretation as a capacitance at low frequencies is secure.
|author=W. M. C. Sansen
|title=Analog Design Essentials
|year= 2006
|page=§0124, p. 13
|publisher=Springer
|location=Dordrechtμ
|isbn=0-387-25746-2
|url=http://worldcat.org/isbn/0387257462}}
</ref>  


<ref name=BSIM>  
==References and notes==
{{reflist|refs=
<ref name=Miller>
{{cite journal |author=John M. Miller |title=Dependence of the input impedance of a three-electrode vacuum tube upon the load in the plate circuit |journal=Scientific Papers of the Bureau of Standards |volume=15 |issue= 351 |pages=pp. 367-385 |year=1920 |url=http://books.google.com/books?id=7u8SAAAAYAAJ&pg=PA367&lpg=PA367}}


A popular MOSFET computer model is BSIM4, see: {{cite web |title=BSIM4.6.4 MOSFET Model: User's manual |work=BSIM 4 web site |publisher=Electrical Engineering and Computer Sciences Department, UC Berkeley |url=http://www-device.eecs.berkeley.edu/~bsim3/BSIM4/BSIM464/BSIM464_Manual.pdf |accessdate=2011-05-20 |date=2009}} This 170-page manual describes a model much more complex than a simple algebraic formula.
</ref>
</ref>


<ref name=Spencer>
{{cite book
|author=R.R. Spencer and M.S. Ghausi
|title=Introduction to electronic circuit design.
|year= 2003
|page=533
|publisher=Prentice Hall/Pearson Education, Inc.
|location=Upper Saddle River NJ
|isbn=0-201-36183-3
|url=http://worldcat.org/isbn/0-201-36183-3}}
</ref>


 
<ref name = PoleSplitting>See article on [[pole splitting]].</ref>




}}
}}

Revision as of 14:57, 23 May 2011

In electronics, the Miller effect accounts for the increase in the equivalent input capacitance of an inverting voltage amplifier due to amplification of the capacitance between the input and output terminals. The additional input capacitance due to the Miller effect is given by

where is the gain of the amplifier and C is the feedback capacitance.

Although the term Miller effect normally refers to capacitance, any impedance connected between the input and another node exhibiting gain can modify the amplifier input impedance via this effect. These properties of Miller effect are generalized by Miller theorem.


History

The Miller effect was named after John Milton Miller.[1] When Miller published his work in 1920, he was working on vacuum tube triodes, however the same theory applies to more modern devices such as bipolar and MOS transistors.

Derivation

Figure 1: Circuit for deriving the Miller effect, with ideal inverting voltage amplifier

Consider an ideal inverting voltage amplifier of gain with an impedance connected between its input and output nodes. The output voltage is therefore . Assuming that the amplifier input draws no current, all of the input current flows through , and is therefore given by

.

The input impedance of the circuit is

If Z represents a capacitor with impedance , the resulting input impedance is

Thus the effective or Miller capacitance CM is the physical C multiplied by the factor [2].

Effects

As most amplifiers are inverting (i.e. ), the effective capacitance at their inputs is increased due to the Miller effect. This can lower the bandwidth of the amplifier, reducing its range of operation to lower frequencies. The tiny junction and stray capacitances between the base and collector terminals of a Darlington transistor, for example, may be drastically increased by the Miller effects due to its high gain, lowering the high frequency response of the device.

It is also important to note that the Miller capacitance is the capacitance seen looking into the input. If looking for all of the RC time constants (poles) it is important to include as well the capacitance seen by the output. The capacitance on the output is often neglected since it sees and amplifier outputs are typically low impedance. However if the amplifier has a high impedance output, such as if a gain stage is also the output stage, then this RC can have a significant impact on the performance of the amplifier. This is when pole splitting techniques are used.

The Miller effect may also be exploited to synthesize larger capacitors from smaller ones. One such example is in the stabilization of feedback amplifiers, where the required capacitance may be too large to practically include in the circuit. This may be particularly important in the design of integrated circuit, where capacitors can consume significant area, increasing costs.

Mitigation

The Miller effect may be undesired in many cases, and approaches may be sought to lower its impact. Several such techniques are used in the design of amplifiers.

A current buffer stage may be added at the output to lower the gain between the input and output terminals of the amplifier (though not necessarily the overall gain). For example, a common base may be used as a current buffer at the output of a common emitter stage, forming a cascode. This will typically reduce the Miller effect and increase the bandwidth of the amplifier.

Alternatively, a voltage buffer may be used before the amplifier input, reducing the effective source impedance seen by the input terminals. This lowers the time constant of the circuit and typically increases the bandwidth.

Impact on frequency response

Figure 2: Operational amplifier with feedback capacitor CC.
Figure 3: Circuit of Figure 2 transformed using Miller's theorem, introducing the Miller capacitance on the input side of the circuit.

Figure 2 shows an example of Figure 1 where the impedance coupling the input to the output is the coupling capacitor CC. A Thévenin voltage source VA drives the circuit with Thévenin resistance RA. At the output a parallel RC-circuit serves as load. (The load is irrelevant to this discussion: it just provides a path for the current to leave the circuit.) In Figure 2, the coupling capacitor delivers a current jωCC( vi - vo ) to the output circuit.

Figure 3 shows a circuit electrically identical to Figure 2 using Miller's theorem. The coupling capacitor is replaced on the input side of the circuit by the Miller capacitance CM, which draws the same current from the driver as the coupling capacitor in Figure 2. Therefore, the driver sees exactly the same loading in both circuits. On the output side, a dependent current source in Figure 3 delivers the same current to the output as does the coupling capacitor in Figure 2. That is, the R-C-load sees the same current in Figure 3 that it does in Figure 2.

In order that the Miller capacitance draw the same current in Figure 3 as the coupling capacitor in Figure 2, the Miller transformation is used to relate CM to CC. In this example, this transformation is equivalent to setting the currents equal, that is

or, rearranging this equation

This result is the same as CM of the Derivation Section.

The present example with Av frequency independent shows the implications of the Miller effect, and therefore of CC, upon the frequency response of this circuit, and is typical of the impact of the Miller effect (see, for example, common source). If CC = 0 F, the output voltage of the circuit is simply Av vA, independent of frequency. However, when CC is not zero, Figure 3 shows the large Miller capacitance appears at the input of the circuit. The voltage output of the circuit now becomes

and rolls off with frequency once frequency is high enough that ω CMRA ≥ 1. It is a low-pass filter. In analog amplifiers this curtailment of frequency response is a major implication of the Miller effect. In this example, the frequency ω3dB such that ω3dB CMRA = 1 marks the end of the low-frequency response region and sets the bandwidth or cutoff frequency of the amplifier.

It is important to notice that the effect of CM upon the amplifier bandwidth is greatly reduced for low impedance drivers (CM RA is small if RA is small). Consequently, one way to minimize the Miller effect upon bandwidth is to use a low-impedance driver, for example, by interposing a voltage follower stage between the driver and the amplifier, which reduces the apparent driver impedance seen by the amplifier.

The output voltage of this simple circuit is always Av vi. However, real amplifiers have output resistance. If the amplifier output resistance is included in the analysis, the output voltage exhibits a more complex frequency response and the impact of the frequency-dependent current source on the output side must be taken into account.[3] Ordinarily these effects show up only at frequencies much higher than the roll-off due to the Miller capacitance, so the analysis presented here is adequate to determine the useful frequency range of an amplifier dominated by the Miller effect.

Miller approximation

This example also assumes Av is frequency independent, but more generally there is frequency dependence of the amplifier contained implicitly in Av. Such frequency dependence of Av also makes the Miller capacitance frequency dependent, so interpretation of CM as a capacitance becomes more difficult. However, ordinarily any frequency dependence of Av arises only at frequencies much higher than the roll-off with frequency caused by the Miller effect, so for frequencies up to the Miller-effect roll-off of the gain, Av is accurately approximated by its low-frequency value. Determination of CM using Av at low frequencies is the so-called Miller approximation.[2] With the Miller approximation, CM becomes frequency independent, and its interpretation as a capacitance at low frequencies is secure.

References and notes

  1. John M. Miller (1920). "Dependence of the input impedance of a three-electrode vacuum tube upon the load in the plate circuit". Scientific Papers of the Bureau of Standards 15 (351): pp. 367-385.
  2. 2.0 2.1 R.R. Spencer and M.S. Ghausi (2003). Introduction to electronic circuit design.. Upper Saddle River NJ: Prentice Hall/Pearson Education, Inc.. ISBN 0-201-36183-3. 
  3. See article on pole splitting.