User:John R. Brews/Sample2: Difference between revisions

From Citizendium
Jump to navigation Jump to search
imported>John R. Brews
No edit summary
imported>John R. Brews
No edit summary
Line 1: Line 1:
{{Image|Negative feedback amplifier.PNG|right|300px|Ideal negative feedback amplifier.}}
{{TOC|right}}
A '''negative feedback amplifier''' (or more commonly simply a '''feedback amplifier''') is an amplifier in which a fraction of its output is combined with the signal at its input that opposes the signal in what is called [[negative feedback]]. The negative feedback improves performance (gain stability, linearity, frequency response, [[step response]]) and reduces sensitivity to parameter variations due to manufacturing or environmental uncertainties. A single feedback loop with unilateral blocks is shown in the figure. Negative feedback is used in this way in many amplifiers and control systems.<ref name=Kuo/>


== Advantages and disadvantages of feedback ==
{{TOC|right}}
Many electronic devices used to provide gain (for example, [[vacuum tube]]s, [[Bipolar transistor]]s, [[MOSFET]]s) are [[nonlinear]].  Negative feedback is a circuit technique that trades gain for higher linearity (reducing [[distortion]]), amongst other things. If not designed correctly amplifiers with negative feedback can become unstable, resulting in unwanted behavior, such as [[oscillation]].  The [[Nyquist stability criterion]] developed by [[Harry Nyquist]] of [[Bell Laboratories]], or the [[Bode plot]] can be used to study the stability of feedback amplifiers.
The '''return ratio''' of a dependent source in a linear electrical circuit is the ''negative'' of the ratio of the current (voltage) returned to the site of the dependent source to the current (voltage) of a replacement independent source. The terms ''loop gain'' and ''return ratio'' are often used interchangeably; however, they are necessarily equivalent only in the case of a single feedback loop system with unilateral blocks. <ref name=Spencer>
 
{{cite book
Feedback amplifiers share these properties:<ref name=Palumbo/>
|author=Richard R Spencer & Ghausi MS
 
|title=Introduction to electronic circuit design
Pros:
|page=p. 723
*Can increase or decrease input impedance (depending on type of feedback)
|year= 2003
*Can increase or decrease output impedance (depending on type of feedback)
|publisher=Prentice Hall/Pearson Education
*Reduces distortion (increases linearity)
|location=Upper Saddle River NJ
*Increases bandwidth (the range of frequencies for which the circuit works)
|isbn=0-201-36183-3
*Desensitizes gain to component variations
|url=http://worldcat.org/isbn/0-201-36183-3}}
*Can control [[step response]] of amplifier
 
Cons:
*May lead to instability if not designed carefully
*The gain of the amplifier decreases
*The input and output impedances of the amplifier with feedback (the '''closed-loop amplifier''') become sensitive to the gain of the amplifier without feedback (the '''open-loop amplifier'''); that exposes these impedances to variations in the open loop gain, for example, due to parameter variations or due to nonlinearity of the open-loop gain
 
==History==
The negative feedback amplifier was invented by [[Harold Stephen Black]] (US patent 2,102,671, filed 1932, issued 1937) <ref name=patent/>) while a passenger on the Lackawanna Ferry (from Hoboken Terminal to Manhattan) on his way to work at [[Bell Laboratories]] on August 2, 1927. Black had been pondering how to reduce distortion in repeater amplifiers used for telephone transmission. On a blank space in his copy of The New York Times, currently on display at Bell Laboratories in Mountainside, New Jersey, he sketched a diagram equivalent to that in the figure, and derived the equations below.<ref name=Waldhauer/>
 
==Classical feedback==
 
=== Voltage amplifiers ===
 
Below, the gain of the amplifier with feedback, the '''closed-loop gain''' ''A<sub>fb</sub>'', is derived in terms of the gain of the amplifier without feedback, the '''open-loop gain''' ''A<sub>OL</sub>'' and the '''feedback factor''' β, which governs how much of the output signal is applied to the input. See the figure, top right. The feedback  parameter β is determined by the feedback network that is connected around the amplifier.
 
Consider a voltage amplifier with voltage feedback.  Without feedback, the output voltage ''V<sub>out</sub> = A<sub>OL</sub>'' ''V<sub>in</sub>'', where the open-loop gain ''A<sub>OL</sub>'' in general may be a function of both frequency and voltage.
 
The open-loop gain ''A<sub>OL</sub>'' is defined by:
 
:<math>A_{OL} = \frac{V_{out}}{V_{in}} \ ,</math>
 
where ''V<sub>in</sub>'' is the input to the amplifier, with no feedback, and ''V<sub>out</sub>'' is the amplifier output, again with no feedback.
 
Suppose we have a feedback loop so that a fraction β'' V<sub>out</sub>'' of the output is subtracted from the input. The input to the amplifier itself is now ''V’<sub>in</sub>'', where
 
:<math>V'_{in} = V_{in} - \beta \cdot V_{out}</math>
 
The gain of the amplifier with feedback, called the closed-loop gain, ''A<sub>fb</sub>'' still is given by,
 
:<math>A_{fb} = \frac{V_{out}}{V_{in}} \ ,</math>
 
but now the signal increased by the gain of the amplifier is ''V’<sub>in</sub>''. That is,
 
:<math>A_{OL}=\frac {V_{out}}{V'_{in}} \ . </math>
 
Rewriting the expression for ''V’<sub>in</sub>'':
 
:<math>\frac{V'_{in}}{V_{out}} = \frac{V_{in}}{V_{out}} - \beta \ ,</math>
 
or:
 
:<math>\frac{1}{A_{OL}} = \frac{1}{A_{fb}} - \beta \ .</math>
 
Solving for ''A<sub>fb</sub>'':
 
:<math>A_{fb} = \frac{A_{OL}}{1 + \beta \cdot A_{OL}} \ .</math>
 
If ''A''<sub>OL</sub> >> 1, then ''A''<sub>fb</sub> ≈ 1 / β and the effective amplification (or closed-loop gain) ''A''<sub>fb</sub> is set by the feedback constant β, and hence set by the feedback network, usually a simple reproducible network, thus making linearizing and stabilizing the amplification characteristics straightforward. Note also that if there are conditions where β ''A''<sub>OL</sub> = −1, the amplifier has infinite amplification – it has become an oscillator, and the system is unstable. The stability characteristics of the gain feedback product β ''A''<sub>OL</sub> are often displayed and investigated on a [[Nyquist plot]] (a polar plot of the gain/phase shift as a parametric function of frequency). A simpler, but less general technique, uses [[Bode_plot#Gain_margin_and_phase_margin|Bode plot]]s.
 
The combination ''L'' = β ''A''<sub>OL</sub> appears commonly in feedback analysis and is called the '''loop gain'''. The combination ( 1 + β ''A''<sub>OL</sub> ) also appears commonly and is variously named as the '''desensitivity factor''' or the '''improvement factor'''.
 
===Bandwidth extension===
{{Image|Bandwidth comparison.PNG|right|300px| Gain ''vs.'' frequency for a single-pole amplifier with and without feedback; corner frequencies are labeled.}}
Feedback can be used to extend the bandwidth of an amplifier (speed it up) at the cost of lowering the amplifier gain.<ref name=Broderson/> The figure at right shows such a comparison. The figure is understood as follows. Without feedback the so-called '''open-loop''' gain in this example has a single time constant frequency response given by
 
::<math> A_{OL}(f) ={A_0}  \frac {1}{ 1+ j \frac{f} {f_C} } \ , </math>
 
where ''A<sub>0</sub>'' is the zero-frequency gain of the amplifier, and ''f<sub>C</sub>'' is the [[cutoff frequency|cutoff]] or [[corner frequency]] of the amplifier. In this example, the gain at zero frequency is ''A<sub>0</sub>'' = 10<sup>5</sup> V/V and the corner frequency is ''f<sub>C</sub>'' = 10<sup>4</sup> Hz. The figure shows the gain is flat out to the corner frequency and then drops. When feedback is present the so-called '''closed-loop''' gain, as shown in the formula of the previous section, becomes,
 
::<math> A_{fb} (f) = \frac { A_{OL} } { 1 + \beta  A_{OL} } </math>
 
::::<math> = \frac { A_0/(1+jf/f_C) } { 1 + \beta A_0/(1+jf/f_C) } </math>
 
::::<math> = \frac {A_0} {1+ jf/f_C + \beta A_0} </math>
 
::::<math> = \frac {A_0} {1 + \beta A_0}\  \frac{1} {1+j \left(\frac {f} {(1+ \beta A_0) f_C }\right) }   
\ .
</math>
 
The last expression shows the feedback amplifier still has a single time constant behavior, but the corner frequency is now increased from the value ''f<sub>C</sub>'' by the improvement factor {{nowrap|(1+β ''A<sub>0</sub>'')}} to become {{nowrap|(1+β ''A<sub>0</sub>'') ''f<sub>C</sub>'',}} and the gain at zero frequency has dropped by exactly the same factor. This reciprocal behavior is called the '''gain-bandwidth tradeoff'''. In Figure 2,  (1 + β ''A<sub>0</sub>'') = 10<sup>3</sup>, so ''A<sub>fb</sub>''(0)= 10<sup>5</sup> / 10<sup>3</sup> = 100 V/V, and ''f<sub>C</sub>'' increases to 10<sup>4</sup> × 10<sup>3</sup> = 10<sup>7</sup> Hz.
 
===Multiple poles===
When the open-loop gain has several poles, rather than the single pole of the above example, feedback can result in complex poles (real and imaginary parts). In a two-pole case, the result is peaking in the frequency response of the feedback amplifier near its corner frequency, and ringing and overshoot in its its [[step response]]. In the case of more than two poles, the feedback amplifier can become unstable, and oscillate. See the discussion of [[Bode_plot#Gain_margin_and_phase_margin|gain margin and phase margin]].<ref name=Sansen/>
 
==Asymptotic gain model==
 
In the above analysis the feedback network is [[Electronic amplifier#Unilateral or bilateral|unilateral]]. However, real feedback networks often exhibit '''feed forward''' as well, that is, they feed a small portion of the input to the output, degrading performance of the feedback amplifier. A more general way to model negative feedback amplifiers including this effect is with the [[asymptotic gain model]].
 
==Feedback and amplifier type==
Amplifiers use current or voltage as input and output, so four types of amplifier are possible, choosing one of the two inputs and pairing with one of the two outputs. Any of these four choices may be the open-loop amplifier used to construct the feedback amplifier.  The objective for the feedback amplifier also may be any one of the four types of amplifier, not necessarily the the same type as the open-loop amplifier. For example, an op amp (voltage amplifier) can be arranged to make a current amplifier instead. The conversion from one type to another is implemented using different feedback connections, usually referred to as series or shunt (parallel) connections.<ref name=Goel/><ref name=Zimmer/> See the table below.
{| class="wikitable" style="background:white;text-align:center "
!Feedback amplifier type
!Input connection
!Output connection
!Ideal feedback 
!Two-port feedback
|-
|-valign="top"
| '''Current'''
| '''Shunt'''
| '''Series'''
| '''CCCS'''
| '''g-parameter'''
 
|-
|-valign="top"
| '''Transresistance'''
| '''Shunt'''
| '''Shunt'''
|  '''VCCS'''
|  '''y-parameter'''
 
|-
|-valign="top"
| '''Transconductance'''
| '''Series'''
| '''Series
| '''CCVS'''
| '''z-parameter'''
 
|-
|-valign="top"
| '''Voltage'''
| '''Series'''
| '''Shunt'''
| '''VCVS'''
| '''h-parameter'''
 
|}
The feedback can be implemented using a [[two-port network]]. There are four types of two-port network, and the selection depends upon the type of feedback. For example, for a current feedback amplifier, current at the output is sampled and combined with current at the input. Therefore, the feedback ideally is performed using an (output) current-controlled current source (CCCS), and its imperfect realization using a two-port network also must incorporate a CCCS, that is, the appropriate choice for feedback network is a g-parameter two-port.
 
==Two-port analysis of feedback==
One approach to feedback is the use of [[return ratio]]. Here an alternative method used in most textbooks is presented by means of an example treated in the article on [[Asymptotic_gain_model#Two-stage_transistor_amplifier|asymptotic gain model]].<ref name=Subramanian/><ref name=Gray-Meyer1/><ref name=Sedra1/>
[[Image:Two-transistor feedback amp.PNG|thumbnail|250px|Figure 3: A ''shunt-series'' feedback amplifier]]
Figure 3 shows a two-transistor amplifier with a feedback resistor ''R<sub>f</sub>''. The aim is to analyze this circuit to find three items: the gain, the output impedance looking into the amplifier from the load, and the input impedance looking into the amplifier from the source.
 
===Replacement of the feedback network with a two-port===
The first step is replacement of the feedback network by a two port. Just what components go into the two port?
 
On the input side of the two-port we have ''R<sub>f</sub>''. If the voltage at the right side of ''R<sub>f</sub>'' changes, it changes the current in ''R<sub>f</sub>'' that is subtracted from the current entering the base of the input transistor. That is, the input side of the two-port is a dependent current source controlled by the voltage at the top of resistor ''R<sub>2</sub>''.
 
One might say the second stage of the amplifier is just a [[voltage follower]], transmitting the voltage at the collector of the input transistor to the top of ''R<sub>2</sub>''. That is, the monitored output signal is really the voltage at the collector of the input transistor. That view is legitimate, but then the voltage follower stage becomes part of the feedback network. That makes analysis of feedback more complicated.
[[Image:G-equivalent circuit.PNG|thumbnail|250px|Figure 4: The g-parameter feedback network]]
An alternative view is that the voltage at the top of ''R<sub>2</sub>'' is set by the emitter current of the output transistor. That view leads to an entirely passive feedback network made up of ''R<sub>2</sub>'' and ''R<sub>f</sub>''. The variable controlling the feedback is the emitter current, so the feedback is a current-controlled current source (CCCS). We search through the four available [[two-port network]]s and find the only one with a CCCS is the g-parameter two-port, shown in Figure 4. The next task is to select the g-parameters so that the two-port of Figure 4 is electrically equivalent to the L-section made up of ''R<sub>2</sub>'' and ''R<sub>f</sub>''. That selection is an algebraic procedure made most simply by looking at two individual cases: the case with ''V<sub>1</sub>'' = 0, which makes the VCVS on the right side of the two-port a short-circuit; and the case with ''I<sub>2</sub>'' = 0. which makes the CCCS on the left side an open circuit. The algebra in these two cases is simple, much easier than solving for all variables at once. The choice of g-parameters that make the two-port and the L-section behave the same way are shown in the table below.
{| class="wikitable" style="background:white;text-align:center "
!g<sub>11</sub>
!g<sub>12</sub>
!g<sub>21</sub>
!g<sub>22</sub>
|-
|-valign="center"
| '''<math>\frac {1} {R_f+R_2}</math>'''
| '''<math> - \frac {R_2}{R_2+R_f}</math>''
| '''<math> \frac {R_2} {R_2+R_f} </math>'''
| '''<math>R_2\mathit{\parallel}R_f \ </math>'''
|}
[[Image:Small-signal current amplifier with feedback.PNG|thumbnail|400px|Figure 5: Small-signal circuit with two-port for feedback network; upper shaded box: main amplifier; lower shaded box: feedback two-port replacing the ''L''-section made up of ''R''<sub>f</sub> and ''R''<sub>2</sub>.]]
 
===Small-signal circuit===
The next step is to draw the small-signal schematic for the amplifier with the two-port in place using the [[hybrid-pi model]] for the transistors. Figure 5 shows the schematic with notation ''R<sub>3</sub>'' = ''R<sub>C2</sub>||R<sub>L</sub>'' and ''R<sub>11</sub>'' = 1 / ''g<sub>11</sub>'', ''R<sub>22</sub>'' = ''g<sub>22</sub>'' .
 
===Loaded open-loop gain===
Figure 3 indicates the output node, but not the choice of output variable. A useful choice is the short-circuit current output of the amplifier (leading to the short-circuit current gain). Because this variable leads simply to any of the other choices (for example, load voltage or load current), the short-circuit current gain is found below.
 
First the loaded '''open-loop gain''' is found. The feedback is turned off by setting ''g<sub>12</sub> = g<sub>21</sub>'' = 0. The idea is to find how much the amplifier gain is changed because of the resistors in the feedback network by themselves, with the feedback turned off. This calculation is pretty easy because ''R<sub>11</sub>, R<sub>B</sub>, and r<sub>π1</sub>'' all are in parallel and ''v<sub>1</sub> = v<sub>π</sub>''. Let ''R<sub>1</sub>'' = ''R<sub>11</sub> // R<sub>B</sub> // r<sub>π1</sub>''. In addition, ''i<sub>2</sub> = −(β+1) i<sub>B</sub>''.  The result for the open-loop current gain ''A<sub>OL</sub>'' is:
 
::<math> A_{OL} = \frac { \beta i_B } {i_S} = g_m R_C \left( \frac { \beta }{ \beta +1} \right)
\left(
\frac {R_1} {R_{22} +
\frac {r_{ \pi 2} + R_C } {\beta + 1 } } \right)  \ . </math>
 
===Gain with feedback===
In the classical approach to feedback, the feedforward represented by the VCVS (that is,  ''g<sub>21</sub> v<sub>1</sub>'') is neglected.<ref name=note1/> That makes the circuit of Figure 5 resemble the block diagram of Figure 1, and the gain with feedback is then:
 
::<math> A_{FB} = \frac { A_{OL} } {1 + { \beta }_{FB} A_{OL} } </math>
:::<math> = \frac {A_{OL} } {1 + \frac {R_2} {R_2+R_f}  A_{OL} } \ , </math>
 
where the feedback factor β<sub>FB</sub> = −g<sub>12</sub>. Notation β<sub>FB</sub> is introduced for the feedback factor to distinguish it from the transistor β.
 
===Input and output resistances===
[[Image:Feedback amplifier input resistance.PNG|thumb|500px|Figure 6: Circuit set-up for finding feedback amplifier input resistance]]
First, a digression on how two-port theory approaches resistance determination, and then its application to the amplifier at hand.
 
====Background on resistance determination====
Figure 6 shows a test circuit for finding the input resistance of a feedback voltage amplifier (left) and for a feedback current amplifier (right). In the case of the voltage amplifier [[Kirchhoff's circuit laws|Kirchhoff's voltage law]] provides:
 
::<math> V_x = I_x R_{in} + \beta v_{out} \ , </math>
 
where ''v''<sub>out</sub> = ''A''<sub>v</sub> ''v''<sub>in</sub> = ''A''<sub>v</sub> ''I''<sub>x</sub> ''R''<sub>in</sub>. Substituting this result in the above equation and solving for the input resistance of the feedback amplifier, the result is:
 
::<math> R_{in}(fb) = \frac {V_x} {I_x} = \left( 1 + \beta A_v \right ) R_{in} \ . </math>
 
The general conclusion to be drawn from this example and a similar example for the output resistance case is:
 
''A series feedback connection at the input (output) increase the input (output) resistance by a factor ( 1 + β ''A''<sub>OL</sub> )'', where ''A''<sub>OL</sub> = open loop gain.
 
On the other hand, for the current amplifier, which uses a shunt input connection, Kirchhoff's laws provide:
 
::<math> I_x = \frac {V_{in}} {R_{in}} + \beta i_{out} \ . </math>
 
where ''i''<sub>out</sub> = ''A''<sub>i</sub> ''i''<sub>in</sub> = ''A''<sub>i</sub> ''V''<sub>x</sub> / ''R''<sub>in</sub>. Substituting this result in the above equation and solving for the input resistance of the feedback amplifier, the result is:
 
::<math> R_{in}(fb) = \frac {V_x} {I_x} = \frac { R_{in}  } { \left( 1 + \beta A_i \right ) } \ . </math>
 
The general conclusion to be drawn from this example and a similar example for the output resistance case is:
 
''A parallel feedback connection at the input (output) decreases the input (output) resistance by a factor ( 1 + β ''A''<sub>OL</sub> )'', where ''A''<sub>OL</sub> = open loop gain.
 
These conclusions can be generalized to treat cases with arbitrary [[Norton's theorem|Norton]] or [[Thevenin's theorem|Thévenin]] drives, arbitrary loads, and general [[two-port network|two-port feedback networks]]. However, the results do depend upon the main amplifier having a representation as a two-port – that is, the results depend on the ''same'' current entering and leaving the input terminals, and likewise, the same current that leaves one output terminal must enter the other output terminal.
 
A broader conclusion to be drawn, independent of the quantitative details, is that feedback can be used to increase or to decrease the input and output impedances.
 
====Application to the example amplifier====
These resistance results now are applied to the amplifier of Figure 3 and Figure 5. The ''improvement factor'' that reduces the gain, namely ( 1 + β<sub>FB</sub> A<sub>OL</sub> ), directly decides the effect of feedback upon the input and output resistances of the amplifier. In the case of a shunt connection, the input impedance is reduced by this factor; and in the case of series connection, the impedance is multiplied by this factor. However, the impedance that is modified by feedback is the impedance of the amplifier in Figure 5 with the feedback turned off, and does include the modifications to impedance caused by the resistors of the feedback network.
 
Therefore, the input impedance seen by the source with feedback turned off is ''R''<sub>in</sub> = ''R''<sub>1</sub> = ''R''<sub>11</sub> // ''R''<sub>B</sub> // ''r''<sub>π1</sub>, and with the feedback turned on (but no feedforward)
 
::<math> R_{in} = \frac {R_1} {1 + { \beta }_{FB} A_{OL} } \ , </math>
 
where ''division'' is used because the input connection is ''shunt'': the feedback two-port is in parallel with the signal source at the input side of the amplifier. A reminder: ''A''<sub>OL</sub> is the  ''loaded'' open loop gain [[Negative_feedback_amplifier#Loaded_open-loop_gain|found above]], as modified by the resistors of the feedback network.
 
The impedance seen by the load needs further discussion. The load in Figure 5 is connected to the collector of the output transistor, and therefore is separated from the body of the amplifier by the infinite impedance of the output current source. Therefore, feedback has no effect on the output impedance, which remains simply ''R<sub>C2</sub>'' as seen by the load resistor ''R<sub>L</sub>'' in Figure 3.<ref name=Jaeger/>
 
If instead we wanted to find the impedance presented at the ''emitter'' of the output transistor (instead of its collector), which is series connected to the feedback network, feedback would increase this resistance by the improvement factor ( 1 + β<sub>FB</sub> A<sub>OL</sub> ).<ref name=note2/>
 
===Load voltage and load current===
The gain derived above is the current gain at the collector of the output transistor. To relate this gain to the gain when voltage is the output of the amplifier, notice that the output voltage at the load ''R<sub>L</sub>'' is related to the collector current by [[Ohm's law]] as ''v<sub>L</sub> = i<sub>C</sub> ( R<sub>C2</sub>||R<sub>L</sub> )''. Consequently, the transresistance gain ''v<sub>L</sub> / i<sub>S</sub>'' is found by multiplying the current gain by ''R<sub>C2</sub>||R<sub>L</sub>'':
 
::<math> \frac {v_L} {i_S} = A_{FB} ( R_{C2}\mathit{\parallel}R_L ) \ . </math>
 
Similarly, if the output of the amplifier is taken to be the current in the load resistor ''R<sub>L</sub>'', [[current division]] determines the load current, and the gain is then:
 
::<math> \frac {i_L} {i_S} = A_{FB} \frac {R_{C2}} {R_{C2} + R_L} \ . </math>
 
=== Is the main amplifier block a two port? ===
[[Image:Two-port ground arrangement.PNG|thumbnail|400px|Figure 7: Amplifier with ground connections labeled by ''G''. The feedback network satisfies the port conditions.]]
Some complications follow, intended for the attentive reader.
 
Figure 7 shows the small-signal schematic with the main amplifier and the feedback two-port in shaded boxes. The two-port satisfies the [[Two-port_network|port conditions]]: at the input port, ''I''<sub>in</sub> enters and leaves the port, and likewise at the output, ''I''<sub>out</sub> enters and leaves. The main amplifier is shown in the upper shaded box. The ground connections are labeled.
 
Figure 7 shows the interesting fact that the main amplifier does not satisfy the port conditions at its input and output unless the ground connections are chosen to make that happen. For example, on the input side, the current entering the main amplifier is ''I''<sub>S</sub>. This current is divided three ways: to the feedback network, to the bias resistor ''R''<sub>B</sub> and to the base resistance of the input transistor ''r''<sub>π</sub>. To satisfy the port condition for the main amplifier, all three components must be returned to the input side of the main amplifier, which means all the ground leads labeled ''G''<sub>1</sub> must be connected, as well as emitter lead ''G''<sub>E1</sub>. Likewise, on the output side, all ground connections ''G''<sub>2</sub> must be connected and also ground connection ''G''<sub>E2</sub>. Then, at the bottom of the schematic, underneath the feedback two-port and outside the amplifier blocks, ''G''<sub>1</sub> is connected to ''G''<sub>2</sub>. That forces the ground currents to divide between the input and output sides as planned. Notice that this connection arrangement ''splits the emitter'' of the input transistor into a base-side and a collector-side – a physically impossible thing to do, but electrically the circuit sees all the ground connections as one node, so this fiction is permitted.
 
Of course, the way the ground leads are connected makes no difference to the amplifier (they are all one node), but it makes a difference to the port conditions. That is a weakness of this approach: the port conditions are needed to justify the method, but the circuit really is unaffected by how currents are traded among ground connections.
 
However, if there is '''no possible arrangement''' of ground conditions that will lead to the port conditions, the circuit might not behave the same way.<ref name=note3/> The improvement factors ( 1 + β<sub>FB</sub> A<sub>OL</sub> ) for determining input and output impedance might not work. This situation is awkward, because a failure to make a two-port may reflect a real problem (it just is not possible), or reflect a lack of imagination (for example, just did not think of splitting the emitter node in two). As a consequence, when the port conditions are in doubt, at least two approaches are possible to establish whether improvement factors are accurate: either simulate an example using [[SPICE|Spice]] and compare results with use of an improvement factor, or calculate the impedance using a test source and compare results.
 
A more radical choice is to drop the two-port approach altogether, and use [[return ratio]]s. That choice might be advisable if small-signal device models are complex, or are not available (for example, the devices are known only numerically, perhaps from measurement or from [[SPICE]] simulations).
 
== References  and notes ==
{{Reflist|refs=
 
<ref name=Broderson>
{{cite web |url=http://bwrc.eecs.berkeley.edu/classes/ee140/Lectures/10_stability.pdf |author=RW Brodersen |title=Analog circuit design: lectures on stability |accessdate=2011-06-24 }}
</ref>
 
<ref name=Goel>
{{cite web |url=http://www.ece.mtu.edu/faculty/goel/EE-4232/Feedback.pdf |author=Ashok K. Goel |title=Feedback topologies |accessdate=2011-06-24}}
</ref>
</ref>


<ref name=Kuo>
==Calculating the return ratio==
[[Image:Bipolar transresistance amplifier.PNG|thumbnail|200px|Figure 1: Collector-to-base biased bipolar amplifier]]
The steps for calculating the return ratio of a source are as follows:<ref name=Gray-Meyer>
{{cite book  
{{cite book  
|author=Kuo, Benjamin C & Farid Golnaraghi
|author=Paul R. Gray, Hurst P J Lewis S H & Meyer RG
|title=Automatic control systems
|title=Analysis and design of analog integrated circuits
|edition=Eighth edition
|page=§8.8 pp. 599-613
|page=p. 46
|year= 2001
|year= 2003
|edition=Fourth Edition
|publisher=Wiley  
|publisher=Wiley  
|location=NY
|location=New York
|isbn=0471134767
|isbn=0-471-32168-0
|url=http://worldcat.org/isbn/0471134767}}
|url=http://worldcat.org/isbn/0-471-32168-0}}
</ref>
 
 
<ref name=Palumbo>
{{cite book
|author=Gaetano Palumbo and Salvatore Pennisi
|title=Feedback amplifiers: theory and design
|chapter=Chapter 3: Feedback
|pages=p. 64
|year= 2002
|publisher=Springer
|isbn=0792376439
|url=http://www.amazon.com/Feedback-Amplifiers-Theory-Gaetano-Palumbo/dp/0792376439/ref=sr_1_1?s=books&ie=UTF8&qid=1308924834&sr=1-1#reader_0792376439}}
</ref>
 
<ref name=patent>
{{cite web
| title=Wave translation system
| url=http://www.google.com/patents?id=tA9EAAAAEBAJ&printsec=abstract&zoom=4&source=gbs_overview_r&cad=0#v=onepage&q&f=false
| accessdate=2011-06-24
}} Patent 2,106,671 refers to "apparatus or systems involving negative feedback".
</ref>
</ref>
#  Set all independent sources to zero.
#  Select the dependent source for which the return ratio is sought.
#  Place an independent source of the same type (voltage or current) and polarity in parallel with the selected dependent source.
#  Move the dependent source to the side of the inserted source and cut the two leads joining the dependent source to the independent source.
#  For a '''voltage source''' the return ratio is minus the ratio of the voltage across the dependent source divided by the voltage of the independent replacement source.
#  For a '''current source''', short-circuit the broken leads of the dependent source. The return ratio is minus the ratio of the resulting short-circuit current to the current of the independent replacement source.


<ref name=Sansen> For a complete discussion, see Sansen.
=== Other Methods ===
The above steps can be implemented in [[SPICE]] simulations by replacing non-linear devices by their small-signal model equivalents. These steps are not feasible where dependent sources inside devices are not directly accessible, for example, when SPICE itself is used to generate the small-signal circuit numerically, or when measuring the return ratio experimentally. When small-signal models cannot be used, an added problem is finding how to break a a loop without affecting the [[bias point]] and altering the results. Middlebrook<ref>[http://www.informaworld.com/smpp/content~content=a771365730~db=all Middlebrook, RD:''Loop gain in feedback systems 1''; Int. J. of Electronics, vol. 38, no. 4, (1975) pp. 485-512 ]</ref>and Rosenstark<ref>[http://www.informaworld.com/smpp/content~content=a777774065~db=all Rosenstark, Sol: ''Loop gain measurement in feedback amplifiers''; Int. J. of Electronics, vol. 57, No. 3 (1984) pp.415-421]</ref> have proposed several methods for experimental evaluation of return ratio (loosely referred to by these authors as simply ''loop gain''), and similar methods have been adapted for use in [[SPICE]] by Hurst.<ref>[http://ieeexplore.ieee.org/xpl/freeabs_all.jsp?arnumber=99170 Hurst, PJ: ''Exact simulation of feedback circuit parameters''; IEEE Trans. on Circuits and Systems, vol. 38, No. 11 (1991) pp.1382-1389]</ref> See [http://www.spectrum-soft.com/news/spring97/loopgain.shtm Spectrum user note] or Roberts, or Sedra, and especially Tuinenga.<ref name=Roberts>
{{cite book  
{{cite book  
|author=Willy M. C. Sansen
|author=Gordon W. Roberts & Sedra AS
|title=Analog design essentials
|title=SPICE
|year= 2006
|edition=Second Edition
|pages=§0513-§0533, p. 155-165
|year= 1997
|publisher=Springer
|pages=Chapter 8; pp. 256-262
|location=New York; Berlin
|publisher=Oxford University Press
|isbn=0-387-25746-2
|location=New York  
|url=http://worldcat.org/isbn/0-387-25746-2}}
|isbn=0-19-510842-6
</ref>
|url=http://worldcat.org/isbn/0-19-510842-6}}
 
</ref><ref name=Sedra>
<ref name=Sedra1>
{{cite book  
{{cite book  
|author=A. S. Sedra and K.C. Smith
|author=Adel S Sedra & Smith KC
|title=Microelectronic Circuits
|title=Microelectronic circuits
|edition=Fifth Edition
|year= 2004
|year= 2004
|edition=Fifth Edition
|pages=Example 8.7; pp. 855--859  
|pages=Example 8.4, pp. 825-829 and PSpice simulation pp. 855-859
|publisher=Oxford University Press
|publisher=Oxford  
|location=New York  
|location=New York  
|isbn=0-19-514251-9
|isbn=0-19-514251-9
|url=http://worldcat.org/isbn/0-19-514251-9}}
|url=http://worldcat.org/isbn/0-19-514251-9}}
</ref><ref name=Tuinenga>
{{cite book
|author=Paul W Tuinenga
|title=SPICE: a guide to circuit simulation and analysis using PSpice
|edition=Third Edition
|year= 1995
|pages=Chapter 8: ''Loop gain analysis''
|publisher=Prentice-Hall
|location=Englewood Cliffs NJ
|isbn=0134360494
|url=http://worldcat.org/isbn/0134360494}}
</ref>
</ref>


<ref name=Subramanian>
==Example: Collector-to-base biased bipolar amplifier==
{{cite web |url=http://organics.eecs.berkeley.edu/~viveks/ee140/lectures/section10p4.pdf |author=Vivek Subramanian |title=Lectures on feedback}}
[[Image:Inserting return ratio source.PNG|700px|thumb|center|Figure 2: Left - small-signal circuit corresponding to Figure 1; center - inserting independent source and marking leads to be cut; right  - cutting the dependent source free and short-circuiting broken leads]]
</ref>


<ref name=Gray-Meyer1>
Figure 1 (top right) shows a bipolar amplifier with feedback bias resistor ''R<sub>f</sub>'' driven by a [[Norton's theorem|Norton signal source]]. Figure 2 (left panel) shows the corresponding small-signal circuit obtained by replacing the transistor with its [[hybrid-pi model]]. The objective is to find the return ratio of the dependent current source in this amplifier.<ref name=Spencer2>
{{cite book  
{{cite book  
|author=P R Gray, P J Hurst, S H Lewis, and R G Meyer
|author=Richard R Spencer & Ghausi MS
|title=Analysis and Design of Analog Integrated Circuits
|title=Example 10.7 pp. 723-724
|year= 2001
|isbn=0-201-36183-3
|page=pp. 586-587
|url=http://worldcat.org/isbn/0-201-36183-3}}
|edition=Fourth Edition
</ref> To reach the objective, the steps outlined above are followed. Figure 2 (center panel) shows the application of these steps up to Step 4, with the dependent source moved to the left of the inserted source of value ''i<sub>t</sub>'', and the leads targeted for cutting marked with an ''x''. Figure 2 (right panel) shows the circuit set up for calculation of the return ratio ''T'', which is
|publisher=Wiley
 
|location=New York
::<math> T = - \frac {i_r} {i_t} \ . </math>
|isbn=0-471-32168-0
 
|url=http://worldcat.org/isbn/0471321680}}
The return current is
</ref>
 
::<math> i_r = g_m v_{\pi} \ . </math>
 
The feedback current in ''R<sub>f</sub>'' is found by [[current division]] to be:
::<math>i_f = \frac {R_D//r_O} {R_D//r_O +R_F +r_{\pi}// R_S} \  i_t \ . </math>
 
The base-emitter voltage ''v<sub>π</sub>'' is then, from [[Ohm's law]]:
 
::<math> v_{\pi} = -i_f \ ( r_{\pi}// R_S ) \ . </math>
 
Consequently,
::<math> T = g_m  (r_{\pi}// R_S ) \  \frac {R_D//r_O} {R_D//r_O +R_F +r_{\pi}// R_S}\ .  </math>
 
=== Application in asymptotic gain model ===
The overall [[Electronic_amplifier#Input_and_output_variables|transresistance gain]] of this amplifier can be shown to be:
 
::<math> G = \frac {v_{out}} {i_{in}} =  \frac {(1-g_m R_F)R_1 R_2} {R_F+R_1+R_2+g_m R_1R_2} \ , </math>
 
with ''R<sub>1</sub> = R<sub>S</sub> || r<sub>π</sub>'' and ''R<sub>2</sub> = R<sub>D</sub> || r<sub>O</sub>''.
 
This expression can be rewritten in the form used by the [[asymptotic gain model]], which expresses the overall gain of a feedback amplifier in terms of several independent factors that often are more easily derived separately than is the gain itself, and that often provide insight into the circuit. This form is:


<ref name=Jaeger>
::<math> G = \ G_{ \infty } \frac {T} {1+T}  + G_0 \frac {1} { 1+T} \ \ , </math>
The use of the improvement factor ( 1 + β<sub>FB</sub> A<sub>OL</sub> ) requires care, particularly for the case of output impedance using series feedback. See {{cite book | title = Microelectronic Circuit Design | author =R.C. Jaeger and T.N. Blalock  | publisher = McGraw-Hill Professional | year = 2006 |edition=Third Edition |page=Example 17.3 pp. 1092-1096| isbn = 978-0-07-319163-8 | url = http://worldcat.org/isbn/978-0-07-319163-8 }}
</ref>


<ref name=note1>
where the so-called '''asymptotic gain''' ''G<sub>&infin;</sub>'' is the gain at infinite ''g<sub>m</sub>'', namely:
If the feedforward is included, its effect is to cause a modification of the open-loop gain, normally so small compared to the open-loop gain itself that it can be dropped. Notice also that the main amplifier block is [[Electronic_amplifier#Unilateral_or_bilateral|unilateral]].
</ref>


<ref name=note2>
::<math> G_{\infty} = - R_F \ , </math>
That is, the impedance found by turning off the signal source ''I<sub>S</sub>'' = 0,  inserting a test current in the emitter lead ''I<sub>x</sub>'',  finding the voltage across the test source ''V<sub>x</sub>'', and finding ''R<sub>out</sub> = V<sub>x</sub> / I<sub>x</sub>''.
</ref>


<ref name=note3>
and the so-called '''feed forward''' or '''direct feedthrough''' ''G<sub>0</sub>'' is the gain for zero ''g<sub>m</sub>'', namely:
The equivalence of the main amplifier block to a two-port network guarantees that performance factors work, but without that equivalence they may work anyway. For example, in some cases the circuit can be shown to be equivalent to another circuit that is a two port, by "cooking up" different circuit parameters that are functions of the original ones. There is no end to creativity!</ref>


<ref name=Waldhauer>
::<math> G_{0} = \frac { R_1 R_2 } {R_F +R_1 +R_2}\ . </math>
{{cite book
|author=Fred F. Waldhauer
|title=Feedback
|url=http://books.google.com/books?ei=F5cETsCRFJDAsAPj9fHYDQ&ct=result&id=EesiAAAAMAAJ&dq=feedback+inauthor%3AFred+inauthor%3AD+inauthor%3AWaldhauer&q=Lackawanna+Ferry#search_anchor
|pages=p. 3, Figure 1.1 p. 4,  p. 8
|year= 1982
|publisher=Wiley
|location=NY
|isbn=0471053198
}}
</ref>


<ref name=Zimmer>
For additional applications of this method, see [[Asymptotic gain model#Return ratio|asymptotic gain model]].
{{cite web |url=http://centrevirtuel.creea.u-bordeaux.fr/ELAB/docs/freebooks.php/virtual/feedback-amplifier/textbook_feedback.html#1.2 |author=Zimmer T & Geoffreoy D |title=Feedback amplifier |accessdate=2011-06-24}}
</ref>


}}
==References==
<references/>

Revision as of 10:19, 30 June 2011

The return ratio of a dependent source in a linear electrical circuit is the negative of the ratio of the current (voltage) returned to the site of the dependent source to the current (voltage) of a replacement independent source. The terms loop gain and return ratio are often used interchangeably; however, they are necessarily equivalent only in the case of a single feedback loop system with unilateral blocks. [1]

Calculating the return ratio

Figure 1: Collector-to-base biased bipolar amplifier

The steps for calculating the return ratio of a source are as follows:[2]

  1. Set all independent sources to zero.
  2. Select the dependent source for which the return ratio is sought.
  3. Place an independent source of the same type (voltage or current) and polarity in parallel with the selected dependent source.
  4. Move the dependent source to the side of the inserted source and cut the two leads joining the dependent source to the independent source.
  5. For a voltage source the return ratio is minus the ratio of the voltage across the dependent source divided by the voltage of the independent replacement source.
  6. For a current source, short-circuit the broken leads of the dependent source. The return ratio is minus the ratio of the resulting short-circuit current to the current of the independent replacement source.

Other Methods

The above steps can be implemented in SPICE simulations by replacing non-linear devices by their small-signal model equivalents. These steps are not feasible where dependent sources inside devices are not directly accessible, for example, when SPICE itself is used to generate the small-signal circuit numerically, or when measuring the return ratio experimentally. When small-signal models cannot be used, an added problem is finding how to break a a loop without affecting the bias point and altering the results. Middlebrook[3]and Rosenstark[4] have proposed several methods for experimental evaluation of return ratio (loosely referred to by these authors as simply loop gain), and similar methods have been adapted for use in SPICE by Hurst.[5] See Spectrum user note or Roberts, or Sedra, and especially Tuinenga.[6][7][8]

Example: Collector-to-base biased bipolar amplifier

Figure 2: Left - small-signal circuit corresponding to Figure 1; center - inserting independent source and marking leads to be cut; right - cutting the dependent source free and short-circuiting broken leads

Figure 1 (top right) shows a bipolar amplifier with feedback bias resistor Rf driven by a Norton signal source. Figure 2 (left panel) shows the corresponding small-signal circuit obtained by replacing the transistor with its hybrid-pi model. The objective is to find the return ratio of the dependent current source in this amplifier.[9] To reach the objective, the steps outlined above are followed. Figure 2 (center panel) shows the application of these steps up to Step 4, with the dependent source moved to the left of the inserted source of value it, and the leads targeted for cutting marked with an x. Figure 2 (right panel) shows the circuit set up for calculation of the return ratio T, which is

The return current is

The feedback current in Rf is found by current division to be:

The base-emitter voltage vπ is then, from Ohm's law:

Consequently,

Application in asymptotic gain model

The overall transresistance gain of this amplifier can be shown to be:

with R1 = RS || rπ and R2 = RD || rO.

This expression can be rewritten in the form used by the asymptotic gain model, which expresses the overall gain of a feedback amplifier in terms of several independent factors that often are more easily derived separately than is the gain itself, and that often provide insight into the circuit. This form is:

where the so-called asymptotic gain G is the gain at infinite gm, namely:

and the so-called feed forward or direct feedthrough G0 is the gain for zero gm, namely:

For additional applications of this method, see asymptotic gain model.

References

  1. Richard R Spencer & Ghausi MS (2003). Introduction to electronic circuit design. Upper Saddle River NJ: Prentice Hall/Pearson Education. ISBN 0-201-36183-3. 
  2. Paul R. Gray, Hurst P J Lewis S H & Meyer RG (2001). Analysis and design of analog integrated circuits, Fourth Edition. New York: Wiley. ISBN 0-471-32168-0. 
  3. Middlebrook, RD:Loop gain in feedback systems 1; Int. J. of Electronics, vol. 38, no. 4, (1975) pp. 485-512
  4. Rosenstark, Sol: Loop gain measurement in feedback amplifiers; Int. J. of Electronics, vol. 57, No. 3 (1984) pp.415-421
  5. Hurst, PJ: Exact simulation of feedback circuit parameters; IEEE Trans. on Circuits and Systems, vol. 38, No. 11 (1991) pp.1382-1389
  6. Gordon W. Roberts & Sedra AS (1997). SPICE, Second Edition. New York: Oxford University Press, Chapter 8; pp. 256-262. ISBN 0-19-510842-6. 
  7. Adel S Sedra & Smith KC (2004). Microelectronic circuits, Fifth Edition. New York: Oxford University Press, Example 8.7; pp. 855--859. ISBN 0-19-514251-9. 
  8. Paul W Tuinenga (1995). SPICE: a guide to circuit simulation and analysis using PSpice, Third Edition. Englewood Cliffs NJ: Prentice-Hall, Chapter 8: Loop gain analysis. ISBN 0134360494. 
  9. Richard R Spencer & Ghausi MS. Example 10.7 pp. 723-724. ISBN 0-201-36183-3.