Sallen–Key topology

(Redirected from Sallen-Key topology)

The Sallen–Key topology is an electronic filter topology used to implement second-order active filters that is particularly valued for its simplicity.[1] It is a degenerate form of a voltage-controlled voltage-source (VCVS) filter topology. It was introduced by R. P. Sallen and E. L. Key of MIT Lincoln Laboratory in 1955.[2]

Explanation of operation edit

A VCVS filter uses a voltage amplifier with practically infinite input impedance and zero output impedance to implement a 2-pole low-pass, high-pass, bandpass, bandstop, or allpass response. The VCVS filter allows high Q factor and passband gain without the use of inductors. A VCVS filter also has the advantage of independence: VCVS filters can be cascaded without the stages affecting each others tuning. A Sallen–Key filter is a variation on a VCVS filter that uses a unity gain amplifier (i.e., a buffer amplifier).

History and implementation edit

In 1955, Sallen and Key used vacuum tube cathode follower amplifiers; the cathode follower is a reasonable approximation to an amplifier with unity voltage gain. Modern analog filter implementations may use operational amplifiers (also called op amps). Because of its high input impedance and easily selectable gain, an operational amplifier in a conventional non-inverting configuration is often used in VCVS implementations.[citation needed] Implementations of Sallen–Key filters often use an op amp configured as a voltage follower; however, emitter or source followers are other common choices for the buffer amplifier.

Sensitivity to component tolerances edit

VCVS filters are relatively resilient to component tolerance, but obtaining high Q factor may require extreme component value spread or high amplifier gain.[1] Higher-order filters can be obtained by cascading two or more stages.

Generic Sallen–Key topology edit

 
Figure 1: The generic Sallen–Key filter topology

The generic unity-gain Sallen–Key filter topology implemented with a unity-gain operational amplifier is shown in Figure 1. The following analysis is based on the assumption that the operational amplifier is ideal.

Because the op amp is in a negative-feedback configuration, its   and   inputs must match (i.e.,  ). However, the inverting input   is connected directly to the output  , and so

 

(1)

By Kirchhoff's current law (KCL) applied at the   node,

 

(2)

By combining equations (1) and (2),

 

Applying equation (1) and KCL at the op amp's non-inverting input   gives

 

which means that

 

(3)

Combining equations (2) and (3) gives

 

(4)

Rearranging equation (4) gives the transfer function

 

(5)

which typically describes a second-order linear time-invariant (LTI) system.

If the   component were connected to ground instead of to  , the filter would be a voltage divider composed of the   and   components cascaded with another voltage divider composed of the   and   components. The buffer amplifier bootstraps the "bottom" of the   component to the output of the filter, which will improve upon the simple two-divider case. This interpretation is the reason why Sallen–Key filters are often drawn with the op amp's non-inverting input below the inverting input, thus emphasizing the similarity between the output and ground.

Branch impedances edit

By choosing different passive components (e.g., resistors and capacitors) for  ,  ,  , and  , the filter can be made with low-pass, bandpass, and high-pass characteristics. In the examples below, recall that a resistor with resistance   has impedance   of

 

and a capacitor with capacitance   has impedance   of

 

where   (here   denotes the imaginary unit) is the complex angular frequency, and   is the frequency of a pure sine-wave input. That is, a capacitor's impedance is frequency-dependent and a resistor's impedance is not.

Application: low-pass filter edit

 
Figure 2: A unity-gain low-pass filter implemented with a Sallen–Key topology

An example of a unity-gain low-pass configuration is shown in Figure 2. An operational amplifier is used as the buffer here, although an emitter follower is also effective. This circuit is equivalent to the generic case above with

 

The transfer function for this second-order unity-gain low-pass filter is

 

where the undamped natural frequency  , attenuation  , Q factor  , and damping ratio  , are given by

 

and

 

So,

 

The   factor determines the height and width of the peak of the frequency response of the filter. As this parameter increases, the filter will tend to "ring" at a single resonant frequency near   (see "LC filter" for a related discussion).

Poles and zeros edit

This transfer function has no (finite) zeros and two poles located in the complex s-plane:

 

There are two zeros at infinity (the transfer function goes to zero for each of the   terms in the denominator).

Design choices edit

A designer must choose the   and   appropriate for their application. The   value is critical in determining the eventual shape. For example, a second-order Butterworth filter, which has maximally flat passband frequency response, has a   of  . By comparison, a value of   corresponds to the series cascade of two identical simple low-pass filters.

Because there are 2 parameters and 4 unknowns, the design procedure typically fixes the ratio between both resistors as well as that between the capacitors. One possibility is to set the ratio between   and   as   versus   and the ratio between   and   as   versus  . So,

 

As a result, the   and   expressions are reduced to

 

and

 
 
Figure 3: A low-pass filter, which is implemented with a Sallen–Key topology, with f0 = 15.9 kHz and Q = 0.5

Starting with a more or less arbitrary choice for e.g.   and  , the appropriate values for   and   can be calculated in favor of the desired   and  . In practice, certain choices of component values will perform better than others due to the non-idealities of real operational amplifiers.[3] As an example, high resistor values will increase the circuit's noise production, whilst contributing to the DC offset voltage on the output of op amps equipped with bipolar input transistors.

Example edit

For example, the circuit in Figure 3 has   and  . The transfer function is given by

 

and, after the substitution, this expression is equal to

 

which shows how every   combination comes with some   combination to provide the same   and   for the low-pass filter. A similar design approach is used for the other filters below.

Input impedance edit

The input impedance of the second-order unity-gain Sallen–Key low-pass filter is also of interest to designers. It is given by Eq. (3) in Cartwright and Kaminsky[4] as

 

where   and  .

Furthermore, for  , there is a minimal value of the magnitude of the impedance, given by Eq. (16) of Cartwright and Kaminsky,[4] which states that

 

Fortunately, this equation is well-approximated by[4]

 

for  . For   values outside of this range, the 0.34 constant has to be modified for minimal error.

Also, the frequency at which the minimal impedance magnitude occurs is given by Eq. (15) of Cartwright and Kaminsky,[4] i.e.,

 

This equation can also be well approximated using Eq. (20) of Cartwright and Kaminsky,[4] which states that

 

Application: high-pass filter edit

 
Figure 4: A specific Sallen–Key high-pass filter with f0 = 72 Hz and Q = 0.5

A second-order unity-gain high-pass filter with   and   is shown in Figure 4.

A second-order unity-gain high-pass filter has the transfer function

 

where undamped natural frequency   and   factor are discussed above in the low-pass filter discussion. The circuit above implements this transfer function by the equations

 

(as before) and

 

So

 

Follow an approach similar to the one used to design the low-pass filter above.

Application: bandpass filter edit

 
Figure 5: A bandpass filter realized with a VCVS topology

An example of a non-unity-gain bandpass filter implemented with a VCVS filter is shown in Figure 5. Although it uses a different topology and an operational amplifier configured to provide non-unity-gain, it can be analyzed using similar methods as with the generic Sallen–Key topology. Its transfer function is given by

 

The center frequency   (i.e., the frequency where the magnitude response has its peak) is given by

 

The Q factor   is given by

 

The voltage divider in the negative feedback loop controls the "inner gain"   of the op amp:

 

If the inner gain   is too high, the filter will oscillate.

See also edit

References edit

  1. ^ a b "EE315A Course Notes - Chapter 2"-B. Murmann Archived 2010-07-16 at the Wayback Machine
  2. ^ Sallen, R. P.; E. L. Key (March 1955). "A Practical Method of Designing RC Active Filters". IRE Transactions on Circuit Theory. 2 (1): 74–85. doi:10.1109/tct.1955.6500159. S2CID 51640910.
  3. ^ Stop-band limitations of the Sallen–Key low-pass filter.
  4. ^ a b c d e Cartwright, K. V.; E. J. Kaminsky (2013). "Finding the minimum input impedance of a second-order unity-gain Sallen-Key low-pass filter without calculus" (PDF). Lat. Am. J. Phys. Educ. 7 (4): 525–535.

External links edit