Staggered tuning

Last updated

Staggered tuning is a technique used in the design of multi-stage tuned amplifiers whereby each stage is tuned to a slightly different frequency. In comparison to synchronous tuning (where each stage is tuned identically) it produces a wider bandwidth at the expense of reduced gain. It also produces a sharper transition from the passband to the stopband. Both staggered tuning and synchronous tuning circuits are easier to tune and manufacture than many other filter types.

Contents

The function of stagger-tuned circuits can be expressed as a rational function and hence they can be designed to any of the major filter responses such as Butterworth and Chebyshev. The poles of the circuit are easy to manipulate to achieve the desired response because of the amplifier buffering between stages.

Applications include television IF amplifiers (mostly 20th century receivers) and wireless LAN.

Rationale

A typical multi-stage tuned amplifier. The amplifier is synchronously tuned if all LC-circuits are tuned at the same frequency, which occurs if all the products Ck * Lk are equal. In staggered tuning, the products Ck * Lk are generally different in each stage. Stagger-tuned 3-stage amplifier.svg
A typical multi-stage tuned amplifier. The amplifier is synchronously tuned if all LC-circuits are tuned at the same frequency, which occurs if all the products Ck * Lk are equal. In staggered tuning, the products Ck * Lk are generally different in each stage.

Staggered tuning improves the bandwidth of a multi-stage tuned amplifier at the expense of the overall gain. Staggered tuning also increases the steepness of passband skirts and hence improves selectivity. [1]

Plot showing the reduction of bandwidth caused by synchronous tuning with increasing number of stages, n. The Q of each stage is 10 in this example. Synchronous tuned plot.svg
Plot showing the reduction of bandwidth caused by synchronous tuning with increasing number of stages, n. The Q of each stage is 10 in this example.

The value of staggered tuning is best explained by first looking at the shortcomings of tuning every stage identically. This method is called synchronous tuning. Each stage of the amplifier will reduce the bandwidth. In an amplifier with multiple identical stages, the 3 dB points of the response after the first stage will become the 6 dB points of the second stage. Each successive stage will add a further 3 dB to what was the band edge of the first stage. Thus the 3 dB bandwidth becomes progressively narrower with each additional stage. [2]

As an example, a four-stage amplifier will have its 3 dB points at the 0.75 dB points of an individual stage. The fractional bandwidth of an LC circuit is given by,

where m is the power ratio of the power at resonance to that at the band edge frequency (equal to 2 for the 3 dB point and 1.19 for the 0.75 dB point) and Q is the quality factor.
Comparison of synchronous and staggered tuning responses Staggered and synchronous comparison.svg
Comparison of synchronous and staggered tuning responses

The bandwidth is thus reduced by a factor of . In terms of the number of stages . [3] Thus, the four stage synchronously tuned amplifier will have a bandwidth of only 19% of a single stage. Even in a two-stage amplifier the bandwidth is reduced to 41% of the original. Staggered tuning allows the bandwidth to be widened at the expense of overall gain. The overall gain is reduced because when any one stage is at resonance (and thus maximum gain) the others are not, unlike synchronous tuning where all stages are at maximum gain at the same frequency. A two-stage stagger-tuned amplifier will have a gain 3 dB less than a synchronously tuned amplifier. [4]

Even in a design that is intended to be synchronously tuned, some staggered tuning effect is inevitable because of the practical impossibility of keeping all tuned circuits perfectly in step and because of feedback effects. This can be a problem in very narrow band applications where essentially only one spot frequency is of interest, such as a local oscillator feed or a wave trap. The overall gain of a synchronously tuned amplifier will always be less than the theoretical maximum because of this. [5]

Both synchronously tuned and stagger-tuned schemes have a number of advantages over schemes that place all the tuning components in a single aggregated filter circuit separate from the amplifier such as ladder networks or coupled resonators. One advantage is that they are easy to tune. Each resonator is buffered from the others by the amplifier stages so have little effect on each other. The resonators in aggregated circuits, on the other hand, will all interact with each other, particularly their nearest neighbours. [6] Another advantage is that the components need not be close to ideal. Every LC resonator is directly working into a resistor which lowers the Q anyway so any losses in the L and C components can be absorbed into this resistor in the design. Aggregated designs usually require high Q resonators. Also, stagger-tuned circuits have resonator components with values that are quite close to each other and in synchronously tuned circuits they can be identical. The spread of component values is thus less in stagger-tuned circuits than in aggregated circuits. [7]

Design

Tuned amplifiers such as the one illustrated at the beginning of this article can be more generically depicted as a chain of transconductance amplifiers each loaded with a tuned circuit.

Generic multi-stage tuned amplifier Stagger tuned amplifier generic.svg
Generic multi-stage tuned amplifier
where for each stage (omitting the suffixes)
gm is the amplifier transconductance
C is the tuned circuit capacitance
L is the tuned circuit inductance
G is the sum of the amplifier output conductance and the input conductance of the next amplifier.

Stage gain

The gain A(s), of one stage of this amplifier is given by;

where s is the complex frequency operator.

This can be written in a more generic form, that is, not assuming that the resonators are the LC type, with the following substitutions,

(the resonant frequency)
(the gain at resonance)
(the stage quality factor)

Resulting in,

Stage bandwidth

The gain expression can be given as a function of (angular) frequency by making the substitution s = where i is the imaginary unit and ω is the angular frequency

The frequency at the band edges, ωc, can be found from this expression by equating the value of the gain at the band edge to the magnitude of the expression,

where m is defined as above and equal to two if the 3 dB points are desired.

Solving this for ωc and taking the difference between the two positive solutions finds the bandwidth Δω,

and the fractional bandwidth B,

Overall response

Gain response of a two-stage stagger-tuned amplifier. The stage 3 dB fractional bandwidth is 0.125, but the overall bandwidth is increased to approximately 0.52. Stagger tuned plot.svg
Gain response of a two-stage stagger-tuned amplifier. The stage 3 dB fractional bandwidth is 0.125, but the overall bandwidth is increased to approximately 0.52.
Gain response of a two-stage stagger-tuned amplifier for various values of stage Q Stagger tuned plot with Q.svg
Gain response of a two-stage stagger-tuned amplifier for various values of stage Q

The overall response of the amplifier is given by the product of the individual stages,

It is desirable to be able to design the filter from a standard low-pass prototype filter of the required specification. Frequently, a smooth Butterworth response will be chosen [8] but other polynomial functions can be used that allow ripple in the response. [9] A popular choice for a polynomial with ripple is the Chebyshev response for its steep skirt. [10] For the purpose of transformation, the stage gain expression can be rewritten in the more suggestive form,

This can be transformed into a low-pass prototype filter with the transform

where ω'c is the cutoff frequency of the low-pass prototype.

This can be done straightforwardly for the complete filter in the case of synchronously tuned amplifiers where every stage has the same ω0 but for a stagger-tuned amplifier there is no simple analytical solution to the transform. Stagger-tuned designs can be approached instead by calculating the poles of a low-pass prototype of the desired form (e.g. Butterworth) and then transforming those poles to a band-pass response. The poles so calculated can then be used to define the tuned circuits of the individual stages.

Poles

The stage gain can be rewritten in terms of the poles by factorising the denominator;

where p, p* are a complex conjugate pair of poles

and the overall response is,

where the ak = A0kω0k/Q0k

From the band-pass to low-pass transform given above, an expression can be found for the poles in terms of the poles of the low-pass prototype, qk,

where ω0B is the desired band-pass centre frequency and Qeff is the effective Q of the overall circuit.

Each pole in the prototype transforms to a complex conjugate pair of poles in the band-pass and corresponds to one stage of the amplifier. This expression is greatly simplified if the cutoff frequency of the prototype, ω'c, is set to the final filter bandwidth ω0B/Qeff.

In the case of a narrowband design ω0q which can be used to make a further simplification with the approximation,

These poles can be inserted into the stage gain expression in terms of poles. By comparing with the stage gain expression in terms of component values, those component values can then be calculated. [11]

Applications

Staggered tuning is of most benefit in wideband applications. It was formerly commonly used in television receiver IF amplifiers. However, SAW filters are more likely to be used in that role nowadays. [12] Staggered tuning has advantages in VLSI for radio applications such as wireless LAN. [13] The low spread of component values make it much easier to implement in integrated circuits than traditional ladder networks. [14]

See also

Related Research Articles

<span class="mw-page-title-main">Resonance</span> Tendency to oscillate at certain frequencies

Resonance is the phenomenon, pertaining to oscillatory dynamical systems, wherein amplitude rises are caused by an external force with time-varying amplitude with the same frequency of variation as the natural frequency of the system. The amplitude rises that occur are a result of the fact that applied external forces at the natural frequency entail a net increase in mechanical energy of the system.

A low-pass filter is a filter that passes signals with a frequency lower than a selected cutoff frequency and attenuates signals with frequencies higher than the cutoff frequency. The exact frequency response of the filter depends on the filter design. The filter is sometimes called a high-cut filter, or treble-cut filter in audio applications. A low-pass filter is the complement of a high-pass filter.

<span class="mw-page-title-main">Bode plot</span> Graph of the frequency response of a control system

In electrical engineering and control theory, a Bode plot is a graph of the frequency response of a system. It is usually a combination of a Bode magnitude plot, expressing the magnitude of the frequency response, and a Bode phase plot, expressing the phase shift.

<i>Q</i> factor Parameter describing the longevity of energy in a resonator relative to its resonant frequency

In physics and engineering, the quality factor or Q factor is a dimensionless parameter that describes how underdamped an oscillator or resonator is. It is defined as the ratio of the initial energy stored in the resonator to the energy lost in one radian of the cycle of oscillation. Q factor is alternatively defined as the ratio of a resonator's centre frequency to its bandwidth when subject to an oscillating driving force. These two definitions give numerically similar, but not identical, results. Higher Q indicates a lower rate of energy loss and the oscillations die out more slowly. A pendulum suspended from a high-quality bearing, oscillating in air, has a high Q, while a pendulum immersed in oil has a low one. Resonators with high quality factors have low damping, so that they ring or vibrate longer.

A resistor–capacitor circuit, or RC filter or RC network, is an electric circuit composed of resistors and capacitors. It may be driven by a voltage or current source and these will produce different responses. A first order RC circuit is composed of one resistor and one capacitor and is the simplest type of RC circuit.

Chebyshev filters are analog or digital filters that have a steeper roll-off than Butterworth filters, and have either passband ripple or stopband ripple. Chebyshev filters have the property that they minimize the error between the idealized and the actual filter characteristic over the operating frequency range of the filter, but they achieve this with ripples in the passband. This type of filter is named after Pafnuty Chebyshev because its mathematical characteristics are derived from Chebyshev polynomials. Type I Chebyshev filters are usually referred to as "Chebyshev filters", while type II filters are usually called "inverse Chebyshev filters". Because of the passband ripple inherent in Chebyshev filters, filters with a smoother response in the passband but a more irregular response in the stopband are preferred for certain applications.

The Sallen–Key topology is an electronic filter topology used to implement second-order active filters that is particularly valued for its simplicity. 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.

<span class="mw-page-title-main">Butterworth filter</span> Type of signal processing filter

The Butterworth filter is a type of signal processing filter designed to have a frequency response that is as flat as possible in the passband. It is also referred to as a maximally flat magnitude filter. It was first described in 1930 by the British engineer and physicist Stephen Butterworth in his paper entitled "On the Theory of Filter Amplifiers".

<span class="mw-page-title-main">LC circuit</span> Electrical "resonator" circuit, consisting of inductive and capacitive elements with no resistance

An LC circuit, also called a resonant circuit, tank circuit, or tuned circuit, is an electric circuit consisting of an inductor, represented by the letter L, and a capacitor, represented by the letter C, connected together. The circuit can act as an electrical resonator, an electrical analogue of a tuning fork, storing energy oscillating at the circuit's resonant frequency.

In electronics 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 cause the amplifier to oscillate, and to control overshoot and ringing in the amplifier's step response. It is also used extensively to improve the bandwidth of single pole systems.

<span class="mw-page-title-main">Common source</span> Electronic amplifier circuit type

In electronics, a common-source amplifier is one of three basic single-stage field-effect transistor (FET) amplifier topologies, typically used as a voltage or transconductance amplifier. The easiest way to tell if a FET is common source, common drain, or common gate is to examine where the signal enters and leaves. The remaining terminal is what is known as "common". In this example, the signal enters the gate, and exits the drain. The only terminal remaining is the source. This is a common-source FET circuit. The analogous bipolar junction transistor circuit may be viewed as a transconductance amplifier or as a voltage amplifier.. As a transconductance amplifier, the input voltage is seen as modulating the current going to the load. As a voltage amplifier, input voltage modulates the current flowing through the FET, changing the voltage across the output resistance according to Ohm's law. However, the FET device's output resistance typically is not high enough for a reasonable transconductance amplifier, nor low enough for a decent voltage amplifier. As seen below in the formula, the voltage gain depends on the load resistance, so it cannot be applied to drive low-resistance devices, such as a speaker. Another major drawback is the amplifier's limited high-frequency response. Therefore, in practice the output often is routed through either a voltage follower, or a current follower, to obtain more favorable output and frequency characteristics. The CS–CG combination is called a cascode amplifier.

A resistor–inductor circuit, or RL filter or RL network, is an electric circuit composed of resistors and inductors driven by a voltage or current source. A first-order RL circuit is composed of one resistor and one inductor, either in series driven by a voltage source or in parallel driven by a current source. It is one of the simplest analogue infinite impulse response electronic filters.

An elliptic filter is a signal processing filter with equalized ripple (equiripple) behavior in both the passband and the stopband. The amount of ripple in each band is independently adjustable, and no other filter of equal order can have a faster transition in gain between the passband and the stopband, for the given values of ripple. Alternatively, one may give up the ability to adjust independently the passband and stopband ripple, and instead design a filter which is maximally insensitive to component variations.

<span class="mw-page-title-main">Gain–bandwidth product</span> Figure of merit for amplifiers

The gain–bandwidth product for an amplifier is a figure of merit calculated by multiplying the amplifier's bandwidth and the gain at which the bandwidth is measured.

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 amplifier's input and output terminals, and is given by

<span class="mw-page-title-main">Parametric oscillator</span> Harmonic oscillator whose parameters oscillate in time

A parametric oscillator is a driven harmonic oscillator in which the oscillations are driven by varying some parameters of the system at some frequencies, typically different from the natural frequency of the oscillator. A simple example of a parametric oscillator is a child pumping a playground swing by periodically standing and squatting to increase the size of the swing's oscillations. The child's motions vary the moment of inertia of the swing as a pendulum. The "pump" motions of the child must be at twice the frequency of the swing's oscillations. Examples of parameters that may be varied are the oscillator's resonance frequency and damping .

An all-pass filter is a signal processing filter that passes all frequencies equally in gain, but changes the phase relationship among various frequencies. Most types of filter reduce the amplitude of the signal applied to it for some values of frequency, whereas the all-pass filter allows all frequencies through without changes in level.

<span class="mw-page-title-main">Electronic filter topology</span> Electronic filter circuits defined by component connection

Electronic filter topology defines electronic filter circuits without taking note of the values of the components used but only the manner in which those components are connected.

<span class="mw-page-title-main">RLC circuit</span> Resistor Inductor Capacitor Circuit

An RLC circuit is an electrical circuit consisting of a resistor (R), an inductor (L), and a capacitor (C), connected in series or in parallel. The name of the circuit is derived from the letters that are used to denote the constituent components of this circuit, where the sequence of the components may vary from RLC.

<span class="mw-page-title-main">Double-tuned amplifier</span>

A double-tuned amplifier is a tuned amplifier with transformer coupling between the amplifier stages in which the inductances of both the primary and secondary windings are tuned separately with a capacitor across each. The scheme results in a wider bandwidth and steeper skirts than a single tuned circuit would achieve.

References

  1. Pederson & Mayaram, p. 259
  2. Sedha, p. 627
  3. Chattopadhyay, p. 195
  4. Maheswari & Anand, p. 500
  5. Pederson & Mayaram, p. 259
  6. Iniewski, pp. 200-201
  7. Wiser, pp. 47-48
  8. Sedha, p. 627
  9. Moxon, pp. 88-89
  10. Iniewski, p. 200
  11. Maheswari & Anand, pp. 499-500
  12. Gulati, p. 147
  13. Wiser, p. vi
  14. Iniewski, p. 200

Bibliography