Sei sulla pagina 1di 15

ECE 1371 Term Paper

LC Oscillators:
Theory, Design and Trend

University of Toronto
Professor K. Phang

April 9, 2001

by
Aichin Chung
LC Oscillators: Theory, Design and Trend
Most electronic signal processing systems require frequency or time reference signals. To use

the full capacity of communication channels, transmitters modulate the baseband message signal

into different parts of the spectrum to exploit better propagation characteristics or to frequency

multiplex several messages, and receivers down convert received high frequency message

signals for demodulation. Oscillators are circuits that provide the accurate frequency reference

signals required by these applications. Specifically, oscillators are autonomous circuits that

produce a stable periodically time-varying waveform. They have at least two states and they

cycle through those states at a constant pace.

Oscillators can be broadly classified into three categories: ring oscillatiors, relaxation oscillators

and harmonic oscillators. Ring oscillators consist of an odd number of single-ended inverters or

an even/odd number of differential inverters with the appropriate connections. Relaxation

oscillators alternately charge and discharge a capacitor with a constant current between two

threshold levels. Harmonic oscillators contain a passive resonator that serves as the frequency

settling element. Although, ring and relaxation oscillators are easy to integrate on monolithic

integrated circuits (ICs) and are very compact, they have poor stability and spectral purity in

comparison to their harmonic counterparts. Additionally, relaxation ring oscillators are typically

very sensitive to noise in the switching thresholds and charging currents.

LC Oscillator Basics

The most common harmonic oscillators are those that use resonant circuits consisting of

inductances and capacitances. These harmonic oscillators that use LC resonant circuits are

known as LC oscillators. LC oscillators have traditionally been hard to integrate on integrated


circuits (ICs) due to the lack of high quality passive inductors in standard IC technologies and

from their large size. However, recent advances in monolithic IC design have allowed inductors

to be fabricated on-chip. As such, entire LC oscillators can now be implemeneted in an IC,

thereby eliminating the need to use external components. Currently, LC oscillators are

frequently used in many analog and radio frequency signal processing systems due to their ease

of implementation and their good phase noise characteristics.

A typical LC oscillator is made up of three distinct blocks as shown in Figure 1. The amplifier

supplies energy to maintain oscillations in the circuit during the intervals between pulses of

excitation energy. The LC tank circuit alternately stores energy in the inductor and the

capacitor. A portion of the output of the LC network is fed back to the input of the amplifier

through the regenerative-feedback network. This regenerative-feedback helps to sustain

oscillation by overcoming the effect of damping caused by the LC tank’s internal and load

resistances. As such, a greater than unity gain needed to sustain oscillation is achieved.

Frequency
L Determining C
LC Tank
Feedback Output
Network

Amplifier

Figure 1: Block diagram of an LC oscillator.


The operating or center frequency of the oscillator is fixed by the resonant frequency of the LC

tank circuit, which is defined by the equation:


1
f res = -------------------
2π LC

LC oscillators may be fixed-frequency oscillators that are mechanically tuned using variable
capacitors, or they may be controllable oscillators that output a range of frequencies. There

exsits two basic types of controlled oscillators: voltage controlled oscillators (VCO) with a

voltage control signal, and current controlled oscillators (ICO) with a current control signal. Of

the two, the VCO is more often used in IC design. The range of output frequencies that an

oscillator oscillates at over the full range of control signal is called the oscillator's tuning range.

Oscillator Topologies

There are several different topologies for LC oscillators. Classical LC oscillator designs were

based on simple single-ended configurations. The four classical LC oscillator configurations

are: Armstrong oscillator, Hartley oscillator, Colpitts oscillator and Clapp oscillator. Each of

the four are distinguished by their method of coupling the feedback signal from the LC tank.

The more recently introduced differential-pair oscillator has also been gaining popularity and

usage. Newer LC oscillator circuit designs currently being introduced are often based on the

basic configurations of the classical single-ended LC oscillators.

Both the Armstrong and the Hartley oscillators are inductively coupled. The Armstrong

oscillator (Figure 2) uses a transformer to sample its feedback signal from the inductor of the

resonant circuit. In the Hartley oscillator (Figure 3), the feedback signal is directly tapped from

R
load
C L1 L2

Figure 2: A simple Arstrong oscillator schematic.


Figure 3: A simple Hartley oscillator schematic.
the inductor. This eliminates the need for another coil. The advantages of using inductively

coupled oscillators is that the frequency can easily be varied by the net value of the capacitance

in the LC tank circuit. The output amplitude remains relatively constant when tuned over the

frequency range. Additionally, the feedback ratio of L1 to L2 remains constant. The

disadvantage, however, is that inductive coupling generates many harmonic frequencies.

Therefore, they are unsuitable when a pure sine wave output is required [1, 2].

The Colpitts and Clapp oscillators are capacitively coupled. Specifically, the Colpitts oscillator

(Figure 4) taps its feedback signal from between two capacitors in series within the LC tank.

Note the similarity between the Hartley and the Colpitts oscillator. The Hartley oscillator taps

the inductor, while the Colpitts oscillator taps the capacitor. Since capacitors, C1 and C2, are in

Figure 4: A simple Colpitts oscillator schematic.


parallel with the input and output capacitances (i.e. the internal junction capacitances between

base, emitter and collector) of the transistor, the intput and output capacitive effect of the

transistor is minimized on the resonator circuit. As a result, better frequency stability is

achieved than in the Armstrong or Hartley oscillators [2].

The Clapp oscillator (Figure 5) is just a variation of the Colpitts oscillator, in that the inductor in

the LC tank is replaced with an inductor in series with a capacitor. This allows frequency to be

tuned via L1 and C3. For a given inductor value, the Clapp configuration has a higher loaded Q

than that of the Colpitts oscillator [3].

R
C1 load
L1

C3 C2

Figure 5: A Clapp oscillator schematic.


In applications such as clock and data recovery, a tunable oscillator with differential outputs is

often required. The differential-pair oscillator provides such complementary outputs, while

additionally maximizing the rejection of common mode disturbances. The key characteristic of

the differential-pair oscillator is a cross-coupled transistor pair that switches current between

two separate LC tanks. Figure 6 shows the schematic of a differential-pair oscillator first

introduced by Nguyen and Meyer in 1992 [4]. The oscillator contains two LC tank circuits with

different resonant frequencies. Loop feedback is applied to each tuned circuit in a classical

Colpitts configuration. The loop transmission applied to each resonator is varied by voltages

Vc+ and Vc-. As such, the oscillation frequency can be varied continuously from the resonant
C1 L1 L2 C2

VC+ Q1 Q2 Q3 Q4 VC-

C3

C4
Q5 Q6

Itail

Figure 6: Schematic of Nguyen and Meyer’s differential LC VCO design[4].


frequency of the left resonator when Q1 is on and Q4 is off, to the resonant frequency of the

right resonator when Q1 is off and Q4 is on.

The basic schematic of a modified differential-pair LC oscillator that is much simpler and has

fewer components is shown in Figure 7. Transistors Q1 and Q2 take turns operating in the cut-

off and linear mode as the two differential output alternate between high and low. It must be

noted that the ideal differential amplitude of this particular circuit can be as large as the product

of the bias current and the equivalent parallel resistance of one LC tank. However, due to

source degeneration of the transistor in linear operating mode, the amplitude of oscillation will

actually be smaller than expected. Another variation of the differential-pair LC oscillator is

presented by Hajimiri and Lee [6]. In this case, only one LC tank circuit is used, with two

additional PMOS transistors.

Analysis Models

Analysis of the LC oscillator can be based on two fundamental models: the feedback model, or

the negative-resistance model. Depending on the oscillator configuration and characteristic, one
L1 L2
C1 C2

Q1 Q2

IQ

Figure 7: Schematic of a simplified differential-pair LC oscillator.


model may be preferred over the other. The feedback model, shown in Figure 8, decomposes an

oscillator circuit into a forward network and a feedback network, both of which are typically

multi-port networks. If the circuit is unstable about its operating point (i.e. poles on the right

half of the S-plane), it can produce an expanding transient when subject to an initial excitation.

As the signal becomes large, the active devices in the circuit behave nonlinearly and limit the

growth of the signal. The linear behaviour of a feedback circuit can be studied using the loop

gain quantity. Supposing the forward transfer function is a(s) and feedback transfer function is

f(s), the loop gain quantity is then defined as:


T (s) = a(s) ⋅ f (s)

The expression: 1-T(s)=0 gives the characteristic equation of the oscillator circuit from which

the poles can be found. The feedback model can be used to analyze any oscillator circuit that

has a well-defined feedback loop.

Forward Network
a(s)

Feedback Network
f(s)

Figure 8: Feedback Model.


Widely used in the design of microwave oscillators, the negative-resistance model, shown in

Figure 9, separates an oscillator circuit into a one-port active circuit and a one-port frequency

determining circuit. The function of the active circuit is to produce a small-signal negative

resistance about the operating point of the oscillator, and to couple with the frequency-

determining circuit in defining the oscillation frequency. The frequency-determining circuit is

usually a linear time-invariant circuit, and is signal independent. The active and frequency-

determining circuits can be characterized in terms of impedance quantities: Za(s) and Zf(s), as

shown in Figure 9, or in terms of admittance quantities: Ya(s) and Yf(s). The characteristic

equation of a negative-resistance oscillator can then be derived from either the expression

Za(s)+Zf(s)=0, or Ya(s)+Yf(s)=0. The negative-resistance model can be used for any oscillator

circuit where a negative resistance can be identified.

Za(s) Zf(s)

Frequency -determining
Active Circuit
Circuit

Figure 9: Negative-resistance model.

Design Considerations

The important characteristics in oscillator design are: operating frequency, phase noise, power

consumption and cost of fabrication. Oscillators being designed today are often tunable,

operating in a range of frequencies. For tunable oscillators, linearity and tuning range are also

important characteristics to be taken into consideration during design. As with the design of any

circuit block, the design of LC oscillators also involves many trade-offs between desirable

features. Research and design continuously seeks to find the optimal compromise in design.
Some examples of the trade-offs and compromises involved in the design of LC oscillators will

be given in this section on design considerations.

The operating frequency of an LC oscillator is primarily determined by the LC tank used in the

circuit. In the microwave region, a phenomenon known as multi-oscillation often appears [7].

The multi-oscillation phenomenon is quite different from that of a multi-mode oscillator. A

multi-mode oscillator uses an arbitrary number of tuned circuits to specify a well-defined set of

oscillation frequencies, and when subject to an injected instruction signal, the circuits oscillates

at only one of these well-defined frequencies [8]. Multi-oscillation, however, is caused by the

effect of parasitic elements in the oscillator circuit [7]. It is characterized by the existence of

two or more simultaneous oscillations, parasitic or unwanted oscillations together with a main

oscillation, during steady state. This causes the net steady-state signal to be severely distorted,

and can result in unwanted spurious effects. Minimizing the effect of parasitic series inductance

by using multiple parallel bond wires and board or substrate layout optimization, can eliminate

the multi-oscillation phenomenon [7].

Much work has and continues to be done in the study of oscillator phase noise, and countless

papers have already been written on this topic. Due to the scope of the topic, only a general

overview in terms of design trade-offs will be covered in this paper. The phase noise of an LC

oscillator is inversely proportional to the square of the resonator Q, which is limited by the

inductor and capacitor Q’s employed in the LC tank circuit [9, 10]. Therefore, using higher-Q

components in the LC tank can lower the phase noise in an LC oscillator.

Typical Q for a simple integrated spiral inductor is around 4-5 at 2GHz on high ohmic silicon

substrates [11]. The main losses and reduction in Q-factor are primarily due to parasitic

capacitance between the inductor and substrate, metal losses of the inductor windings at lower
frequencies, skin effect at high frequencies, eddy currents, and substrate losses due to a

relatively highly doped substrate [12]. To overcome these problems at high frequencies, smaller

inductor areas are favoured. Additionally, the inner turns of the coils can be omitted, and only

the top-level thick metal layer should be used for inductors [13]. Inductors with ploy fingers

under the winding also reduces substrate loss through capacitive coupling [14]. The alternative

also exists to reduce the total capacitance in the LC tank so as to improve phase noise by

decreasing loss. This, however, comes at the expense of making the oscillation frequency and

amplitude sensitive to parasitic capacitances that are in parallel with the LC tank.

To a first-order approximation, the phase noise can also be improved by using a higher

oscillation amplitude or a larger signal swing across the resonator [15]. This can be achieved by

increasing the bias current at the expense of greater power consumption. Note that the signal

swing cannot be arbitrarily increased in a practical VCO due to headroom limitations set by the

active device operating conditions, and by the supply voltage level used.

Reducing power consumption in an LC oscillator requires the current drain in the oscillator to be

reduced. The oscillator current drain can be lowered by using as large a load impedance as is

possible at the oscillation frequency [16]. Assuming the inductor dominates the quality factor,

Q, of the LC tuned circuit load, the load impedance is then largely determined by the inductor.

Using a larger inductance however, leads to an increase in self-capacitance which causes the

inductor to self-resonate below the target oscillation frequency. In this respect, it is most

desirable to use spiral structures with the lowest self-capacitance per unit inductance.

The scaling of transistors in the oscillator circuit should be done keeping in mind noise issues,

parasitic capacitances, and current gain and density. Device noise sources include thermal, shot,

and 1/f noise, which dominates at lower frequencies. The phase noise of an oscillator is partly
due to the 1/f noise that is upconverted into the operating frequency band. Lowering thermal

noise due to the base reistance in bipolar junction transistors (BJT) requires a larger transistor at

the expense of higher current consumption [13]. Large transistors may also suffer from the

degradation in noise performance at gigahertz frequencies because the transistor current gain

begins to decrease.

Conclusion

There is a continuing need for better purity and stability of oscillators particularly as the demand

for frequencies of operation continue to extend even higher. The has lead designers to look for

alternative solutions towards improving the current standards. In 1999, J. T. Hwang, Woo, M.

W. Hwang and Cho introduced the multiple-nested-feedback ring type LC osicllator that gave

significant improvements in speed, noise and supply voltage [17]. A variation of the nested-

feedback ring LC oscillator was used by Kim and Kim in their design that also saw

improvements in noise, as well as immunity to device mismatches and production of multi-

phase outputs [18].

Also in the realms of circuit design, a compact detector circuit for stabilizing LC oscillators was

introduced by Filanovsky [19]. Behbahani and Abidi also introduced switched tuning methods

using voltage-dependent junction capacitors at the source or drain, and voltage-dependent MOS

capacitors (source and drain shorted MOSFETs) to obtain a wide tuning characteristic with

continuously tuned segments to accommodate large shifts in the frequency of fully integrated

oscillators cause by lot-to-lot process spreads [16].

Finally, the continual advances in IC technology have also lead to advances in offshoot areas

such as oscillator design. Fundamental improvements towards low power was achieved with an

LC oscillator design implemented in an advanced bipolar Silicon-on-anything (SOA) process


[20]. LC oscillator circuits have also been implemented in various new process technologies

such as SiGe heterojunction bipolar transistors (HBT) and InP/InGaAs HBTs. In such cases,

center frequencies up to 47 to 67 GHz were attained [21, 22].

In principle, LC oscillators can be mode more and more stable by better control of the physical

elements that determine the frequency of oscillation. However, the presence of phase noise on

an oscillator is fundamentally inescapable. The price to be paid for better phase noise is higher

power, or at least higher stored energy in the oscillator system. However, both these factors are

eventually limited by physical breakdown of components.

References

[1] http://www.electronics-tutorials.com/oscillators/oscillators.htm
[2] http://www.tpub.com/neets/book9/index.htm
[3] R. W. Rhea, Oscillator Design and Computer Simulation. 2nd Edition. New York:
McGraw-Hill, 1997.
[4] N. M. Nguyen and R. G. Meyer, "A 1.8-GHz monolithic LC voltage-controlled oscillator,"
IEEE Journal of Solid-State Circuits, vol. 27, no. 3, pp. 444-450, March 1992.
[5] C.-W. Lo and H. C. Luong, "2-V 900-MHz quadrature coupled LC oscillators with
improved amplitude and phase matchings," Proceedings of the 1999 IEEE International
Symposium on Circuits and Systems, vol. 2, pp.585-588, 1999.
[6] A. Hajimiri and T. H. Lee, "Design Issues in CMOS differential LC oscillators," IEEE
Journal of Solid-State Circuits, vol. 34, no. 5, pp. 717-724, May 1999.
[7] N. M. Nguyen and R. G. Meyer, "Start-up and frequency stability in high-frequency
oscillators," IEEE Journal of Solid-State Circuits, vol. 27, no. 5, May 1992.
[8] W. A. Edson, "Frequency memory in multi-mode oscillators, " IRE Transactions on Circuit
Theory, pp. 58-66, March 1955.
[9] M. Soyuer, J. N. Burghartz, H. A. Ainspan, K. A. Jenkins, P. Xiao, A. R. Shahani, M. S.
Dolan and D. L. Harame, "An 11-GHz 3-V SiGe voltage controlled oscillator with integrated
resonator," IEEE Journal of Solid-State Circuits, vol. 32, no. 9, pp. 1451-1454, September 1997.
[10] D. B. Leeson, "A simple model of feedback oscillator noise spectrum," Proceedings of the
IEEE, vol. 54, no. 2, pp. 329-330, February 1966.
[11] F. Mernyei, M. Pardoen, W. Hob, F. Darrer, "Fully integrated RF VCO for wireless
transcievers," 1998 URSI International Symposium on Signals, Systems, and Electronics, pp. 79 -
82, 1998.
[12] J. Craninckx and M. S. J. Steyaert, "A 1.8-GHz low-phase-noise CMOS VCO using
optimized nollow spiral inductors, " IEEE Journal of Solid-State Circuits, vol. 32, no. 5, pp. 736-
744, May 1997.
[13] N. T. Tchamov, T. Niemi and N. Mikkola, "High-performance differential VCO based on
Armstrong oscillator topology," IEEE Journal of Solid-State Circuits, vol. 36, no. 1, pp. 139-
141, January 2001.
[14] T.-P. Liu, "1.5 V 10-12.5 GHz integrated CMOS oscillators," 1999 Symposium on VLSI
Circuits Digest of Technical Papers, pp. 55-56, 1999.
[15] G. D. Vendelin, A. M. Pavio, and U. L. Rhode, Microwave Circuit Design Using Linear
and Nonlinear Techniques. New York: Wiley, 1990.
[16] A. Kral, F. Behbahani and A. A. Abidi, "RF-CMOS oscillators with switched tuning,"
Proceedings of the IEEE 1998 Custom Integrated Circuits Conference, pp. 555-558, 1998.
[17] J. T. Jwang, S. H. Woo, M. W. Hwang and G. H. Cho, " Wide tunable CMOS multiple-
nested-feedback ring type LC oscillator, " Electronic letters, vol. 35, no. 2, pp. 144-145, January
21, 1999.
[18] J. J. Kim and B. Kim, "A low-phase-noise CMOS LC oscillator with a ring structure," 2000
IEEE International Solid-State Circuits Conference Digest of Technical Papers, pp. 430-431,
2000.
[19] I. M. Filanovsky, "Designer’s Casebook: Compact detector stabilizes LC oscillators," IEEE
Circuits and Devices Magazine, pp. 12, vol. 9 4, July 1993.
[20] J. van der Tang, D. Kasperkovitz, "Monolothic microwave LC oscillator based on a novel
phase noise characterization method," Proceedings of the 1999 Joint Meeting of the European
Frequency and Time Forum and the IEEE International Frequency Control Symposium, vol. 1 ,
pp. 429 -432, 1999.
[21] C. N. Rheinfelder, K. M. Strohm, L. Metzger, H. Kibbel, J.-F. Luy and W. Heinrich, "47
GHz SiGe-MMIC oscillator," 1999 IEEE MTT-S International Microwave Symposium Digest,
vol. 1, pp. 5-8, 1999.
[22] V. Schwarz, T. Morf, A. Huber, H. Jackel and H.-R. Benedickter, "Differential InP-HBT
current controlled LC oscillators with centre frequencies of 43 and 67 GHz," Electronic Letters,
vol. 35, no. 14, pp. 1197-1198, July 8, 1999.

Potrebbero piacerti anche