. First-order frequency ∆-Σ modulation Dr. Scient thesis Mats E. Høvin January 2000 . . I will greatfully thank my superviser Tor Sverre Lande and the following persons for making this thesis possible (in alphabetical order) Yngvar Berg Ole Herman Bjor Monica Finrsud Saife Kiaei Jan Tore Mairenborg Nice guys at NFR Sigbjørn Næss Alf Olsen Tanja Paetow Trond Sæther Chris Toumazou Dag Trygve Wisland . Contents Introduction 6 1 Conventional ∆-Σ noise-shaping 10 1.1 ∆-Σ analog-to-digital conversion . . . . . . . . . . . . . . . . . . . . . . . . . 10 1.2 The first-order ∆-Σ modulator . . . . . . . . . . . . . . . . . . . . . . . . . . 11 1.3 Errors and noise sources in the first-order ∆-Σ modulator . . . . . . . . . . . 13 2 Basic FDSM techniques 2.1 The non-feedback ∆-Σ modulator . . . . . . 2.2 The first-order FDSM principle . . . . . . . 2.3 The basic modulo-2n FDSM . . . . . . . . . 2.4 The D flip-flop FDSM . . . . . . . . . . . . 2.5 The pointer-FDSM . . . . . . . . . . . . . . 2.6 Feature and performance characterization of . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . the first-order FDSM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18 18 20 22 26 28 32 3 Extended FDSM techniques 3.1 FM signal undersampling . . . . . . . . . . 3.1.1 Time-domain analysis . . . . . . . . 3.1.2 Frequency-domain analysis . . . . . 3.2 The sampled-clock FDSM . . . . . . . . . . 3.2.1 Virtual frequencies . . . . . . . . . . 3.2.2 The SC D flip-flop FDSM . . . . . . 3.2.3 Resolution comparison for SS vs SC 3.3 Synchrounization . . . . . . . . . . . . . . . 3.4 Triangularly weighted zero-cross detection . 3.5 The single-bit pointer FDSM . . . . . . . . 3.6 Paralell conversion . . . . . . . . . . . . . . 3.6.1 Parallelisation . . . . . . . . . . . . . 3.6.2 Open-ended delay-line parallelisation 3.6.3 A 7-bit parallel audio FDSM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36 36 36 38 45 47 48 54 55 57 60 63 63 65 72 4 Excess noise 4.1 Phase Noise . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.1.1 Equivalent frequency noise . . . . . . . . . . . . . . . . 4.1.2 Output noise power . . . . . . . . . . . . . . . . . . . . 4.1.3 Simulation . . . . . . . . . . . . . . . . . . . . . . . . . 4.1.4 Comments on phase noise in the sampled-signal FDSM 4.2 Pattern noise . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76 76 78 79 80 80 83 4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.3 4.2.1 Background . . . . . . . . . . . . . . . . 4.2.2 A theoretical model for pattern-noise . . 4.2.3 Verification of the FDSM equivalence by 4.2.4 Valley utilization . . . . . . . . . . . . . Time-domain dithering . . . . . . . . . . . . . 4.3.1 Measurements . . . . . . . . . . . . . . 5 Conclusion and 5.1 Summary . 5.2 Conclusion 5.3 Future work . . . . . . . . . . . . . . simulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83 84 86 86 88 89 future work 92 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95 6 Publications 6.1 IEE Electronics Letters no. 1 . . . . 6.2 ISCAS no. 1 . . . . . . . . . . . . . 6.3 IEEE Journal of Solid State Circuits 6.4 ISCAS no. 2 . . . . . . . . . . . . . 6.5 IEE Electronics Letters no. 2 . . . . 6.6 Low Power . . . . . . . . . . . . . . 6.7 NorChip . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98 98 102 108 120 126 130 136 A PSD definitions 148 B C simulators 150 5 Introduction For 20-30 years ago, the word digital was used by only a small group of specialized engineers and scientists. Nowadays this word is heard almost everywhere and is used with the greatest self-confidence by everyone from the little boy in the street (nowadays more likely found in front of a play-station), to old aunt Mary. I will not suggest that the understanding of what the word digital means, is more widespread today, but who cares as long as it usually is synonymous with cheaper products with stunning new features. It is not difficult to see why digital technologies are so popular today. For those of us who are into digital design, we know that digital signal processing means high resolution, almost no problems with noise and interference, the signal can easily be stored electronically in memory, opening up for a new world of filtering techniques, and so on. In a marked driven by low-cost, short time-to-marked and high flexibility, we do not achieve much by referring to the beauty of analog solutions, unfortunately. So, the digital way of thinking is definitely here to stay, at least for a while. In many cases digital systems needs to communicate with the natural world, which we normally refer to as analog, and in this case a signal converter is needed. For several reasons, it normally pays off to let the digital domain expand in all directions, reducing the use of analog electronics to a minimum. As an example, for analog-to-digital conversion, traditional Nyquist-rate converters are now often replaced by ∆-Σ converters [1, 2, 3] reducing the problems with analog matching and high order analog anti-aliasing filtering in VLSI. The result is high quality, low-cost converters providing up to 24-bit resolution. Another example can be found in some RF demodulators where bandpass ∆-Σ converters [4, 5] have moved the digital domain closer to the antenna allowing for digital demodulation. In this way the flexibility of the RF receiver is significantly increased opening up for multi-standard communication. The main reason for the success of the ∆-Σ converter is, that most of the conversion challenge is shifted from the analog into the digital domain where it is easily taken care of by high speed digital filters. It seems however, to be a trend nowadays to focus on higher-order ∆-Σ modulators [3, 6] to further increase the resolution, or alternatively, increase the signal bandwidth. The reason for this is usually found from a limitation in the maximum sampling speed. In traditional switch-capacitor ∆-Σ modulators the sampling speed is strongly limited by slew-rate effects and finite op-amp gain. However, by increasing the order of the ∆-Σ modulator, the challenge is actually shifted back towards the analog domain again, where problems due to matching and/or stability occurs [3]. On the other hand, there have been a reason to avoid low-order modulators, particularly first-order modulators due to an inherent problem called pattern noise. In 1994, the work on the FDSM (Frequency Delta-Sigma Modulator) concept was started. This is based on a ∆-Σ modulator where the input signal is an ordinary FM signal. In this way, a frequency-to-digital converter with ∆-Σ noise-shaping was made. The first-order 6 FDSM was a purely digital circuit, and by adding a frequency modulator to it we had an alternative ∆-Σ analog-to-digital converter. A frequency modulator can be implemented in many ways. In reference [7], we have showed an example where the frequency modulator was made by a standard CMOS ringoscillator where the frequency was modulated by varying the inverter power supply voltage. In other words, the analog input signal to the ring-oscillator was the inverter power supply voltage. In this way a ∆-Σ analog-to-digital converter was implemented only by standard digital gates, and the border of the digital domain was pushed even further. The ringoscillator FDSM is particularly interesting for integrated sensor applications as the inverter delays can be directly modulated by a physical signal, with a suitable arrangement. Due to high inverter gain and noise-shaping, there are no need for an analog pre-amplifier. In this way, the digital domain is about to meet the natural world directly with almost no analog circuitry in between. A major advantage of a “digital” ∆-Σ modulator is that it will directly benefit from improvements in digital processes and thus scales with technology. When the supply voltage is reduced to 1 volt or probably less, we believe that the right way to go is low modulator order, high sampling speed, time-domain quantization as we will discuss later, and the use of standard digital gates. There are, of course, also problems with such simple solutions. By using a first-order ∆-Σ modulator, a very high oversampling ratio is normally needed to achieve a high digital resolution. But by implementing the ∆-Σ modulator only by digital gates, the sampling speed can be much higher than in a traditional switched-capacitor modulator which means increased digital resolution. As we have shown in this thesis, the resolution of a first-order FDSM may be almost 24-bit in the audio range when a high sampling speed is used combined with multi-bit conversion. Another problem usually seen in first-order ∆-Σ modulators is pattern noise. In this thesis, we have analyzed this effect extensively and proposed three different techniques that can be used to reduce or eliminate this problem. By using a very simple frequency modulation technique, the nonlinearly of the ringoscillator transfer function may be a problem. However, for low relative frequency deviations the total harmonic distortion (THD) may be quite low. As an example, in reference [7] the THD is measured to -80dB for a specific Vdd modulated ring-oscillator FDSM implemented in 1.2µm CMOS VLSI. On the other hand, by comparing this non-linearity problem to the problem with a non-linear DAC in a traditional feedback ∆-Σ modulator, the more deterministic non-linear transfer function of a ring-oscillator will be easier to linearize digitally than the more randomly given function caused by DAC level mismatch. A fundamental difference between a switched-capacitor ∆-Σ modulator and the FDSM is that quantization in the FDSM is done in time or phase and not in amplitude. In this way, we have a new axis to follow when seeking for high resolution which is more suitable for low supply voltage applications. In this thesis, we have proposed a new concept of analog/frequency to digital conversion based on parallelism. This idea utilizes both time domain quantization and the small size and simplicity of the D flip-flop FDSM. The new idea is to make a large number of very small parallel converters where the outputs are added together to make a multi-bit ∆-Σ modulator. Each converter operates on different phase or time information from the intermediate FM signal. The phase information can, for frequency-to-digital conversion applications be obtained from a delay-line, and for analogto-digital conversion applications by a modulated ring-oscillator. To summarize this overview, the overall aim of this thesis can be stated in the following way: 7 The main aim of this thesis have been to develop techniques that make it possible to reduce the analog complexity of ∆-Σ converters and in this way reduce the amount of necessary analog circuitry separating the digital domain and the natural world. Finally, two other research groups currently working with ∆-Σ frequency-to-digital conversion shold be mentioned. The first group is supervised by Professor Ian Galton at the University of Califiornia, San Diego and includes William Huff, Paolo Carbone and Eric Siragusa. The other group consists of Walt T. Bax, Thomas A. D. Riley, Miles A. Copeland, Tom A. D. Riley, Norman M. Filiol and Calvin Plett working at Carleton University, Ottawa in Canada. Both gropus have developed their own ∆-Σ FDC solutions based mainly on digital circuits. The main focus for both groups are higher order modulators [8, 9], and their results seem promising. However, for first-order ∆-Σ frequency-to-digital conversion there are no simpler solution than the D flop-flop FDSM, as this circuit directly utilize the close relationship between a hard-limited FM signal and the first-order ∆-Σ bitstream. Organization of this thesis The thesis starts by introducing the basic ideas of traditional ∆-Σ noise-shaping. Then we have chosen to include some of the work from my Cand.Scient thesis in the introduction, to make it easier for readers that are unfamiliar with the FDSM concept. Chapter 2 “Basic FDSM techniques” is therefore a summary of first-order techniques selected from my Cand.Scient thesis. In Chapter 3 “Extended FDSM techniques”, we start by introducing the undersampled FDSM. It is described how to extend the useable range of the D flip-flop FDSM which is the most simple and compact implementation of a frequency-to-digital converter with ∆-Σ noise-shaping. Then we introduce the sampled-clock FDSM, which will increase the digital resolution for certain application. In Section 3.5 we present the triangularly weighted zero-cross detector which is a mathematical model of the FDSM. In Section 3.6 we present a simplification of the pointer FDSM which is the basis for parallel conversion. Then we continue with the parallel FDSM converter, both based on a ring-oscillator solution targeted at ADC applications, and also based on an open-ended delay-line suited for FDC applications. From several technical discussions with representatives from the industry it have become clear that FM signal phase noise is a major problem in many FM demodulation applications. It have been quite difficult to convince them that phase-noise really is noise-shaped in the FDSM, as it is only intuitively argued for in my Cand.Scient thesis. Section 4.1 is therefore devoted to the analysis of phase noise in a practical FDSM system. The effect of phase noise, both on the FM signal, and also on the clock signal have been analyzed, both analytically, by simulations and by measurements. It is confirmed that phase noise, both on the FM signal and on the clock signal, is first-order shaped by the FDSM. In Section 4.3 the effect of pattern noise is analyzed. And finally, in Section 4.3 we present the method of time-domain dithering. Seven publications are attached to this thesis, but to make it more comprehensive we have chosen to include two publications published during my Cand.Scient work. The first two publications “Novel second-order ∆-Σ modulator/frequency-to-digital converter” and “Delta-Sigma Converters using Frequency-Modulated Intermediate Values” are therefore not considered to be a part of this Dr.Scient work. The next paper “Delta-Sigma Converters using Frequency-Modulated Intermediate Values”, published in the Journal of Solid State Circuits, are also mostly based on results from my Cand.Scient period but the paper is created and written as a part of this Dr.Scient work. 8 9 Chapter 1 Conventional ∆-Σ noise-shaping 1.1 ∆-Σ analog-to-digital conversion ∆-Σ AD conversion techniques have become very popular as they avoid many of the difficulties encountered with conventional methods for AD conversion. Nyquist rate converters, as illustrated in Fig. 1.1 top, have attributes that make them difficult to implement in VLSI technology. Chief among these is the use of analog anti-aliasing filters, the need for high precision analog circuits, and their vulnerability to noise and interference. The ∆-Σ converter, illustrated in Fig. 1.1 bottom, is amenable for VLSI implementation because circuit precision requirements can be significantly relaxed by a high oversampling (fs /fN ) ratio, noise shaping, and feedback. High oversampling eliminated the need for abrupt cutoffs in the anti-aliasing filter, and due to the noise-shaping, a low resolution ADC may be used without precluding a high overall signal-to-quantization noise-ratio (SQNR). In that sense, the ∆-Σ converter can trade resolution in time for resolution in amplitude. Figure 1.1: Top - Nyquist rate analog-to-digital converter. Bottom - ∆-Σ oversampled analog-todigital converter The first part of the ∆-Σ converter is the ∆-Σ modulator. This circuit modulates the 10 analog input into a high frequency digital code shifting the noise from the coarse ADC (usually one-bit), out of the frequency band of the signal. The output of the ∆-Σ modulator is a high frequency bit- or words-stream followed by a digital decimation filter. This filter will, dependent on the oversampling ratio, and the modulator order, be able to provide a high resolution digital output at a low sampling rate. High frequency components and out of band quantization noise will be removed by this filter before it could fold down to the signal band. From Fig. 1.1 one may say that the ∆-Σ technique splits the Nyquist rate conversion challenge in to parts, the decimation filter, and the smaller and simpler implemented analog ∆-Σ modulator. In this way most of the conversion challenge is shifted into the digital domain. Because the sampling rate in a ∆-Σ modulator usually needs to be much higher than in a Nyquist rate converter, conventional ∆-Σ methods are best suited for low signal bandwidth applications. They have found use in applications such as digital audio, communication and instrumentation. 1.2 The first-order ∆-Σ modulator Quantization of amplitude and sampling in time are at the heart of all ADCs. The transfer function of an ADC or quantizer is nonlinear (Fig. 1.2), and in order to achieve a simple mathematical description of the modulator, we may treat the quantization error as a random variable. output quantization levels quantization error thresholds ∆ input Figure 1.2: The input-output relationship of a quantizer For most busy input signals, we may consider this random variable to be uncorrelated with the input signal and to have a constant probability density function (pdf) over the interval ±∆/2 where ∆ is the quantization level spacing. In many cases, reported experiments have confirmed these properties, but there are two important instances where they may not apply; when the input is constant, and when it changes regularly by a multiple or sub-multiple of ∆, as can happen in feedback circuits. As a constant pdf random variable, the mean square value of the quantization error en ∈ [−∆/2, ∆/2 will be given by e2rms 1 = ∆ ∆/2 e2 de = −∆/2 ∆2 12 (1.1) and we will assume that the spectral density is constant for all frequencies. In Fig. 1.3a, the first-order ∆-Σ modulator principle is illustrated. The input to the circuit is fed to the quantizer via an integrator, and the quantized output is then fed back and subtracted from the input. 11 fs in + ∫( )dt _ a) ADC out DAC en xn b) + + _ z-1 + yn + quantizer Figure 1.3: a) The first-order ∆-Σ modulator principle. b) Linearized mathematical model The feedback forces the average value of the quantized output to track the average input. Any differences between them accumulates in the integrator and corrects itself. Unless otherwise specified, the quantizer gain, defined as the level spacing divided by the threshold spacing, will be unity for all quantizers described in this thesis. In Fig. 1.3b a commonly used simplified model of the ∆-Σ modulator is shown. The circuit is linearized by representing the quantizer by the additive noise source en , and the integrator is represented as an analog discrete-time accumulator. By considering the ideal behavior, the DAC may be disregarded. From the figure, we see that the modulator output may be expressed by yn = xn−1 + en − en−1 (1.2) in other words, the signal passes unchanged except from a unit delay while the quantization error is differentiated. Due to the differentiation, low frequency quantization noise components will be attenuated. The spectral density of the output quantization noise is given by eon = en − en−1 (1.3) and may be expressed as N (f ) = 2erms f 2/fs sin(π ) fs (1.4) In Fig. 1.4a the theoretical output noise power spectral density (psd) is shown for an example where ∆=1 and the sampling frequency is 2MHz. As we see, the quantization noise is shifted up in frequency. For the maximum signal frequency fmax , the total in-band noise power is given by integrating Eq. 1.4 over the frequency range 0 − fmax. For fs fmax , the resulting in-band noise power will, by using Eq. 1.1, be n20 ∆2 π 2 = 36 2fmax fs 3 (1.5) In Fig. 1.4b n0 is plotted as a function of fmax for ∆=1 and fs =2MHz. As we see, the in-band noise power is significantly reduced for low signal bandwidths. The improvement in 12 50 0 0 50 0 Inband power (W) (dB) Output spectral density (W/Hz) (dB) 100 1 150 200 2 100 1 150 200 2 250 250 300 0 10 1 10 2 10 3 4 10 10 Frequency (Hz) 5 10 300 6 10 0 10 a) 1 10 2 10 3 4 10 10 Max signal frequency (Hz) 5 10 6 10 b) Figure 1.4: a) Qantization noise psd plot. b) In-band quantization noise power as a function of fmax . Both figures - fs =2MHz, ∆=1. 0: Unshaped noise, 1: first-order shaped noise, 2: secondorder shaped noise resolution over ordinary oversampling do however require that the ∆-Σ modulator output is decimated by a sharp digital low-pass filter. From Eq. 1.5, the SQNR of the modulator will be 3 SRo ∆2 π 2 2fmax √ SQN R = 20 log − 10 log (1.6) 36 fs 2 2 where SRo is the output signal range which by the use of a unity gain quantizer will be equal to the input signal range. From Eq. 1.6 we have that each doubling of the oversampling ratio (OSR) increases the SQNR by ≈9dB. An alternative way to increase the SQNR is by decreasing the quantization spacing ∆, or increasing the input signal range. By doing so, we have to increase the number of quantization thresholds, and the result is a SQNR gain of ≈6dB for each doubling of the number of effective quantization thresholds which in turn doubles the output word-length of the modulator. In Fig. 1.5 the output psd from a single-bit ∆-Σ modulator simulation is shown. The input signal is a sinusoidal with maximum amplitude and a frequency of 100Hz. The plot is normalized to let 0dB correspond to the maximum signal amplitude, where the input signal range is limited to [−∆/2, ∆/2]. The sampling frequency was set to 2MHz, and we notize that the noise power is shifted up in frequency. There is also some pattern noise at lower frequencies. The power spectral density is estimated by a modified periodogram estimator, see Appendix A. 1.3 Errors and noise sources in the first-order ∆-Σ modulator So far, we have treated the quantization error as uncorrelated with the input signal. In a practical modulator, this assumption will not always be sufficient. In addition, the ∆-Σ modulator is an analog device where the effect of inaccurate components must be considered. In the following sections, the main limiting error and noise sources will be presented. 13 Figure 1.5: a) FFT analysis of simulated ∆-Σ modulator output sequence, fs =2MHz, ∆=1 Pattern noise For DC or slowly varying inputs, the output will bounce between two levels, keeping its mean equal to the input. The output sequence may be repetitive and the frequency of this sequence will be determined by the input DC level. When the repetition frequency lies in the signal band, the modulation is noisy, but when it does not, the modulation is quiet. Fig. 1.6 top [3] shows how the in-band noise power depends on the value of the DC input for a ∆-Σ modulator with quantization levels at ±1 and a low OSR. Fig. 1.6 bottom (scanned from reference [10]) shows the corresponding noise for a high OSR, and quantization levels at 0 and 1. There are peaks of noise adjacent to integer divisions of the space between levels, elsewhere the noise is small. This structure of the quantization error is called pattern noise, and it will appear as spurious tones/peaks in the modulator output spectrum. The larger peaks in Fig. 1.6 can far exceed the expected noise level calculated from Eq. 1.5, and the following properties of pattern noise are noteworthy: I) The height and width of each peak are inversely proportional to the OSR. The height and width are also inversely proportional to the denominator of the fraction that describe the position of the peak within the quantization interval relative to the level spacing. II) About half the total power is in the end peaks, one-sixteenth in the center peak. Pattern noise may however be randomized by injecting a signal with frequency well above the signal band at the input. This signal, also called a dither signal ”smears out” the output power spectrum, and as a consequence, it reduces the energy of the noise peaks/tones. For one-bit, first-order ∆-Σ conversion, the pattern noise problem may be significant as the inband pattern noise power may be large compared to the signal power. If not the full output signal range is used, the output signal range may however be biased and located well away from the main noise peaks to reduce this problem. DAC output levels The main problem for multi-bit ∆-Σ modulator operation is inaccurate DAC output levels. From Fig. 1.3 we notice that all errors introduced in the feedback path will add directly to the signal. In addition, since yn is close to xn in a multi-bit ∆-Σ modulator, 14 In-band noise power (dB) DC input level DC input level Figure 1.6: Measured in-band noise v.s input signal DC value. Arrows indicate theoretical in-band noise from Eq. 1.5. Top - low OSR. Bottom - High OSR the correlation between the input and the DAC error will be large. Thus the DAC error will usually have a large baseband energy content and will generate harmonic distortion. The DAC output levels must therefore be positioned with an accuracy better than the overall SQNR which for high resolution conversion normally will require, either external components and/or trimming, or the use of more or less complex correction techniques [11, 12, 13, 14, 15]. The use of external components or trimming will, however, raise the cost of the system considerably. The fact that the DAC output settling time must be very low (less than 1/fs ), eliminates the use of most ordinary high resolution DAC architectures. For single-bit ∆-Σ modulator operation only two DAC output levels are required, and any misplacement is equivalent to a bias in the modulator input signal. The only requirement will therefore be properly placed DAC output levels to accommodate the input signal range. ADC thresholds The quantizer in the multi-bit ∆-Σ modulator usually takes the form of a flash ADC. Misplaced ADC thresholds may be regarded as a non-linearity of its gain, but due to the high gain feedback, such errors have little effect on the baseband properties of the modulator. Misplacing the thresholds by as much as a quarter of their spacing is usually tolerable. For single-bit ∆-Σ modulator operation, the only requirement on the single threshold placement is to prevent overloading in the integrator. 15 ADC overloading The internal signal swing in the integrator uses up some of the dynamic range of the circuit, and if the ADC is not to overload, the input signal range (SRi ) must not exceed the maximum and minimum ADC output levels. If a large input causes the ADC to overload, the modulator noise will increase rapidly. Simulations show that excess noise introduced by larger signals appears as odd-order harmonic distortion of sine waves. The excess noise will decrease with increased OSR but increases with the frequency of the applied sine wave. Integrator errors In the integrator, there are several noise sources, but small errors introduced by the integrator will generally due to the high gain feedback have little impact on the overall modulator performance. For most applications a 10% change of integrator gain will be tolerable. Integrator leakage decreasing the integrator DC gain to at least the OSR will not be a problem. The integrator slew rate will however be a limiting factor for high sampling frequency operation. Most ∆-Σ modulator integrators are implemented by the use of switched capacitor (SC) techniques, and linear capacitors and a non-overlapping two-phase clock is normally required. To reduce the effect of power supply noise, charge injection and clock feedtrough, a fully differential architecture is normally used. The SC technique is very little sensitive to clock jitter, but to achieve a low kT /C level, there will be a minimum limit on the capacitor values. For high SNR operation, the minimum value may typically be in the 10pF range. Due to this constraint, there will be a trade-off between high speed operation combined with a low kT/C level, and low power consumption. It should be mentioned that in a ∆-Σ modulator, the main power consumer will normally be the integrator, but in a complete ∆-Σ converter, the decimator power consumption may dominate. Implementation constraints As previously stated, by using SC techniques in the ∆-Σ modulator, linear capacitances are required, and the SC-∆-Σ modulator can generally not be implemented in the least expensive standard digital CMOS processes, since an extra poly-layer will be required. In addition, as a voltage-mode device, the SC-∆-Σ modulator is not well-suited for low power supply voltage operation. With power supply voltages close to 1V likely to become a future industrial standard, process parameters such as threshold voltage will be chosen to optimize digital performance so voltage domain behavior will suffer as a consequence. To overcome these limitations, switched current (SI) techniques have been proposed [16], but so far no high-resolution SI-∆-Σ modulators have been reported due to several implementation difficulties associated with this new technique. 16 17 Chapter 2 Basic FDSM techniques The frequency delta-sigma modulator is a feed-forward device with no feedback. Usually ∆-Σ modulation is synonymous with feedback, but in the next section we will prove that the traditional first-order ∆-Σ modulator is equivalent to a cascade of an integrator and a differentiator where the quantizer is located in between. This is shown without linearizing the quantizer function. 2.1 The non-feedback ∆-Σ modulator For now, it will be convenient to have a mathematical definition of the quantizing function. For a practical ∆-Σ modulator, where the number of quantizer output levels is restricted, the level placements must be properly chosen to accommodate the input signal range due to the feedback. In this discussion, the number of levels is assumed to be large, and therefore, the level placements will not be critical. The unity-gain quantizing function q(x) with quantizing thresholds integer multiplies of l will therefore simply be defined as q(x) = x − modl (x) (2.1) where modl (x) is the remainder of x after division with l. Compared to the quantizing function illustrated in Fig. 1.2, q(x) is shifted down by ∆/2 as illustrated in Fig. 2.1. The quantization error en will now be restricted to the interval −1, 0], but by considering its new mean value as a bias in the input signal, the root mean square (rms) value will be the same as for the former used quantization function, and the theoretical results from Chaper 1 will therefore still be valid for q(x). The Z-transformed representation of a conventional discrete-time first-order ∆-Σ modulator is shown in Fig. 2.2. In this circuit, the ADC or quantizer is represented by the complex quantization function Q, defined by Q[U (z)] = Z{q(un )}. The output may then be expressed as −1 z z −1 Y (z) = Q X(z) − Y (z) (2.2) 1 − z −1 1 − z −1 In [7] it is proven that in general, for two real sequences fn and gn , the quantization function Q posses the property Q[F (z) + P · G(z)] = Q[F (z)] + P · G(z), 18 if q(gn ) = gn (2.3) q(x) output levels quantization error thresholds x ∆ Figure 2.1: The q(x) quantizer function X(z) + _ z−1 + U(z) Y(z) Q( ) Figure 2.2: Z-transformed representation of the traditional first-order ∆-Σ modulator if the sequence gn is already quantized. In this statement, P is a general complex polynomial of the form z −a /(1 − z −1 )b where a and b are positive integers. By using this property, referred to as the resolving property, the Y(z) terms of Eq. 2.2 may be resolved from the quantizer function, and the equation may be expressed as −1 z −1 Y (z) = (1 − z )Q X(Z) (2.4) 1 − z −1 X(z) z-1 + + Q( ) + z-1 z-1 Y(z) - Figure 2.3: Equivalent first-order ∆-Σ modulator But Eq. 2.4 is an expression describing a different circuit with the same mathematical behavior, and it may be sketched by Fig. 2.3. The corresponding time domain representation is n n−1 yn = q xi−1 − q xi−1 (2.5) i=−∞ i=−∞ This new circuit reveals the simple principle of the first-order ∆-Σ modulator. Except from the unit delay, the modulator operates by first integrating the signal, then quantizing it, and finally differentiating the signal and the quantization error to restore the signal. Since the quantization error is not integrated, it will be differentiated, and the quantization noise 19 will be shifted up in frequency and out of the signal band. By considering the quantization error as uncorrelated with the input signal, the signal will pass unchanged except for the unit delay. To achieve a better understanding of the first-order ∆-Σ noise shaping principle we may think of the modulator as a cascade of a continuous-time integrator and a derivator. Their equivalent gains will be 1/Ts and Ts respectively, where Ts = 1/fs is the sampling interval. In between, the quantization noise is entering the circuit as illustrated in Fig. 2.4. e x ∫ ( )dτ + + 1/Ts Ts d () dt y Figure 2.4: Another look at the first-order ∆-Σ noise-shaping principle In this circuit, we have two different mechanism both reducing the in-band noise in the output signal. First, by letting the error enter in between the integrator and the derivator, the output noise will be given by the derivative of the error. This mechanism shapes the noise by suppressing low frequency in-band error components. The next noise suppressing effect follows from the fact that the error enters in between the two complementary gain factors 1/Ts and Ts which scales up the signal compared to the error by a factor 1/Ts for all frequencies. By doubling the sampling frequency 1/Ts , we actually double the magnitude of the input signal compared to the magnitude of the error, and this gives a ≈6dB increase in SQNR. It is further a well known result that by sampling a signal mixed with white noise, the in-band SNR increases with ≈3dB for each doubling of the sampling frequency. Together these two factors explains the increase in SQNR by 6+3=9dB for each doubling the sampling frequency in the traditionally ∆-Σ modulator. 2.2 The first-order FDSM principle By considering the theoretical circuit in Fig. 2.3, there are no feedback and thus no DAC, and this eliminates all problems due to misplaced DAC output levels. As the output of the quantizer is digitally differentiated, all baseband noise introduced by misplaced quantizer thresholds will be heavily suppressed. These two features open up for straightforward multibit quantization. The drawback is however that due to the lack of feedback, inaccuracies in the integrator adds directly to the signal. The modulator SNR will therefore be restricted to the SNR of the integrator, which for high SQNR operation must be very accurate. In addition, without feedback, a conventional accumulator will at some point normally saturate, and this is why the FM representation of signals become interesting. A FM signal be expressed as [17] f m(t) = sin[θ(t)] where the instantaneous angle is given by θ(t) = 2π (2.6) t −∞ (fc + kx(τ ))dτ (2.7) In this expression fc represents the carrier frequency, and k the frequency sensitivity to the modulating signal x(t). From Eq. 2.7 we notice that the FM signal or more exactly the 20 frequency modulator source by its angle θ(t) is an integrator with respect to the modulating signal x(t). This property do also follow from the fact that the FM carrier may be considered as a phase modulated carrier where the modulating signal is pre-integrated. The frequency modulator may be very useful as an ∆-Σ modulator integrator in applications where the feedback is removed. It will never saturate as it may be considered as a modulo integrator where θ(t) represents a rotating pointer where the angular speed is modulated by the input x(t). The modulo operation on the angle is carried out by the sin() function. The integrator SNR will correspond to the SNR of the FM signal which may be high, depending on the architecture of the frequency modulator. The frequency modulator is however a continuoustime integrator, and by exchanging the discrete-time integrator in Fig. 2.3 with a frequency modulator as illustrated in Fig. 2.5, the resulting FDSM will be mathematically equivalent to a conventional ∆-Σ modulator with a continuous-time loop integrator and input Ts (fc + kx(t)) [7]. fs θ(t) fm(t) θn/2π detector x(t) θn/2π + q( ) _ yn z-1 frequency modulator Figure 2.5: The general first-order FDSM But for a high OSR we may from Eq. 2.7 approximate θn as θn ≈ 2πTs n (fc + kxi ) (2.8) i=−∞ and the FDSM overall operation may, in principle, be described by applying this sampled FM variable directly to the quantizer. Since there is no unit delay, the output will from Eq. 2.5 be n n−1 yn = q T s (fc + kxi ) − q Ts (fc + kxi ) (2.9) i=−∞ i=−∞ By representing the quantizer by the additive noise en , we have yn = T s n i=−∞ (fc + kxi ) + en − Ts n−1 (fc + kxi ) − en−1 = Ts (fc + kxn ) + en − en−1 (2.10) i=−∞ As we see, the modulating input signal xi is just scaled and biased, while the quantization error is differentiated. The overall circuit will therefore, except from the scaling and biasing factors, provide equivalent ∆-Σ noise-shaping with respect to x(t). We may also consider it as delta modulation [18] with respect to θn since the sub-circuit handling the FM signal actually is a delta modulator. Buried in the output bit or word stream, the effective output signal range will be SRo = 2∆f /fs = Ts kSRi 21 (2.11) where ∆f is the maximum frequency deviation caused by the input signal range SRi . By doubling the sampling frequency in the first-order FDSM, the in-band quantization noise is reduced by ≈9dB. However, from Eq. 2.11 we see that the signal range is also reduced by a factor of two. The net gain in SQNR will, therefore only be ≈3dB. In FD applications, the input signal will already be given as a FM carrier i.e. the frequency modulator is implemented in the signal source. For these applications, the FDSM system must be considered to include the signal source integrator even if it is implemented as an oscillating sensor or as a traditional frequency modulator several miles away in a radio transmitter. In these applications, the integrator SNR is already given by the SNR of the received FM signal. Any noise introduced to the FM signal, including noise generated by the θn detector, will be first-order shaped as the FM signal is the integral of x(t). Due to the high noise immunity of the FM signal, the accuracy requirement of the θn detector will be very low. 2.3 The basic modulo-2n FDSM As in the traditional ∆-Σ modulator, the ADC thresholds placement and spacing may be varied, but in the FDSM system, we will see that the implementation will be particularly simple by locating the quantization thresholds to integer values by letting the module l of q(x) equal 1. θn detection By looking at the sampled FM variable θn /2π, the quantizer input may be expressed as θn = pn + ϕn (2.12) 2π where pn is an integer representing the number of received FM periods or rising FM edges at time nTs , and ϕn ∈ [0, 1 represent the phase difference between the previous rising FM edge and the sampling edge scaled by 1/2π as illustrated in Fig. 2.6. sample pn-1 fm(t) pn 2πϕn θn=pn2π+2πϕn pn+1 ϕn 2πpn Figure 2.6: The splitted angle representation The modulator output may now be expressed by θn θn−1 yn = q −q = q[pn + ϕn ] − q[pn−1 + ϕn−1 ] 2π 2π (2.13) By the use of integer quantization thresholds, pn and pn−1 will be already quantized, and according to the resolving lemma [7], the output will be equivalent to yn = pn − pn−1 + q[ϕn ] − q[ϕn−1 ] 22 (2.14) But since ϕn is restricted to the interval [0, 1 and we are using integer quantization thresholds, the quantization error will be −ϕn , and the output from the quantizer function will always be zero which let us reduce Eq. 2.14 to yn = pn − pn−1 (2.15) This is simply the received number of rising FM edges during the sampling interval Ts . From this discussion we realize that the most straightforward first-order FDSM may be implemented simply as a frequency modulator followed by a count and dump FD converter as shown in Fig. 2.7. fs fm(t) x(t) reset period counter reg θ(t) yn count frequency modulator Figure 2.7: The most straightforward FDSM implementation To put it in another way, we have shown that by introducing oversampling and proper decimating in a traditional count and dump FDC system, we achieve multi-bit ∆-Σ noise shaping with respect to the modulating signal. Since the quantization error is −ϕn , the output from this FDSM may from Eq. 2.10 also be expressed as yn = Ts (fc + kxn ) − ϕn + ϕn−1 (2.16) As a ∆-Σ analog-to-digital-to-frequency converter, the SQNR will from Eq. 2.11 and Eq. 1.6 be 3 kSRi π 2 2fmax √ SQN R = 20 log − 10 log (2.17) 36 fs 2 2fs As we see, this expression is identical to the traditional ∆-Σ modulator expression, except from the scaling of the input signal by k/fs . Since multi-bit quantization is inherent in the FDSM concept, a significantly higher SQNR may be achieved in the FDSM by the use of a high k value. A high k value can be obtained by the use of a high frequency deviation. Since the non-linearity of the frequency modulator normally will increase with the ∆f /fc ratio, the use of a high carrier frequency may be the best way to go. If the frequency modulator is already implemented in the signal source, we may from Eq. 2.11, more easily express the SQNR of the FD subcircuit by 3 ∆f π2 2fmax SQN R = 20 log √ − 10 log (2.18) 36 fs 2fs It will, of course, be possible to increase the dynamic range by counting both rising and falling FM edges. This will directly correspond to letting θn in Eq. 2.12 represent the received number of halve periods at time nTs , and ϕn ∈ [0, 1 be the phase difference between the last 23 FM edge and the current sampling edge scaled by π. Since the quantization error rms value will be unchanged while the signal range is doubled, an increase in SQNR of ≈6dB will result. Probably, the most practical implementation may then by the use of a separate positive and negative edge triggered counter and simply add their outputs. The first objection may be that an accurate FM signal duty-cycle of 50/50 may be required, but simulations show minor SNR degradations for minor FM duty-cycle deviations. This is reasonable since the dynamic range is independent of the duty cycle while the quantization error range will increase from [0, 1 to [0, 2 when the duty cycle approach 0/100 or 100/0. By reducing the duty-cycle to 0/100 or 100/0, the 6dB increase in SQNR will therefore gradually be lost. The modulo-2n counting principle The counter resetting operation will be a limiting factor for high-speed operation. According to Eq. 2.15 and by considering the count and dump circuit as an ideal counter with no upper limit followed by a digital differentiator, the ideal counter may be realized as a modulo-2n counter as shown in Fig. 2.8. fs fm(t) modulo-2n counter n-bit reg θ(t) x(t) + _ reg frequency modulator n-bit output word-stream Figure 2.8: The basic modulo-2n FDSM with a positive edge triggered modulo-2n counter By disregarding the borrow bit from the differentiator, the modulator output will be equivalent to the output from the modulator in Fig. 2.7 assuming that the maximum number of received rising FM edges during Ts is smaller than the module of the counter. In most applications, non-varying bias components in the measured signal have no information content. In the FDSM, the carrier will normally be of no interest, and by using modulo-2n arithmetic, the only restriction on the size of the module is that it have to be larger than the maximum variation of the number of received rising FM edges during Ts to avoid signal aliasing. The counter may therefore pass through several cycles during the sampling interval as long as the maximum difference in its outcome is less than 2n . This property let us reduce the internal word-length in the modulator without loosing any information about the signal itself. By using a smaller module than the maximum number of received rising FM edges during Ts , the modulator output will be represented in modulo-2n arithmetic. In reference [3] it is shown that an efficient way to implement the first stages of the decimator may also be done in modulo-2n arithmetic. Therefore, if a decimator where the first stages are based on modulo-2n arithmetic is chosen, no interfacing is required. The modulo-2n output word-stream may also easily be converted to a standard representation by adding a proper bias and disregarding the carry bit. The minimum number of received rising edges during Ts = 1/fs will be the integer part of (fc − ∆f )/fs . If (fc + ∆f )/fs is not an integer, which is only of mathematically interest, the maximum number of received rising edges will, as illustrated in Fig. 2.9, be the integer part of (fc + ∆f )/fs + 1 as the sampling period may be shifted in time. If 1/fs is close to an integer multiple of 1/fc, the minimum variation in the counter outcome, may therefore, even for very small signal ranges be 3. For the basic modulo-2n 24 1/(fc-∆f) a) fm(t) 0 1 2 3 1/fs 1/(fc+∆f) fm(t) 0 b) 1 2 3 4 4 5 1/fs 1 2 3 Figure 2.9: a) Minimum number of received positive FM edges during 1/fs . b) Maximum number of edges during 1/fs FDSM, the module may in general, be chosen to be the smallest integer larger than int fc + ∆f fs + 1 − int fc − ∆f fs ≤ int 2∆f fs + 2 = int(SRo ) + 2 (2.19) where int(x) is the integer part of x. The minimum internal word-length will then be log2 [int(SRo ) + 2] (2.20) and the output bias component due to fc will be clipped down to mod2n (fc /fs ). For low output signal ranges the shortest internal word length will therefore generally be two bit. No counter resetting is needed, and the sampling speed may be considerably higher than by using the traditional count, dump and reset converter. Counter readout The counter reading operation may be considered as a data transfer between two independently clocked systems as the counter is asynchronous to the rest of the system. When a multi-bit word is to be transferred between two such systems, great care must be taken to avoid situations when the receiving system tries to read the data at the same time as more than one bit changes. In that sense, the binary data representation is not well suited, and if a binary counter is to be used, a separate synchronizing circuit must be applied. There are however other ways to overcome this problem, and one way is the use of Gray code, either by a modulo-2n Gray code counter or a binary counter followed by a binary to Gray code converter. Both solutions will require some more logic, and a Gray to binary converter must also be applied. Another solution may be using a ring-counter followed by a binary encoder. If the counter length is large, this solution will however require a large encoder and a lot of wiring. System-level simulations To verify the theory, and simulate the ideal behavior, several system level simulations have been carried out. Details of the simulator are given in [7]. There are several more or less sophisticated ways to estimate the output power spectral density from the simulated word-streams. Throughout this thesis, the most straight-forward and commonly used FFT 25 technique where the psd is estimated by the periodogram estimator [19] (Appendix A), is used. To better illustrate the modulator dynamic range, the estimated psd is scaled to let 0dB represent the maximum input signal amplitude for all plots throughout this thesis. When working with high SNR ∆-Σ output spectrums, a high window sidelobe suppression factor will be desirable, and a 6-term Blackman-Harris-Hodie window (see Appendix A), which is specially suited for high SNR analysis is therefore used in all computations. Figure 2.10: A 218 point FFT analysis of a ideal modulo-22 FDSM output. Max signal amplitude, signal frequency 100Hz, fc =10MHz In Fig. 2.10 the estimated output psd for a basic modulo-22 FDSM where fs = 2MHz, fc =10MHz and the input is a single sinusoid with maximum amplitude and frequency of 100Hz is shown. The maximum input signal amplitude is defined as the amplitude that produces a maximum intermediate frequency deviation of ±10%. By this choise of circuit parameters, the modulator SRo will from Eq. 2.11 be 2 · 1 · 106 /2 · 106 = 1, which is the same as for the traditional first-order ∆-Σ modulator simulation presented in Chapter 1. The result in Fig. 2.10 may therefore directly be compared to the result in Fig. 1.5. As we see, the two spectrums are quite similar with the same ≈20dB/decade slope, and the ∆-Σ noise-shaping behavior of the FDSM is confirmed. Compared to Fig. 1.5, the FDSM spectrum is smoother with less pattern noise. This is only a consequence of the different quantization level location. For the traditional ∆-Σ modulator simulation, the sinusoidal input signal have its peak values equal to the quantization levels. The signal will then, for a longer period be in the large pattern noise end-peak sections, illustrated by Fig. 1.6, than the FDSM output signal, which have its integer quantization level in the middle of its output signal range of [4.5,5.5]. For the FDSM, a higher SQNR may be obtained by increasing the sensitivity k of the frequency modulator. 2.4 The D flip-flop FDSM From the previous discussion, by using a module less than the maximum received number of FM edges during Ts , the minimum FDSM internal word-length will, in general, be two 26 bit. But if the maximum received number of FM edges during Ts is less than two, we may, without risking signal aliasing use modulo-21 arithmetic. This follows from the fact that the maximum difference in the counter outcome will be two, which will be accommodated by modulo-21 arithmetic. By using modulo-21 arithmetic and disregarding the borrow bit from the subtractor, the subtractor may be implemented simply as a XOR gate. A onebit modulo counter will, easiest be implemented as a double edge counter, since a one-bit modulo counter triggered by both rising and falling edges is equivalent to a D flip-flop. The entire FDSM may therefore be implemented as a frequency modulator connected to a D flip-flop followed by a one-bit register and a XOR gate as illustrated in Fig. 2.11 θ(t) x(t) fm(t) fs CK D Q reg reg output bit-stream fs frequency modulator Figure 2.11: The D flip-flop FDSM FM a) out FM b) out Figure 2.12: The D flip-flop FDSM intermeadiate/output signals. a) Modest fs /fc ratio. b) High fs /fc ratio The operation of the D flip-flop FDSM may also be considered as a frequency modulator connected to a synchronizer or direct bit-stream converter as illustrated in Fig. 2.12. From this discussion, we notice the close relationship between the FM representation of signals and the first-order ∆-Σ noise-shaped bit-stream. By rising the fs /fc ratio as illustrated in Fig. 2.12b, the bit-stream will approach an exact representation of the analog information which in the FM signal is given by the zero crossings or edge positions. In that sense, the first-order ∆-Σ noise-shaped bit-stream may also be considered as a quantized FM signal. Since the D flip-flop FDSM is counting both rising and falling edges, its SQNR will be given by 3 k · SRi π 2 2fmax SQN R = 20 log √ − 10 log (2.21) 36 fs 2fs If the frequency modulator is already implemented in the signal source, we may for FD applications, express it as 3 √ ∆f π2 2fmax SQN R = 20 log 2 − 10 log (2.22) fs 36 fs Compared to the traditional ∆-Σ modulator, the output signal range in the D flip-flop FDSM will for a high fs /fc ratio be very small, and by locating the output signal range well 27 away from the main pattern noise peaks illustrated in Fig. 1.6, we may achieve a slightly higher SQNR than the theoretical value since the pattern noise problem in this way may be eliminated. System-level simulations A high-level D flip-flop FDSM simulator have been made (see Appendix B), and by following the same procedure as in the previous section, the psd is estimated by FFT analysis of the ideal output bit-stream. Several simulations have been carried out, and in Fig. 2.13b, the estimated output psd is shown for a D flip-flop FDSM with fs =100MHz, fc =40MHz and a maximum frequency deviation of ±10%. a) b) Figure 2.13: A 218 point FFT analysis of the D flip-flop FDSM output for max input amplitude. Input signal frequency 3439Hz, fs =100MHz, fc =40MHz. a) 35/65 FM duty cycle. b) 50/50 FM duty cycle As we see, the noise spectrum is shaped according to the theory with a slope of ≈20dB/decade. There are some high-frequency pattern noise, but for low and medium frequencies, the spectrum is quite smooth. Compared to the basic modulo-22 spectrum shown in Fig. 2.10, the SQNR will from Eq. 2.18 and Eq. 2.22 be increased by ≈35dB, and we should expect the noise floor to be lowered by 35-10·log(100MHz/2MHz)≈18dB. As we see, the noise floor is further lowered by almost 10dB, which is a consequence of the signal range location in the quantization level interval and the low signal range. To illustrate the effect of duty-cycle deviations, the same simulation was done with a FM duty cycle of 35/65, and as we see from Fig. 2.13a, the noise floor is just slightly raised. 2.5 The pointer-FDSM In some applications the frequency modulator can be implemented as a ring oscillator, where the propagation delay of each inverter is modulated by the input signal. Dependent on the architecture, the modulating signal may be given either as a current or as a voltage or even as a directly measured physical parameter. Examples are: Current controlled bipolar ring oscillators [20], acceleration modulated ring oscillators, and pressure modulated ring oscillators where the inverters are located on a silicon membrane. A particularly simple 28 voltage controlled oscillator (VCO) may also be implemented by standard CMOS inverters connected in ring where the frequency is modulated by the inverter power supply voltage. Although the exact relationship between the power supply voltage and the loop frequency is quite complex, Spice simulations indicates that the loop frequency is a surprisingly linear function of the power supply voltage over a large range as shown in Fig. 2.14. 6 14 x 10 12 Frequency (Hz) 10 8 6 4 2 0 1 1.5 2 2.5 3 3.5 Input (Volt) 4 4.5 5 5.5 Figure 2.14: The frequency of a 15-inverter CMOS ring-oscillator as a function of the inverter power supply voltage simulated by Spice. The straight line is fitted to the simulated data by linear regression In such a VCO, there will be no power supply noise in the traditional sense. By considering the ring oscillator itself as a modulo-m counter, we may both simplify the FDSM, and increase the resolution. The ring oscillator node values may be considered as a cycling binary state vector with 2m different states where m is the number of nodes or inverters as illustrated in Fig. 2.15. 0 1 0 1 0 0 1 0 3 2 1 1 1 1 0 0 0 1 1 1 0 Figure 2.15: The different states of a 3-inverter ring-oscillator The operation of the ring oscillator may be described by a transition edge running through all inverters in sequence, and by looking at the state diagram, we realize that at a given time instant, all neighboring nodes will, apart from the two at each side of the edge, have their complementary values. The logical XNOR or equivalence function between each neighboring nodes will therefore provide a active high ”pointer” output which will run through all nodes in sequence with a loop frequency twice the overall ring oscillator frequency. By sampling the node values by D flip-flops, and generating the logical XNOR between each neighboring nodes, the m bit pointer vector consisting of one ”1” and m − 1 29 binary encoder ’0”s may be fed to a binary encoder, providing a sampled modulo-m representation of the state of the oscillator Fig. 2.16. 4-bit reg borrow _ + _ + 4-bit modulo-15 output word-stream 4-bit Figure 2.16: A 15-inverter pointer-FDSM. The outer ring symbolize the individual D flip-flop and XNOR units The output may then be differentiated to form a modulo-m representation of the number of pointer shifts during Ts . Compared to counting each rising FM edge, the pointer-FDSM will therefore, in addition, provide quantized phase information by counting 2m times for each overall FM period and thereby increase the resolution. The maximum number of pointer shifts during Ts will be Ts /(τ0 − τ0 δτ ) where δτ is the maximum deviation in propagation delay relative to the unmodulated propagation delay of one inverter τ0 . Together with the minimum number of pointer shifts Ts /(τ0 + τ0 δτ ) the effective output signal range will be SRo = Ts Ts 2δτ − = τ0 (1 − δτ ) τ0 (1 + δτ ) fs τ0 (1 − δτ 2 ) (2.23) and for δτ 1 it may be simplified to SRo = 2δτ fs τ0 (2.24) As we see, the SRo is proportional to the maximum relative propagation delay deviation, and inversely proportional to the sampling frequency and the unmodulated inverter propagation delay. From Eq. 2.24 we also notice that for a given inverter propagation delay, the resolution of the pointer-FDSM is independent of the number of inverters. The minimum number of inverters or module must however still be larger than int(SRo + 2) to avoid signal aliasing. The ring-oscillator carrier frequency may be expressed as fc = 1 2τ0 m (2.25) Since the maximum relative frequency deviation ∆f /fc will be equal to δτ , we may also express the output signal range as SRo = 2m∆f fs (2.26) which is 2m times larger than for the basic modulo-2n FDSM. However, since the number of inverters m, in a ring oscillator must be odd, the output of the binary encoder can not be represented in modulo-2n arithmetic, and a standard binary 30 subtractor can not be applied. To overcome this obstacle, some extra logic is required. By using 2n − 1 inverters, and two binary subtractors as illustrated in Fig. 2.16, we achieve a modulo-m represented output difference, and by finally adding a proper bias modulo-2n to the result, we arrive with a standard represented output. The modulo-m subtraction may probably be implemented in a simpler way, but this proposal just illustrates the principle. In a practical implementation, there will be process dependent deviations amongst the different inverter delays and within the D flip-flop thresholds. Such errors may be considered as phase offsets and will as duty-cycle deviations be systematic errors. Simulations indicates that the effect of phase offsets will, in the overall SNR, correspond to the effect seen by duty cycle diversions in the D flip-flop FDSM, and will thus be of less significance. If the unmodulated and maximum inverter delay is given, the SQNR of the pointer-FDSM will be 3 δτ π2 2fmax √ SQN R = 20 log (2.27) − 10 log 36 fs fs τ0 2 If the ring oscillator carrier frequency and max frequency deviation is given, we have 3 √ ∆f π 2 2fmax SQN R = 20 log m 2 − 10 log (2.28) fs 36 fs For the same fc and ∆f , we notice that the dynamic range is increased by m compared to the D flip-flop FDSM. System-level simulations Several system-level simulations have been carried out to verify the theory and the ideal behavior of the pointer-FDSM. In Fig. 2.17a, the estimated psd for a 15-inverter pointerFDSM output is shown. A 3.33ns inverter propagation delay is chosen which will provide a carrier frequency of ≈10MHz. The input is a single sinusoidal signal with maximum amplitude and a frequency of 100Hz. The maximum frequency deviation is set to 5%. As we see, the spectrum is shaped according to the theory. Compared to the basic modulo-2n FDSM spectrum in Fig. 2.10, the noise floor should from Eq. 2.18 and Eq. 2.28 be lowered by ≈24dB which is close to the simulated result. In Fig. 2.17b, the same simulation is carried out for randomly chosen phase offsets in the range [0, 1 rad. As we see from the figure, the noise floor is not significantly raised, and due to the dithering effect it is slightly smoother than for the ideal simulation. An application example For the specific pressure sensor (hydrophone) developed by Schlumberger-Geco-Prackla, the sensor was implemented as a ring-oscillator mounted on a silicon membrane. The membrane is then exposed to a differential pressure causing stress in the silicon substrate. This stress modulates the transistor β parameter, causing a modulation in the ring-oscillator frequency. Sensor output specifications are: digital resolution 20bit, signal bandwidth 400Hz, maximum sampling frequency 2MHz and relative frequency deviation ∆f /fc . For a pointerFDSM with ∆f /fc=0.1, the SQNR will from Eq. 2.28 be 128dB, and from the simulated result in Fig. 2.17a we may conclude that this ring oscillator will provide the necessary resolution. 31 a) b) Figure 2.17: A 218 -point FFT analysis of a 15-inverter pointer-FDSM output. Input signal frequency 100Hz, fc =10MHz, fs =2MHz and ∆f =500KHz. a) Ideal modulator. b) Modualtor with random phase offsets Differential operation To reduce the impact from common mode noise, differential architectures are widely used. In the FDSM system, a differential realization seems to require two independent frequency modulators, and the single FDSM output may then be implemented by subtracting the output from the original counter from the reference counter output. The new quantization error will then be the difference between the original quantization error and the reference signal quantization error. Since there are no correlation between the two quantization errors, the new quantization error pdf will have a√triangular shape spanning from -1 to +1, and its mean square value will be a factor of 2 times the original mean square value. For differential operation, the output signal amplitude will be doubled and the result is a net SQNR gain of ≈3dB. On the other hand, if the reference oscillator is not modulated, a loss of ≈3dB will result. 2.6 Feature and performance characterization of the firstorder FDSM In the following sections, the different features of the first-order FDSM system will be compared to the traditional ∆-Σ modulator. Pattern noise As we have seen, the FDSM is mathematically equivalent to a traditional ∆-Σ modulator where the input signal is scaled and biased. Therefore, the pattern noise properties are expected to be the same as in the traditional ∆-Σ modulator. For the basic modulo-2n and the pointer-FDSM, the output signal rms value may however, due to multi-bit quantization, be large compared to the pattern noise. For the D flip-flop FDSM, the output center value will be fc /fs , and by locating this value well away from the main DC pattern noise peaks, the pattern noise problem is expected to be eliminated for small signal ranges. 32 DAC output levels Due to the absence of a DAC, problems due to misplaced DAC output levels will be eliminated, and multi-bit quantization is straightforward. ADC thresholds In the FDSM, the counter value must be considered as a quantized analog value as the counter is asynchronous. By sampling the counter value, both quantization and sampling have been executed, and an equivalent AD conversion have been carried out. In the basic modulo-2n FDSM, the single counter threshold which is used to decide if the FM signal is high or low, may be placed anywhere it will be convenient from a digital consideration. In the D flip-flop and the pointer-FDSM, there will in a practical implementation, be some duty-cycle and phase deviations from the ideal. These deviations may be considered as misplaced ADC thresholds, but due to the differentiation, the baseband noise will be heavily suppressed. As opposed to the traditional ∆-Σ modulator, these errors will be very little correlated with the input signal, and as the simulations indicates, have less practical significance. Such errors should however be minimized as large offsets may slightly degrade the overall SNR. ADC overloading Since the ADC is implemented by a counter and the number of output levels may easily be chosen larger than the output signal range requires by increasing the internal word length, problems due to ADC overloading is eliminated. Integrator errors In a frequency modulator, there will be no leakage, charge injection and slew-rate limitations, and for the FDSM, the only limiting factor for high sampling speed, will be metastability effects in the synhrounizing circuits [21, 22]. The main problem is however the fact that frequency modulating noise in the frequency modulator will add directly to the signal. Dependent on the architecture of the frequency modulator and the application, harmonic distortion may be a problem. By reducing ∆f /fc, the maximum non-linearity of a frequency modulator will normally decrease, and due to the high resolution of the ∆-Σ modulator, a low ∆f /fc may be utilized to reduce harmonic distortion. If the frequency modulator is already implemented in the signal source, the FDSM concept may therefore be used to improve harmonic distortion in a FD system by enabling the use of a lower ∆f /fc ratio. External noise In the FDSM system as in the traditional ∆-Σ modulator, several external noise factors will be present. As we have both the signal and its integral represented in the circuit, we may differ between noise that affects the signal (frequency modulating noise) and noise that affect the integral of the signal (phase modulating noise). In general, phase modulating noise will due to the differentiation be first-order noise shaped, and will normally be of no significance compared to the quantization error. In this category we have sampling clock jitter, power supply noise in the θn detector/quantizer and digital/transmission noise deteriorating the intermediate FM signal. Frequency modulating noise will however directly add to the signal, and must be carefully considered. In this category, we may, depended on the architecture, have power supply noise in the frequency modulator, and low frequency noise caused by 33 the frequency modulator sensitivity to temperature drift. By using a reference frequency modulator, the impact from such factors may be significantly decreased. Another noise source that should not be overlooked, is frequency modulating noise in the sampling clock oscillator. By using an intermediate FM signal, the sampling clock frequency must be considered as a reference signal, and a frequency stable clock oscillator must be used. By using a reference modulator we convey the stability requirement from the clock oscillator over to the reference input signal. The integrator-differentiator combination As in a conventional ∆-Σ modulator, by using an analog loop filter, the frequency response of the continuous-time integrator - discrete-time differentiator combination is not constant. The impulse response of a continuous-time integrator is given by H i (jω) = 1/jω, and in the frequency interval |ω| < 2πfs /2, the Fourier transform of the sampled impulse response sequence is Y (ejωTs ) = fs H i (jω). Together with the accumulator frequency response H a (ejωTs ) = (1 − e−jωTs ), the overall frequency response will be H(ejωTs ) = fs H i (jω)H a (ejωTs ) = fs (1 − e−jωTs ) , jω |ω| ≤ πfs (2.29) and the magnitude will be sin(ω/(2fs )) , |ω| ≤ πfs (2.30) ω Due to the single pole at zero, the maximum in band signal suppression will be seen for the maximum signal frequency, and the low pass filtering effect will be most noticeable for systems with a low OSR. But as an example, for fs = 1M Hz and fmax = 20KHz which corresponds to an OSR as low as 25, the maximum signal amplitude deviation in the passband will be |H(ejωTs )| = 2fs |H(ej2π0·Ts )| − |H(ej2πfmax Ts )| = −145dB (2.31) As we see, the continuous-time integrator - discrete-time differentiator combination causes no practical problems. Implementation The implementation of a FDSM system is very different from the implementation of a traditional ∆-Σ modulator, and the frequency modulator may, dependent on the application, be implemented in several ways. If the input signal is already given as a FM carrier, the necessary extra circuitry will be purely digital. The CMOS pointer-FDSM where the carrier frequency is modulated by the inverter power supply voltage, is an example which illustrates an AD converter build exclusively by digital components, and a standard digital CMOS process may be used. By, in this way, using only a few digital components, a low power supply voltage operation and a reduced power consumption is possible. 34 35 Chapter 3 Extended FDSM techniques In this Chapter four novel FDSM techniques will be presented and analyzed. We will start by looking at the undersampled FDSM. 3.1 FM signal undersampling In some frequency demodulation applications it may be desirable to move the frequency demodulator and the ADC as close as possible to the antenna. In a FDSM based system this means that the incoming frequency to the FDSM may be very high. If the digital resolution requirement is low, the chosen sampling clock frequency may be much lower than the FM carrier frequency. For systems where the sampling frequency is lower than 2(fc + ∆f ), a multi-bit FDSM should be used according to the discussion in Section 2.4. In a multi-bit FDSM running at very high frequencies, the maximum operation frequency will be limited by synchronizer metastability effects [21, 22]. If a very high frequency operation is required, the synchronizing elements of the FDSM should therefore be kept as simple as possible, and the D flip-flop FDSM will normally be the best candidate. From Eq. 2.18 and Eq. 2.22 we also notice that the D flip-flop FDSM provides ≈6dB higher SQNR than the standard counter-based FDSM since both FM edges are counted. On the other hand, by using a D flip-flop FDSM and increasing the sampling frequency to 2(fc + ∆f ), this frequency is now more than twice the FM frequency and if we try to push the limit of the processing technology, the sampling frequency may now be the limiting factor. In this way, the power consumption, particularly in the decimator, will also be much higher than necessary. In the following, we will propose a solution to this problem by proving that the simple D flip-flop FDSM may also be used in applications where the sampling frequency is lower than 2(fc + ∆f ) by using the alias of the FM signal. To illustrate this idea we will start by looking at the undersampling FDSM from a time-domain view, and continue by analyzing the converter in the frequency-domain. 3.1.1 Time-domain analysis The D flip-flop FDSM is based on both modulo-2 edge counting and modulo-2 differentiation. The modulo-2 edge counter output is represented by the “logical” value of the limited FM signal, and the modulo-2 differentiator is implemented by a XOR gate as illustrated in Fig. 3.1. 36 fclk fm fclk clk clk DQ DQ out Figure 3.1: The standard D flip-flop FDSM This modulo-2 wrap-around error cancellation scheme works well as long as the counted number of FM edges during the sampling interval is 0 or 1, which is true for sampling clock frequencies higher than 2(fc + ∆f ). However, by using a multi-bit FDSM running at a sampling frequency lower than 2(fc + ∆f ), the output will, for low FM deviations, be m or m + 1 where m is an integer given by the sampling clock frequency to the FM carrier ratio. In other words, for low FM deviations, the number of counted edges during 1/fclk will be limited to two adjacent integers. But in this case, if we do not need the bias component m, modulo-2 arithmetic will accommodate the FDSM output signal and we may use a D flip-flop FDSM. From this intuitive discussion we realize that for certain fclk /fc and ∆f combinations we may use a D flip-flop FDSM even if the sampling frequency is less than 2(fc + ∆f ). To put it in another way, what we should do is to make sure that the sampling clock frequency is chosen so that we omit certain (fc ± ∆f )/fclk ratios. To find these ratios we notice that the highest possible variation in FM edge counts will occur when the previous sampling edge is close to an FM edge (Fig. 3.2). This is because a small phase variation on the sampling signal will either include or exclude the first FM edge. fm(t) clk(t) Figure 3.2: FM signal relative to sampling clock. Shaded areas indicate ambiguous regions. In Fig. 3.2 the shaded regions illustrate the maximum phase deviation of the FM signal. So, depending on the modulating signal, the FM zero-crossings will be restricted to these intervals. To restrict the FDSM outcome to m or m + 1 we therefore must ensure that the next sampling edge do not occur in one of these shaded regions. Otherwise the number of FM edges during 1/ff clk may be m, m + 1 or m + 2 as the previous sampling edge may be located both on the left and on the right side of the corresponding FM edge. Since we have lined the previous sampling edge up to a FM edge, this frequency constraint can be formulated in the following way; the number of FM edges per second must not be divisible by the sampling frequency or 2(fc + δf ) −∆f < δf < ∆f = m, (3.1) m∈N fclk This statement may be formulated as fclk ∈ [2(fc − ∆f )/m, 2(fc + ∆f )/m], m ∈ N . 37 (3.2) For a given m, the allowed sampling frequencies will then be all frequencies in between the shaded intervals in Fig. 3.2 which from Eq. 3.2 is fclk ∈ 2(fc + ∆f )/(m + 1), 2(fc − ∆f )/m, m ∈ N . (3.3) By reducing the sampling frequency for a fixed fc , m increases and the intervals in between the shaded regions decreases. For a certain m, the lower limit and the higher limit in Eq. 3.3 will be switched, and there will no longer be room to locate the sampling edge. To find this maximum m or ‘squeezing’ limit the following statement must be true 2(fc + ∆f )/(m + 1) < 2(fc − ∆f )/m, m ∈ N (3.4) and we have the following limitation on m m< fc −∆f 2∆f , m ∈ N. (3.5) For a D flip-flop FDSM, the minimum sampling frequency will then from Eq. 3.3 be 2(fc + ∆f ) . (3.6) mmax + 1 By approximating mmax by (fc − ∆f )/2∆f , and assuming fc ∆f , the minimum sampling clock frequency may be approximated by min fclk = min fclk ≈ 4∆f (3.7) It should be emphasized that this is a theoretical limit. In a practical implementation, one should also make room for phase noise. In Fig. 3.3 the D flip-flop clock usable range is illustrated for two different ∆f . Upper graph - fc =400MHz, ∆f =10MHz. Lower graph - fc =400MHz, ∆f =5MHz. The inclining line segments illustrate the usable clock range given by Eq. 3.3. Above 2(fc + ∆f ) the range is comprehensive as the converter operates in FM oversampling mode. For the case where ∆f =10MHz the “squeezing” limit frequency, illustrated by point g, is twice as high as for ∆f =5MHz. 3.1.2 Frequency-domain analysis The undersampling operation may also be analyzed in the frequency domain by looking at the frequency content of the FM signal. The most intuitive point of view will be to look at a very low frequency modulated carrier and model it as a single tone varying in frequency with the amplitude of the modulating signal (Fig. 3.4). By this approach the maximum and minimum frequency will be given by fc + ∆f and fc − ∆f respectively. Since the FM signal is immediately digitized by the first D flip-flop, we may equivalently model this operation by assuming a square wave FM signal and the D flip-flop to be an analog-to-digital converter with infinite amplitude resolution. In this case the continuous-time, square wave FM signal spectrum will consist of odd harmonic components as illustated in Fig. 3.4. The maximum deviation from the center frequency in each harmonic group will be proportional to the harmonic center frequency as illustrated in the figure. For modulating signals at medium or high frequencies we should rather look at the FM spectrum as a carrier frequency with two symmetrical sidebands given by the modulating signal amplitude and frequency. For sinusoidal modulation, the modulating index [17] is defined as 38 130 125 a 120 cb d f SQNR (dB) 115 e 110 g 105 100 90 4∆f1 4∆f2 fc 100MHz Sampling frequency fclk (Hz) 2fc+∆f1 2fc+∆f2 95 1GHz Figure 3.3: Usable sampling clock ranges for the D flip-flop FDSM. Upper graph - fc =400MHz, ∆f =10MHz. Lower graph - fc =400MHz, ∆f =5MHz. a,b,c) - FM oversampling mode, d,e,f ) undersampling mode, g) “squeezing” limit β= ∆f fm (3.8) where fm is the frequency of the modulating sinusoidal. For β > 1 the FM signal is described as wide-band and the required bandwidth is 2∆f according to Carlsons rule [17]. For β < 1 the FM signal is described as narrow-band and the required bandwidth is 2fm . For nonsinusoidal modulation we normally refer to the deviation ratio given by D= ∆f fmax (3.9) where fmax is the highest frequency component in the modulating signal. By using the deviation ratio we implicitly assume that the FM signal may be wide-band modulated as the modulating signal may contain all frequencies from 0 − fmax . In Fig. 3.5 top, the spectrum of a wide-band modulated square-wave FM signal is illustrated. To simplify the picture, both sidebands are symbolized as a square box also containing the carrier frequency. For higher harmonics the bandwidth increases proportional to the harmonics number, while the power density is proportional to −20 log(k) where k is the harmonic number. By sampling the square-wave FM signal, all harmonics above half the sampling frequency will fold down to the sampling interval as illustrated in Fig. 3.5 bottom. If the sampling frequency is lowered, the position of these harmonic bands will change, and 39 PSD ∆f 3∆f fc 0 5∆f 3fc 7∆f 5fc 9∆f 7fc 9fc f Figure 3.4: The continuos-time square-wave FM signal spectrum represented as a single tone varying in frequency with the amplitude of the modulating signal 0 PSD (dB) -10 -20 fc 3fc 5fc 7fc 9fc 0 -10 -20 fc fclk-fc fclk Figure 3.5: Top - the continuous-time square-wave FM signal spectrum represented by a carrier component with two symmetrical sidebands (all in one box) and its higher-order harmonic components. Bottom - the sampled FM signal at certain fclk /fc ratios the high-power lower harmonics bands, will interfere with the original FM band. By looking closely at the situation for different fclk /fc ratios, it seems like this effect corresponds very closely to the effect of pattern noise. In this way pattern noise is described as interference between the original FM band and its higher harmonic bands. It should be mentioned that this relationship have not been proven analytically, but threre are indications that this can be an alternative model for pattern noise. In Fig. 3.6 the sampled FM signal power spectrum is illustrated for different sampling frequencies corresponding to point a-g in Fig. 3.3. In the following, each point will be commented. • Point a: The sampling frequency is 1.3GHz corresponding to point a in Fig. 3.3. In this case the FM signal is well oversampled and we notice some low power harmonic interference. • Point b: The clock frequency is lowered to 860MHz. The FM signal is still oversampled, and the upper mirror signal band is approaching the original main signal band at fc − ∆f, fc + ∆f . There is no noticeable harmonic interference although the high-power second harmonic signal bands are quite close. • Point c: The sampling frequency is lowered to 2(fc + ∆f )=820MH which is the minimum FM oversampling frequency. In Fig. 3.6 c we notice that the main signal band is touching its mirror band at fc + ∆f, fc + 3∆f and there is heavy interference with all harmonic signal bands. 40 • Point d: The FM signal is undersampled at 668MHz and the main signal band has exchanged position with its mirror band. In this case, the FDSM will demodulate the FM mirror band. By further lowering the clock frequency the FM mirror band will move towards zero frequency while the original FM band will be closer to the sampling frequency. • Point e: The clock frequency is 410MHz corresponding to the lower end of the line segment if Fig. 3.3. From the time-domain analysis given by Eq. 3.3 this is the lowest clock frequency where the FDSM output is restricted to m=1 and m=1+1. From Fig. 3.6e we notice that this is the point were the original FM signal band at fc − ∆f, fc + ∆f starts to interfere with the mirror band at fclk + (fclk − [fc + ∆f ]), fclk + (fclk − (fc − ∆f )). There is heavy harmonic interference. • Point f: The clock frequency is 362MHz and the FDSM output is restricted to m=2 and m=2+1. From the frequency spectrum we see that in the clock frequency band, there are two mirrors of the original FM signal band, and the FDSM will now demodulate these two images instead of the original FM band at fc − ∆f, fc + ∆f . We also notice that the two FM mirror bands and the harmonic signal bands occupy a larger fraction ot the sampling clock band. In Fig. 3.7 the sampling clock frequency is lowered to 41MHz which is close to the squeezing limit shown as point g in Fig. 3.3. From the frequency spectrum we see that the squeezing limit in the time-domain corresponds in the frequency-domain to the situation where there is no longer room for narrowing the sampling interval as the two FM mirror bands occupies the complete interval. A further decrease in fclk will now make the two signal band interfere with each other causing signal distortion. To decribe the resulting distortion in the demodulated signal caused by FM mirror band interference is a complex affair, as the demodulation process is non-linear. We may however now look at the time-domain behavior by inspecting the demodulated signal. In Fig. 3.8 the demodulated signal is plotted both against time and frequency for fclk =82.3MHz. In this case the FM signal is moderately undersampled. There are no visible distortion as the FDSM output is restricted to m=9 and m=10. The SQNR was estimated to 107dB which is very close to its theoretical value of 108dB (Eq. 2.22). In Fig. 3.9 right, the sampling frequency was set to 400MHz which is in center of the ambiguous interval fc − ∆f, fc + ∆f from Eq. 3.2. The demodulated signal is now maximally distorted, which we see corresponds to a rectifying operation. In Fig. 3.9 left the sampling frequency is slightly biased away from 400MHz but is still in the ambiguous region. In this case, we see the distortion as biased rectifying. Finally, in Fig. 3.10 a 1·106 point FFT analysis of a sampled square-wave FM signal is shown. In this example the carrier frequency was set to 400MHz, the maximum frequency deviation - 10MHz and a 2KHz modulating sinusoidal was used. The sampling frequency was set to 668MH which make the parameters identical to point d in Fig. 3.6. The plot is shown for the 0-fclk /2 range, and we notice that both the FM mirror band and its major harmonics bands are present at the same places and with the same power densities that in Fig. 3.6 d. Comment on FM bandwidth In this frequency analysis we have assumed a wideband modulated FM signal, as the modulating signal generally may contain low frequency components. For narrowband FM signals the FM bandwidth will be given by fc − fm , fc + fm where fm is the lowest 41 frequency component of the modulating signal. In this case the FM signal band and its mirror bands will occupy a smaller fraction of the sampling interval and the squeezing limit will be somewhat lower. This will also slightly increase the usable frequency intervals of Eq. 3.3. 42 0 a) -10 -20 fclk 2fclk 3fclk 0 b) -10 -20 fclk 2fclk 3fclk 4fclk 0 PSD (dB) c) -10 -20 fclk 2fclk 3fclk 4fclk 5fclk 0 d) -10 -20 fclk 2fclk 3fclk 4fclk 5fclk 0 e) -10 -20 fclk 2fclk 3fclk 4fclk 5fclk 6fclk 7fclk 8fclk 9fclk 0 f) -10 -20 fclk 2fclk 3fclk 4fclk 5fclk 6fclk 7fclk 8fclk 9fclk 10fclk 11fclk PSD (dB) Figure 3.6: The sampled FM signal spectrum for different sampling frequenies 0 -10 -20 fclk 2fclk Figure 3.7: The sampled FM signal spectrum for fclk =41MHz illustrating the squeezing limit 43 500 450 Amplitude 400 350 300 250 200 150 100 50 2.6 2.65 2.7 2.75 Sample no. 2.8 2.85 4 x 10 Figure 3.8: Left - demodulated signal, fclk =82.3MHz. Right - estimated spectrum 350 250 300 200 Amplitude Amplitude 250 200 150 150 100 100 50 50 0 3.34 3.36 3.38 3.4 3.42 3.44 3.46 3.48 3.5 3.52 3.54 4 Sample no. x 10 0 2.54 2.56 2.58 2.6 2.62 2.64 Sample no. 2.66 2.68 4 x 10 Figure 3.9: Right - demodulated signal, fclk =400MHz. Left - fclk slightly increased Figure 3.10: A 1·106 point FFT analysis of a sampled square-wave FM signal. fc =400MHz, ∆f =10MHz, fm =2KHz and fclk =668MHz 44 3.2 The sampled-clock FDSM Although the first-order FDSM provides first-order ∆-Σ noise-shaping, the SQNR is only increased by ≈3dB for each doubling of the sampling frequency (Eq. 2.18). This is because, a doubling of the sampling frequency decreases the noise power by ≈9dB, but in addition, the FDSM signal range is decreased by ≈6dB resulting in a net SQNR gain of ≈3dB. In a FDSM, a more effective way to increase the resolution is to increase the FM deviation. By doubling the deviation, the SQNR will increase by ≈6dB. If the FM signal source has a constant ∆f /fc ratio, as it may be in a modulated ring-oscillator, a doubling of fc will now double the SQNR. However, in many applications both the FM carrier frequency and its deviation is already given and the only thing to play with is the sampling clock frequency. In Fig. 3.11 a typical FDSM frequency-to-digital system is shown. The front-end is the FDSM followed by a decimator. In the other end of the system, we have the user of the demodulated digital signal. In many cases, the user is a non-human digital system running at the same system clock as the FDSM. In other cases the user is a human listening to or looking at the data provided via an ADC or visualized on the computer screen. From now on we will denote this kind of system as a “sampled signal” system as we are sampling the FM signal with the system clock. mod-2n counter fm reg fclk mod-2n diff cnt decimator user reg FDSM Figure 3.11: A standard modulo-2n “sampled signal” FDSM system In a FDSM system as illustrated by Fig. 3.11, the information given to the user is high quality information about the ratio of the instantaneous FM frequency to the system clock frequency, as this ratio is proportional to the modulating signal. If the FDSM module is chosen larger than the maximum number of positive edge counts during 1/fclk , the FDSM output mean value will, from Eq. 2.16 be ȳn = f¯FM fc + kx̄n = fclk fclk (3.10) assuming low-frequency modulation. Here, xn is the modulating signal. From the previous discussion we saw that by doubling the sampling clock frequency, the SQNR increase was only ≈3dB, and by doubling the signal frequency fc and ∆f on the counter input, the SQNR increase was ≈6dB. If the target is high SQNR, this is good for applications where we are free to choose a high FM deviation. However, if the FM deviation is already given, but we are able to use a very high sampling frequency we do not achieve to much. 45 As the system output represents the ratio of the instantaneous FM frequency to the sampling frequency we may ask if it could be possible to exchange the FM and clock input to the system and still keep information about the modulating signal. By doing so, we may utilize the higher gain in SQNR resulting from the high frequency clock signal on the counter input. By exchanging the two signals the FDSM output will, now with high quality, represent the ratio of the clock frequency to the FM frequency, provided that the FDSM still shapes its quantization error. Since the modulating signal now is in the denominator, the FDSM transfer function will be non-linear, and the output mean value will be fclk fclk ȳn,sc = ¯ = f fFM c + kx̄n (3.11) Max relative non-linearity error (dB) By fitting this function to an ideal linear characteristic, the maximum relative non-linearity is plotted against ∆f /fc in Fig. 3.12. -50 -60 -70 -80 -90 -100 -110 0.001 0.1 1 10 ∆f / fc (%) Figure 3.12: Maximum non-linearity error relative to signal range versus ∆f /fc As we see from the figure, for low ∆f /fc ratios the non-linearity distortion is quite low. For example if the FM carrier frequency is 40MHz and the FM deviation is 100KHz, the relative non-linearity will be ≈-103dB. On the other hand, for high ∆f /fc ratios the nonlinear error will be deterministic and given by Eq. 3.11. In contrast to stochastic non-linear errors given by process variations, temperature drift an so on, this kind of error is relativerly easy to compensate for by following the decimator by a non-linear digital filter employing the inverse function. Another factor that must be taken into consideration it that by applying the FM signal to the system clock input, the entire system will operate on a time-varying or non-uniform timebase as shown in Fig. 3.13. To distinguish this kind of system from the sampled-signal system we will refer to it as a sampled-clock system, as we are sampling the constant clock frequency with the FM signal. Depending on the user this may be acceptable or not. If the user is a large digital system where the FDC is only a minor component this could be a problem. However, if the FDC is a main part of the system this can, in many applications, be acceptable. Finally, to be able to use the sampled-clock system we need to show that this system also shapes the quantization error according to ∆-Σ theory. It is possible to prove that the system employs ∆-Σ noise-shaping by modeling the system on a non-uniform time-base. 46 mod-2n counter fclk reg fm mod-2n diff cnt decimator user reg FDSM Figure 3.13: A modulo-2n “sampled clock” FDSM system. However, this is not an easy way to go. In the following section, we will therefore show the ∆-Σ equivalence simply by introducing the concept of virtual frequencies. 3.2.1 Virtual frequencies What is time? This not a simple question to answer. Most people take it for granted that time is linear and timesteps are uniformly distributed. On the other hand, a distribution must always be in reference to something else. By looking at the modulated FM carrier we say that its zero-crossings are modulated or non-uniform with reference to the clock signal. In the following, we will change the point of view and say that the clock signal’s zerocrossings are modulated with reference to the FM signal which we consider to be constant. This change of time-base is possible because we are dealing with a digital or sampled-time system where nothing is happening in between the sampling time-steps. From the sampledclock FDSM system this new time-base makes it easy to model the behaviour of the system. With this new time-base, the FM frequency is now virtually constant and the clock signal on the FDSM input is virtually varying with time. However, to be able to utilize this simplified description we need a model of the virtual frequency on the FDSM input. As a frequency modulated signal, this virtual frequency will be on the form fclk + k xn , where k is a constant and xn is the virtually modulating signal. To proceed with this analysis we will make use of the following invariant: The ratio between the real clock frequency and the real FM signal frequency must equal the ratio between the virtual clock signal frequency and the virtual constant FM frequency. This invariant is fundamental for the idea of a change in time-base. Whitout this invariant, there will be no use in a timebase shift as the system analysis would not be simplified. So, what we have is fclk fclk + k xn = . fc + kxn fc (3.12) The virtual clock signal frequency will therefore be fclk + k xn = fclk fc . fc + kxn (3.13) The sampled clock FDSM will now count each positive edge in the virtual clock signal and thereby extract a quantized representation of the virtual clock signal phase θn /2π producing 47 a quantization error φ ∈[0, 1. The difference between this system and a sampled-signal FDSM can now be modeled as a non-linearity in the modulating signal xn given by Eq. 3.13, thus the ∆-Σ noise-shaping equivalence is verified. The new output signal range SRo will now be given by the difference between the maximum and minimum number of counts during the sampling interval which is SRo = fclk fclk 2fclk ∆f − = 2 , fc − ∆f fc + ∆f fc − ∆f 2 (3.14) In most cases ∆f will be reasonably lower than fc making ∆f 2 fc2 . In this case the output signal range may be written as SRo ≈ 2fclk ∆f /fc2 (3.15) n The resolution of the sampled clock modulo-2 FDSM will now be 3/2 SRo π fmax √ SQN R ≈ 20 log − 20 log 2 6 fc 2 2 (3.16) By doubling the constant clock frequency fclk we notice that the SQN R is now increased by ≈6dB. 3.2.2 The SC D flip-flop FDSM For applications aiming at high SQNR where there is available a high frequency system clock, the sampled-clock FDSM may be a better choice than the sampled-signal converter. As in the sampled-signal converter, the maximum operation frequency will be limited by metastability in the parts of the converter that are asynchronous to the system clock. For this reason the FDSM implementation should be kept as simple as possible, and the D flipflop solution should be used if possible. In addition, the resolution of the D flip-flop FDSM is ≈6dB higher than for the modulo-2n FDSM as it counts both positive and negative edges in the input signal. By using the sampled-clock D flip-flop FDSM, a very high clock frequency may be used providing a high SQNR. As the sampling frequency is given by the FM frequency and not by the high-speed clock frequency, the FDSM output data rate may be kept at manageable level, keeping the decimator power consumption down. One must keep in mind that for applications where the minimum FM frequency fc − ∆f is lower than two times the clock frequency fclk , the sampled-clock converter operates in undersampling mode. fm fm fclk clk clk DQ DQ out Figure 3.14: The sampled-clock D flip-flop FDSM As for the sampled-signal D flip-flop FDSM, if the converter is used in undersampling mode, there are certain fclk /fc ratios which must be avoided to keep the modulo-2 wraparound error cancellation scheme work properly. To find these frequencies we will start by analyzing the converter in the time-domain. 48 Time-domain analysis As in the sampled-signal case, to ensure proper operation, the FDSM output must be restricted to m and m + 1 where m is an integer given by the number of received signal edges during the sampling interval. In the sampled-clock FDSM we are now dealing with the number of received clock edges during the FM signal period. clk(t) fm(t) Figure 3.15: FM signal relative to sampled clock signal In Fig. 3.15 the clock signal is plotted against the sampling FM signal and the FM deviation is shown by a shaded region at the positive sampling edge. If the number of clock edges during the FM period is to be limited to m and m + 1, there must be no clock edge in the shaded region as the previous sampling edge may be at either side of the corresponding clock edge. In other words, the double clock frequency must not be divisible by the FM frequency or 2fclk −∆f < δf < ∆f = m, (3.17) m∈N fc ± δf Equivalently fclk ∈ [ m m (fc − ∆f ), (fc + ∆f )], 2 2 m ∈ N, (3.18) or m m+1 (fc + ∆f ), (fc − ∆f ), m ∈ N. (3.19) 2 2 As for the sampled-signal D flip-flop FDSM, we need to ensure that the left limit in Eq. 3.19 is lower than the right limit, and this will restrict m to fclk ∈ m< fc − ∆f , 2∆f m ∈ N, (3.20) The maximum constant clock frequency will now, from Eq. 3.19 be mmax + 1 (fc − ∆f ). 2 For applications where ∆f fc , the maximum constant clock frequency will be max fclk = max fclk ≈ fc2 4∆f (3.21) (3.22) As we see, in the sampled-clock D flip-flop FDSM the, maximum clock frequency is proportional to the square of the FM carrier frequency and inversely proportional to the FM deviation. In Fig. 3.16 this result is visualized by a plot of the SQNR against constant clock frequency. The upper graph illustrates the usable clock range for an application with fc =400MHz, fmax =5KHz and a maximum frequency deviation of 10MHz. In the bottom graph, the deviation is reduced to 5MHz. We notice that although the resolution is reduced 49 with ≈6dB, the maximum resolution is the same as for the previous case as the squeezing limit is increased. However, it will be difficult to utilize the new squeezing limit as the usable clock intervals are now quite narrow, even if the plot is compressed for high frequencies. A slight deviation in the fclk /fc ratio may now push the converter out of the usable range. To confirm the result of this analysis and gain more insight in the sampling process we will now continue with a frequency-domain analysis of the converter. 140 e 130 d SQNR (dB) 120 110 b c a 100 Hz M 10 = ∆f Hz M =5 f ∆ 90 80 10 Figure 3.16: 100 1000 Constant clock frequency (MHz) 10 000 SQNR for a sampled-clock D flip-flop FDSM application with fc =400MHz, fmax =5KHz Frequency-domain analysis To find the bandwidth of the virtual FM signal we assume that the signal is wideband modulated. By using Carlsons rule, we need to find the maximum deviation of the signal. The virtual frequency devation on the clock signal will be the maximum deviation minus the minimum deviation divided by two, which from Eq. 3.13 is fclk fc fclk fc fclk fc ∆f ∆f = − /2 = 2 (3.23) fc − ∆f fc + ∆f fc − ∆f 2 For most applications the carrier frequency is somewhat higher than the deviation, giving ∆f ≈ fclk ∆f fc (3.24) So, the bandwidth of the virtual FM signal will be fclk −∆f , fclk +∆f and the bandwidth will increase with the ffclk ratio. The corresponding square-wave signal will have harmonic c bands at odd harmonic frequencies as illustrated in Fig. 3.17, top. By sampling the virtual 50 square-wave FM signal with the virtual constant frequency fc , the converter will operate in undersampling mode if fc − ∆f < 2fclk . In Fig. 3.17 a-e, the virtual FM spectrum is illustrated for different clock frequencies, and in the following, each case will be commented. • Point a In this case fclk =75MHz and the converter operates in FM oversampling mode. All harmonic signal bands are folded down to 0, fc , but there is no interference between the main signal band and the major harmonic bands. • Point b In this case fclk =195MHz, and the converter operates just on the limit of FM oversampling mode. We notice that the main signal band is touching its mirror band at fc − (fclk + ∆f , fc − (fclk − ∆f ). There is heavy harmonic interference. • Point c In this case fclk =205MHz, and the converter operates in undersampling mode. From Eq. 3.19 we see that this clock frequency is in the first usable interval given by m=1, also corresponding to point b in Fig. 3.16. From Fig. 3.17 b, we notice that the main signal band has exchanged position with the mirror band and the converter will now demodulate the FM mirror signal. There is heavy harmonic interference. • Point d In this case fclk =850MHz, and the converter operates in undersampling mode. From Eq. 3.19 the clock frequency is in the fourth usable interval given by m=4. • Point e In this case fclk is set to 1.3GHz illustrating the sqeezing limit in the frequency domain. This limit correspond to the case where the two signal bands occupies the whole sampling interval and there are no room for a further increase in fclk as the signal bandwidth is proportional to fclk . Simulations When a ∆-Σ modulator is to be simulated care must be taken to control the effect of numerical errors. Depending of the location of the error it will be noise-shaped or not. The sampled-signal FDSM was simulated [7], Appendix B, in a loop where each time-step n · 1/fclk , corresponds to the sampling clock edge. The FM phase θn is then calculated from Eq. 2.7, which for sinusoidal modulation is given by θn = 2πfc n/fclk + ∆f sin(2πfm n/fclk ) fm (3.25) In this way there is no need to numerically estimate the time for the sampling clock edge, and all errors will be present as round-off θn errors. As round-off θn errors will not accumulate they will represent phase-noise, which we will from Section 4.1, see is first-order noise-shaped. By using long-double constants in C, there have not been observed any simulation quality degradation or break-down even for simulations running through hundreds of millions of samples. However, to simulate the sampled-clock FDSM, a different approach must be taken. If each time-step should correspond to each FM sampling edge, the simulator needs to know the time for each positive FM edge. Unfortunately, there exists no analytical expression for Eq. 2.7 resolved with respect to t, so each time-step would have to be numerically approximated. To simulate a FDSM where the oversampling ratio can be of more than 100 000 times, there is a need for a high number of samples to achieve sufficient frequency resolution in the signal-band. For a simulation including more than a million samples, the estimation of the time-step will be time consuming, and the number of numerical errors will increase. To overcome this problem another approach is taken. 51 0 -10 -20 fclk 3fclk 5fclk 7fclk 0 a) -10 -20 fclk 3fclk fc 2fc 3fc fc 2fc 3fc fc 2fc 3fc fc 2fc fclk 3fc fc 2fc 3fc Normalized PDS (dB) 0 b) -10 -20 fclk 0 c) -10 -20 fclk 0 d) -10 -20 0 e) -10 -20 Figure 3.17: The virtual FM spectrum for different clock frequencies By looking at Fig. 3.18 and Fig. 3.14 we notice that for each time there is a positive FM edge the clock signal must be sampled, and this sampled sequence must then be modulo-2 differentiated. Even if the FM edge is in continuous time, there is actually no need to know the exact position of the edge if we know which halve fclk period it occurs in. What we need to know is which fclk value the edge is sampling, one, or zero. As long as we know the correct sampling value, the exact position of the sampling edge does not matter in a non real-time simulation. The D flip-flop sampled-clock FDSM is therefore simulated in a loop where each time-step is given by n · 1/(2fclk ). The time-step is then inserted in the expression for the FM angle, and if the angle have past a 2π period, the value of the clock signal is fed to the decimator. If the FM angle have not past a new 2π period, nothing is done. In this way the time-step do not have to be estimated, and the θn round-off error will, even now be present as phase-noise. In Fig. 3.19 the D flip-flop sampled-clock FDSM is simulated for a modulating sinusoid 52 n=k+10 n=k+9 n=k+1 n=k T0=1/(2fclk) clk(t) fm(t) Figure 3.18: FM signal relative to the sampled clock signal of 233Hz. The FM carrier frequency was 712kHz with a frequency deviation ∆f , of 7100Hz. By applying a clock frequency of 12636220Hz, the FDSM output will be restricted to m=35 and m=36. A number of 32 millions clock samples was simulated decimated by 2, and the result is seen in the figure. The inherent non-linearity of this converter is clearly seen by the 2’th and 3’th harmonics components. To verify the resolution, we may set the signal band to 0-400Hz and calculate the SQNR in this region. From Eq. 3.16, the ideal resolution should then be 13.6bit. The calculated SQNR from the simulation matches almost exact with 13.8bit. A higher resolution than the theoretical model predicts, is possible due to pattern noise effects. From the figure we also notice some high frequency pattern noise. In Appendix B the sampled-clock, D flip-flop FDSM simulator is listed. 0 −20 Normalized PSD (dB) −40 −60 −80 −100 −120 −140 1 10 2 10 3 10 Frequency (Hz) 4 10 5 10 Figure 3.19: Simulated output psd of a D flip-flop sampled-clock FDSM with a single sinusoidally modulated input signal at 233Hz. fclk =12636220Hz, fc =710kHz, ∆f =7100Hz 53 3.2.3 Resolution comparison for SS vs SC In general, the first-order FDSM can be implemented in two ways; as a sampled-signal converter, and as a sampled-clock converter. The sampled-signal converter is inherently linear, and provides a constant output data-rate. On the other side, the sampled-clock converter employs a 1/(k + x) transfer function and provides a time-varying output datarate, where the maximum variation in time is proportional to ∆f /fc. In both converters the quantization error is first-order noise-shaped providing a decrease in in-band noise power of ≈9dB for each doubling of the sampling frequency. In the sampled-signal converter the dynamic range is however inversely proportional to the sampling frequency fclk and therefore, the overall increase in SQNR is only ≈3dB for a doubling of fclk . This is not the case for the signal bandwidth fmax , and by reducing fmax by a factor of two, the net SQNR will now increase by ≈9dB. In the sampled-clock converter the dynamic range is inversely proportional to the square of the sampling frequency fc . This makes the sampled clock converter behave in a somewhat strange manner because now the net SQNR will increase by ≈3dB for each halving of the sampling frequency. In other words, if the FM signal can be mixed down to a lower frequency before it is fed to the FDSM, the SQNR will be increased. A main advantage of the sampledclock converter is that the net SQNR is proportional to the constant clock frequency fclk . For both converters, the net SQNR is proportional to the deviation of the sampling frequency. By re-arranging Eq. 2.18 and Eq. 3.16, the resolution of both converters can be compared SQNRss SQNRsc ∆f = k + 20 log + 10 log(fclk ) f 1.5 max fclk ∆f ) = k + 20 log + 10 log(fclk ) + 10 log( 1.5 fmax fc (3.26) (3.27) SQNRss is the resolution of the sampled-signal FDSM, and SQNRsc the resolution of the sampled-clock FDSM. The constant k is 20 log(3/(2π)). From the last equations we notice that the only difference between the two, are the term given by fclk /fc . We therefore have that, if fc > fclk , the sampled-signal FDSM will provide a higher resolution than the sampled-clock FDSM. If fclk > fc , the sampled-clock FDSM will provide the highest resolution. In other words, to achieve the highest SQNR, we should use the lowest frequency as the sampling frequency. 1.5 Another interesting observation is that the resolution is proportional to the term ∆f /fclk . This term is quite similar to the deviation ratio introduced in Section 3.1.2. The difference is that here, the maximum in-band frequency fclk is raised to the power of 1.5. Since the deviation ratio is a measure of the maximum phase deviation of the FM signal for the highest in-band modulation frequency fmax , we notice that the resolution of the FDSM is strongly dependent on the phase deviation of the signal. For sinusoidal modulation we should replace fclk with the frequency fm of the sinusoid. Now the resolution is proportional to the term 1.5 ∆f /fm which is a modification of the FM modulation index. It can also be shown that L+0.5 the resolution of a L’th order FDSM is proportional to the factor ∆f /fm . From this, it may be tempting to define an order of the modulation index for the FDSM application. L+0.5 The L’th order modulation index will then be defined by ∆f /fm . Both the sampled-signal and the sampled-clock FDSM may be implemented as a singlebit D flop-flop FDSM. The D flop-flop FDSM counts both edges in the incoming signal and will therefore provide ≈6dB higher resolution than a single-edge converter. If the D flop-flop FDSM operates in FM undersampling mode we must avoid certain fc /fclk ratios. 54 In Fig. 3.20 the resolution of both the sampled-signal and the sampled-clock modulo2n FDSM is plotted against the constant clock frequency for an example with fc =80MHz, ∆f =2MHz and a signal bandwidth of 0-5KHz. 120 110 SQNR (dB) 100 90 SS D SS m 80 lop flip-f lo-n odu 70 p -flo 60 SC D SC flip n lo- du mo 50 1 10 100 Constant clock frequency (MHz) 1000 Figure 3.20: Theoretical resolution of sampled-signal and sampled-clock FDSM plotted against constant clock frequency for an example with fc =80MHz, ∆f =2MHz and a signal bandwidth of 05KHz. The resolution of the D flip-flop FDSM’s are ≈6dB higher than the corresponding modulo-2n FDSM’s as both signal edges are counted The resolution of the corresponding D flop-flop converters are also shown. As we see, the resolution of the sampled-clock FDSM increases with ≈6dB for each doubling of fclk while for the sampled-signal FDSM the increase is only ≈3dB. Above the point fclk = fc the resolution of the sampled-clock FDSM is higher than for the sampled-signal converter. For these ∆f and fc values it may however be difficult to use the D flop-flop converters close to their squeezing limits as the usable intervals are quite narrow. 3.3 Synchrounization For many applications, the use of a sampled-clock FDSM will introduce a problem due to the time-varying data rate. There are however techniques to resample data streams with non-integer rate factors [19, 30], and in this way synchronize the output data sequence with the system clock. Resampling can be carried out both on the Nyquist rate signal and also on the high speed bit-stream. Resampling of the sampled-clock FDSM have not been fully investigated yet, but there is reason to believe that the most simple approach is to synchronize the bit-stream with a slightly higher frequency signal derived from the main system clock. Generally the bit-stream can not be resampled by the clock signal due to alias effects, but for some frequency combinations this can be done without additional filtering. 55 0 0 -20 -20 -40 -40 Normalized PSD (dB) Normalized PSD (dB) In Fig. 3.21 right, a sampled-clock FDSM simulation was carried out where the output bitstream was sampled by the same clock frequency that was fed to the D flip-flop input. By doing so, a very interesting result was observed. By sampling the time varying bit-stream with the system clock it will be introduced harmonic distortion to the demodulated signal. This distortion was found to be the inverse of the inherent 1/(k − x) transfer function of the sampled-clock system. As we see from the psd spectrum compared to the standard sampledclock psd in Fig. 3.21 left, the harmonic distortion is significantly reduced as the two effects cancels each other. (Care was taken to make sure that the synchronized circuit did not equal a sampled-signal system). From the simulation, the S/(N+D) was found to be 60dB in the synchronized circuit and 50.2dB in the original circuit. Compared to the theoretical value of 58dB, the synchronized circuit worked well. For this successful experiment the following values was used: fclk =10.15MHz, fc =530kHz, ∆f =917Hz, fm =120Hz. However, for other carrier frequencies the performance degraded significantly, and it was never found possible to achieve proper operation for high FM deviations. The simulator is listed in Appendix B. As already stated only preliminary work have been carried out on synchronizing the sampled-clock FDSM, but these ad-hoc experiments indicate that it may be possible both to cancel the inherent non-linearity, and synchronize the sampled-clock FDSM in the same operation. -60 -80 -100 -60 -80 -100 -120 -120 100 1000 100 Frequency (Hz) 1000 Frequency (Hz) Figure 3.21: Simulated psd. Left: sampled-clock FDSM. Right: Synchronized sampled-clock FDSM. fclk =10.15MHz, fc =530kHz, ∆f =917Hz, fm =120Hz. Synchronizer frequency = fclk 56 3.4 Triangularly weighted zero-cross detection The FDSM concept combines traditional ∆-Σ theory with FM theory to obtain high resolution frequency-to-digital converters. As a new concept combining two different fields, the understanding of the FDSM operation can be viewed in different ways. In this section we will look at the oversampled D flip-flop FDSM from a traditional ZC counting frequency discriminator point of view. By doing so, we will find a simple relationship between these two converters modeled by a difference in an inherent data window. The traditional ZC frequency discriminator The traditional ZC frequency discriminator operates by counting the number of zerocrossings in the FM signal during a fixed time window T0 . The number of zero-crossings is then dumped to the output as a digital word while the counter is reset. Usually the frequency of the time or data window 1/T0 is chosen slightly higher than the Nyquist frequency of the modulating signal. The digital resolution of these kind of converters are normally very low and are heavily dependent on the modulating signal bandwidth. As an example, consider the following case; fc = 425KHz, ∆f = 550Hz, fmax =500Hz. The output signal range (digital resolution) for fm close to 0Hz and 1/T0 set to four times the Nyquist frequency will now be outmax − outmin = 2(425KHz + 550Hz) 2(425KHz − 550Hz) − = 0.55 4000Hz 4000Hz (3.28) or less than 1bit. To increase the resolution, the sampling frequency could be reduced to the Nyquist frequency but this may cause problems due to aliasing of out-of band noise. The oversampled D flip-flop FDSM In most applications, the D flip-flop FDSM must be followed by a standard ∆-Σ decimator as illustrated in Fig. 3.22. Normally a multi-rate architecture is chosen to reduce the physical complexity of the decimator. For first-order modulation, a sinc2 based decimator is found to be a proper choise [23] for the high-frequency stage. fclk FM fclk clk clk DQ DQ FDSMout 1 m fd m m m REG REG xn REG REG m yn Figure 3.22: A D flip-flop FDSM followed by a modulo-m sinc2 decimator stage 57 The output of the FDSM will, in FM oversampling mode, be HIGH when a FM edge is detected and LOW elsewhere as illustrated in Fig. 3.23. The input to the high-speed decimator stage xn will now be a synchronous representation of the FM edge positions. FM FDSM out Figure 3.23: FDSM output for FM oversampling mode From [23] the output of a sinc2 decimator stage is given by yn = D D xD(n−2)+k+l (3.29) k=1 l=1 where D is the decimation ratio. As an example, for D=4 the output at n=1 will be y1 = 1x−2 + 2x−1 + 3x0 + 4x1 + 3x2 + 2x3 + 1x4 (3.30) In other words, the decimator output is a triangularly weighted sum of the input sequence. If the decimator output frequency fd is chosen four times the Nyquist frequency, the width of the triangular window will be two times 1/fd or 1/(2fN ). To see the relationship between the traditional frequency-discriminator and the FDSM, we notice that counting zero-crossings during a fixed time window 1/(4fN ) is mathematically equivalent to dividing the time window into small non-overlapping sub-intervals, then count the number of zero-crossings in each sub-interval, mutiply each number by one, and finally add the result together to form the overall sum. The overall sum will now equal the total number of zero-crossings during the original interval 1/(4fN ). This operation is illustrated in the upper part of Fig. 3.24. weigth FM weigth 1/fd 1/fclk Figure 3.24: Upper part: Uniformly windowed zero-cross-counting. Lower part: Bartlett windowed zero-cross counting Since the number of zero-crossings in all sub-intervals is multiplied with one, there is an inherent uniform time window involved in the operation. In the traditional frequencydiscriminator this is seen by the fact that the position of each zero-crossings in the original 58 time window 1/(4fN ) does not affect the converter output. The only thing that matters is the total number of zero-crossings inside the main time window. The operation of the FDSM together with the sinc2 decimator can now be decsribed as follows; the FDSM divides the original time window 1/(fd ) into fclk /fd number of nonoverlapping sub-windows of length 1/fclk . Then the number of zero-crossings during each sub-intervals are counted. Since the FM signal is oversampled, the number of zero-counts in each sub-interval will be restricted to zero and one. The decimator input xn will now represent the number of zero-crossings during each sub-interval. Due to the triangularly weighted windowing function of the decimator, the decimator output will now be the weighted sum of each number of zero-crossing in each sub-interval as illustrated in the lower part of Fig. 3.24. Due to the triangular or Bartlett data window, not only the number of zero-crossings during 1/(fd ) is measured, but also the position of each zero-cross inside the main time-interval will now be of significance in calculating the output data sequence. In this way we are also suppressing the significance of zero-crossings close to the edges of each time interval since the FM period represented by these crossings contains a large phase error due to cutoff effects. What we have seen now is that by exchanging the inherent uniform data window in the traditional zero-cross frequency discriminator with a Bartlett window, ∆-Σ noise-shaping results. Unfortunately there are no indications that the noise-shaping order will be further increased by using a more sophisticated window. 59 3.5 The single-bit pointer FDSM In Section 2.5 the pointer-FDSM was introduced. This FDSM variant includes the frequency modulator as a modulated ring-oscillator. The applications for this kind of converter is analog-to-digital conversion where the analog input signal can be any kind of physical or electrical signal modulating the inverter delays. The pointer-FDSM is therefore suitable for integrated sensor applications as it will immediately convey the measured signal to the digital domain. To obtain the highest resolution for a given ring-oscillator the sampling clock frequency should be set as high as possible. Since the converter operates by sampling the state of the ring-oscillator, there will be a certain clock frequency, for which the use of a higher frequency, the change in state will be restricted to zero or one. This clock frequency will be given by fsb = 1/(τ0 − ∆τ ) (3.31) where ∆τ is the maximum delay deviation from the nominal inverter delay. For fclk > fsb , by using the implementation shown in Fig. 2.16, the FDSM output will be a bit-stream. If the number of inverters in the oscillator is high and/or a very high sampling frequency is used, the implementation shown in Fig. 2.16 based on a binary encoder and a non-binary differentiator, is rather complex and may be the limiting speed factor. For high speed operation where fclk > fsb , the converter used in Fig. 2.16 may be replaced by a simpler converter detecting whether a change of state during 1/fclk has occurred or not. 01 0 0 1 1 0 state n 0 01 0 1 1 1 0 state n+1 0 Figure 3.25: The difference between two adjacent states of a length k ring-oscillator In Fig. 3.25 the values of the top nodes in a k-bit ring-oscillator is shown for two adjacent states. As the number of inverters is odd there will be a specific location in the ring where two adjacent nodes have the same value, elsewhere all nodes are stable at 0-1-0-1... The difference between two adjacent states will be given by the value of the two equal nodes. Therefore, by finding the number of logical “1”s in a k-inverter ring-oscillator, this number will oscillate between (k − 1)/2 and (k − 1)/2 + 1. The new simplified FDSM may now be implemented by a circuit, first sampling the logical node values, then adding all nodes values together, and finally differentiating the sum to detect whether a new state has arrived or not. As the adder output is restricted to (k − 1)/2 and (k − 1)/2 + 1, there are no need for carry calculation and the adder may be implemented simply as a k-input XOR gate as illustrated in Fig. 3.26. For applications where k is large and a high sampling frequency is used, the XOR gate may be splitted into a pipelined, two-input XOR tree to increase speed. As in the original pointer FDSM there are no problems due to race conditions or glitches in the oscillator readout circuit as the difference between two adjacent states is the value of a single bit. By using a pipelined XOR tree, and a multistage decimator, the maximum sampling speed will be limited by metastability failure in the sampling D-flip-flops [21, 22]. For a modern high speed digital process, where the sampling D flip-flops are implemented as high speed sampled comparators, the maximum sampling speed may be at several GHz. Another factor that must be considered when such a high sampling frequency is used is phase-noise on the system clock (see Section 4.1). 60 k D Q yn clk clk Q D clk Q D clk Q D clk x(t) x(t) x(t) Figure 3.26: The single-bit pointer FDSM From Section 2.5 we saw that for a given inverter delay, the number of inverter used do not affect the SQNR. However, in [24] an interesting result is documented where it is found that in a MOS transistor, flicker noise is reset when the current through the transistor is turned off. In a CMOS inverter-based ring-oscillator, the MOS transistors are most of the time switched off, causing the flicker noise to loose its correlation with itself from each time the transistor is active to next time it is active. In this way flicker noise will loose its 1/f spectral density shape and be significantly reduced for low frequencies. This result explains why there was not found any noise floor down to 7.5Hz in the pointer FDSM measurement carried out in [7]. Due to this effect the pointer FDSM will be an interesting device used in low-noise sensor applications. When it comes to transistor thermal noise it is reason to believe that increasing the number of inverters will decrease the overall output noise as the thermal noise is not correlated from transistor to transistor. By doubling the √ number of inverters, the overall thermal noise rms value should therefore be decreased by 2. However, this result is not yet documented. On the other hand, increasing the number of inverters will not increase the power consumption in the ring-oscillator itself, but there will be more power burnt in the read-out circuit and in the XOR gate. Finally it should be mentioned that, as in the ordinary D flip-flop FDSM, the single-bit pointer FDSM may also be used in FM undersampling mode by restricting the sampling frequency to an interval where the change in state is limited to m and m + 1 where m is an integer. Fig. 3.27 shows the ouput psd of a simulation of a 7-inverter FDSM based on the architecture shown in Fig. 3.26. The nominal inverter delay was set to 180ps. The delay was then sinusoidally modulated to ±7%, and a 1GHz clock frequency was used. For this frequency, the number of state changes from sample to sample will be limited to 5 and 6. The SQNR was found to be 129dB for the audio band 0-20kHz. From Eq. 2.28 the theoretical value for a 7-inverter pointer FDSM is 121dB. The difference is assumed to be due to pattern noise effects, see Section 4.2. 61 Normalized PSD (dB) 0 -50 -100 -150 3 10 4 5 10 10 6 10 Frequency (Hz) Figure 3.27: The simulated output psd of a 7-inverter, single-bit pointer FDSM. fclk =1GHz, τ0 =180ps, ∆τ =13pS, fmax =20kHz 62 3.6 Paralell conversion In ∆-Σ modulators, there are several ways to increase the SQNR. In traditional switchedcapacitor modulators the use of very high sampling frequencies are troublesome due to slewrate and finite bandwidth limitations. One way to go is to increase the modulator order, but now the conversion challenge is shifted back towards the analog domain again and problems due to stability and/or matching will arise. A currently popular research field is therefore multi-bit modulators where the oversampling ratio and the modulator order may be kept low, although a high resolution may be achieved due to the reduced quantization error rms value. Compared to a single-bit modulator of the same order, a multi-bit modulator will also reduce the pattern noise and for orders higher than two, increase stability. In traditionally ∆-Σ modulators where the signal amplitude is represented by a voltage, multi-bit modulators are struggling with non-linearity problems in the corresponding multi-bit feedback DACs. These problems may however be overcomed or reduced by a number of more or less complex correction techniques [11, 12, 13, 14, 15]. In FDSM converters, the situation is quite different. Here, there are no feedback DAC and thus no need for DAC non-linearity reduction techniques. Another factor is that quantization in the FDSM is not done in voltage or current [16], but rather in time by quantizing the FM phase. As an example, we may consider the pointer FDSM where the use of several phase shifted versions of the FM signal increases the θn estimation accuracy. Since quantization is done in time, the FDSM concept is suitable for parallelisation where the output from several identical modulator structures are added together to produce a multi-bit ∆-Σ modulator. In this way the overall resolution is increased without increasing the oversampling ratio, FM deviation or the modulator order. To illustrate this point we will start by showing that the single-bit pointer FDSM can be viewed as a ring-oscillator connected to a group of parallel D flip-flop FDSM’s. 3.6.1 Parallelisation The output signal from the single-bit modulator in Fig. 3.26 can be modelled as yn = (Q1n ⊕ Q2n ... ⊕ Qkn ) ⊕ (Q1n−1 ⊕ Q2n−1 ... ⊕ Qkn−1 ) (3.32) where Q1n ...Qkn are the digital outputs of D flip-flop no. 1...k at time n/fclk . However, the XOR operator is distributive [25], and the equation can equally be written as yn = (Q1n ⊕ Q1n−1 ) ⊕ (Q2n ⊕ Q2n−1 )... ⊕ (Qkn ⊕ Qkn−1 ) (3.33) This equation describes a parallel modulator scheme as the content of each parenthesis is equivalent to the output of a separate D flip-flop FDSM. Each D flop-flop FDSM output is then added together in a k input XOR gate to form the final output. As each D flipflop FDSM operates on a phase shifted version of the overall FM signal, the resolution is improved. This result is to some extent intuitive as the output from a ∆-Σ modulator can be viewed as a low frequency in-band signal plus quantization noise. By adding several slightly, continuos-time shifted versions of the signal together, the underlying signal will be strongly correlated while the quantization noise is to some degree negative correlated. Due to the negative correlation, the SQNR is increased by ≈6dB for each doubling of the number 63 of channels compared to ≈3dB that would have been the result if the quantization error in each channel had been totally uncorrelated. It is quite easy to mathematically describe the operation of the single-bit pointer FDSM if the phase shifts are equally distributed in the [0, 2π interval. In this case the underlying FM signal at node l can be modeled as t f m(t)l = sin 2πfc t + 2πk (3.34) x(τ )dτ + ϕl 0 where ϕl = 2πl/6 + ((−1)l − 1)π is the phase difference given by the tap number l. In Fig. 3.28 the hard-limited continuous-time FM node values in a 3-inverter ring-oscillator are illustrated. 2π FM1 2π/3 FM2 FM3 FMs FDSM out Figure 3.28: Timing diagram of a 3-inverter, single-bit, pointer FDSM. Top signal FM1 to FM3 , oscillator node values. Signal FMs - equivalent FM signal for single-channel conversion The output of the single-bit pointer FDSM will now be logical HIGH when the state of the ring-oscillator has changed, elsewhere the output is LOW. A change of state has taken place when there have been a logical change in one of the three nodes. The bottom signal in the figure illustrates the FDSM output. By looking closer at the timing of the different signals we notice that the FDSM output is identical to the output of a standard single channel D flip-flop FDSM where the input FM signal is given by the signal FMs in the figure. The signal FMs is actually the continuous-time XOR sum of the three node values, and the signal itself is an ordinary limited FM signal with three times the frequency of the FM signal in one of the ring-oscillator nodes. This result make the analysis of pattern noise in the pointer FDSM much easier as the pattern noise will be identical to that in a ordinary single-channel D flop-flop FDSM (see Section 4.2) but where the input carrier and deviation are multiplied by k. Multi-bit conversion The parallel pointer FDSM described by Eq. 3.33 consists of k ordinary ∆-Σ bit-streams added together in a XOR gate. As each signal by itself is a true ∆-Σ bit-stream they can also be added together in a standard digital adder. By doing so, the FDSM output will be a word-stream if the output dynamic range exceeds 1. In Fig. 3.29 the parallel multi-bit modulator is illustrated. If the dynamic range of the bit-stream in each channel is close to 1, the FDSM output dynamic range will be close to k, yelding a log2 k-bit output signal. For equal inverter delays 64 clk x(t) clk clk D Q D Q 1 clk clk 1 D Q D Q x(t) n yn k 1 x(t) clk clk D Q D Q Figure 3.29: A multi-bit pointer-FDSM based on a k-inverter ring-oscillator and k parallel D flip-flop FDSM’s it can be shown that the FDSM output signal is equivalent to that of a single modulo-2n FDSM, but where the input FM carrier frequency and deviation is multiplied by k. 3.6.2 Open-ended delay-line parallelisation In frequency-to-digital conversion applications where the input to the converter is an externally generated FM signal, there will normally not be available any phase shifted FM information. In this case the simplest way to achieve additional phase information is to feed the FM signal through an open ended delay line with constant delays as illustrated in Fig. 3.30. FM clk clk clk D Q D Q 1 clk clk 1 D Q D Q n yn k 1 clk clk D Q D Q Figure 3.30: A multi-bit pointer-FDSM based on a k-inverter delay-line and k parallel D flip-flop FDSM’s By doing so, we can no longer expect the phase shifts to be uniformly distributed over the [0, π interval, as will be the case in a symmetrical ring-oscillator where all inverter 65 delays are equally modulated. In principle, it is possible to adjust the FM carrier frequency to match the inverter delays to make the phase shifts uniformly distributed over the [0, π interval, but when the FM signal is modulated to a higher or lower frequency, the symmetry will be disturbed. However, for modulating signals of very small amplitudes the phase distribution may, in principle be almost equal to that in a closed ring-oscillator for a properly chosen carrier to delay ratio. On the other hand, by making the delay line ring-connected, modulated by the FM signal, and included in a phase locked loop, it may be possible to distribute the phase shifts almost uniformly even for high FM deviations, but this possibility have not been investigated yet. To analyze the effect of non-uniformly distributed phase shifts, we may start by looking at the FDSM output for a situation where the total delay through the line is less than half a FM period. This situation is illustrated in Fig. 3.31. π FM0 FM1 FM2 ϕ ϕ ϕ ϕ FDSM out Figure 3.31: Timing diagram of a two-inverter delay-line FDSM with two non-uniform phase shifts. In this example a delay-line with three tappings are used where the phase shift of each inverter is ϕ. The FDSM output will now consist of periods of pulse-trains corresponding to the zero-crossings in each channel. To find the degradation in SQNR caused by non-uniform phase delays, we may start by noticing that by adding k phase-shifted bit-streams together, the output signal power will be k times the signal power in each bit-stream if the phase shifts are small compared to the frequency of the modulating signal. This is because the amplitude of the underlying signal is almost equal in each bit-stream causing the signal correlation to be close to 1. The correlation of the quantization error in each bit-stream will, on the other hand be far from 1, and the output quantization noise power can be found by expressing the total quantization error as the sum of the quantization error in each bit-stream as follows etot n = k−1 l=0 eln = k−1 modπ (θn − ϕl ) (3.35) l=0 In this expression eln is the quantization error in bit-stream number l at sample number n. This is illustrated in Fig. 3.32 for a delay-line with three tappings sampled twise. To find the total quantization noise power at the FDSM output we may start by looking closer at the sum given by Eq. 3.35. First, we notice that the total quantization error is dependent on the position of the sampling edge relative to the FM period. From Fig. 3.32 we see that by shifting the position of the sampling edge to the right, the total quantization error will increase until the error in one of the channels has reached its maximum of 1 corresponding to a phase difference of π. When this situation occurs, the estimated FM angle in that channel increases with 1 (one ZC count) and the quantization error in that channel is reset to zero. When the sampling edge is further shifted to the right, the quantization 66 sample π e'0 FM0 e0 FM1 FM2 sample' e1 e'1 e'2 ϕ ϕ e2 Figure 3.32: The quantization errors in a three tapping delay line FDSM for two different samples 4 4 3.5 3.5 3 3 Total quantization error Total quantization error error in each channel will increase again until the error in one of the other channels reaches its maximum and the process is repeated. In Fig. 3.33 the total quantization error in a three-inverter delay-line FDSM is shown as the sampling edge is varied over an interval of 5/4 times the FM period. 2.5 2 1.5 1 2 1.5 1 0.5 0.5 0 2.5 0 π 2π Sampling edge position relative to FM phase π 2π Sampling edge position relative to FM phase Figure 3.33: Total quantization error in a 3-inverter delay-line FDSM versus position of sampling edge. Equal delays for each inverter. Left picture: Blue line - no inverter delay. Red line - phase shift for last tapping 0.4π: Black line - uniformly distributed phase shifts in the [0, π interval. Right picture: Blue line - phase shift for last tapping 2π. Red line - phase shift for last tapping 2.4π. Black line - phase shift for last tapping 3π In the left picture, the blue graph illustrates the total error for the case where the inverter delays are zero. In other words, the FM signals at each tappings are equal. In this case, the total quantization error will be four times the error in each channel. As we see from the graph the total quantization error varies between 0 and 4 as the sampling edge position is shifted. In this case both the signal power and the quantization noise power will be increased by 4 times compared to that in a single D flip-flop FDSM. In other words, the signal to quantization noise ratio will be the same. This should not be surprising, as there are no additional information added to the system by copying the FM signal into 4 equal versions. The quantization noise power can, to a first degree of accuracy, be calculated by assuming equal possibility for the sampling edge to occur anywhere in the FM period. From this 67 20 20 18 18 16 16 Total quantization error Total quantization error assumption the quantization probability density function (pdf) will be uniform over the √ [0, 4 interval, and the rms value will be 4/ 12. In the same picture, the red graph illustrates the total quantization error for a situation where the total phase difference between the input and the output of the line is 0.4π. In this case the total quantization error will be limited to the interval [0.6, 3.4]. The pdf for this case will be divided into three different intervals each with a uniform distribution, but by visual inspection we can expect the noise rms value to be lower than for the zero delay case. Finally, the black graph illustrates the optimal case where the phase delays are symmetrically distributed over the [0, π interval. In this case the total quantization error is limited to√the interval [1.5, 2.5] and the pdf is uniform. The output noise rms value will now be 1/ 12, and the SQNR will now be 4 times higher than for the single D flip-flop FDSM. In Fig. 3.33 right, the total error is calculated for the case where the total delay through the line is greater than half the FM period. For the blue graph the delay from the input of the line to the last tapping is one FM period. In this case the error is increased to the [1, 3] interval as the ZC distribution symmetry is disturbed. By increasing the line delay further to 2.4π, the error interval is somewhat decreased as illustrated by the red graph. Finally, for a total line delay of 3π the phase delays are again symmetrical, and the total error is restricted to the [1.5, 2.5] interval. 14 12 10 8 6 14 12 10 8 6 4 4 2 2 0 0 π 2π Sampling edge position relative to FM phase π 2π Sampling edge position relative to FM phase Figure 3.34: Total quantization error in a 90-inverter delay-line FDSM versus position of sampling edge. Equal delays for each inverter. Left picture: Blue line - no inverter delay. Red line - phase shift for last tapping 0.4π: Black line - uniformly distributed phase shifts in the [0, π interval. Right picture: Blue line - phase shift for last tapping 2π. Red line - phase shift for last tapping 2.4π. Black line - phase shift for last tapping 3π In Fig. 3.34 the same procedure is repeated for a 90-inverter delay line. Compared to the 3-inverter case we notice that the error rms value is significantly reduced even for cases where the total line delay is more than halve the FM period. As the number of inverters is increased, the overall phase distribution will be more complex and seems to be more evenly distributed causing a lower noise rms value. For a delay line with a high number of inverters the relationship between the overall quantization noise power and the total line delay will be quite complex. To get a impression of this relationship the noise power for different delays are simulated in Matlab [26] and illustrated in Fig. 3.35. In the left figure, the quantization noise power is calculated as a function of the overal line delay for a line of 30 inverters. The noise power is normalized to make zero correspond 68 Normalized total quantization noise power (dB) Normalized total quantization noise power (dB) 0 -5 -10 -15 -20 -25 -30 π 2π 3π 4π 5π 6π Delay of last tapping (rad) 7π 8π 0 -5 -10 -15 -20 -25 -30 -35 -40 π 2π 3π 4π 5π 6π Delay of last tapping (rad) 7π 8π Figure 3.35: Total quantization noise power as a function of line delay. Left picture - 30-inverter delay-line. Right picture -90-inverter delay-line to the noise power in a single D flip-flop FDSM. As the inverter delays are increased from zero the noise power is decreased following a sinc like function with the lower extremes at multiple locations of π. The bottom line illustrates the theoretical noise level at -20log(30) corresponding to a perfect symmetrical distribution of the phase delays in the interval [0, π. We notice that the first minimum at π is located in a very narrow range, and just a slight delay deviation from this point will increase the noise power significantly. In theory, it would be possible to choose the necessary delay / FM carrier ratio to utilize this point and obtain a high resolution for small signal modulation. However, for large modulating signals the overal noise power will increase significantly as the phase delay distribution will be heavily varyed around the π point in the figure. Particularly, when the modulating signal is at its lower extrema, the FM signal frequency will be low, causing the maximum phase delay to be less than π which is the region where the noise power increases most rapidly. In a practical implementation, it will also be quite difficult to utilize this point as just a slight process deviation in the transistor parameter β [27] or a drift in the FM carrier frequency will move the operation point out of range. However, if it is possible to dynamically adjust the FM carrier frequency to make the phase distribution uniform, this point may be utilized to increase the resolution for low amplitude signals. A probably better range to locate the modulator in will be at, for example 4π where the valley is wider allowing more drift in the FM carrier frequency and relaxed delay accuracy. For higher line delays the sinc like function is dominated by periodic noise peaks with triangular shapes. If the operation point is to be located in these regions, one must ensure that there are enough room for drift and process deviations to avoid the main peaks in the picture. In Fig. 3.35 right, the same noise power is calculated for a 90-inverter delay line. The depth of the first valley is now increased as the power of an ideal phase distribution is -20log(90), illustrated by the horizontal line at the bottom of the figure. Also in this case, the noise power follows a sinc like function, and even if the number of inverters is increased, the magnitude of this function is the same as for the 30-inverter line. However, the periodic noise peaks at higher line delays are significantly reduced. This makes it easier to locate the operation point in these regions. In a practical implementation it is hard to make the inverter delays perfectly equal due to β mismatch effects. The delay matching can for CMOS implementations be significantly improved by increasing the gate area of each transistor [28]. On the other hand, for a 69 Normalized total quantization noise power (dB) Normalized total quantization noise power (dB) 0 -5 -10 -15 -20 -25 -30 π 2π 3π 4π 5π 6π Delay of last tapping (rad) 7π 8π 0 -5 -10 -15 -20 -25 -30 π 2π 3π 4π 5π 6π Delay of last tapping (rad) 7π 8π Figure 3.36: Total quantization noise power as a function of line delay. 10% random delay deviation added to each inverter. Left picture - 30-inverter delay-line. Right picture - 90-inverter delay-line delay line consisting of a large number of inverters where the total delay is longer than the FM period, we have seen that the sum of all quantization errors follows a quite complex pattern. Small delay deviations along the line due to mismatch effects will now increase the complexity further, and there may be a point where complexity is so high that the total quantization error is almost independed of small deviations from the operation point. So, instead of trying to minimize delay mismatch, it can be interesting to increase the mismatch, and in this way make the overall noise power less sensitive to drift of the operation point. This will also decrease the noise power for modulating signals of high amplitude. In Fig. 3.36 the noise power is calculated for the same lines as in Fig. 3.35 but now there are added a 10% relative mismatch to the inverter delays. From the figure we notice that the main sinc shape is still present but the periodic noise peaks at higher line delays are significantly reduced. As we expected, the noise function is much more convergent for higher line delays, and there are more improvement seen in the SQNR when the number of inverters is increased. The noise power in Fig. 3.35 and Fig. 3.36 are calculated for different constant FM period to delay distributions. The situation for a sinusoidally modulated FM signal with high deviation will be quite more complex as the phase distribution will vary with the deviation of the FM signal. At the point where the FM signal frequency is at its maximum and minimum, the noise power can be described by Fig. 3.35 and Fig. 3.36, but since the FM frequency, and thus the phase distribution is constantly varying the picture will be much more complex. A more accurate way to find the SQNR in this case is by a simulation of the complete modulator. In Fig. 3.37-left, a delay-line FDSM is simulated for different line length and inverter delays, and the SQNR is calculated. As the number of inverters is increased, the inverter delay is decreased to make the phase distribution uniform over the [0, π interval. This locates the operating point to the first and deepest valley in Fig. 3.35 and Fig. 3.36 for all simulations. The first point to the left in the figure for each curve, is the SQNR for a single D flop-flop FDSM. The next point corresponds to a simulation with 3 inverters where the inverter delay is 4.17ns. The last point to the right for each curve corresponds to a 128-inverter line where the inverter delay is 130ps. The FM carrier frequency was set to 30MHz and a sinusoidal modulating signal at 3MHz deviation was applied. By using such a high deviation the phase distribution is heavily disturbed by the modulating signal. The sampling frequency was 200MHz and an audio application was considered with a signal 70 130 140 125 rin 115 al ide 110 no mismatch 10% mismatch 105 78ns 130 SQNR (dB) SQNR (dB) 120 or lat cil s g-o 120 s 0n 15 110 al ide s 110n 100 100 90 95 80 10 1 100 Numb of taps 10 100 Numb of taps 1000 Figure 3.37: Simulated SQNR versus number of inverters in a delay-line FDSM. fc =30MHz, ∆f =3MHz, fclk =200MHz, signal bandwidth 20kHz. Left figure - inverter delay choosen to make phase distribution uniform over the [0, π interval for zero modulation. Top curve: ideal SQNR for ring-ocillator FDSM, next curve: simulated result for ring-ocillator FDSM. Bottom curves: open delay-line FDSM with equal inverter delays and with 10% random delay mismatch. Right figure simulated SQNR for open delay-line FDSM with different inverter delays bandwidth of 20KHz. The top curve in the figure illustrates the ideal SQNR for the case where the phase is always perfectly distributed as it will be in a ring-oscillator based FDSM. The next line illustrates the simulated SQNR for a practical ring-oscillator FDSM with no delay mismatch. The SQNR is at average ≈4dB lower than the ideal curve. This is assumed to be caused by pattern noise. The next two curves shows the SQNR for the open delay-line FDSM both with no inverter delay mismatch and with 10% relative delay mismatch. There are no significant differences between these two cases as the number of inverters increases, and we may conclude that the phase distribution is already so complex in these cases that a random delay mismatch does not affect the overall SQNR signifficantly. However, as the number of inverters increases the difference in SQNR from the ideal curve increases as the first valley in Fig. 3.35 and Fig. 3.36 gets narrowed. From this simulation we may therefore conclude that if the operating point is to be located in the first valley of Fig. 3.35 and Fig. 3.36, there is no significant increase in SQNR by increasing the number of inverters to more than ≈80. In Fig. 3.37-left, another situation is shown. In this case the inverter delay is kept constant while the number of inverters in the line is increased. To make the simulation more realistic, a 5% random delay mismatch is added to each inverter. The SQNR is found for three different inverter delays, 150ps, 110ps and 78ps which are achievable values in a modern CMOS process. In the left part of the figure the phase distribution is restricted to only a fraction of the [0, π interval, and there are almost no increase in SQNR seen by increasing the number of inverters. In the middle of the figure the phase distribution starts to cover the [0, π interval and the SQNR increases. The SQNR for the 150ps and the 110ps case are however far from the ideal curve even when a high number of inverters is used. From this simulation, we can conclude that for FM signals with high deviations the SQNR may be significantly lower than for a perfectly phase distributed ring-oscillator FDSM. However, the SQNR is still increased by increasing the number of inverters in the line, although the exact resolution is a complex function of different parameters. So far we have only simulated the delay-line FDSM for FM signals with a high relative 71 deviation. Now we will verify that the delay-line FDSM can provide a much higher performance for small relative deviations by locating the operating point in one of the valleys shown in Fig. 3.35 and Fig. 3.36. In the following example, the first valley was chosen. In Fig. 3.38 the SQNR is found from a simulation of a 128-inverter delay-line FDSM where the FM carrier frequency is set to 30MHz, the sampling frequency 200MHz and the signal bandwidth set to the audio range of 0-20kHz. From Eq. 2.28 the theoretical SQNR for a sinusoidal modulating signal causing a 3Hz FM deviation should be 5.3dB. From the simulation the SQNR was calculated to 20dB, or almost 4bit higher than the theoretical value. The reason for this result is that the FDSM dynamic range is located in a patter noise valley, as will be described in Section 4.2. In Appendix B, the simulator used in this example is listed. Nomalized power Spectral Density (dB) -80 -100 -120 -140 -160 -180 200 1000 10 000 100 000 Frequency (Hz) 1000 000 Figure 3.38: Simulated power spectral density for a 128-inverter delay-line FDSM. fc =30MHz, ∆f =3Hz, fclk =200MHz, signal bandwidth 20kHz. 3.6.3 A 7-bit parallel audio FDSM An experimental VLSI implementation of a delay-line FDSM have been made by the DeltaSigma Group1 at IFI, UIO. The test converter is implemented in a CMOS, single poly 0.6µm AMS process, and are based on a delay-line of 1024 inverters. The converter is one of two solutions targeted at a microphone interface application in an audio noise analyzer. The other solution is based on a second-order FDSM and will not be described in this thesis. For each solution there are scheduled several test chips to achieve more experience and understanding of the FDSM concept. The system microphone is based on a low noise LC oscillator and the output is a FM signal with a carrier frequency of 30MHz. The maximum frequency deviation is 3MHz, and the goal is to achieve a dynamic range of 24bit in the 0-20kHz audio range. The first FDSM test chip consist of 1024 parallel D flip-flop FDSM’s 1 The Delta-Sigma group consists of Dag T.E. Wisland, Jan T. Marienborg, Tor S. Lande, Yngvar Berg, Monica Finsrud and the author. The group is currently working in collaboration with the company NorSonic in Lier, Norway, to develop high resolution digital audio noise analyzer. 72 128-inverter delay line sinc2 decimator as shown in Fig. 3.30, and the output bit-stream of each FDSM is summed together in a pipelined adder treee. By using 1024 parallel modulators, the size of the chip can be quite large, and to make the implementation more compact the system is grouped into four equal parts consisting of 128 inverters with corresponding D flip-flop FDSM’s and a bit-stream adder. The output from these four groups are then added together to make the final FDSM output. The overall FDSM output is 10-bit, but for a maximum amplitude audio signal only 7-bits will be active. adder tree final adder Figure 3.39: A multi-bit pointer-FDSM based on a 4·256-inverter ring-oscillator and 4·256 parallel D flip-flop FDSM’s In Fig. 3.39 the layout of the chip is shown. The delay line is divided in four parts and located as a ring just inside the pads. The adder tree output for each section is at the center of the chip, and the final adder is located on the outer side of the right delay-line section. A 30-bit wide sinc2 decimator is also included on the chip. The total chip area without pads is 7mm2 . As a test chip there have been no efforts taken to reduce area in the digital circuits which are designed in house. The decision to locate the analog delay-line along the edge of the circuit where substrate coupled switching noise from the pads are at its highest, have been carefully analyzed. There have been numerous simulations indicating that switching noise modulating the inverter delay has almost no effect on the overall signal to noise ratio if the fundamental noise frequency is a multiple of the decimator output frequency. Even so, the delay-line Vdd and Gnd terminals are implemented with separate routing and pads. In the next section we will also see that phase noise in general is shaped by the FDSM. The delay-line inverters were implemented with minimum gate lengths, and the delay is simulated to 130ps. Under normal operation the power consumption is simulated to 73 maximum 1W. The SQNR for this circuit is high-level simulated for sinusoidal modulation at different modulating signal amplitudes as illustrated in Fig. 3.40. For a maximum amplitude modulating sinusoidal, the SQNR was found to be 132dB or ≈22-bit. However, as the FM deviation is lowered the quantization noise is reduced, and the dynamic range of the converter is found to be more than 140dB or almost 24-bit. 140 120 SQNR (dB) 100 80 60 40 20 0 -7 10 10 -6 -5 -4 -3 -2 10 10 10 10 Amplitude of modulating sinusoudal 10 -1 10 0 Figure 3.40: SQNR for the 1024-inverter delay-line FDSM for versus modulating signal amplitude. Straight line - theoretic resolution based on the white quantiztion noise model 74 75 Chapter 4 Excess noise 4.1 Phase Noise In many analog demodulation systems where a high signal to noise ratio is essential, phase noise can be a serious problem. Phase noise will be present both on the analog modulated signal, and for sampled-time systems it will, in addition, be present on the sampling clock. This will also be the case for a practical FDSM system where the phase noise present on the FM signal and on the clock signal is illustrated in Fig. 4.1. If a high demodulated signal to noise ratio is to be achieved, the effect of phase noise must be carefully analyzed. In the following, the effect of phase noise both on the FM signal and on the clock signal will be modeled. Since the phase of a FM signal is the integral of the FM signal frequency, phase noise can also be modeled as frequency noise by considering a 1/s relationship. It turns out that phase noise on the clock signal is easier modeled by considering its frequency noise equivalent, and the comparison between FM and clock noise will therefore be carried out on their frequency noise equivalents. To simplify the analysis, only the sampled-signal FDSM is considered. However, as the sampled-clock FDSM is an equivalent FDSM system the analysis for this kind of modulators will be identical by considering a change in time-base (Section 3.2.1). fclk + ϕ'n ϕn FM + FDSM yn Figure 4.1: Phase noise on FM and clock inputs Phase noise on the FM signal We will begin the noise analysis be looking at phase noise on the FM signal. First, we will recapitulate and look at the operation of a ZC counting, multi-bit FDSM which consists of an asynchronous ZC counter followed by a synchronizing register and a differentiator as 76 illustrated in Fig. 4.2. This circuit can be viewed as an extension of the D flip-flop FDSM as the two outputs will be identical as long as the ∆f /fclk ratio and the FM carrier frequency is chosen to make the output limited to m and m + 1, where m is an integer. In a practical realization, modulo arithmetic will be used both in the counter and in the differentiator to limit the necessary word-length. However, as long as the modulo arithmetic is working properly it will not affect the converter operation from a mathematical point of view. By disregarding the modulo operation and considering infinite word-length, the asynchronous ZC counter will represent the total FM phase θ(t) by the total number of received zerocrossings. In this way the FM phase is scaled and quantized since it is represented by integers (number of halve FM periods). As an example, if the total FM phase θ(t) is (247π + 0.03), it will be represented by the integer 247. The scaling factor is π and the quantization error e is −0.03/π. counter fm reg fclk θn diff cnt yn = θn-θn-1 reg Figure 4.2: A basic multi-bit FDSM The counter output can now be considered as an estimate of the total phase given by θ(t) + e(t), −1 < e(t) ≤ 0 (4.1) π Notice that the counter output represents the FM phase in continuous time but in quantized amplitude. In a FSDM the variation in kx(τ ) is small during the sampling interval 1/fclk due to oversampling. Therefore Eq. 2.7 can be approximated by Eq. 2.8 which is repeated here for convenience θ̂(t) = θn = 2π fclk n (fc + kxm ). (4.2) m=−∞ The counter output is then synchronized with the system clock and differentiated. From Eq. 4.1 and Eq. 4.2 the FDSM output θ̂n − θ̂n−1 will be yn = 2 fc + kxn + en − en−1 , fclk −1 < en ≤ 0 (4.3) Phase noise on the FM signal can now be represented by an additive error signal ϕn on the FM phase fm(t) = sin[θ(t) + ϕ(t)]. (4.4) The FDSM input is a limited FM signal, thus amplitude noise on the original FM signal may also modulate the FM edge positions. However, this disturbance can also be modeled as phase noise included in ϕ(t). The FDSM counter output θ̂n will now be an estimate of the sum of the original phase θn and the phase noise ϕn given by 77 (θn + ϕn ) + en , −1 < en < 0 π this value is then differentiated and from Eq. 4.2 we get θ̂n = yn = 2 fc + kxn 1 + (en − en−1 ) + (ϕn − ϕn−1 ). fclk π (4.5) (4.6) In other words, FM signal phase noise will enter the system in the same way as the quantization error and thus be first-order noise-shaped. This is an important result as the FDSM concept otherwise would had been unsuitable for high quality conversion. Phase noise on the system clock The effect of phase-noise on the sampling clock is similar to the previous case as we may change the reference point and consider the noisy clock signal as constant with virtual phase noise on the FM signal. However, by making a change in time-base, we need a description of the virtual FM noise. The power of the virtual noise will not necessarily equal the power in the original clock noise, and the systems sensitivity to clock noise may therefore be different from its sensitivity to FM signal noise. The easiest way to compare these effects will be to look at their equivalent frequency noise representations. 4.1.1 Equivalent frequency noise Starting with FM signal phase noise, from Eq. 4.6 the FDSM output can be expressed as yn = 2 (fc + kxn + fnϕ ) + en − en−1 fclk (4.7) where fnϕ is a resulting frequency error given by fnϕ = fclk (ϕn − ϕn−1 )/π (4.8) We notice that the frequency error is the difference of the phase error as the FM frequency is the derivative of the FM phase. The spectral density of the frequency error will therefore be given by a shaped and scaled version of the spectral density of the phase error. Phase noise on the sampling clock can now be modeled as a frequency error in the clock frequency yn = 2 fc + kxn (fclk + fnϕ ) + en − en−1 (4.9) We can now compare the FDSM sensitivity to FM signal and clock frequency errors by considering Eq. 4.7 and Eq. 4.9 for the case where there are no modulating signal, xn = 0 and the frequency error is small compared to fclk and fc . The output sensitivity to a small frequency error dfnϕ on the FM signal can be found from the derivative of Eq. 4.7 with respect to fnϕ . Notice that we are looking at the derivative of yn with respect to fnϕ and not with respect to time. So, the FDSM sensitivity to a small error in the FM frequency will be dyn 2 = . dfnϕ fclk (4.10) Similarly, for a small frequency error dfnϕ on the clock signal we have that the FDSM sensitivity is given by the derivative of Eq. 4.9 with respect to fnϕ which is 78 dyn dfnϕ =− 2fc (fclk + fnϕ )2 , (4.11) thus the output sensitivity to FM signal phase noise is 2 df ϕ , fclk n and the systems sensitivity to clock signal phase noise is dyn = dyn = − fc 2 df ϕ fclk fclk n (4.12) (4.13) assuming fclk fnϕ . From this analysis we can conclude that if the FM carrier frequency is higher than the sampling frequency, the FDSM is more sensitive to frequency and thus phase noise on the clock signal than on the FM signal. If the FM carrier frequency is lower than the sampling frequency, the FDSM is more sensitive to phase noise on the FM signal. In other words, the FDSM is most sensitive to phase noise on the signal with the lowest frequency. These results are for absolute frequency or phase errors. However, if the error specified is relative to the corresponding signal center frequency, the situation is different. If the relative FM and clock frequency errors are equal, we have dfnϕ /fc = dfnϕ /fclk , and in this case, |dy| in Eq. 4.12 will equal |dy| in Eq. 4.13. In other words, the FDSM is equally sensitive to relative frequency noise on the clock and on the FM signal. 4.1.2 Output noise power To further illustrate the relationship between phase and quantization noise in the demodulated signal, we may simplify the analysis by consider no correlation between the quantization error, phase errors and and the modulating signal. From Eq. 4.6, Eq. 4.12 and Eq. 4.13 we then have the power spectral density of the sampled-signal FDSM output given by f2 k2 Sy (f ) = 4 20 δ(0) + 4 2 Sx f f clk clk carrier signal + |(1 − e −j2πf /fclk )|2 Se quantization noise 1 + 2 |(1 − e−j2πf /fclk )|2 Sϕ π FM signal phase noise f2 + 2 c2 |(1 − e−j2πf /fclk )|2 Sϕ π f clk clock signal phase noise (4.14) In this equation Sx is the psd of the modulating signal xn , and Se is the psd of the quantization error, assumed to be uniformly distributed in the −1, 0] interval. Sϕ is the psd of the FM signal phase noise and Sϕ is the psd of the clock signal phase noise (in both cases the phase error is given in radians). 79 To calculate the in-band noise power in the demodulated signal we may consider the simplest case where both the quantization noise and the phase noise is white. Let the variance of the quantization error be σe2 = 1/12 and the variance of the phase noises be σϕ2 and σϕ2 respectively. The resulting in-band noise can then be found by integrating Eq. 4.14 over the signal band. For the quantization error, we already know the result which from Eq. 2.18 is Pye = σe2 π2 3 2fmax fclk 3 π2 = 36 2fmax fclk 3 , (4.15) For FM signal phase noise we get Ptϕ σϕ2 = 36 fmax fclk 3 . (4.16) Finally, for clock signal phase noise we get Pyϕ = σϕ fc 2 2 36 fclk 2 fmax fclk 3 . (4.17) The output power of a sinusoidal modulating signal with maximum amplitude will be Pyx = 4 k2 ∆f 2 1 P = 4 x 2 2 2. fclk fclk (4.18) By comparing Eq. 4.18 to Eq. 4.16 or to Eq. 4.17 the FDSM output signal to white-phase noise ratio can be found. 4.1.3 Simulation To verify Eq. 4.14 for FM signal noise and modulating signal power, the FDSM have been simulated for a sinusoidal modulating signal at 67.4Hz with two harmonic FM phase noise components added. The amplitude of the harmonic noise components was both set equal to 0.01rad, and the frequency to 2 and 0.5 times the modulating signal frequency respectively. The frequency deviation caused by the modulating signal was 425Hz which from Eq. 4.18 should produce a signal power of Pyx = −82dB. The corresponding phase noise power for the high frequency harmonic will from Eq. 4.14 be Pyϕ = −132dB, and for the low frequency harmonic, Pyϕ = −145dB. By using the periodogram psd estimator (Appendix A), the corresponding psd harmonic peaks will be Syϕ (134.6Hz) = −106.5dB, and Syϕ (33.75Hz) = −118.6dB respectively. For the modulating signal, we have Syx (67.3)Hz) = −56.6dB. In Fig. 4.3 the estimated psd of the simulated bit-stream is shown. By inspecting the level of the three harmonics there is found a good match with the theory. 4.1.4 Comments on phase noise in the sampled-signal FDSM We have shown that phase noise on both the FM signal and on the clock signal will be first-order noise shaped and therefore suppressed for low frequencies. The theory has been confirmed for phase noise on the FM signal by simulation. The sensitivity to absolute phase noise measured in radians is not the same at the FM and clock inputs. At the clock input the sensitivity is −fc /fclk times the sensitivity at the FM input. However, in some systems the phase noise power is relative to the period of the corresponding signal. In this case, we should keep in mind that the FDSM is equally sensitive to relative phase noise on the 80 −60 Power spectral density (dB) −80 −100 −120 −140 −160 −180 1 10 2 10 Frequency (Hz) 3 10 Figure 4.3: Simulated FDSM output psd for two equal amplitude harmonic phase noise components. Left peak - harmonic phase noise component, amplitude: 0.01 radians, frequency 33.75Hz. Middle peak - frequency modulating signal itself. Right peak - harmonic phase noise component, amplitude: 0.01 radians, frequency 134Hz. FM and on the clock input. The noise shaping effect on phase noise can be illustrated by 4 different examples I) White phase noise If the phase noise is white, its power will be uniformly distributed in the [0, fclk interval. In this case a doubling of the oversampling ratio (fclk for a constant fmax ) will reduce the in-band phase noise power by ≈9B. However, the signal range is reduced by a factor of two reducing the signal power by ≈6dB. Therefore, by increasing fclk we will have a ≈3dB improvement in the signal to phase noise ratio in the demodulated signal. II) Harmonic phase noise If harmonic phase noise is present, the frequency of the noise will determine the resulting output noise power. Since phase noise is first-order shaped, low frequency harmonics in the signal band are heavily suppressed. Tones outside the signal band will be removed by the decimating filter. If a harmonic noise tone is located in the signal band, an increase in the sampling frequency will not improve the signal to phase noise ratio in the demodulated signal. III) Flicker phase noise (f −1 ) The flicker phase noise floor increases by 10dB/decade as we go down in frequency. Due to the noise shaping the resulting psd at the FDSM output will now decrease by 20dB-10dB = 10dB/decade as we go down in frequency. An increase in the sampling frequency will not improve the signal to phase noise ratio in the demodulated signal. 81 IV) f −2 phase noise (FM noise) Noise directly on the input signal x(t) will be present as f −2 phase noise or FM noise. In this case the -20dB/decade slope of the phase noise will be equalized by the inverse shape of the noise-shaping transfer function. Due to the noise-shaping, f −2 phase noise corresponds to white noise in the modulating signal x(t). 82 4.2 Pattern noise Although first-order ∆-Σ modulators are significantly simpler to implement in VLSI than higher-order modulators, they are rarely used due to two problems: 1. In traditional SC ∆-Σ modulators the oversampling ratio is limited by slew-rate effects and finite op-amp gain. As the noise-shaping is only first-order, the digital resolution will normally not be sufficient high. 2. The quantization error in a first-order ∆-Σ modulator is highly correlated with the input signal which may cause problems due to pattern noise. The FDSM concept is, so far, based on first- and second-order noise-shaping. Particularly the D flip-flop FDSM solutions are suitable for very high sampling speed operation, and can therefore provide a high digital resolution even if the noise shaping is only first-order. In this way problem no.1 may be overcomed. By using multi-bit quantization, implemented either by parallel modulators, or the modulo-2n FDSM, problem no.2 can be significantly reduced as the worst-case signal to pattern noise ratio is reduced by the number of bits in the quantizer. In some applications it will however be desirable to use a single D flipflop FDSM with a bit-stream output, and in this case the effect of pattern noise must be carefully analyzed. In this section we will start by verifying that the FDSM is equivalent to a traditional ∆-Σ modulator with respect to pattern noise. However, since the internal signal range in the FDSM may be much lower than in a traditional ∆-Σ modulator, and the range is defined by frequency ratios and not by voltages, there can be a significant difference in performance. 4.2.1 Background The basis for Eq. 2.18 which predicts the resolution of the delta-sigma modulator is based on the assumption that the quantization noise is white, and thus uncorrelated with the input signal. This assumption is suitable for most busy input signals. However, for DC or slowly varying inputs, the white-noise model is far from exact as the quantization error will be heavily correlated with the input signal. When the input signal is DC, the deltasigma modulator output will bounce between two levels keeping its mean equal to the input signal. For certain DC input values the output sequence will be repetitive. If the repetition frequency lies in the signal band, the modulation will be noisy, if not, it will be quiet. In Fig. 4.4 (top) the input to a traditional delta-sigma modulator is swept over different DC values and the resulting in-band noise power is measured [10]. The horizontal line represent the calculated in-band noise level given by the white quantization noise model. As we see from the figure, there are certain input DC values which produce an output noise power far above the calculated level. For other input values the measured power is far below the white-noise level. This inherent feature is called pattern-noise. From [10] half the total power is found to be in the end peaks while 1/16 in the central ones. Most traditional delta-sigma modulators need to utilize most of their signal range to reduce the effect of internal noise sources. It is therefore normally necessary to let the signal range overlap many of the pattern noise peaks in the figure. As we soon will see, the FDSM is equivalent with respect to pattern noise, but here the internal signal range can be very small compared to the output levels. As the signal range location is defined by the fc /fclk 83 Total in-band noise power (dB) DC value 20 25 30 35 40 45 0 0.1 0.2 0.3 0.4 0.5 DC value 0.6 0.7 0.8 0.9 1 20 25 30 35 40 45 0 0.5 1 1.5 2 2.5 Constant frequency (MHz) 3 3.5 4 Figure 4.4: Top) Measured DC scan, traditional delta-sigma modulator. Middle) Theoretical DC model. Bottom) Simulated FDSM frequency scan. Horizontal lines - white noise model. For all plots - oversampling ratios = 9.14 ratio, we may for small signal ranges, locate the signal range in a pattern noise valley and achieve a significantly higher resolution than the white-noise model predicts. A necessary requirement for doing so, is that we have control over practical effects such as temperature drift and aging which may affect the fc /fclk ratio. The pattern noise picture is a direct function of the oversampling ratio. In Fig. 4.5 a theoretical DC scan is carried out for three different oversampling ratios. We notice that both the heights of each peak and the valley depth is increased as the oversampling ratio is increased. In addition, the number of visible peaks increases. However, since the widths of the peaks are reduced, the power in each peak is inversely proportional to the oversampling ratio cubed. 4.2.2 A theoretical model for pattern-noise In [10] an analytical model for the output of a first-order delta-sigma modulator with DC input is given. The analysis is carried out for a continuos-time modulator, which is shown 84 20 25 fmax = 437KHz 30 35 45 0 0.5 1 1.5 2 2.5 Constant frequency (MHz) 3 3.5 4 0 0.5 1 1.5 2 2.5 Constant frequency (MHz) 3 3.5 4 0 0.5 1 1.5 2 2.5 Constant frequency (MHz) 3 3.5 4 30 35 40 fmax = 133KHz Total in-band noise power (dB) 40 45 50 55 60 65 50 60 fmax = 5000Hz 70 80 90 100 110 Figure 4.5: Theoretic DC model for different oversampling ratios, fclk = 8MHz. Horizontal lines white noise model. to be equivalent to a discrete-time model. For a DC input of amplitude x, the output components that lie in half the frequency band can be expressed as yx (t) = x + 2 ∞ sin(πlx) l=1 πl cos(2π frac[lx]tfclk ) (4.19) Where frac[x] is the fractional roundoff of x, that is x minus the closest integer to x. In Eq. 4.19 the first part is recognized to be the input itself, while the sum represents the quantization noise. As we see, the quantization noise consists of scaled harmonic components with a frequency dependent on the index l. For a harmonic component to lie in the signal band 0 < f < fmax we have f fmax = |frac[lx]| < fclk fclk (4.20) and the power associated with that component will from Eq. 4.19 be Pl (x) = 2 sin2 (π frac[lx]) (πl)2 85 (4.21) The total noise power in the signal band is then found by adding together the power of all harmonics where the index l satisfies Eq. 4.20. The two equations Eq. 4.20 and Eq. 4.21 are the basis for the theoretic model used in this thesis. In Fig. 4.4 (middle) the theoretic model is used to predict the pattern noise picture for the same oversampling ratio that was used in the measurement. Although the measured result is more smoothed, the model resembles the main shape. 4.2.3 Verification of the FDSM equivalence by simulation To be able to use the presented pattern noise model to describe pattern noise in the FDSM, we must first verify that the FDSM behaves equivalent to a traditional delta-sigma modulator with respect to pattern noise. Sweeping the component 2(fc + kxn )/fclk in Eq. 4.3 over the interval 0-1 is equivalent to sweeping the input x in the traditional modulator over the interval 0-1, as the traditional modulator output can be expressed yn = xn + en − en−1 . A simulated frequency scan of a D flip-flop FDSM with fc =425KHz and fclk =8MHz was carried out. The signal bandwidth was set to 437KHz to match the two other plots in Fig. 4.4, and the input frequency was swept from 0 to 4MHz in 500 steps. In Fig. 4.4 (bottom) the resulting in-band noise-power is shown. The match with the theoretical model is almost exact. The D flip-flop FDSM was simulated in C while the output bit-stream analyzed in Matlab (Appendix B). In Fig. 4.6 the input frequency is swept from 0-12MHz to illustrate the pattern noise picture for multi-bit conversion. By this we can conclude that the model given by Eq. 4.20 and Eq. 4.21 is well suited to describe pattern noise in the FDSM for constant inputs. By constant inputs in the FDSM we refer to constant or unmodulated input frequencies. BW noise power (dB) 20 25 30 35 40 45 0 2 4 6 8 Constant frequency (MHz) 10 12 Figure 4.6: Theoretic DC model for multi-bit conversion fclk = 8MHz, oversampling ratio = 9.14 4.2.4 Valley utilization For systems where the FDSM signal range is small, the range may be located in a pattern noise valley by choosing a proper fc /fclk ratio. As an example, for a specific FM demodulation system the carrier frequency was 425KHz with maximum deviation 550Hz, and a 8MHz system clock was applied. In this case the signal range is given by 4∆f /fclk = 0.00025 or 0.025% of the quantization level spacing illustrated in Fig. 4.5. In Fig. 4.7 a zoom of the quantization level interval is shown where the signal range is illustrated by a short horizontal line on top of the peak/valley landscape. The oversampling ratio is 10 000 and there are peaks at ≈20dB above the white noise level. We also notice valleys far below this level. If the fc /fclk ratio can be properly chosen and fixed we may gain a significantly amount of resolution by locating the range in one of these valleys. 86 BW noise power (dB) In Fig. 4.9 the picture is zoomed down to the size of the dynamic range, and the range is now located over one of the highest peaks in Fig. 4.7. Now the structure of the peak is revealed showing a symmetrical bat shape. The top is ≈20dB higher than the white noise level while the area close to its side walls are >30dB lower than the white noise level. In many systems it is not possible to keep a fixed fc /fclk ratio due to temperature drift and process deviations. In this case the dynamic range may accidentally be located over one of the peaks in the pattern noise landscape, and a significantly performance reduction will result. One way to overcome this problem may be to use a feedback arrangement where the mean value of the decimator output is used to tune either fc or fclk to achieve a proper fc /fclk ratio. Another solution is to use a dither signal to smooth out the pattern noise landscape. In this way the peaks are removed at the cost of shallower valleys as we will see next. 100 110 120 130 140 3.8 3.9 4 4.1 4.2 4.3 4.4 Constant frequency (Hz) 4.5 4.6 4.7 4.8 5 x 10 Figure 4.7: Frequency scan, theoretic model. Zoom of Fig. 4.5 (bottom). Short horizontal line in center/top indicates FDSM dynamic range. Long horizontal line - white noise model. 87 4.3 Time-domain dithering A frequently used technique to reduce the effect of pattern noise in traditional delta-sigma modulators is dithering [3]. There are two main topologies for dithering. I) Adding a pseudo random signal to the input of the modulator. II) Adding a noise shaped pseudo random signal to the input. The first method may reduce the signal to noise ratio significantly while the second method preserves the resolution by adding mainly out of band noise to the modulator. This noise will then be removed by the decimator. Adding noise-shaped dither to the input is equivalent to adding white noise to the quantizer input as illustrated in Fig. 4.8. For the FDSM, this is equivalent to adding white phase noise to the FM (or clock) signal. In applications where it is difficult to locate the signal range in a pattern noise valley, white phase noise can be added to the FM / clock signals to reduce the correlation between the signal and the noise floor. dn xn dn z -1 Q() yn z -1 z -1 Figure 4.8: Two different dithering strategies 100 100 110 110 120 120 BW noise power (dB) BW noise power (dB) In a traditional delta-sigma modulator, the recommended dither magnitude is 1-2LSB of the decimated word. If a maximum-length pseudo random sequence generator [29] is used, a sequence length of > 226 is recommended together with a 8 level quantization of the pseudo random output [3]. These values are also found to be convenient in the FDSM by simulations and measurements. 130 140 130 140 150 150 160 160 170 4.44 4.442 4.444 4.446 Constant frequency (Hz) 4.448 170 4.45 5 x 10 4.44 4.442 4.444 4.446 Constant frequency (Hz) 4.448 4.45 5 x 10 Figure 4.9: Simulated frequency scan. Zoom of Fig. 4.7 down to dynamic range of FDSM, centered over right twin-bat. Horizontal line - white noise model. Left figure: No phase noise added, right figure 0-30ns white phase noise added The same frequency scan that produced the result in Fig. 4.9 left have been repeated for the same modulator, but now a 0-0.73rad pseudo random dither signal is added to the 88 FM phase. The simulation result is shown in Fig. 4.9 right. The bat magnitude is now reduced to maximum ≈0.5dB above the white noise level at the cost of a higher noise level in the neighborhood. In this case the FDSM output noise floor will be almost independent on the modulating signal amplitude and the signal range location which can be a desirable situation. 4.3.1 Measurements To verify the theory a test board was build where the FDSM, decimator and pseudo random generator was implemented in Altera FLEX 10K20 Fig. 4.10. White phase noise was added to the FM signal by the use of a 74AS04 based delay line. By selecting one tapping at a time through a pseudo randomly controlled multiplexer the time shift was approximately uniformly distributed in the 0-30ns interval. A SUN workstation was used to collect and analyze the output data, and the FM signal was made by a GPIB controlled HP33120A signal generator. Signal gen HP33120A Pseudo random generator ALTERA FLEX9000 Matlab sinc2 decimator bus interface and diver SUN workstation GPIB FDSM S16D interface 74AS04 MUX Crystal oscillator Figure 4.10: The FDSM test board In Fig. 4.11 the output psd is shown for two measurements where the FM carrier frequency is 425kHz. The maximum deviation caused by the 67Hz modulating sinusoidal is ∆f =550Hz, and a sampling clock frequency of 8MHz was used. In the left figure no extra phase noise is added to the system. The noise floor is dominated by spurious noise tones which indicates that the quantization noise is far from white. In the right figure a 0-30ns pseudo random phase noise is added to the FM signal. We notice that the noise floor is slightly increased and more smooth as the quantization noise is whitened. In Fig. 4.12 two measured amplitude scans are shown for a carrier frequency fc of 425KHz, and a clock frequency fclk of 12MHz. A number of 75300864 clock samples where taken for each measurement. 1000 measurements are carried out for different constant FM deviations caused by a sinusoidal. As the amplitude of the sinusoidal increases, the S/(N+D) follows. A decimation ratio of 2048 is choosen. The upper graph represent the S/(N+D) ratio where no dithering is applied. For all signal amplitudes the S/(N+D) ratio is well above the white noise model shown by the bottom line. This indicates that the dynamic range is located in a pattern noise walley. For certain signal amplitudes the extremal values of the modulating 89 0 Normalized PSD (dB) Normalized PSD (dB) 0 -50 -100 -150 10 100 Frequency (Hz) -50 -100 -150 1000 10 100 1000 Frequency (Hz) Figure 4.11: Measured result, fc = 425KHz, fclk = 8MHz, ∆f =550Hz, fmax =500Hz. Left figure no extra phase noise is added. Right figure - 0-30ns pseudo random phase noise is added to the FM signal sinusoid are close to a pattern noise peak producing excess noise. This is clearly seen for deviations slightly above 200Hz where the curve drops rapidly. In the next measurement, 0-30ns dither is added to the FM signal. Now the curve is an almost ideal straight line increasing by 20dB/decade with the FM deviation. The most interesting point is where the dithered curve exceeds the non-dithered curve at ≈225Hz. This shows that phase noise can be used to increase the resolution for certain input levels confirming the simulated result in Fig. 4.9. Finally, just to illustrate that the D flip-flop FDSM also can be used in high resolution applications, it is included a measurement where the frequency deviation is set to 2.6MHz and the corresponding FM carrier frequency is raised to 7MHz. The psd of the measured output signal is shown in Fig. 4.13. The measured harmonic distortion is considered to be in the frequency modulator and not in the FDSM, as it corresponds well to the specifications of the FM signal source which is in the -80dB range. By disregarding the 2’th harmonic component, the resolution is found to be almost 20bit in the 0-500Hz signal band. 90 60 55 1000 measurements, each 73536000 fclk cycles sinc2 decimator, dec ratio 2048 50 45 S/N+D (dB) d de 40 ise ad s tep o en as h op 35 s 0n N ise ase 30 m do FM eu 25 Ps 8s 0-3 no ph l de ise an r do in mo o en hit W 20 15 10 1 2 10 10 FM deviation (Hz) Figure 4.12: Measured result. Amplitude (deviation) scan. fc = 425KHz, fclk = 12MHz, fmax =500Hz S/N = 118dB (19.7bit), S/D = 83dB, f0 = 6.4MHz, FMdev = 2 .6MHz, fm = 179Hz, fclk=32.7MHz 0 −20 Normalized PSD (dB) −40 −60 −80 −100 −120 −140 −160 −180 1 10 2 10 Frequency (Hz) 3 10 Figure 4.13: Measured result illustrating high frequency operation. Deviation 2.6MHz, decimation ratio 2048 91 Chapter 5 Conclusion and future work In this thesis four novel first-order FDSM techniques have been proposed and analyzed. Practical effects such as phase noise and pattern noise have been modeled and evaluated. Compared to previously existing FDSM techniques, some of the techniques presented here will increase the digital resolution significantly for certain applications. Other techniques make the FDSM concept more suitable for very high frequency operation by extending the usable range of the D flip-flop FDSM. 5.1 Summary In the following the main contribution of this thesis to the FDSM research field is summarized. Triangularly weighted ZC counting The most straight-forward way to do frequency-to-digital conversion is to count the number of FM zero-crossings during a fixed time interval, and then let this number be the digital representation of the FM frequency. This is done in the classical zero-cross counting frequency discriminator. In this converter the position of each zero-count within the counting time interval do not affect the output signal as the converter employs a uniform window on the input data. In this thesis we have seen that by using a triangularly weighted data window, equivalent first-order ∆-Σ noise-shaping results. In this way the difference between the ordinary zero-cross counting frequency discriminator and the FDSM can be modeled as a difference in an inherent data window, and the understanding of the FDSM operation is simplified. Undersampling In this thesis, we have shown that the D flip-flop FDSM can be operated in FM undersampling mode where the sampling frequency is lower than 2(fc + ∆f ). This means that the D flip-flop FDSM can replace the modulo-2n FDSM in applications where the input FM frequency is very high. Since the maximum input frequency in a FDSM is limited by synchronizer metastability effects, it is important to keep the synchronizing elements as simple as possible. From this point of view the D flip-flop FDSM is an ideal choise as there are no simpler way to synchronize a signal than using a single D flip-flop. In addition, the D 92 flip-flop FDSM provides ≈6dB higher digital resolution than the modulo-2n FDSM as it counts both positive and negative FM signal edges. The operation of the FDSM in undersampling mode is analyzed both in the time and in the frequency domain. In the frequency domain analysis, the D flip-flop FDSM is modeled as a converter digitizing the FM signal, directly followed by a digital demodulator. In other words, the FDSM is now modeled without considering the phase of the FM signal. By doing so we have seen, without proving it mathematically, that pattern noise can be modeled as interference between the FM signal band and higher harmonics of the FM band that are folded down in frequency. The sampled clock FDSM By doubling the sampling frequency in the traditional first-order FDSM, the in-band quantization noise is reduced by ≈9dB. However, the signal range is also reduced by a factor of two. The net gain will therefore only be ≈3dB. In this thesis we have shown that by exchanging the clock and FM input signal to the FDSM, a similar noise-shaping converter results. This converter operates by counting the number of zero-crossings in the clock signal during the time varying FM period, thus the name sampled-clock FDSM. In this converter the SQNR is increased by ≈6dB for each doubling of the constant frequency, and a higher digital resolution can be achieved for applications where the clock frequency is higher than the FM carrier frequency. To analyze this circuit mathematically, the concept of virtual frequencies have been introduced. The sampled-clock FDSM can be realized both as a modulo-2n FDSM, and also as a D flip-flop FDSM. The sampled-clock D flip-flop FDSM have also been analyzed in undersampling mode, both in time and in frequency. The sampled-clock FDSM is inherent non-linear with a 1/(k +x) signal transfer function. The non-linearity is given by the ∆f /fc ratio. For low relative deviations the THD can be in the <-100dB range. For applications with high relative deviations, the non-linearity can be significantly reduced by digital correction since its transfer function is well defined. The ouput data-rate of the sampled-clock FDSM is given by the FM signal frequency and is thus time varying. It have been shown that resampling of such time-varying data-rates are possible [19, 30]. A single experiment carried out in this thesis indicates that it may also be possible to remove most of the non-linearity in the resampling operation. The D flip-flop ring-oscillator FDSM It is shown that the pointer FDSM read-out circuit can be simplified by replacing it with a parallel connection of D flip-flop FDSM’s. In this way the maximum sampling frequency can be somewhat increased, increasing the SQNR. The new circuit is a single-bit FDSM but by replacing the output XOR gate with a binary adder, the concept of parallel conversion is introduced. It is recently proved [24] that flicker noise is reset and thus significantly reduced in a CMOS ring-oscillator. This result combined with the fact that phase-noise is noiseshaped in a FDSM (se next), makes the ring-oscillator FDSM a very interesting candidate for low-noise integrated sensor applications. Parallel conversion With the D flip-flop ring-oscillator FDSM, the concept of parallel conversion was introduced. Parallelisation is based on the use of several signal channels carrying different FM phase information. However, if the input to the converter is a single FM signal, no extra phase information is supplied, but one way to obtain additional phase information is by the use 93 of a open-ended delay line. This solution have been analyzed both theoretically and by simulations. For small relative FM deviations, the open-ended delay-line FDSM can be biased to behave almost exactly like the ring-oscillator FDSM with respect to quantization noise. However, for high relative FM deviations the distribution of the phase information over the channels will be disturbed causing a degradation in the overall SQNR. For a high number of channels where inverter delay mismatch is present, the overall effect of this disturbance will follow a very complex pattern and should rather be modeled by statistical methods. This is not done yet, but by simulations it is found that for a high number of poorly matched inverters, where the total line delay is more than ≈5 times the FM period, the overall SQNR tends to follow the SQNR of the ring-oscillator FDSM with a maximum loss of ≈15dB. From these results, the delay-line FDSM seems like an interesting way to go for increasing the digital resolution in first-order FDSM’s. To verify the analysis, a 1024-inverter delay-line FDSM have been implemented in CMOS VLSI by a collaboration between the author and the ∆-Σ group at IFI. The chip is targeted at 24-bit audio conversion and will soon be returned from the AMS foundry. Phase noise analysis Phase noise is a major problem in many FM demodulation systems. In this thesis the effect of phase noise, both on the FM signal, and also on the clock signal have been analyzed, both analytically, by simulations and by measurements. It is found that phase noise, both on the FM signal and on the clock signal, is first-order shaped by the FDSM. By considering FM signal phase noise, clock signal phase noise and the quantization noise as uncorrelated, a model for the resulting in-band noise power in the demodulated signal is given. As the phase noise is shaped, the resulting output power from a harmonic phase noise component will be dependent on its frequency. Low frequency components in the signal band will be heavily suppressed while out-of-band components will be removed by the decimator. White phase noise may actually improve the performance of the FDSM as it will act as a noise-shaped dither signal and suppress pattern noise. Pattern noise analysis A major problem in traditional first-order ∆-Σ modulators is pattern noise. In this work, we have found that the FDSM is equivalent to a traditional ∆-Σ modulator with respect to pattern noise. By using a well established model for pattern noise [10], we have been able to model pattern noise in the FDSM to a high degree of accuracy. Then we have discussed three strategies that may eliminate this problem. First, we have presented the multi-bit parallel FDSM concept. In a multi-bit ∆-Σ modulator the worst case signal to pattern noise ratio is reduced by the number of bits used. As an example, we have designed a delay-line FDSM with an effective word length of 7-bit. In this modulator problems due to pattern noise is expected to be almost eliminated. Next, for applications where the ∆f /fclk ratio is small, the FDSM internal dynamic range will be small, and by controlling the location of the dynamic range in the quantization level interval, given by the fc /fclk ratio, we can locate the dynamic range in a pattern noise valley. In this way the problem with pattern noise is reversed to an advantage making the effective digital resolution higher than the white quantization noise model predicts. This result is verified both by simulations and measurements. In some applications it will be difficult to maintain a fixed fc /fclk ratio due to various practical effects such as temperature drift, aging and so on. In this case, it can be impossible 94 to locate the dynamic range in a pattern noise valley, and in the worst case, the range can be accidentally located at the top of a pattern noise peak. For these cases, it will be possible to add phase noise to decorrelate the quantization error and the modulating signal as we will see next. Time-domain dithering In this thesis it is shown that adding white noise to the quantizer input in a traditional ∆-Σ modulator is equivalent to adding white phase noise to the FM signal in the FDSM. In this way, the most frequently used dithering technique in the traditional ∆-Σ modulator can be applied to the FDSM. A simple way to add white phase noise to the FM signal, is to feed the FM signal through a delay-line and select one tapping at a time with a multiplexer controlled by a pseudo random sequence. This technique have been verified both by simulations and measurements. 5.2 Conclusion The main aim of this thesis was to develop techniques making it possible to reduce the analog complexity of ∆-Σ converters by the use of first-order FDSM solutions. This aim is met in the following way: 1. Modeling and understanding Through the concept of triangularly weighting and the work done on phase and pattern noise, the behavior of the FDSM is now extensively modeled and understood. This is a necessary requirement for using the FDSM in practical applications. 2. Pattern and phase noise The problem with pattern noise is significantly reduced or eliminated through the multi-bit parallel FDSM concept. Alternatively, time domain dithering and pattern noise valley utilization have been proposed to reduce or eliminate this problem. Phase noise is found to be noise-shaped by the FDSM. By this, it is shown that the first-order FDSM can be used in practical high resolution applications. 3. Resolution By the sampled-clock and the parallel FDSM concept, the resolution of the FDSM is significantly increased making it possible to use the first-order FDSM even in high-bandwidth / high-resolution applications. 4. Extended use By the concept of FM undersampling the usable input frequency range of the most simple D flip-flop FDSM is extended allowing for use in high frequency RF applications. By the D flip-flop ring-oscillator FDSM we have presented an example where a ∆-Σ analogto-digital converter can be made only by standard digital gates. In this case the only analog node in the converter is the power-supply line to the ring-oscillator inverters. As an analogto-digital converter it is not possible to reduce the analog circuitry further. 5.3 Future work The FDSM research field is still in its infancy, and there are many leads to follow, both to increase the performance in already developed solutions, but also by using the FDSM idea in other applications. In the following a brief list of topics are given. 95 Massive parallelism Since quantization in the FDSM is done in time or phase, we have seen that the FDSM can be implemented by a delay line and a parallel collection of D flip-flops plus an adder. Since each D flip-flop FDSM is very compact, the main chip area will be dominated by the adder tree. The adder tree can be significantly simplified by extensive use of modulo adders or XOR gates, and it will be possible to make converters with a very high number of parallel channels. Another interesting feature is that the parallel FDSM is quite fault tolerant, as stuck at 1 or stuck at 0 errors in a few channels will almost not be of no significance at the output since this effect only represents a bias in the output signal. The resulting quantization noise from a delay-line FDSM with a large number of channels, including random inverter delay mismatches, have not been statistically modeled yet. It would be interesting to know more about the noise propertied of a ∆-Σ modulator based on a “sea of D flip-flops”. Low noise integrated sensors As proved in [24], flicker noise is reset and thus significantly reduced in a CMOS ringoscillator. In addition, we have seen that phase-noise is noise-shaped in a FDSM, and this makes the ring-oscillator FDSM very interesting for low-noise integrated sensor applications. This result explains why there was not found any noise floor down to 7.5Hz in the pointer FDSM measurement carried out in [7]. When it comes to transistor thermal noise, we have reason to believe that increasing the number of inverters will decrease the overall output noise as the thermal noise is not correlated from transistor to transistor. By doubling the number of inverters, the overall thermal noise rms value should therefore be decreased by √ a factor of ≈ 2. However, this result is not yet documented but it may be an interesting way to go. Linearity improvement in ADC applications The FDSM concept can be used both in ADC and FDC applications. By replacing a traditional feedback ADC ∆-Σ modulator by a FDSM, the overall linearity of the converter will be given by the linearity of the frequency modulator in the actual frequency range. Deterministic non-linearity given by a non-linear frequency modulator transfer function can be compensated for digitally at low sampling frequency if the transfer function is known. On the other hand, non-linearity given by random variation in process parameters and temperature drift requires a more complicated calibration process before digitally correction can take place. To improve linearity in ADC applications one may therefore try to design the frequency modulator in a way that keeps the random or stochastic part of the non-linearity as small as possible. A differential solution may be a good starting point for this work. New applications Most activity in the ∆-Σ research field today is focused on fine tuning more or less classical SC circuits to squeeze out another bit in resolution. We will not revile this activity as this is outmost important from an industrial point of view. However, from an academical point of view it is more interesting to catch the basic ideas behind noise-shaping. By understanding the underlying basics we may be able to use the technique in other, more distant fields where the name ∆-Σ is still unknown. The FDSM concept have, to some extent achieved this, as it is based on viewing the ∆-Σ process as a cascade of a modulo integrator and a modulo differentiator with a quantizer in between. By peeking into the RF field, the main feature of the ∆-Σ noise shaping idea, the modulo integrator was first found in the standard FM signal represented by the FM phase. 96 When this was done, all we needed was a quantizer (ZC counter or D flip-flop) and a digital modulo differentiator. Then the way to high quality FM demodulators were relatively short. Noise-shaping is actually a more general technique than described here. The very basic idea is to suppress the effect of an interfering signal on another signal. The signal that we want to shelter is first protected from the interference by changing its structure in one or another way. When the interference is over, the structure of the original signal is changed back again. When the structure is changed back, the structure of the interference is changed similarly. So what we are aiming for is to change the structure of the interference to make it less troublesome in the final signal. In low-pass ∆-Σ modulators the structure reversal is done by digital high-pass filtering, and the interference is quantization noise. By using a high-pass filter, the interference in the signal band is reduced. In bandpass ∆-Σ modulators [4, 5] it can be shown that the underlying operation can be described by a cascade of a resonator (bandpass filter) and an anti-resonator (bandstop filter) where the quantization error is entering in between. In this case the final filter suppresses the interference in the signal band that now are located some where up in frequency. So, one way to change the structure of the original signal is by linear filtering, but in principle, the original signal can be protected in many different ways. In the RF field today, there are numerous modulating techniques FSK, QAM, SSB, ASK.. [31]. There have not yet been searched in any other modulation technique for noise-shaping features, but there is reason to believe that there may be something out there. 97 Chapter 6 Publications 6.1 IEE Electronics Letters no. 1 Vol.31, no.2, pp.81-82, 1995 98 99 6.2 ISCAS no. 1 Proceedings IEEE ISCAS’95, pp.175-178 102 103 6.3 IEEE Journal of Solid State Circuits IEEE Journal of Solid State Circuits, vol.32, no.1, pp.13-22, January 1997 108 109 6.4 ISCAS no. 2 Proceedings IEEE ISCAS’97, pp.77-80 120 121 6.5 IEE Electronics Letters no. 2 Vol.33, no.13, pp.1121-1122, 1997 126 127 6.6 Low Power Proceedings on the 1997 International Symposium on Low Power Electronics and Design, Monterey, CA, USA, pp.52-55, August 1997 130 131 6.7 NorChip Proceedings 17’th IEEE NorChip’99, pp.391-398 136 137 Appendix A PSD definitions The following definitions are used througout this thesis. The Wiener Khintchine theorem Sx (f ) = rxx (k)e−jk2πf (A.1) k The total power definition Ptot = f clk /2 1 fclk Sx (f )df (A.2) −fclk /2 White noise Sx (f ) = σx2 (A.3) Ptot = σx2 (A.4) Harmonic component A sin(2πfm n) Sx (f ) = fclk A2 [δ(f − fm ) + δ(f + fm )] 2 Ptot = , −fclk fclk <f < 2 2 A2 2 (A.5) (A.6) Sx (f ) estimation The modified periodogram estimator 1 Ŝx (f ) = NU N −1 2 −jn2πf wn xn e n=0 148 (A.7) U= N −1 1 |wn |2 N n=0 (A.8) The expected value E{Ŝx (f )} = 1 Px (f ) ∗ |W (f )|2 2πN U (A.9) Ŝx (fm ) versus Ptot Ŝx (fm ) = Ptot = Ptot N − loss 4 (A.10) 4 (Ŝx (fm ) + loss) N (A.11) Loss: Rectangular window: 0dB, Hamming: 1.6dB, Blackman: 2.4dB, BHH: 4.2dB. The Blackman-Harris-Hoodie window wk = w0 − 6 wm cos m=1 w0 =0.6164032131405 w1 =0.98537119272586 w2 =0.49603771622007 w3 =0.14992232793243 w4 =0.02458719103474 w5 =0.0017660465148687 w6 =0.000031581188567106 149 2mπk N (−1)m (A.12) Appendix B C simulators 150 159 Bibliography [1] C.C. Cutler: “Transmission synstem employing quantization”, U.S. Patent No. 2.927.962, filed 1954 [2] H. Inose, Y. Yasuda and J. Murakami: “A telemetering system code modulation ∆-Σ modulation”, IRE Tranaction on Space Electronic Telemetry, vol. SET-8, pp.204-209, September 1962 [3] Steven R. Norsworthy, Richard Scheier and Gabor C. Temes: Delta-Sigma Data Converters, IEEE Press, New York, pp.1-43, 1997. [4] P.H. Gailus, W.J. Turney and F.R. Yester Jr.: “Method and arrangement for a sigma delta converter for bandpass signals”, U.S. Patent No. 4.857.928, filed Januar 28, 1988, issued August 15 1989 [5] Richard Schreier and W.M. Snelgrove: “Bandpass sigma-delta modulation”, Electronics Letters, vol. 25, no. 23, pp 1560-1561, November 1989 [6] Crystal Seminconductors Corporation, Austin Texas, USA, http://www.crystal.com, “7’th order 24-bit audio ∆-Σ ADC”, On the marked in 1998 [7] Mats Høvin: Novel ∆ − Σ modulators using frequency modulated intermediate values. Master thesis, Departments of Informatics, University of Oslo, 1995 [8] Ian Galton, Wiliam Huff, Paolo Carbone and Eric Siragusa: “A Delta-Sigma PLL for 14-b, 50kSample/s Frequency-to-Digital Conversion of a 10MHz FM Signal”, IEEE Journal of Solid State Circuits, vol. 33, no. 12, December , 1998 [9] Thomas A. D. Riley, Norman M. Filiol, M.A. Copeland and Calvin Plett: “A twoloop third-order multistage Delta-Sigma frequency-to-digital converter”, Proc. IEEE ISCAS’98, vol.1, pp.408-411 [10] James C. Candy and Oconnell J. Benjamin: “The Structure of Quantization Noise from Sigma-Delta Modulation”, IEEE Trans. Communication, vol. COM-29, pp. 1316-1323, 1981. [11] B.H. Leung and S. Sutarja: “Multi-bit ∆-Σ A/D converter incorporating a novel class of dynamic element matchning”, IEEE Transaction on Circuits and Systems II, vol. 39, pp35-51, January 1992 [12] Y. Sakina: “Multi-bit ∆-Σ Analog-to-Digital Conversion with Nonlinearity Correction using Dynamic Barrel Shifting”, Electronics Research Lab, College of Engineering, University of California, Berkely CA, Memorandum No. UCB/ERL M93/63, 1993 160 [13] R. Schreier and B. Zhang: “Noise-shaped multi-bit D/A converter employing unit elements”, Electronics Letters, vol. 31, no. 20, pp. 1712-1713, September 1995 [14] L.E. Larsen, T. Caltepe and G.C. Temes: “Multi-bit oversampled ∆-Σ A/D converter with digital error correction”, Electronics Letters, vol. 24, pp. 1051-1052, August 1988 [15] T.C. Leslie and B. Singh: “An improved sigma-delta modulator architecture”, IEEE Proceedings ISCAS’89, vol.1, pp.372-375, May 1990 [16] J.B. Hughes and W.R. White: “Switched-current limitations and non-ideal behaviour”, in C. Toumazou, J.B. Hughes and N.C. Battersby (Ed.): Switched-current, an analog technique for digital technology, P. Pereginus Ltd. pp. 71-75, 1993 [17] S. Haykin: Communication systems, John Wiley and Sons, 1987, pp.179-183 [18] R. Steele: Delta Modulation Systems, John Wiley and Sons, 1975 [19] Alan V. Oppenheim and Ronald W. Scafer: Discrete-time signal processing, Prentice Hall Inc, New Jersey, 1989 [20] H. Reichl, H.J. Hwang and H. Riedel: “Frequency-Analog-Sensors using the i2 L Technique”, Sensors and Actuartors, vol. 4 pp.247-254, 1983 [21] L.S. Kim and R.W. Dutton: “Metastability of CMOS Lath/Flip-flop”, IEEE Journal of Solid State Circuits, vol. 25, no. 4, August 1990 [22] J.U. Horstman, H.W. Eichel and R.L. Coates: “ Metastability behavious of CMOD ASIC Flip-flops in theory and test” IEEE Journal of Solid State Circuits, vol. 24, no. 1, Februar 1989 [23] S. Chu and C. S. Burrus: “Multirate Filter Designs using Comb Filters”, IEEE Trans. on Circuits and Systems, vol. CAS-31, pp. 913-924, Nov. 1984. [24] Bram Nauta, Eric Klumperink, Sander Gierkink and Arnoud vd Wel: “Reduction of 1/f Noise in MOSFETSs by Switched Bias Techniques”, Proceedings NorChip’99, pp.60-72 [25] M. Morris Mano: Digital Design, Prentice Hall Inc, New Jersey, 1991 [26] Matlab reference guide, the Math Wwork Inc, 1993 [27] N. Weste and K. Eshraghian: Principles of CMOS VLSI Design, Addison Wesley, 1988 [28] K. Lakshmikumar, M. Copeland and R. Hadaway: “Characterization and Modelling of Mismatch in MOS Transistors for Precision Analog Design”, IEEE Journal of Solid State Circuits, vol.SC-21, pp.1057-1066, December 1989 [29] Frank F. Tsui: LSI / VLSI Testability design, McGraw-Hill Book Company, New York, 1986. [30] Wiliam Huff and Ian Galton: “Nonuniform-to-uniform decimation for delta-sigma frequency-to-digital conversion”, 1998 [31] Bernard Sklar: Digital Communication, Prentice Hall Inc, New Jersey, 1988 [32] James C. Candy: “Decimation for Sigma-Delta Modulation”, IEEE Trans. Communication, vol. COM-34, pp. 72-76, 1986. 161 [33] J. Yuan and C. Svensson: “High-Speed CMOS Circuit Techique”, IEEE Journal of Solid State Circuits, vol. 24, February, 1989. [34] M. H. Hayes: Statistical digital signal processing and modelling, John Wiley and Sons, Inc, New York, 1996. 162