Next Article in Journal
Optoelectronic Angular Displacement Measurement Technology for 2-Dimensional Mirror Galvanometer
Next Article in Special Issue
Validation of Embedded State Estimator Modules for Decentralized Monitoring of Power Distribution Systems Using IoT Components
Previous Article in Journal
The Large-Scale Physical Model Tests of the Passing Ship Effect on a Ship Moored at the Solid-Type Berth
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

An 11.8-fJ/Conversion-Step Noise Shaping SAR ADC with Embedded Passive Gain for Energy-Efficient IoT Sensors

Information and Communication System-on-Chip (SoC) Research Center, School of Electronics and Information, Kyung Hee University, Yongin 17104, Korea
*
Author to whom correspondence should be addressed.
Sensors 2022, 22(3), 869; https://doi.org/10.3390/s22030869
Submission received: 10 December 2021 / Revised: 28 December 2021 / Accepted: 21 January 2022 / Published: 24 January 2022
(This article belongs to the Special Issue Embedded Systems for the New Generation of IoT Sensors)

Abstract

:
Herein, we present a noise shaping successive-approximation-register (SAR) analog-to-digital converter (ADC) with an embedded passive gain multiplication technique. The noise shaping moves the in-band quantization noise from the signal band to out-of-band for improved signal-to-noise ratio (SNR). The proposed approach tackles the drawback of the previous active noise shaping (increased power and extra noise) and passive noise shaping (limited noise suppression and signal loss). Both noise shaping and gain multiplication are realized on-chip in an energy-efficient manner without an opamp. This approach uses only capacitors and switches in the finite impulse response (FIR) and infinite impulse response (IIR) filters. A comparator suppressing kickback noise is presented to handle the tradeoff between noise suppression and the filter capacitor size. The energy-efficient merged-capacitor switching (MCS) technique is effectively combined with rail-to-rail swing comparator and thermometer-coded capacitor array, which reduces the settling error in the digital to analog converter (DAC). The process-induced mismatch effect in the capacitive DAC is investigated using a behavioral model of the ADC. Additionally, we propose dynamic element matching (DEM) for the thermometer-coded capacitor array. The ADC is fabricated using a 0.18 μm CMOS process in an area of 0.26 mm2. Consuming 4.1 μW, the ADC achieves a signal-to-noise and distortion ratio (SNDR) of 66.5 dB and a spurious-free dynamic range (SFDR) of 79.1 dB. The figure-of-merit (FoM) of the ADC is 11.8 fJ/conversion-step.

1. Introduction

Demands for energy-efficient applications, such as the Internet of Things (IoT), battery-operated sensors, and wearable electronics, are continuously increasing. Ultra-low power consumption is required in these systems for signal sensing and processing to provide a long battery life. An analog-to-digital converter (ADC) is a key component in the processing of sensor output [1,2,3] and wireless communication [4,5]. Among various ADCs, successive approximation register (SAR) ADC is suitable for achieving high energy efficiency with low power consumption [6].
Typical SAR ADC consists of a digital-to-analog converter (DAC) realized using a capacitor array, a comparator, and SAR logic. The digital output for the analog input is obtained through charge redistribution in the capacitive DAC (CDAC). The SAR ADC provides medium resolution using very low power since the clocked comparator and capacitive switching consume only the dynamic power. One drawback of the SAR ADC is that the area of the CDAC needed to realize the binary weight increases rapidly with the resolution. When the number of CDAC bits is increased for high resolution, routing becomes more complicated in the SAR ADC. Additionally, the comparator’s input-referred noise and quantization noise limit ADC performance; designing high-resolution SAR ADC with low complexity is a challenging task.
The noise shaping technique has been actively investigated to address the challenge [7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22]. This technique moves the in-band quantization noise from the signal band to out-of-band for improved signal-to-noise ratio (SNR). The number of capacitors in the DAC can be reduced using noise shaping, simplifying the practical implementation of the SAR ADC. The previous work on the SAR ADC realizes the noise shaping filter using opamp and achieves a 10-bit effective number of bits (ENOB) using 8-bit CDAC [7]. The filter consists of finite impulse response (FIR) and infinite impulse response (IIR) filters. This approach shows that a relatively good noise shaping can be achieved even with a low-quality integrator for the IIR filter.
The residue remaining on the DAC after completing the digital conversion is the difference between the sampled input and a digital estimate. An opamp is used to process this small voltage [7,9,10]; the opamp consumes static power and introduces extra noise. A dynamic amplifier is used for the noise shaping filter to handle this issue [11,12]. A dynamic structure realizing the passive FIR and IIR filters can remove the static power consumption; however, the gain of the dynamic amplifier can be sensitive to supply voltage and temperature, and additional calibration may be needed [12]. Additionally, power consumption using this approach is still high, for example, 460 [11] and 84 μW [12]. Alternatively, a voltage–time–voltage converter can be used to achieve process-insensitive active residue processing [8]. Because there are two signal components, DAC output and filtered residue at the comparator input, the comparator with multi-input pairs is used [7,15,16,17]. To handle the small residue, the differential input pair for the residue is sized larger than the one other receiving the DAC output. This approach provides the advantage of the increased gain for processing the residue; however, the kickback noise of the comparator is proportionally increased with the size of the input pair (or the capacitance). Additionally, a multi-input comparator increases the input-referred noise.
The passive residue summation using a single input pair can be an alternative solution [18]; however, this approach achieves relatively weak suppression of the in-band quantization noise, and signal loss problems remain. In work [15], two capacitors added in the integration path increase the zero of the noise transfer function (NTF) to 0.75; however, the capacitor performing the residue sampling is reset after each conversion cycle, degrading the integration effect. Therefore, the previous approaches suffer from the tradeoff between gain, kickback, and input-referred noise. These results indicate that the noise shaping technique suitable for simple and power-efficient SAR ADC has not been fully investigated.
This paper proposes a simple and power-efficient noise shaping technique, which reduces the number of capacitors in the DAC. We embed a charge pump in the filter for passive gain multiplication to deal with the residue attenuation in the previous passive noise shaping. This approach uses only capacitors and switches in the FIR and IIR filters. Thus, noise shaping and gain multiplication are realized on-chip in an energy-efficient manner without an opamp. To handle the tradeoff between noise suppression and chip area, we propose a comparator canceling the kickback noise. The energy-efficient merged-capacitor switching (MCS) technique is effectively combined with the rail-to-rail comparator and the thermometer-coded capacitor array, which reduces the settling error in the DAC. The process-induced mismatch effect in the CDAC is investigated using a behavioral model of the ADC, and we propose a dynamic element matching (DEM) technique for the noise-shaping ADC. The proposed ADC fabricated in 180 nm CMOS demonstrates that the passive noise shaping technique enables ADC operation with an effective number of bits (ENOB) of 10.8-bit using a 9-bit CDAC. The measured result shows a significant improvement in the signal-to-noise and distortion ratio (SNDR) and spurious-free dynamic range (SFDR). Consuming 4.1 μW, the ADC achieves an SNDR of 66.5 dB and an SFDR of 79.1 dB with a figure-of-merit (FoM) of 11.8 fJ/conversion-step.

2. Design

2.1. ADC Operation

Figure 1 shows the functional signal-flow diagram of the proposed ADC. After the sampling and conversion, the residue VRES, which is the difference between the analog input Vin and the digital estimate Dout, remains on the top plate of the CDAC. VRES is integrated by the FIR and IIR filters. The ADC feedforwards Vin to the quantizer, and the integrated residue VINT is added with the Vin to generate Dout [9]. Considering the quantization noise QN and the comparator noise VN,COMP, Dout can be expressed as
D out = V in + 1 1 + L ( z ) ( Q N + V N , COMP )
where L(z) = VINT(z)/VRES(z) is the filter transfer function. Using the proper NTF = 1/[1+ L(z)], both QN and VN,COMP can be reduced at the expense of bandwidth. Because VRES is less than one least significant bit (LSB), proper processing of VRES is important to achieve noise shaping. In this work, VRES is boosted by passive multiplication inside the FIR filter. The multiplication is realized using the capacitive charge pumping. Switches are controlled to sample VRES in parallel, and the connection is changed to series to achieve the multiplication of n, which is the number of FIR capacitors. The IIR filter is realized using a single capacitor for integrating the output of the FIR filter.
Figure 2a shows the block diagram of the proposed ADC. Top-plate sampling is performed using a bootstrapped switch [23]. The MCS technique is used for the DAC, chosen for its high energy efficiency and constant common-mode (CM) operation [24]. To realize noise shaping, the FIR and IIR filters are located between the CDAC and the comparator. The integrated residue is handled using the residue-summation technique [18], which allows processing the residue using the comparator having a single input pair.
Figure 2b shows the schematic of the proposed ADC with the related timing waveforms. The VDAC,p and VDAC,n are the top plate voltages of the positive and negative DAC, respectively. The settling error in the DAC can be reduced using the thermometer coding, which is used for the upper 3-bit. Binary coding is used for the remaining 6-bit; the DAC consists of seven thermometer-coded capacitor array Ci (i = 6 to 12) and six binary-weighted array Cj (j = 0 to 5). We note that the seven thermometer-coded elements represent a 3-bit binary code. Therefore, overall DAC consists of a 9-bit. When the comparator determines the LSB, the result of the last decision (ninth decision) is not fed back to the DAC. This operation explains why the 9-bit DAC generates the digital output having 8-bit accuracy. Additionally, a residue remains at the top plate of the DAC, which is the difference between the sampled input and an 8-bit digital estimate [7]. Each FIR filter consists of residue sampling capacitors CRES. Each IIR filter consists of an integrating capacitor CINT. The sampling clock CLKS is used for the bootstrapped switch, and the ADC operates synchronously with the clock signal CLK. After sampling and conversion operations are performed, the noise shaping (NS) cycle follows. Residue processing is performed using the two-phase signals, ΦRES for residue sampling and ΦINT for residue integration.

2.2. Noise Shaping Operation

Figure 3a shows the sampling and conversion operations in the (k − 1)th cycle. During conversion, VDAC,p and VDAC,n change around the CM voltage VCM. After the digital conversion, there are residue voltages, VRES,p and VRES,n, on the positive and negative DAC, respectively, which is the difference between the sampled input and an 8-bit digital estimate. The previous conversion cycle sets the voltage VINT[k − 1] across CINT. Switches are controlled to connect CRES in series, which is in parallel with CINT. Then, the voltage VCRES[k − 1] across each CRES is VINT[k − 1]/n. Here, n is the number of CRES, and the case of n = 3 is shown.
Figure 3b shows the operation during the NS cycle when ΦRES is high (ΦINT is low). During this time, the residue is captured. The residue 2VRES = (VRES,pVRES,n) on the top plate of the differential DAC is sampled on CRES. At this time, six CRES are connected in parallel with the DAC. The VRES is transferred from the DAC to CRES by charge redistribution. Therefore, VRES is scaled by a factor α, which is the ratio of CDAC and nCRES as
α = C DAC C DAC + n C RES
where CDAC is the sum of the DAC capacitors. To obtain VCRES, we need to consider another charge from CINT. In the previous cycle, we have VCRES[k − 1] = VINT[k − 1]/n. Considering that the charge from CINT is shared between nCRES and CDAC, we can express VCRES as
V CRES [ k ] = C DAC C DAC + n C RES 2 V RES [ k 1 ]   + n C RES C DAC + n C RES V INT [ k 1 ] n = 2 α V RES [ k 1 ]   + ( 1 α ) V INT [ k 1 ] n
The first term considers the charge transferred from CDAC to CRES. The second term accounts for the charge sharing between nCRES and CDAC, occurring when the charge stored in CINT is transferred to nCRES.
Figure 3c shows the operation during the NS cycle when ΦRES is low (ΦINT is high). During this time, both voltage multiplication and residue integration are performed. After the residue capture, switches are controlled to connect nCRES in series. Then, VCRES is charge pumped and multiplied by n. The boosted voltage is scaled by the factor β, which accounts for the charge sharing between n series-connected CRES and CINT as
β = ( 1 / n ) C RES C INT + ( 1 / n ) C RES .
The integration with a gain of β is performed during the high ΦINT cycle. By adding the value VINT[k − 1] of the previous cycle, which is charge shared between CINT and CRES/n, we can express VINT[k] of the kth cycle as
V INT [ k ] = C INT C INT + ( 1 / n ) C RES V INT [ k 1 ]   + ( 1 / n ) C RES C INT + ( 1 / n ) C RES n V CRES [ k ] = ( 1 β ) V INT [ k 1 ] + β n V CRES [ k ] .
Using (3) and (5), we obtain
V INT [ k ] = ( 1 α β ) V INT [ k 1 ]   + 2 n ( α β ) V RES [ k 1 ]   .
The L(z) is obtained by rearranging (6) as
L ( z ) = 2 n α β z 1   1 ( 1 α β ) z 1 .
After the integration is finished during the NS cycle, the next kth cycle for the sampling and conversion starts. At this time, the integrated residue is added to the CDAC at the comparator input.
Figure 4 shows the flowchart of the proposed ADC operation. After sampling the analog input, the DAC is determined by the binary search algorithm. Using the comparator output, the DAC switch is connected to either VREF or gnd, repeated seven times for the thermometer-coded capacitor array Ci (i = 6 to 12) and six times for the binary-weighted array Cj (j = 0 to 5). Then, the noise shaping cycle follows, consisting of one cycle for residue capture (ΦRES = high) and another cycle for residue integration (ΦINT = high). After the NS cycle, the integrated residue is added to the CDAC at the comparator input during the next kth cycle for the sampling and conversion.

2.3. Analysis of Noise Suppression

Using (7), we obtain the NTF as
NTF = 1 ( 1 α β ) z 1 1 + [ ( 2 n + 1 ) α β 1 ] z 1 .
Using the magnitude of NTF, we obtain the in-band quantization noise reduced by noise shaping. Considering the tradeoff between the chip area and the passive gain, we investigate the two cases of n = 2 and n = 3. Figure 5 shows the noise suppression calculated at (fS/fin) = 0.1 (See Figure 6). Here, fS is the sampling frequency of CLK, and fin is the input frequency. The result shows the improved noise suppression (2–3 dB) achieved using n = 3.
The noise suppression increases with α and β values; however, it saturates with increased values. When we consider the residual kickback from a clocked comparator, the size of CRES cannot be reduced (increased α) to an arbitrarily small value. For the given CDAC, the kickback effect on VDAC increases with α. Additionally, the stability condition (the pole of NTF should be inside the unit circle in the z-domain) sets the upper limit for α and β values. Because CRES is fixed by the selected α value, CINT is reduced with increasing β. When CINT is reduced, the kickback noise increases. Additionally, CINT should be sized considering the kT/C noise [7] and the charge sharing with the CDAC. Because there is no external charge supplied into the passive filter, the tradeoff is inherent in the ADC based on passive noise shaping. Using circuit simulations, we investigate the kickback noise and choose n = 3, α = 0.7, and β = 0.3 so that the noise is less than 0.5 LSB. Noise suppression up to 15 dB is achieved at low fin using these parameters.
Table 1 shows the various NTF expression and the calculated noise suppression. Figure 6 shows the comparison of the magnitude of NTF. The result shows that our approach achieves 7.23 dB and 3.81 dB better noise suppression than the previous works [13,18], respectively.
Using (6), we implement the behavioral model of the noise shaping ADC, as shown in Figure 7. The charge pump is modeled using an amplifier with a gain of n. Comparator and kT/C noise are not considered as they experience the same NTF as the quantization noise [7]. Effect of process variations in the CDAC can be considered by including random mismatch rate. Simulations are performed to investigate the performance improvement by the proposed noise shaping technique. Figure 8 shows the output spectrum of the proposed ADC obtained from the fast Fourier transform (FFT) spectrum with 8192 points. The result confirms the first-order noise shaping achieved by the proposed method. When noise shaping is enabled, the SNR and SNDR increase by 7.2 and 9.2 dB, respectively.
The performance improvement by noise shaping can be affected by the CDAC mismatch. We investigate the random mismatch effect in the CDAC using the behavioral ADC model. Figure 9 shows the probability distributions of the ENOB for different CDAC mismatch rates obtained by 1000 Monte Carlo simulations. When the mismatch increases from 1% to 2%, the average ENOB decreases from 11.5 to 10.8 bits. Considering the mismatch effect, we determine the unit capacitor size (C0 = 21 fF) to keep the mismatch less than 1%. The linearity characteristics affected by the CDAC mismatch can be further improved using foreground calibration [17]. Figure 10a shows the output spectrum of the previous work [18], which uses a 13-bit DAC (10-bit CDAC, 2-bit for redundancy, and 1-bit for noise shaping). Because of additional capacitor switching for noise shaping, three extra cycles are needed for A/D conversion. The results are obtained from the FFT spectrum with 4096 points. Figure 10b shows the output spectrum of the proposed work, which uses a 9-bit CDAC and a passive filter. Our work uses only one additional clock for A/D conversion. Compared to the previous work [18], our work achieves increased zero value in the NTF. The results show that our work using 1-bit smaller DAC achieves increased SNR and SNDR by 3.2 and 4.9 dB, respectively.

2.4. Comparator for Reduced Kickback

The previous work uses cascoding transistors to reduce the kickback noise [25]. Because the comparator is designed for the monotonic switching algorithm for the SAR ADC, it is implemented with a PMOS differential input pair. When the MCS algorithm is used, the VCM of the DAC is fixed during the conversion. When the previous comparator is used for MCS, it can result in a relatively large offset at the input of the comparator, especially during LSB conversion. Figure 11 shows the schematic of the comparator used in this work. The cascoding transistors are removed, and complementary differential input pairs are used, allowing rail-to-rail input range. We note that the comparator does not have a separate input pair for the residue. The proposed comparator uses two clock signals, CLK and CLKB. Consider the VDAC,n on the negative branch DAC, connected to the negative terminal V- of the input pair. The CLK and CLKB signals generate two kickback noise components. Because CLKB is an inverted signal of CLK, the kickback noise in VDAC,n(CLKB) is the inverted version of the noise in VDAC,p(CLK). Because the complementary input pair generates the two kickback noise in opposite directions, they can be canceled out. Similarly, the kickback noise on VDAC,p connected to the positive terminal V+ of the input pair is canceled. The residual kickback noise depends on capacitance matching between the two signal paths.

3. Measured Results

Figure 12 shows the microphotograph of the ADC fabricated in the 0.18 μm CMOS process. The core area is 0.26 mm2. The overall power consumption is 4.1 μW, including 1.2 μW for the reference buffers. Analog, digital, and SAR logic consume 82.4%, 9.3%, and 8.3%, respectively. The measurement setup is also shown. The power supplies for the analog and digital blocks of the ADC are separated. They are stabilized using 1000 μF bypass capacitors and low-dropout (LDO) regulators. A field-programmable gate array (FPGA) board collects the ADC output.
Figure 13a shows the measured output spectrum using fin = 1.33 kHz and fS = 52 kS/s. The result is obtained from the FFT spectrum with 8192 points. The peak SNDR, SFDR, and ENOB are 66.5, 79.1, and 10.8 bits, respectively. Figure 13b shows the measured output spectrum at increased fin = 8 kHz and fS = 180 kS/s. Figure 14a shows the measured SNDR and SFDR as a function of fS. The result shows that the dynamic ADC performance is relatively constant, up to 180 kS/s. Figure 14b shows the measured SNDR and SFDR as a function of fin for two sampling rates. The result shows that the dynamic performance gradually increases with the oversampling ratio (OSR). Figure 15 shows the measured dynamic range at fin = 1.33 kHz and fS = 52 kHz. Peak SNR and SNDR are measured with an input amplitude of −0.4 dBFS. Figure 16 shows the static linearity of the ADC. The result is obtained using a histogram test of 260,000 samples. The peak differential non-linearity (DNL) is +1.34/−1.05 LSB, and the peak integral non-linearity (INL) is +0.89/−0.96 LSB. Because the capacitors in the IIR and FIR filter are dynamically reconfigured, the exact binary weight condition cannot be satisfied for the CDAC. The result indicates the tradeoff in the design of the noise shaping ADC; the static performance is traded for improved dynamic performance.
The mismatch in the CDAC can affect the ADC linearity, and the DEM technique can be used to address the issue [9,10,19]. Either random or cyclic selection can realize the DEM. The cyclic selection uses the output of each conversion determined by the cumulated sum of the elements that are cyclically selected [26]. Two building blocks are usually used [27]. The first is the pointer that indicates the unit element used as the starting point for the DAC operation. The second is a decoder that maps the relationship between the thermometer-code and DAC unit elements. The pointer can be realized using an accumulator and a register. To reduce the implementation complexity, we use a binary counter to implement the pointer. Because the mismatch effect increases with the capacitor size, the DEM is used for the thermometer-coded capacitor array [23]. The binary-weighted arrays are not used for DEM; this approach requires sufficient intrinsic linearity for binary-weighted capacitors.
Figure 17a shows the block diagram of the noise-shaping ADC with the DEM logic. The thermometer-coded capacitor arrays are controlled using the output VD [6:0] of the 3 to 7 decoder. A binary counter, clocked by the comparator output CMP_OUT, is used as a pointer that determines the unit capacitor in the DAC. When CMP_OUT becomes high, the pointer is increased. The decoder receives the 3-bit output from the counter and decides the connection sequence of the thermometer-coded capacitors. The DEM is enabled only for seven clocks after input sampling. For this reason, we use a separate DEM control logic instead of the SAR logic. Figure 17b shows the related timing waveform. The CLK_DEM is enabled when CMP_OUT becomes high, increasing the pointer. The rising edge of the decoder output VD [6:0] triggers the DEM control logic to switch the bottom plate of the capacitors.
We implement the behavioral model of the noise-shaping ADC with the DEM logic. Figure 18 shows the dynamic performance of the ADC with and without DEM, obtained using a 1% CDAC mismatch. Without the DEM, the third harmonic level is located at around −67 dB, which is reduced to −84 dB using the DEM. Figure 19 shows the static performance with and without DEM. A total of 260,000 samples are used. The peak DNL is +0.66/−0.61 LSB, and the peak INL is +0.4/−0.61 LSB without the DEM. Using the DEM, peak DNL is reduced to +0.47/−0.62 LSB, and the peak INL is reduced to +0.25/−0.42 LSB. The results show that the linearity of the noise-shaping ADC can be improved using the DEM.
Table 2 shows the comparison with the previous works. Schreier’s figure-of-merit (FOMs) is defined as
FOM S = SNDR + 10 log 10 ( BW / Power )   [ dB ]
where the bandwidth is defined as BW = fS/(2∙OSR). Walden’s figure-of-merit (FOMW) is defined as
FOM W = Power 2 ERBW 2 ENOB   [ J / conv . ]
where effective resolution bandwidth (ERBW) is approximately half of the sampling frequency. The work [9] achieves a relatively good performance using the DAC mismatching error shaping. The SNR is increased from 69 to 97.9 dB using a relatively high OSR = 512; however, the opamp in the noise shaping filter consumes static power, leading to a relatively low FOMW. All the works except ours [12,18] use a multi-path comparator having an additional input pair for residue processing. The increased input-referred noise of the comparator can limit the achievable ADC performance [7]. The authors of [10,12,18] use 28, 40, and 14 nm CMOS processes and achieve a FOMW better than ours; however, the power consumption of the SAR ADC usually decreases with the CMOS process scaling. Therefore, direct comparison is difficult. The DEM technique addresses the mismatch problem [9,19]. These works show slightly better FOMS than ours, while our work achieves better FOMW. The work [19] uses the passive noise shaping filter; however, the comparator having three input branches increases the power and noise. Works [16,17,18,19,20] consume power > 100 μW, and it is difficult to use these works for the IoT demanding an ultra-low power.
Realized using the noise shaping filter with passive gain multiplication, the proposed ADC consumes the lowest power of 4.1 μW, leading to a favorable FOMW of 11.8 fJ/conversion-step. Our work presents the effectiveness of the DEM using a behavioral model, which can further increase SNDR. The result shows that the proposed approach of noise shaping is promising for improving the performance of the SAR ADC. Although the proposed ADC achieves a moderate FoMS, low power consumption at a medium conversation rate is suitable for IoT. The FoMS can be further enhanced by implementing a more advanced CMOS process. There are many application scenarios of the proposed ADC since sensing analog signals is necessary for various IoT systems. For the sensor interface in these applications, very low power consumption is required to provide a long battery life. Examples include various battery-operated sensing systems [28], deployed in various biomedical, home, industrial, and environment monitoring objects.

4. Conclusions

We propose a noise-shaping SAR ADC featuring a passive gain multiplication technique and successfully verify the approach using a chip fabricated in a 0.18 μm CMOS process. We embed the charge pump in the noise shaping filter to boost the gain without static power consumption, which effectively deals with the residue voltage attenuation. The proposed approach consists of a few capacitors and switches, allowing noise shaping implemented with low power and small area. We present the comparator with reduced kickback noise that effectively handles the tradeoff between noise suppression and chip area. The energy-efficient MCS technique is effectively combined with thermometer-coded CDAC, which reduces the settling error in the DAC. The effect of filter capacitor size and process-induced mismatch in the CDAC is investigated using a behavioral model of the ADC. Additionally, we propose a simple DEM implementation, confirmed using the behavioral simulations. The ADC is fabricated using a 0.18 μm CMOS process. Measured data show the successful operation of the proposed noise shaping technique. The ADC achieves measured SNDR of 66.5 dB and SFDR of 79.1 dB with FoM of 11.8 fJ/conversion-step. The main contribution of this paper is validating a simple and power-efficient noise shaping technique for the SAR ADC using the embedded passive gain multiplication. The proposed approach tackles the drawback of increased power and extra noise of the active noise shaping and limited noise suppression of the passive noise shaping. Future research direction will be implementing the SAR ADC using an advanced CMOS node to increase the bandwidth. Experimental validation of the proposed DEM is also demanded. The result will be useful for realizing a power-efficient SAR ADC for various IoT sensor systems.

Author Contributions

C.C. designed the ADC, performed the experimental work, and wrote the manuscript. J.-W.L. conceived the project, organized the paper content, and edited the manuscript. Corresponding author: J.-W.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported in part by the Basic Science Research Program through the National Research Foundation of Korea (No. 2021R1A2B5B01001475) and in part by National R&D Program through the National Research Foundation of Korea funded by the Ministry of Science and ICT (No. 2020M3H2A1076786). The chip fabrication and CAD tools were supported by the IDEC (IC Design Education Center).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Talens, J.B.; Pelegri-Sebastia, J.; Canet, M.J. Low complexity system on chip design to acquire signals from MOS gas sensor applications. Sensors 2021, 21, 6552. [Google Scholar] [CrossRef]
  2. Ro, D.; Um, M.; Lee, H.-M. A soft-error-tolerant SAR ADC with dual-capacitor sample-and-hold control for sensor systems. Sensors 2021, 21, 4768. [Google Scholar] [CrossRef] [PubMed]
  3. Lee, S.; Jin, J.; Baek, J.; Lee, J.; Chae, H. Readout integrated circuit for small-sized and low-power gas sensor based on HEMT device. Sensors 2021, 21, 5637. [Google Scholar] [CrossRef] [PubMed]
  4. Seong, K.; Jung, D.-K.; Yoon, D.-H.; Han, J.-S.; Kim, J.-E.; Kim, T.T.-H.; Lee, W.; Baek, K.-H. Time-interleaved SAR ADC with background timing-skew calibration for UWB wireless communication in IoT systems. Sensors 2020, 20, 2430. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Cho, S.; Park, D. Robust intra-body communication using SHA1-CRC inversion-based protection and error correction for securing electronic authentication. Sensors 2020, 20, 6056. [Google Scholar] [CrossRef]
  6. Shehzad, K.; Verma, D.; Khan, D.; Ain, Q.U.; Basim, M.; Kim, S.J.; Rikan, B.S.; Pu, Y.G.; Hwang, K.C.; Yang, Y.; et al. A low-power 12-bit 20 MS/s asynchronously controlled SAR ADC for WAVE ITS sensor based applications. Sensors 2021, 21, 2260. [Google Scholar] [CrossRef]
  7. Fredenburg, J.A.; Flynn, M.P. A 90-MS/s 11-MHz-bandwidth 62-dB SNDR noise shaping SAR ADC. IEEE J. Solid-State Circuits 2012, 47, 2898–2904. [Google Scholar] [CrossRef]
  8. Chen, C.-C.; Hsieh, C.-C. A 12-ENOB second-order noise shaping SAR ADC with PVT-insensitive voltage-time-voltage converter. In Proceedings of the IEEE Asian Solid-State Circuits Conference (A-SSCC), Busan, Korea, 7–10 November 2021; pp. 1–3. [Google Scholar]
  9. Shu, Y.-S.; Kuo, L.-T.; Lo, T.-Y. An oversampling SAR ADC with DAC mismatch error shaping achieving 105 dB SFDR and 101 dB SNDR over 1 kHz BW in 55 nm CMOS. IEEE J. Solid-State Circuits 2016, 51, 2928–2940. [Google Scholar] [CrossRef]
  10. Obarta, K.; Matsukawa, K.; Miki, T.; Tsukamoto, Y.; Sushihara, K. A 97.99 dB SNDR, 2 kHz BW, 371 μW noise-shaping SAR ADC with dynamic element matching and modulation dither effect. In Proceedings of the 2016 IEEE Symposium on VLSI Circuits (VLSI-Circuits), Kyoto, Japan, 15–17 June 2016; pp. 1–2. [Google Scholar]
  11. Liu, C.C.; Huang, M.C. A 0.46 mW 5MHz-BW 79.7 dB-SNDR noise-shaping SAR ADC with dynamic-amplifier-based FIR-IIR filter. In Proceedings of the 2017 IEEE International Solid-State Circuits Conference (ISSCC), San Francisco, CA, USA, 5–9 February 2017; pp. 466–467. [Google Scholar]
  12. Li, S.; Qiao, B.; Gandara, M.; Pan, D.Z.; Sun, N. A 13-ENOB second order noise-shaping SAR ADC realizing optimized NTF zeros using an error-feedback structure. IEEE J. Solid-State Circuits 2018, 53, 3484–3496. [Google Scholar] [CrossRef]
  13. Chen, Z.; Miyahara, M.; Matsuzawa, A. A 9.35-ENOB, 14.8 fJ/conv.-step fully-passive noise-shaping SAR ADC. In Proceedings of the IEEE Symposium VLSI Circuits, Kyoto, Japan, 17–19 June 2015; pp. C64–C65. [Google Scholar]
  14. Jie, L.; Zheng, B.; Flynn, M.P. A 500 MHz-bandwidth 70.4 dB-SNDR calibration-free time-interleaved 4th-order noise-shaping ADC. In Proceedings of the IEEE International Solid-State Circuits Conference: Digest of Technical Papers, San Francisco, CA, USA, 5–9 February 2019; pp. 332–333. [Google Scholar]
  15. Guo, W.; Sun, N. A 12 b-ENOB 61 μW noise-shaping SAR ADC with a passive integrator. In Proceedings of the European Solid-State Circuits Conference (ESSCIRC), Lausanne, Switzerland, 12–15 September 2016; pp. 405–408. [Google Scholar]
  16. Chen, Z.; Miyahara, M.; Matsuzawa, A. A 2nd order fully-passive noise-shaping SAR ADC with embedded passive gain. In Proceedings of the IEEE Asian Solid-State Circuits Conference, Toyama, Japan, 7–9 November 2016; pp. 309–312. [Google Scholar]
  17. Guo, W.; Zhuang, H.; Sun, N. A 13 b-ENOB 173 dB-FoM 2nd-order NS SAR ADC with passive integrators. In Proceedings of the Symposium on VLSI Circuits, Kyoto, Japan, 5–8 June 2017; pp. C236–C237. [Google Scholar]
  18. Lin, Y.Z.; Lin, C.Y.; Tsou, S.C.; Tsai, C.H.; Lu, C.H. A 40 MHz-BW 320 MS/s passive noise-shaping SAR ADC with passive signal-residue summation in 14 nm FinFET. In Proceedings of the IEEE International. Solid-State Circuits Conference (ISSCC), San Francisco, CA, USA, 17–21 February 2019; pp. 330–332. [Google Scholar]
  19. Hwang, Y.H.; Song, Y.S.; Park, J.E.; Jeong, D.K. A 0.6-to-1V 10k-to-100kHz BW 11.7b-ENOB noise shaping SAR ADC for IoT sensor applications in 28-nm CMOS. In Proceedings of the IEEE Asian Solid-State Circuits Conference (A-SSCC), Tainan, Taiwan, 5–7 November 2018; pp. 247–248. [Google Scholar]
  20. Song, Y.; Chan, C.H.; Zhu, Y.; Geng, L.; Seng-Pan, U.; Martins, R.P. Passive noise shaping in SAR ADC with improved efficiency. IEEE Trans. Very Large Scale Integ. (VLSI) Syst. 2018, 26, 416–420. [Google Scholar] [CrossRef]
  21. Lin, Y.-Z.; Tsai, C.-H.; Tsou, S.-C.; Chu, R.-X.; Lu, C.-H. A 2.4-mW 25-MHz BW 300-MS/s passive noise shaping SAR ADC with noise quantizer technique in 14-nm CMOS. In Proceedings of the IEEE Symposium VLSI Circuits, Kyoto, Japan, 5–8 June 2017; pp. C234–C235. [Google Scholar]
  22. Liu, J.; Li, D.; Zhong, Y.; Tang, X.; Sun, N. A 250kHz-BW 93dB-SNDR 4th-order noise-shaping SAR using capacitor stacking and dynamic buffering. In Proceedings of the IEEE International Solid-State Circuits Conference—(ISSCC) Digest of Technical Papers, San Francisco, CA, USA, 22–26 February 2021; pp. 369–371. [Google Scholar]
  23. Lee, J.-H.; Park, D.; Cho, W.; Phan, H.N.; Nguyen, C.L.; Lee, J.-W. A 1.15 μW, 200 kS/s 10-b monotonic SAR ADC using dual on-chip calibrations and accuracy enhancement techniques. Sensors 2018, 18, 3486. [Google Scholar] [CrossRef] [Green Version]
  24. Hariprasath, V.; Guerber, J.; Lee, S.-H.; Moon, U. Merged capacitor switching based SAR ADC with highest switching energy-efficiency. Electron. Lett. 2010, 46, 620–621. [Google Scholar] [CrossRef] [Green Version]
  25. Liu, C.; Chang, S.; Huang, G.; Lin, Y. A 10-bit 50-MS/s SAR ADC with a monotonic capacitor switching procedure. IEEE J. Solid-State Circuits 2010, 45, 731–740. [Google Scholar] [CrossRef]
  26. Neitola, M.; Rahkonen, T. A generalized data-weighted averaging algorithm. IEEE Trans. Circuits Syst. II Exp. Briefs 2010, 57, 115–119. [Google Scholar] [CrossRef]
  27. Miller, M.R.; Petrie, C.S. A multibit sigma-delta ADC for multimode receivers. IEEE J. Solid-State Circuits 2003, 38, 475–482. [Google Scholar] [CrossRef]
  28. Duong, H.V.; Hieu, N.X.; Park, D.; Lee, J.-W. A battery-assisted passive EPC Gen-2 RFID sensor tag IC with efficient battery power management and RF energy harvesting. IEEE Trans. Ind. Electron. 2016, 63, 7112–7123. [Google Scholar] [CrossRef]
Figure 1. Functional representation of the proposed ADC.
Figure 1. Functional representation of the proposed ADC.
Sensors 22 00869 g001
Figure 2. (a) Block diagram of the proposed ADC; (b) schematic of the proposed noise-shaping SAR ADC with timing waveforms. The VREF = 1.8 V is the reference voltage.
Figure 2. (a) Block diagram of the proposed ADC; (b) schematic of the proposed noise-shaping SAR ADC with timing waveforms. The VREF = 1.8 V is the reference voltage.
Sensors 22 00869 g002aSensors 22 00869 g002b
Figure 3. Operation of the noise shaping ADC. (a) (k − 1)th cycle for sampling and digital conversion, (b) (k − 1)th cycle for residue capture, and (c) kth cycle for charge pumping and residue integration.
Figure 3. Operation of the noise shaping ADC. (a) (k − 1)th cycle for sampling and digital conversion, (b) (k − 1)th cycle for residue capture, and (c) kth cycle for charge pumping and residue integration.
Sensors 22 00869 g003
Figure 4. Flowchart of the proposed ADC operation during (k − 1)th cycle.
Figure 4. Flowchart of the proposed ADC operation during (k − 1)th cycle.
Sensors 22 00869 g004
Figure 5. Noise suppression as a function of β for various α values.
Figure 5. Noise suppression as a function of β for various α values.
Sensors 22 00869 g005
Figure 6. Calculated NTF as a function of normalized frequency.
Figure 6. Calculated NTF as a function of normalized frequency.
Sensors 22 00869 g006
Figure 7. Behavioral model of the proposed ADC.
Figure 7. Behavioral model of the proposed ADC.
Sensors 22 00869 g007
Figure 8. Output spectra of the ADC (a) without noise shaping and (b) with noise shaping for the CDAC mismatch rate of 0.5%. fin = 1.33 kHz and fS = 52 kHz.
Figure 8. Output spectra of the ADC (a) without noise shaping and (b) with noise shaping for the CDAC mismatch rate of 0.5%. fin = 1.33 kHz and fS = 52 kHz.
Sensors 22 00869 g008
Figure 9. Probability distributions of the ENOB for different CDAC mismatch rate.
Figure 9. Probability distributions of the ENOB for different CDAC mismatch rate.
Sensors 22 00869 g009
Figure 10. Output spectra of the ADC. (a) Previous work and (b) proposed work. fin = 1.33 kHz, OSR = 10.
Figure 10. Output spectra of the ADC. (a) Previous work and (b) proposed work. fin = 1.33 kHz, OSR = 10.
Sensors 22 00869 g010
Figure 11. Schematic of the comparator having complementary differential input pairs. VDD = 1.8 V.
Figure 11. Schematic of the comparator having complementary differential input pairs. VDD = 1.8 V.
Sensors 22 00869 g011
Figure 12. Microphotograph of the fabricated ADC. Measurement setup is also shown.
Figure 12. Microphotograph of the fabricated ADC. Measurement setup is also shown.
Sensors 22 00869 g012
Figure 13. Measured output spectrum of the ADC. (a) fin = 1.33 kHz and fS = 52 kS/s and (b) fin = 8 kHz and fS = 180 kS/s.
Figure 13. Measured output spectrum of the ADC. (a) fin = 1.33 kHz and fS = 52 kS/s and (b) fin = 8 kHz and fS = 180 kS/s.
Sensors 22 00869 g013
Figure 14. (a) Measured SNDR and SFDR as a function of the sampling rate. (b) Measured SNDR and SFDR as a function of the input frequency.
Figure 14. (a) Measured SNDR and SFDR as a function of the sampling rate. (b) Measured SNDR and SFDR as a function of the input frequency.
Sensors 22 00869 g014
Figure 15. Measured dynamic range.
Figure 15. Measured dynamic range.
Sensors 22 00869 g015
Figure 16. Measured static performance of the ADC.
Figure 16. Measured static performance of the ADC.
Sensors 22 00869 g016
Figure 17. (a) Block diagram of the noise shaping ADC with the DEM. (b) Timing waveform of the DEM logic.
Figure 17. (a) Block diagram of the noise shaping ADC with the DEM. (b) Timing waveform of the DEM logic.
Sensors 22 00869 g017
Figure 18. Comparison of the dynamic performance with and without the DEM.
Figure 18. Comparison of the dynamic performance with and without the DEM.
Sensors 22 00869 g018
Figure 19. Comparison of the static performance of the ADC with and without the DEM. (a) DNL, (b) INL.
Figure 19. Comparison of the static performance of the ADC with and without the DEM. (a) DNL, (b) INL.
Sensors 22 00869 g019
Table 1. List of noise transfer function.
Table 1. List of noise transfer function.
NTFNoise Suppression
Ideal1 − z−1−10 dB
[13]1 − 0.5z−1−5.26 dB
[18](1 − 0.5z−1)/(1 + 0.5z−1)−8.68 dB
Proposed(1 − 0.79z−1)/(1 + 0.45z−1)−12.49 dB
Table 2. Performance comparison.
Table 2. Performance comparison.
[9][10][12][15][16][17][18][19][20] *This Work
Filter typeActiveActiveActivePassivePassivePassivePassivePassivePassivePassive
OP-amp freeNoNoYesYesYesYesYesYesYesYes
Filter order1321221211
Extra input for comparator (No.)Yes (2)Yes (2)NoYes (2)Yes (2)Yes (3)NoYes (3)Yes (2)No
CDAC (bit)1212910891010109
Supply (V)1.21.55/0.751.11.211.10.91.011.8
Bandwidth (kHz)12625125800026240,00010020003
OSR51225884164162520
Power (μW)15.737.1846125314312501185614.1
Process (nm)5528401306540142865180
FoMS (dB)180175178167169173171.7173176.8170
FoMW
(fJ/conv.-step)
8559710.9338.92511611.8
* Simulation results.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Choi, C.; Lee, J.-W. An 11.8-fJ/Conversion-Step Noise Shaping SAR ADC with Embedded Passive Gain for Energy-Efficient IoT Sensors. Sensors 2022, 22, 869. https://doi.org/10.3390/s22030869

AMA Style

Choi C, Lee J-W. An 11.8-fJ/Conversion-Step Noise Shaping SAR ADC with Embedded Passive Gain for Energy-Efficient IoT Sensors. Sensors. 2022; 22(3):869. https://doi.org/10.3390/s22030869

Chicago/Turabian Style

Choi, Changhyung, and Jong-Wook Lee. 2022. "An 11.8-fJ/Conversion-Step Noise Shaping SAR ADC with Embedded Passive Gain for Energy-Efficient IoT Sensors" Sensors 22, no. 3: 869. https://doi.org/10.3390/s22030869

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop