Next Article in Journal
Exploitability of Maritime Fleet-Based 5G Network Extension
Previous Article in Journal
Cognitive Insights into Museum Engagement: A Mobile Eye-Tracking Study on Visual Attention Distribution and Learning Experience
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A 0.4 V CMOS Current-Controlled Tunable Ring Oscillator for Low-Power IoT and Biomedical Applications

by
Md Anas Abdullah
1,
Mohamed B. Elamien
1,* and
M. Jamal Deen
1,2
1
Department of Electrical and Computer Engineering, McMaster University, 1280 Main Street West, Hamilton, ON L8S 4K1, Canada
2
School of Biomedical Engineering, McMaster University, 1280 Main Street West, Hamilton, ON L8S 4K1, Canada
*
Author to whom correspondence should be addressed.
Electronics 2025, 14(11), 2209; https://doi.org/10.3390/electronics14112209
Submission received: 22 April 2025 / Revised: 22 May 2025 / Accepted: 26 May 2025 / Published: 29 May 2025

Abstract

This work presents a current-controlled CMOS ring oscillator (CCRO) optimized for ultra-low-voltage applications in next-generation energy-constrained systems. Leveraging bulk voltage tuning in 22 nm FDSOI differential inverter stages, the topology enables frequency adjustment while operating MOSFETs in the subthreshold region—critical for minimizing power in sub-1 V environments. Simulations at 0.4 V supply demonstrate robust performance: a three-stage oscillator achieves a 537–800 MHz tuning range with bias current (IBIAS) modulation from 30–130 nA, while a four-stage configuration spans 388–587 MHz. At 70 nA IBIAS, the three-stage design delivers a nominal frequency of 666.8 MHz with just 10.23 µW power dissipation, underscoring its suitability for ultra-low-power IoT and biomedical applications. The oscillator’s linear frequency sensitivity (2.63 MHz/nA) allows precise, dynamic control over performance–power tradeoffs. To address diverse application needs, the design integrates three tunability mechanisms: programmable capacitor arrays for coarse frequency adjustments, configurable stage counts (three- or four-stage topologies), and supply voltage scaling. This multi-modal approach extends the operational range to 1 MHz–1 GHz, ensuring compatibility with low-speed sensor interfaces and high-speed edge-computing tasks. The CCRO’s subthreshold operation at 0.4 V—coupled with nanoampere-level current consumption—makes it uniquely suited for battery-less systems, wearable health monitors, and implantable medical devices where energy efficiency and adaptive clocking are paramount. By eliminating traditional voltage-controlled oscillators’ complexity, this topology offers a compact, scalable solution for emerging ultra-low-power technologies.

1. Introduction

The rapid growth of the Internet of Things (IoT), wireless sensor networks, and energy-harvested microsystems continues to present significant challenges while driving demand for power-efficient analog and mixed-signal integrated circuit (IC) solutions using CMOS technology [1,2,3,4,5,6,7]. In such applications like implantable and wearable biomedical devices, body sensor networks, and other compact systems, controlled oscillators (COs) are essential due to their need for extremely low power consumption, low phase noise, minimal layout area, and wide frequency-tuning capability to accommodate process and temperature variations [7,8,9,10,11]. COs also play a vital role in analog-to-digital converters [12,13,14], time-to-digital converters [15,16,17], and phase-locked loops (PLLs), which serve as the time reference for control systems, RFID systems, clock generation, and other clock-dependent applications, including switching power converters and beyond [18,19,20,21,22].
CMOS-controlled oscillators (COs) are broadly classified into two main categories: LC tank oscillators and ring oscillators, where LC oscillators are preferred in applications demanding both high-quality factors (Q) and low-phase noise. However, their reliance on spiral inductors leads to large silicon area consumption and higher power consumption, making them unsuitable for ultra-low-power integrated circuits or applications where compact physical area is essential [23,24,25]. In contrast, ring oscillators (ROs) consisted of an odd number of typically identical delay stages connected in a loop, with the output of the last stage fed back to the input of the first. Due to their fully digital nature, ROs offer several advantages over LC oscillators, including smaller area, simpler design, better scalability with CMOS technology, and ease of integration. These properties make ROs well-suited for applications that prioritize low power and high integration, such as in wearable biomedical systems, IoT edge devices, and sensor nodes.
Based on the tuning mechanism, ring oscillators can be further classified as voltage-controlled or current-controlled. Controlled ring oscillator (CRO) designs offer additional benefits such as a wide frequency-tuning range, minimal power consumption, reduced design complexity, small layout area, and full compatibility with CMOS processes. While relaxation oscillators can achieve a wide tuning range but ring oscillators are notably more efficient in terms of power usage [26].
Several design strategies for low-power controlled oscillators (COs) have been studied in the literature to enhance the phase-noise and frequency-tuning range performance. One notable technique combines current-starving, commonly used in current-controlled oscillators, with a negatively skewed delay strategy to enhance the power-delay product, defined as the product of power consumption and the delay of a single gate [27]. Another approach involves a traditional voltage-controlled oscillator (VCO) enhanced with negative resistance, a multi-gate structure, and a bypass capacitor to reduce higher-order harmonics [28]. Additionally, a digital control circuit was designed to regulate the oscillation frequency [1]. An alternative design employs positive feedback to every stage, enabling operation with just two stages rather than three, which helps lower power consumption [29]. A delay cell for frequency tuning, composed of a PMOS and an NMOS transistor forming a transmission gate, has been utilized to adjust the oscillation frequency by modulating the gate voltage [30]. Finally, a dynamic threshold MOS (DTMOS) technique has been proposed to lower the Vth-threshold voltage of NMOS transistors in the delay cell, allowing for faster transitions and achieving a higher operating frequency [19].
The concept of controlling the ring oscillator (RO) frequency through the bulk terminal was explored, including the impact of bulk voltage fluctuations on phase-noise performance [31]. Additionally, an RO was integrated into an adaptive body-bias (ABB) generator for low-voltage CMOS VLSI circuits to facilitate delay characterization of CMOS gates [32]. However, most prior works do not explore body-biasing as a reliable current-controlled tuning method for sub-0.4 V operation, especially on differential architecture.
This work introduces a body-bias-controlled current-controlled ring oscillator (CCRO) that is optimized for ultra-low-power and ultra-low-voltage operation in 22 nm FDSOI technology. It was initially applied in the analog domain [33,34,35], including logic gates [36,37], and later adapted for setting the quiescent current in inverter stages [38] to design a current-controlled oscillator (CCO) that operates with low power and low voltage. The proposed design maintains a static output voltage at half the supply, while the bias current controls the oscillation frequency.
The effectiveness of body biasing also depends on the choice of forward or reverse biasing conditions, each impacting circuit performance differently. Forward biasing lowers the threshold voltage (Vth) and increases transconductance (gm) for improved speed and efficiency [39]. In contrast, reverse biasing higher the threshold voltage (Vth), which reduces leakage but may introduce noise [39,40]. In noise-sensitive analog designs, forward biasing is often preferred [39,41,42]. Finally, employing a systematic body-biasing loop for delay control, the inverter stages achieve reliable cascading and eliminate input–output voltage transfer offsets. This approach significantly broadens the tuning range and enhances the robustness of RO against process, voltage, and temperature (PVT) variations. Moreover, the proposed CCRO demonstrates robust operation even at low supply voltages, comparable to the 0.4 V operation observed in 22 nm FDSOI technologies, thus ensuring energy-efficient performance. This also achieves favorable phase noise and a competitive figure of merit (FoM) compared to prior works, reinforcing its potential for low-power applications.
The manuscript is structured as follows: Section 2 provides a detailed description of the proposed body-bias-controlled current-controlled ring oscillator (CCRO) architecture, including its circuit-level implementation and design considerations. Section 3 presents comprehensive simulation results, including performance analysis under process, voltage, and temperature (PVT) variations, as well as phase noise, power consumption, and tuning characteristics. Finally, Section 4 concludes the paper by summarizing the key findings, highlighting the advantages of the proposed design, and outlining potential directions for future research.

2. Proposed Design

A ring oscillator consists of N inverter stages (delay stages) connected in a closed loop, producing square wave outputs. The oscillation frequency is determined by the number of inverter stages (N) and the delay time of each inverter (τ) following the relationship.
f o s c = 1 2 N . τ = I D 2 N . C i . V
where τ is the delay per stage, C i is the load capacitance at the output node, V is the voltage swing (~ V D D / 2 ), and I D is the current through each differential output node.
The delay of an inverter is influenced by its load capacitance and the biasing current. Consequently, adjusting the bias current allows for modification of the inverter delay, thereby altering the oscillation frequency.
The proposed current-controlled ring oscillator (CCRO), as shown in Figure 1, is designed with an N-stage ring oscillator. Each stage consists of a differential CMOS inverter with controllable bulk terminals. The advantage of the differential CMOS inverter is that it inherently relaxes the constraint of using an odd number of stages for designing a ring oscillator [43,44]. Unlike single-ended designs, differential ring oscillators can use an even number of stages with minimal changes in design [45,46,47].
Furthermore, to address the mismatch caused by the longer path between the first- and last-stage feedback, the delay cell can be arranged in a twisted configuration [15,48]. This problem especially happened when we required more delay cells (N-stage) in the design. Additionally, an output buffer is used to ensure balanced loading across delay cells and enhance output drive strength. While not shown in the schematic, it is highlighted in the layout.
To establish the typical oscillation frequency and mitigate sensitivity to parasitic effects, we include an output capacitor (Ci) at the output of each stage, as shown in Figure 1a. Further details of this design are discussed in the following section. The body bias voltages VBP and VBN for the transistors are produced using the auxiliary circuit shown in Figure 2 [38,49]. This auxiliary circuit maximizes the current in the reference inverter (MB1-MB2) when the input voltage reaches the logic threshold of VDD/2. The reference inverter input is biased at VDD/2 during quiescent operation, and this condition is mirrored at the output through the negative feedback provided by the error amplifier.
The quiescent current is controlled through the bulk terminals, where VBP and VBN are applied to the p-channel and n-channel transistors, respectively. These voltages are generated by amplifiers A1 and A2 using a method introduced in [33] and later adopted in [38].
The transistors MB1 and MB2 serve as reference devices, each with a VGS value set to VDD/2. Their quiescent drain current is maintained at IBIAS through a local feedback loop controlled by the A1 and A2 auxiliary amplifiers, which provide the necessary bulk voltages VBP and VBN while applying the following conditions:
(a)
assigned aspect ratios of MB1 and MB2
(b)
ID1,2 = kIBIAS
(c)
VSGR1= VGSR2= VDD/2
(d)
VSDR1= VDSR2= VDD/2 [Assuming virtual ground]
The aspect ratios of MB1 and MB2 should be carefully selected to ensure that the bulk voltages remain within the VDD/2 and ground limits. A precise layout is essential to reduce mirroring errors between the bias and reference branches. Any offset or mismatch in the error amplifiers (A1 and A2) could affect VBN and VBP, potentially leading to variations in the current flowing through the inverters. To mitigate this, careful layout techniques, such as common-centroid placement and symmetrical routing, are employed. Proper device sizing is also maintained to reduce mismatch sensitivity. This includes using similar channel lengths for all MOSFETs and choosing even-numbered multiplications for wider devices, which helps achieve better matching and a more compact layout. Auxiliary amplifiers A1 and A2 are required to deliver a rail-to-rail output, whereas the input common-mode range is non-critical due to the input voltage being fixed at VDD/2. This allows the use of simple symmetrical operational transconductance amplifiers (OTAs) operating in the subthreshold region. A practical implementation of such an amplifier can be found in [38,50] and is shown in Figure 3.
The voltages VBN and VBP are supplied to the delay cells (inverters) forming the ring oscillator illustrated in Figure 1c,d. These voltages limit the maximum current flowing through the inverters to a specific value when the input voltage reaches the threshold. For example, the transistor MN1 in the first inverter stage, shown in Figure 1a, operates under quiescent conditions where VIN is maintained at VDD/2. Under these conditions, MB2 and MN1 share the same source, gate, and bulk voltages, leading to MN1’s drain current being directly mirrored from MB2.
I D 1 ,     N 1 = ( W / L ) N 1 ( W / L ) M R 2 I B I A S
This equality is rigorously maintained due to the source-drain voltage of MN1 being precisely biased at VDD/2. The same principle applies to all the transistors within the ring oscillator. Essentially, both PMOS and NMOS devices have their currents regulated by IBIAS through current–mirror relationships.
I D 1 , P i = ( W / L ) P i ( W / L ) M R 1 I B I A S
I D 1 , N i = ( W / L ) N i ( W / L ) M R 2 I B I A S
The aspect ratios of the respective p-type and n-type MOSFETs within the ROs are indicated by the terms (W/L)Pi and (W/L)Ni, where i = 1,2,…, N.

3. Simulation Results

The proposed design for the three-stage and four-stage current-controlled ring oscillators (CCROs) is implemented using 22 nm FDSOI technology and verified through schematic-level simulations. A regular threshold voltage Vth (RVT) device is used for the design to achieve symmetrical inverter behavior, balanced body effect coefficients, and uniform control bulk voltage ranges. The circuit is operated power supply of 0.4 V and a bias current of 70 nA. The operating temperature range, suitable for wearable and implantable devices, was set between 0 °C and 60 °C. Table 1 provides the transistor dimensions and other component parameters.
In OPAMP design, a slightly higher length of 60 nm is used to mitigate the short-channel effects. Based on the selected design parameters, the mirroring coefficient (Ki) and the transistors’ dimensions ratios, as expressed in Equations (2)–(4), are unified to a value of one. This design choice ensures that the nominal quiescent current per branch, equal to the short-circuit current of 70 nA, leads to a total nominal quiescent current of N × 70 nA, where N represents the number of delay cells or inverter stages. A coarse-tuning capacitor (Ci) of 1 fF is uniformly applied across all stages. The auxiliary amplifiers (A1 and A2) with transistors operating in the subthreshold region delivered an open-loop gain of approximately 33 dB and a phase margin of 94.8°. Figure 4 presents the Bode plots, illustrating both the magnitude and phase responses of the amplifier’s open-loop gain at the nominal temperature of 27 °C.
The stability of the quiescent conditions is confirmed through validation. The nominal bulk voltages, VBP and VBN, produced by the circuit shown in Figure 2, are measured at 202.34 mV and 202.4 mV, respectively. Monte Carlo (MC) simulation results show that the bias-generator circuit in Figure 2 delivers an average current of 678 nA, with a standard deviation of 32.15 nA.
Figure 5 illustrates how the body-bias voltages (VBP and VBN) of the transistors in the reference-inverting gate are impacted by the biasing current (IBIAS), which is swept across a range of 30 nA to 130 nA. These voltages remain within the supply rails, and their symmetrical behavior indicates proper sizing and efficient use of the control voltage’s full dynamic range. Additionally, currents flowing into the transistor body terminal are analyzed, which are negligible (<<1 nA).
The increase in bias current (IBIAS) from 30 nA to 130 nA in the differential delay cell leads to a proportional rise in drain current, which reduces the propagation delay of each stage. As shown in Equation (1), the oscillation frequency is inversely related to this delay, resulting in higher operating frequencies. Moreover, the higher current improves the transconductance (gm) of the differential pairs, enhancing the switching speed and strengthening the regenerative behavior of each stage. These effects collectively contribute to the broader tuning range and higher achievable operating frequencies. This relationship is clearly demonstrated in the simulation results shown in Figure 6, where the oscillator frequency scales effectively with changes in IBIAS.
The three-stage CCRO output signal with Ci = 1 fF is illustrated in Figure 7 for three different biasing current values: 30 nA, 70 nA, and 130 nA, corresponding to the minimum, nominal, and maximum values, respectively.
The relationship between oscillation frequency and biasing current for Ci = 1 fF is illustrated in Figure 8. The oscillation frequency range is achieved from 537 MHz to 800 MHz, with a tuning sensitivity of approximately 2.63 MHz/nA, calculated as (fMAXfMIN)/(IBIAS, MAXIBIAS, MIN). This high tuning sensitivity provides wide frequency adjustability while consuming minimal current, making it highly efficient for low-power applications. The 2.63 MHz/nA slope enables significant frequency shifts with small changes in bias current, allowing precise control with low power overhead. Similarly, we have received the oscillation frequency ranges 388 MHz to 587 MHz for the four-stage RO with the sensitivity of 1.99 MHz/nA. While the predicted behavior suggests a linear relationship as derived from Equation (1), the observed results indicate a logarithmic dependence on IBIAS. The oscillation frequency is highly sensitive to IBIAS, which is externally supplied in this design for tuning the frequency. To ensure stable operation, a well-regulated and low-noise IBIAS source is required to minimize the effects of temperature, supply noise, and process variations. While bandgap references or adaptive tuning can help stabilize IBIAS, their effectiveness depends on the external circuitry [51,52]. Future improvements could explore on-chip compensation to further enhance robustness.
Phase noise is a critical parameter in a ring oscillator (RO), originating from thermal noise, flicker noise, and supply variations, including downscaling transistors at low-frequency noise [37], which influences the threshold voltage (Vth) and oscillation stability. Body-bias tuning further influences Vth, transconductance (gm), and flicker noise, which directly impacts the phase-noise performance. However, body-biasing optimization, differential architectures, and advanced layout techniques (such as guard rings and symmetric design) have already been adopted in the design to improve the phase noise. Future designs can consider advanced techniques for phase-noise reduction, such as refining low-noise biasing circuits, improved power supply regulation through low-dropout regulators (LDOs), and decoupling capacitors. Additionally, optimizing transistor sizing and using noise-filtering techniques in the control circuitry can further suppress unwanted noise sources.
Finally, Figure 9 illustrates the phase-noise performance as a function of relative frequency (∆f). Phase noise is simulated across the different process corners, varying from −80.69 to −84.97 dBc/Hz, Table 2. This low-phase noise performance enhances signal integrity and improves overall system reliability, making it well-suited for noise-sensitive applications.
However, this tuning flexibility comes with design trade-offs. A higher tuning range generally requires higher bias currents, leading to increased power consumption. Conversely, lower power consumption limits the tuning range, requiring a careful balance based on application requirements. Additionally, the capacitor size (Ci) significantly impacts oscillator performance: increasing (Ci) reduces phase noise but at the cost of a narrower and lower tuning range and higher power dissipation and area. Conversely, a smaller (Ci) extends the tuning range but may result in higher phase noise.
Furthermore, the number of stages affects oscillator frequency, power, and area. Increasing the number of stages generally lowers the oscillation frequency while increasing power consumption and chip area. Conversely, reducing the number of delay cells results in a higher frequency but may degrade phase noise due to reduced averaging. These trade-offs must be carefully considered based on power, area, and performance constraints for specific applications.
Corner analysis results for the three-stage ring oscillator operating at a nominal bias current (IBIAS) of 70 nA are presented in Table 2. The analysis reveals a notable variation in oscillator frequency across PVT conditions. However, this variation can be effectively mitigated by fine-tuning the bias current (IBIAS), offering a practical solution for achieving more stable performance under varying conditions. In real-world systems, adjusting IBIAS effectively compensates for corner-induced frequency shifts. This flexible tuning capability allows the design to maintain stable operation without complex calibration circuits.
Figure 10 highlights the outcomes of mismatch Monte Carlo (MC) simulations conducted under standard conditions (VDD = 0.4 V and T = 27 °C) at IBIAS = 70 nA. The simulations provide insights into the impact of device mismatches on oscillator performance, specifically analyzing frequency variation due to inherent process variations. These results help assess the statistical robustness of the design and its ability to maintain consistent performance under real-world manufacturing conditions.
The layout of the three-stage CCRO with an output buffer is illustrated in Figure 11, occupying the area of 4.8 µm × 2.3 µm. This extremely compact layout is beneficial for large-scale system-on-chip (SoC) integration, making it ideal for space-constrained applications such as implantable devices and wearable devices. Post-layout simulations are performed to ensure the reliability of the presented results. The primary impact of the layout is the addition of parasitic capacitance of ~1 fF (to be more specific, 0.82 fF to 0.87 fF) at the CCRO’s output node. By compensating for this effect through a reduction in the stated tuning capacitance by an equivalent amount and the post-layout simulation results closely match those obtained at the schematic level.
A comparison of differential structure-based controlled oscillators (COs) topologies from the literature with the proposed design is presented in Table 3. The proposed CCRO outperforms others in terms of power consumption, achieving the best figure of merit.
(FoM) as defined by Equation (5), where the lower FoM (higher negative values) indicates better performance. This is a widely used benchmark for evaluating oscillator efficiency. These results demonstrate the superior trade-off between power efficiency and spectral purity, making the proposed design highly suitable for low-power and high-performance applications.
F o M = L f + 10 log P ( m W ) 20 l o g ( f o s c f )
where L f is the phase noise at a particular frequency offset (Δf), measured in dBc/Hz, P ( m W ) , power consumption normalized to 1 mW, f o s c f is the normalized frequency offset, where f o s c is the oscillation frequency.
The proposed design achieves an FoM of −161.1 dBc/Hz. This value represents the best performance among the evaluated oscillators, highlighting the effectiveness of the current control and body-biasing techniques used in this work. The extremely low power consumption of 10.23 µW, combined with a wide tuning range (537–800 MHz), contributes to this superior FoM. These features enable efficient frequency control with minimal power overhead, making the design highly suitable for low-power IoT and biomedical applications. Compared to state-of-the-art designs, the improved FoM demonstrates a better balance between phase noise and power efficiency, validating the design approach’s effectiveness. This performance advantage emphasizes the potential of the proposed CCRO for advanced SoC integration and space-constrained applications. However, further improvements could be explored for sub-0.4 V operation, as the current biasing approach may face limitations at ultra-low voltages. Especially, the op-amp may not be functioning properly at very low voltages. Future work could explore scalable methods for current control without excessive power consumption and examine alternative technologies or devices that support lower voltage operations.

4. Conclusions

This paper presents an innovative technique for adjusting the delay of the differential inverter cell in current-controlled ring oscillators (CCROs). The method allows precise control of the maximum current in each inverter through a body bias loop and facilitates tuning of the oscillation frequency by modulating a reference current.
CCROs with three and four stages are developed using a 22 nm commercial CMOS technology operating at a supply voltage of 0.4 V. Simulation results demonstrate a wide tuning range from 537 MHz to 800 MHz, extremely low power consumption of 10.23 µW, and a strong figure of merit (FoM) comparable to state-of-the-art designs. This balance of wide tunability, low power, and robust performance highlights the effectiveness of the proposed approach. Additionally, extensive corner and parametric simulations confirm the high resilience of the proposed CCROs to variations in process, voltage, and temperature (PVT), even when utilizing an advanced short-channel technology node.
Future work could explore implementing fractional-N PLL blocks to enhance frequency synthesis flexibility, and temperature compensation schemes could be investigated to improve stability under varying environmental conditions.

Author Contributions

Conceptualization, M.A.A., M.B.E. and M.J.D.; validation, M.A.A.; formal analysis, M.A.A.; investigation, M.B.E.; resources, M.B.E. and M.J.D.; writing—original draft preparation, M.A.A.; writing—review and editing, M.A.A., M.B.E. and M.J.D.; supervision, M.B.E. and M.J.D. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by a Discovery Grant from the Natural Science and Engineering Research Council of Canada (NSERC).

Data Availability Statement

Data is contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Srikram, P.; Ambalathankandy, P.; Motomura, M.; Ikebe, M. A 0.5 V modified pseudo-differential current-starved ring-VCO with linearity improvement for IoT devices. In Proceedings of the 2023 International Electrical Engineering Congress (iEECON), Krabi, Thailand, 8–10 March 2023; pp. 219–223. [Google Scholar]
  2. Corres-Matamoros, A.; Martínez-Guerrero, E.; Rayas-Sanchez, J.E. A programmable CMOS voltage controlled ring oscillator for radio-frequency diathermy on-chip circuit. In Proceedings of the 2017 International Caribbean Conference on Devices, Circuits and Systems (ICCDCS), Cozumel, Mexico, 5–7 June 2017; pp. 65–68. [Google Scholar]
  3. Alioto, M. (Ed.) Enabling the Internet of Things: From Integrated Circuits to Integrated Systems; Springer International Publishing: Berlin/Heidelberg, Germany, 2017. [Google Scholar]
  4. Margarit, M.A.; Tham, J.L.; Meyer, R.G.; Deen, M.J. A low-noise, low-power VCO with automatic amplitude control for wireless applications. IEEE J. Solid-State Circuits 1999, 34, 761–771. [Google Scholar] [CrossRef]
  5. Meng, X.; Li, X.; Cheng, L.; Tsui, C.-Y.; Ki, W.-H. A Low-Power Relaxation Oscillator with Switched-Capacitor Frequency-Locked Loop for Wireless Sensor Node Applications. IEEE Solid-State Circuits Lett. 2019, 2, 281–284. [Google Scholar] [CrossRef]
  6. Hassan, A.; Hamed, E.; Badr, E.; Elsharqawy, O.; Ismail, T.; Gabran, S.R.I.; Ismail, Y.; Mostafa, H. A VCO-Based MPPT Circuit for Lo-Voltage Energy Harvesters. In Proceedings of the 2017 IEEE Computer Society Annual Symposium on VLSI (ISVLSI), Bochum, Germany, 3–5 July 2017; pp. 580–584. [Google Scholar]
  7. Ballo, A.; Bottaro, M.; Grasso, A.D. A Review of Power Management Integrated Circuits for Ultrasound-Based Energy Harvesting in Implantable Medical Devices. Appl. Sci. 2021, 11, 2487. [Google Scholar] [CrossRef]
  8. Ghafari, B.; Koushaeian, L.; Goodarzy, F.; Evans, R.; Skafidas, E. An ultra-low-power and low-noise voltage-controlled ring oscillator for biomedical applications. In Proceedings of the IEEE 2013 Tencon—Spring, Sydney, NSW, Australia, 17–19 April 2013; pp. 20–24. [Google Scholar]
  9. Zambrano, B.; Garzón, E.; Strangio, S.; Crupi, F.; Lanuzza, M. A 0.05 mm2, 350 mV, 14 nW Fully-Integrated Temperature Sensor in 180-nm CMOS. IEEE Trans. Circuits Syst. II Express Briefs 2021, 69, 749–753. [Google Scholar] [CrossRef]
  10. Ballo, A.; Bruno, G.; Grasso, A.D.; Vaiana, M.G.G. A Compact Temperature Sensor with a Resolution FoM of 1.82 pJ.K2. IEEE Trans. Instrum. Meas. 2020, 69, 8571–8579. [Google Scholar]
  11. Lo, Y.L.; Chiu, Y.-T. A high-accuracy, high-resolution, and low-cost all-digital temperature sensor using a voltage compensation ring oscillator. IEEE Sens. J. 2016, 16, 43–52. [Google Scholar] [CrossRef]
  12. Kim, J.; Jang, T.K.; Yoon, Y.G.; Cho, S. Analysis and design of voltage-controlled oscillator based analog-to-digital converter. IEEE Trans. Circuits Syst. I 2010, 57, 18–30. [Google Scholar]
  13. Xing, X.; Gielen, G.G.E. A 42 fJ/Step-FoM Two-Step VCO-Based Delta-Sigma ADC in 40 nm CMOS. IEEE J. Solid-State Circuits 2015, 50, 714–723. [Google Scholar] [CrossRef]
  14. Gutierrez, E.; Rombouts, P.; Hernandez, L. Why and how VCO-based ADCs can improve instrumentation applications. In Proceedings of the 25th IEEE International Conference on Electronics, Circuits and Systems (ICECS), Bordeaux, France, 9–12 December 2018; pp. 101–104. [Google Scholar]
  15. Cheng, Z.; Deen, M.J.; Peng, H. A Low-Power Gateable Vernier Ring Oscillator Time-to-Digital Converter for Biomedical Imaging Applications. IEEE Trans. Biomed. Circuits Syst. 2016, 10, 445–454. [Google Scholar] [CrossRef]
  16. Vornicu, I.; Carmona-Galán, R.; Rodríguez-Vázquez, Á. Arrayable voltage-controlled ring-oscillator for direct time-of-flight image sensors. IEEE Trans. Circuits Syst. I Regul. Pap. 2017, 64, 2821–2834. [Google Scholar] [CrossRef]
  17. Elshazly, A.; Rao, S.; Young, B.; Hanumolu, P.K. A Noise-Shaping Time-to-Digital Converter Using Switched-Ring Oscillators—Analysis, Design, and Measurement Techniques. IEEE J. Solid State Circuits 2014, 49, 1184–1197. [Google Scholar] [CrossRef]
  18. Bae, J.; Radhapuram, S.; Jo, I.; Kihara, T.; Matsuoka, T. A low-voltage design of controller-based ADPLL for implantable biomedical devices. In Proceedings of the 2015 IEEE Biomedical Circuits and Systems Conference (BioCAS), Atlanta, GA, USA, 22–24 October 2015; pp. 1–4. [Google Scholar]
  19. Ranjan, R.; Raman, A.; Kashyap, N. Low Power and High Frequency Voltage Controlled Oscillator for PLL Application. In Proceedings of the 2019 6th International Conference on Signal Processing and Integrated Networks (SPIN), Noida, India, 7–8 March 2019; pp. 212–214. [Google Scholar]
  20. Moon, J.-W.; Choi, K.-C.; Choi, W.-Y. A 0.4-V, 90~350-MHz PLL With an Active Loop-Filter Charge Pump. IEEE Trans. Circuits Syst. II Express Briefs 2014, 61, 319–323. [Google Scholar]
  21. Deen, M.J.; Murji, R.; Fakhr, A.; Jafferali, N.; Ngan, W.L. Low power CMOS integrated circuits for radio frequency applications. IEE Proc. Circuit Device Syst. 2005, 152, 509–522. [Google Scholar] [CrossRef]
  22. Stornelli, V.; Barile, G.; Pantoli, L.; Scarsella, M.; Ferri, G.; Centurelli, F.; Tommasino, P.; Trifiletti, A. A New VCII Application: Sinusoidal Oscillators. J. Low Power Electron. Appl. 2021, 11, 30. [Google Scholar] [CrossRef]
  23. Razavi, B. A study of phase noise in CMOS oscillators. IEEE J. Solid-State Circuits 1996, 31, 331–343. [Google Scholar] [CrossRef]
  24. Sivaraaj, N.R.; Majeed, K.K.A. A Comparative Study of Ring VCO and LC-VCO: Design, Performance Analysis, and Future Trends. IEEE Access 2023, 11, 127987–128017. [Google Scholar] [CrossRef]
  25. Elamien, M.B.; Maundy, B.J.; Belostotski, L.; Elwakil, A.S. On chip 0.5 V 2 GHz four-output quadrature-phase oscillator. AEU-Int. J. Electron. Commun. 2020, 126, 153393. [Google Scholar] [CrossRef]
  26. Zaman, K.S.; Reaz, M.I.; Haque, F.; Arsad, N.; Ali, S.H.M. Optimization of WiFi Communication System using Low Power Ring Oscillator Delay Cell. In Proceedings of the 2020 IEEE 8th Conference on Systems, Process and Control (ICSPC), Melaka, Malaysia, 11–12 December 2020; pp. 91–94. [Google Scholar]
  27. Nayak, R.; Kianpoor, I.; Bahubalindruni, P.G. Low power ring oscillator for IoT applications. Analog. Integr. Circuits Signal Process. 2017, 93, 257–263. [Google Scholar] [CrossRef]
  28. Lee, S.-Y.; Hsieh, J.-Y. Analysis and Implementation of a 0.9-V Voltage-Controlled Oscillator with Low Phase Noise and Low Power Dissipation. IEEE Trans. Circuits Syst. II Express Briefs 2008, 55, 624–627. [Google Scholar]
  29. Reddy, N.; Pattanaik, M.; Rajput, S.S. 0.4 V CMOS based low power voltage controlled ring oscillator for medical applications. In Proceedings of the TENCON 2008–2008 IEEE Region 10 Conference, Hyderabad, India, 19–21 November 2008; pp. 1–5. [Google Scholar]
  30. Chuang, Y.-H.; Jang, S.-L.; Lee, J.-F.; Lee, S.-H. A low voltage 900 MHz voltage controlled ring oscillator with wide tuning range. In Proceedings of the 2004 IEEE Asia-Pacific Conference on Circuits and Systems, Tainan, Taiwan, 6–9 December 2004; Volume 1, pp. 301–304. [Google Scholar]
  31. Srivastava, A.; Zhang, C. An Adaptive Body-Bias Generator for Low Voltage CMOS VLSI Circuits. Int. J. Distrib. Sens. Netw. 2008, 4, 213–222. [Google Scholar] [CrossRef]
  32. Deen, M.J.; Kazemeini, M.H.; Naseh, S. Ultra-low power VCOs-performance characteristics and modeling. In Proceedings of the Fourth IEEE International Caracas Conference on Devices, Circuits and Systems (Cat. No. 02TH8611), Oranjestad, The Netherlands, 19 April 2002; p. C033. [Google Scholar]
  33. Monsurró, P.; Pennisi, S.; Scotti, G.; Trifiletti, A. Exploiting the Body of MOS Devices for High Performance Analog Design. IEEE Circuits Syst. Mag. 2011, 11, 8–23. [Google Scholar] [CrossRef]
  34. Palumbo, G.; Scotti, G. A Novel Standard-Cell-Based Implementation of the Digital OTA Suitable for Automatic Place and Route. J. Low Power Electron. Appl. 2021, 11, 42. [Google Scholar] [CrossRef]
  35. Wang, Z.; Savci, H.S.; Griggs, J.; Dogan, N.S. 1-V Ultra-Low-Power CMOS LC VCO with Dynamic Body Biasing. In Proceedings of the 2007 International Symposium on Signals, Circuits and Systems, Iasi, Romania, 13–14 July 2007; pp. 1–4. [Google Scholar]
  36. Deen, M.J.; Naseh, S.; Marinov, O.; Kazemeini, M.H. Very Low-Voltage Operation Capability of CMOS Ring Oscillators and Logic Gates. J. Vac. Sci. Tech. A (Spec. Issue) 2006, 24, 763–769. [Google Scholar] [CrossRef]
  37. Mostafa, H.; Anis, M.; Elmasry, M. Adaptive Body Bias for Reducing the Impacts of NBTI and Process Variations on 6T SRAM Cells. IEEE Trans. Circuits Syst. I Reg. Pap. 2011, 58, 2859–2871. [Google Scholar] [CrossRef]
  38. Ballo, A.; Pennisi, S.; Scotti, G. 0.5 V CMOS Inverter-Based Transconductance Amplifier with Quiescent Current Control. J. Low Power Electron. Appl. 2021, 11, 37. [Google Scholar] [CrossRef]
  39. Jiménez-P, A.; De la Hidalga-W, F.J.; Deen, M.J. Modelling of the Dynamic Threshold MOSFET. IEE Proc. Circuits Devices Syst. 2005, 152, 502–508. [Google Scholar] [CrossRef]
  40. Shrivastava, A.; Saxena, A.; Akashe, S. High performance of low voltage controlled ring oscillator with reverse body bias technology. Front. Optoelectron. 2013, 6, 338–345. [Google Scholar] [CrossRef]
  41. Von Arnim, K.; Borinski, E.; Seegebrecht, P.; Fiedler, H.; Brederlow, R.; Thewes, R.; Berthold, J.; Pacha, C. Efficiency of body biasing in 90-nm CMOS for low-power digital circuits. IEEE J. Solid-State Circuits 2005, 40, 1549–1556. [Google Scholar] [CrossRef]
  42. Marinov, O.; Deen, M.J.; Jiménez-Tejada, J.A. Low-frequency noise in downscaled silicon transistors: Trends, theory and practice. Phys. Rep. 2022, 990, 1–179. [Google Scholar] [CrossRef]
  43. Tianwang, L.; Jiang, J.; Bo, Y.; Xingcheng, H. Ultra low voltage, wide tuning range voltage controlled ring oscillator. In Proceedings of the 2011 9th IEEE International Conference on ASIC, Xiamen, China, 25–28 October 2011; pp. 824–827. [Google Scholar]
  44. Saheb, Z.; El-Masry, E.; Bousquet, J.-F. Ultra-low voltage and low power ring oscillator for wireless sensor network using CMOS varactor. In Proceedings of the 2016 IEEE Canadian Conference on Electrical and Computer Engineering (CCECE), Vancouver, BC, Canada, 15–18 May 2016; pp. 1–5. [Google Scholar]
  45. Abdollahvand, S.; Oliveira, L.B.; Gomes, L.; Goes, J. A low-voltage voltage-controlled ring-oscillator employing dynamic-threshold-MOS and body-biasing techniques. In Proceedings of the 2015 IEEE International Symposium on Circuits and Systems (ISCAS), Lisbon, Portugal, 24–27 May 2015; pp. 1294–1297. [Google Scholar]
  46. Jiang, T.; Yin, J.; Mak, P.I.; Martins, R.P. A 0.5-V 0.4-to-1.6-GHz 8-phase bootstrap ring-VCO using inherent non-overlapping clocks achieving a 162.2-dBc/Hz FoM. IEEE Trans. Circuits Syst. II Express Briefs 2018, 66, 157–161. [Google Scholar] [CrossRef]
  47. Chiariello, M.; La Gala, A.; Malanchini, M.; Gelmi, L.; De Matteis, M. A 900-MHz Hardening-by-Design Voltage Controlled Oscillator in 28nm CMOS. In Proceedings of the 2024 31st IEEE International Conference on Electronics, Circuits and Systems (ICECS), Nancy, France, 18–20 November 2024; pp. 1–4. [Google Scholar]
  48. Liu, W.; Li, W.; Ren, P.; Lin, C.; Zhang, S.; Wang, Y. A PVT tolerant 10 to 500 MHz all-digital phase-locked loop with coupled TDC and DCO. IEEE J. Solid-State Circuits 2010, 45, 314–321. [Google Scholar] [CrossRef]
  49. Ballo, A.; Pennisi, S.; Scotti, G.; Venezia, C. A 0.5 V Sub-Threshold CMOS Current-Controlled Ring Oscillator for IoT and Implantable Devices. J. Low Power Electron. Appl. 2022, 12, 16. [Google Scholar] [CrossRef]
  50. Pérez-Nicoli, P.; Veirano, F.; Rossi-Aicardi, C.; Aguirre, P. Design method for an ultra low power, low offset, symmetric OTA. In Proceedings of the 2013 7th Argentine School of Micro-Nanoelectronics, Technology and Applications, Buenos Aires, Argentina, 15–16 August 2013; pp. 38–43. [Google Scholar]
  51. Sundaresan, K.; Allen, P.E.; Ayazi, F. Process and temperature compensation in a 7-MHz CMOS clock oscillator. IEEE J. Solid State Circuits 2006, 41, 433–442. [Google Scholar] [CrossRef]
  52. Luk, T.; Potts, D.C. Statistical design characterization of analog circuits. In Proceedings of the 2008 9th International Conference on Solid-State and Integrated-Circuit Technology, Beijing, China, 20–23 October 2008; pp. 1705–1708. [Google Scholar]
Figure 1. (a) Transistor implementation of the differential delay element, (b) conventional ring oscillator, (c) proposed twisted-manner arrangement of the RO (3-stage), and (d) proposed twisted-manner arrangement of the RO (4-stage). (VINP—positive input, VINN—negative input, VOUTP—positive output, VOUTN—negative output, VBN—body-bias voltage for NMOS, VBP—body-bias voltage for PMOS, τ—delay of each stage).
Figure 1. (a) Transistor implementation of the differential delay element, (b) conventional ring oscillator, (c) proposed twisted-manner arrangement of the RO (3-stage), and (d) proposed twisted-manner arrangement of the RO (4-stage). (VINP—positive input, VINN—negative input, VOUTP—positive output, VOUTN—negative output, VBN—body-bias voltage for NMOS, VBP—body-bias voltage for PMOS, τ—delay of each stage).
Electronics 14 02209 g001
Figure 2. Implemented schematic diagram of the body bias generator for VBP, VBN.
Figure 2. Implemented schematic diagram of the body bias generator for VBP, VBN.
Electronics 14 02209 g002
Figure 3. Simplified OTA schematic implemented in A1 and A2.
Figure 3. Simplified OTA schematic implemented in A1 and A2.
Electronics 14 02209 g003
Figure 4. Boad plots of the amplifier’s open-loop gain and phase.
Figure 4. Boad plots of the amplifier’s open-loop gain and phase.
Electronics 14 02209 g004
Figure 5. Body-biased voltage (VB) vs. current (IBIAS).
Figure 5. Body-biased voltage (VB) vs. current (IBIAS).
Electronics 14 02209 g005
Figure 6. Inverter cell static currents vs. IBIAS (T = 27 °C).
Figure 6. Inverter cell static currents vs. IBIAS (T = 27 °C).
Electronics 14 02209 g006
Figure 7. Output signal of three-stage CCRO for IBIAS = 30 nA, 70 nA, and 130 nA (T = 27 °C).
Figure 7. Output signal of three-stage CCRO for IBIAS = 30 nA, 70 nA, and 130 nA (T = 27 °C).
Electronics 14 02209 g007
Figure 8. Output frequency vs. IBIAS (T = 27 °C) for three-stage and four-stage RO.
Figure 8. Output frequency vs. IBIAS (T = 27 °C) for three-stage and four-stage RO.
Electronics 14 02209 g008
Figure 9. Phase noise vs. relative frequency.
Figure 9. Phase noise vs. relative frequency.
Electronics 14 02209 g009
Figure 10. Monte Carlo simulations of the oscillation frequency.
Figure 10. Monte Carlo simulations of the oscillation frequency.
Electronics 14 02209 g010
Figure 11. Layout of three-stage CCRO.
Figure 11. Layout of three-stage CCRO.
Electronics 14 02209 g011
Table 1. Summary of design parameters employed in simulations.
Table 1. Summary of design parameters employed in simulations.
ParameterValue
VDD0.4 V
IBIAS70 # nA
(W/L)B1960 nm/28 nm
(W/L)B2200 nm/28 nm
A1, A233 dB
(W/L) Ni; Pi500 nm/20 nm; 600 nm/20 nm
Ci1 fF
#: This value will be changed over the frequency range.
Table 2. Corner analysis of the three-stage CCRO across different temperatures (TT—Typical–Typical, FF—Fast–Fast, SF—Slow–Fast, FS—Fast–Slow, SS—Slow–Slow).
Table 2. Corner analysis of the three-stage CCRO across different temperatures (TT—Typical–Typical, FF—Fast–Fast, SF—Slow–Fast, FS—Fast–Slow, SS—Slow–Slow).
Corner at T = 0 °CTTFFSFFSSS
Osc. Frequency (MHz)608.7919.8511.6572.8254.8
Average Power Consumption (µW)9.5115.9410.717.544.96
Phase noise @1 MHz (dBc/Hz)−83.43−81.43−82.49−83.81−84.97
Corner at T = 27 °CTTFFSFFSSS
Osc. Frequency (MHz)666.81149576.5745.3373.4
Average Power Consumption (µW)10.2316.0211.29.326.82
Phase Noise @1 MHz (dBc/Hz)−83.12−81.13−82.1−83.13−84.02
Corner at T = 60 °CTTFFSFFSSS
Osc. Frequency (MHz)751.91225598.3800.6496.9
Average Power Consumption (µW)11.2516.2710.879.336.87
Phase Noise @1 MHz (dBc/Hz)−82.71−80.69−81.65−82.47−83.19
Table 3. State-of-the-art comparison.
Table 3. State-of-the-art comparison.
Reference2023, [1]2016, [44]2019, [46]2024, [47]This Work
Tech. Node (nm)18065652822
Type of ControlVoltageVoltageVoltageVoltageCurrent
VDD (V)0.50.60.510.4
N-Stages43443
StructureDiff.Diff.Diff.Diff.Diff.
Osc. Frequency (MHz)3.59–13.74250–800340–1360734–975537–800
Power Consumption (µW)0.221146.2260322.610.23
Phase Noise @1 MHz
(dBc/Hz)
−95.01−86.38−90.2−83.78−83.12
FoM−154.32−153.2−158.7−148.47−161.1
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Abdullah, M.A.; Elamien, M.B.; Deen, M.J. A 0.4 V CMOS Current-Controlled Tunable Ring Oscillator for Low-Power IoT and Biomedical Applications. Electronics 2025, 14, 2209. https://doi.org/10.3390/electronics14112209

AMA Style

Abdullah MA, Elamien MB, Deen MJ. A 0.4 V CMOS Current-Controlled Tunable Ring Oscillator for Low-Power IoT and Biomedical Applications. Electronics. 2025; 14(11):2209. https://doi.org/10.3390/electronics14112209

Chicago/Turabian Style

Abdullah, Md Anas, Mohamed B. Elamien, and M. Jamal Deen. 2025. "A 0.4 V CMOS Current-Controlled Tunable Ring Oscillator for Low-Power IoT and Biomedical Applications" Electronics 14, no. 11: 2209. https://doi.org/10.3390/electronics14112209

APA Style

Abdullah, M. A., Elamien, M. B., & Deen, M. J. (2025). A 0.4 V CMOS Current-Controlled Tunable Ring Oscillator for Low-Power IoT and Biomedical Applications. Electronics, 14(11), 2209. https://doi.org/10.3390/electronics14112209

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