Next Article in Journal
Exploring the Effectiveness of Road Maintenance Interventions on IRI Value Using Crowdsourced Connected Vehicle Data
Previous Article in Journal
Rapid Diagnosis of Distributed Acoustic Sensing Vibration Signals Using Mel-Frequency Cepstral Coefficients and Liquid Neural Networks
Previous Article in Special Issue
Design and Realization of a High-Q Grounded Tunable Active Inductor for 5G NR (FR1) Transceiver Front-End Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Compact and Tunable Active Inductor-Based Bandpass Filter with High Dynamic Range for UHF Band Applications

1
Technopôle de Château Gombert, Aix Marseille Université, IM2NP UMR 7334, 13397 Marseille, France
2
Dynamics RF/mmW Consulting and Design, 39 avenue du Vercors, 38600 Grenoble, France
*
Author to whom correspondence should be addressed.
Sensors 2025, 25(10), 3089; https://doi.org/10.3390/s25103089
Submission received: 3 April 2025 / Revised: 10 May 2025 / Accepted: 11 May 2025 / Published: 13 May 2025
(This article belongs to the Special Issue Feature Papers in Electronic Sensors 2025)

Abstract

:
This paper presents a fully integrated bandpass filter (BPF) with high tunability based on a novel differential active inductor (DAI), designed for sensor interface circuits operating in the ultra-high frequency (UHF) band. The design of the proposed DAI is based on a symmetrical configuration, utilizing a differential amplifier for the feedforward transconductance and a common-source (CS) transistor for the feedback transconductance. By integrating a cascode scheme with a feedback resistor, the quality factor of the active inductor is significantly improved, leading to enhanced mid-band gain for the bandpass filter. To facilitate independent tuning of the BPF‘s center frequency and mid-band gain, an active resistor adjustment and bias voltage control are employed, providing precise control over the filter’s operational parameters. Post-layout simulations and process corner results are conducted with 0.13 µm CMOS technology at 1.2 V supply voltage. The proposed second order BPF achieves a broad tuning range of 280 MHz to 2.426 GHz, with a passband gain between 8.9 dB and 16.54 dB. The design demonstrates a maximum noise figure of 16.54 dB at 280 MHz, an input-referred 1 dB compression point of −3.78 dBm, and a third-order input intercept point (IIP3) of −0.897 dBm. Additionally, the BPF occupies an active area of only 68.2×30 µm2, including impedance-matching part, and consumes a DC power of 14–20 mW. The compact size and low power consumption of the design make it highly suitable for integration into modern wireless sensor interfaces where performance and area efficiency are critical.

1. Introduction

The rapid expansion of wireless communication systems has significantly transformed modern technological fields, driven by the increasing demand for high-speed, reliable, and scalable data transmission. The frequency range between 300 MHz and 3 GHz represents a crucial spectrum for a diverse set of radio frequency (RF) applications, owing to its favorable propagation characteristics and versatility.
This frequency range supports critical applications across communication, navigation, industrial, and consumer domains [1], and plays an essential role in enabling a broad spectrum of sensor-based systems and wireless sensing applications. In smart wireless communication systems, frequencies within this band are extensively employed in applications such as cellular networks [2], industrial telemetry [3,4], and experimental wireless platforms [5,6], many of which integrate embedded sensing capabilities for real-time data acquisition, environmental monitoring, and system diagnostics.
Cellular bands (e.g., 700 MHz, 850 MHz, 1.8 GHz, and 2.1 GHz) constitute the foundation of modern internet of things (IoT) and machine-type communication (MTC) infrastructures, supporting various sensor-driven applications including smart metering, asset tracking, remote health diagnostics, and infrastructure condition monitoring. Industrial telemetry systems operating within the UHF band (e.g., 400–470 MHz) are widely deployed for remote sensing of pressure, temperature, and vibration in utility grids, manufacturing environments, and energy facilities. Furthermore, the globally unlicensed industrial, scientific, and medical (ISM) band at 2.4 GHz enables widely adopted short-range communication protocols such as Wi-Fi, Bluetooth, BAN, Wireless HART, Z-wave, and ZigBee, which are the foundation of wireless sensor networks (WSNs) used in building automation, smart agriculture, healthcare wearables, and industrial IoT systems. Experimental sensor deployments have also leveraged amateur radio bands (e.g., 433 MHz, 902–928 MHz), particularly in educational and research contexts for applications like remote weather stations, aerial sensing with balloons and drones, and emergency telemetry. This frequency range also supports sensor-integrated systems in navigation and radar applications. For instance, global navigation satellite systems (GNSS) operating in the L-band (1.2–1.6 GHz) provide high-precision positioning for sensors in autonomous vehicles, drones, and geodetic instruments. Similarly, weather and maritime radar systems in the S-band (2.7–3.0 GHz) deliver real-time environmental data essential for climate monitoring, air traffic control, and disaster early warning systems. Additionally, low-power wide-area networks (LPWANs) operating in sub-GHz bands (e.g., 433 MHz, 868 MHz, and 915 MHz) offer long-range, low-energy solutions for distributed sensing networks in smart cities, logistics, environmental sensing, and precision farming [7]. The military and aerospace sectors further exploit this spectrum for tactical sensor networks, including target tracking, surveillance, and secure telemetry systems.
With the ongoing development of wireless communication technology, the design of reconfigurable RF integrated circuits (ICs) has become a crucial focus area for next-generation wireless sensing platforms. One of the most essential components of every wireless transceiver is the bandpass filter (BPF). Figure 1 presents the transceiver front-end circuit positioning a BPF as the first component in the processing chain. The BPF is directly integrated with the antenna without the inclusion of intermediary components, ensuring seamless and unimpeded interaction with the received signal. Upon reception, the antenna transmits the sensed signal to the BPF, which either selectively passes the desired frequency range or attenuates unwanted signals outside the target bandwidth. This filtering process ensures the integrity of the signal passed to the subsequent processing stages.
Consequently, the performance characteristics of the BPF, such as its bandwidth, insertion loss, and out-of-band rejection, significantly influence the overall performance of both the receiver and the transmitted signal [8].
Several integrated RF-BPF filter topologies have been reported in the literature, including MOSFET-C filters [9], active-RC filters [10], switched-capacitor filters [11,12], Gm-C filters [13], and Q-enhanced LC filters [14]. Each filter’s configuration suffers from one or more limitations that restrict their use in gigahertz-frequency integrated applications. The continuous-time filters based on MOSFET-C filters and active-RC integrators have limited high-frequency operation due to the finite gain-bandwidth product (GBWP) of the operational amplifiers (op-amps) used in the feedback circuit. Meanwhile, switched-capacitor filters suffer from clock feedthrough problems.
Among these, Gm-C and Q-enhanced LC filters have shown greater promise for implementation at high frequencies. Q-enhanced LC topology utilizes lossy LC-tank resonators and improves their effective Q-factor through negative resistance compensation circuits. However, their practical implementation at RF frequencies is hindered by challenges such as the bulky nature of inductors, limited scalability, moderate quality factors of on-chip monolithic inductors (typically below 12) [15], and various parasitic losses, including ohmic losses, substrate-induced eddy currents, and resistive losses. Furthermore, these filters suffer from limited dynamic range and tunability, as well as a dependency on the Q-factor of the tank inductors, which directly impacts overall filter performance. Efforts to improve the Q-factor of monolithic inductors (Q < 40) using fabrication process modifications have shown some promise. However, such methods often increase manufacturing complexity and costs, and reduce the reliability of large-scale production [16,17,18].
To overcome the challenges associated with monolithic inductors in standard silicon processes, active inductors (AIs) have become increasingly preferred by many RF circuit designers, owing to their numerous advantageous properties. Compared to monolithic inductors, active inductors offer several advantages, including a high and tunable quality factor, a higher self-resonance frequency, and a reduced chip area.
Active inductors can be categorized into three primary types: amplifier-based AIs [19,20], current conveyor-based AIs [21,22], and gyrator-C AIs [23]. The first two types require many passive and active components, which often result in larger area requirements. Additionally, these AIs are associated with drawbacks such as limited tunability, significant power dissipation, and a constrained frequency range. On the other hand, gyrator-C-based AIs provide several advantages, including a higher Q-factor, greater self-resonance frequency, improved tunability, and reduced size [24].
Various studies have introduced techniques to enhance the performance of gyrator-C-based active inductors. These methods target improvements in key parameters, such as the operating frequency range, tunability, Q-factor, power efficiency, and linearity. Leveraging a cascode or regulated cascode configuration [25,26] in the feedback path enhances the Q-factor and extends the frequency range of AIs by suppressing zero frequency. Exploiting a feedback resistor technique [27,28] leads to an increase in Q-factor and inductance simultaneously. Flipped and cascode-flipped [29,30,31], additive capacitor [32], DC level shifter [33], double feedback [34], and distortion reduction [35] are among the various techniques reported in the literature to enhance the performance of AIs. Class AB active inductors were introduced in [36,37]. However, these inductors consume significant power and are inherently noisy.
For optimal performance in radio frequency applications, differential active inductor structures are superior to single-ended designs due to their wider dynamic range and higher quality factor [38]. Single-ended AIs lack symmetry, resulting in inconsistent port characteristics that deviate from ideal inductor behavior. Despite progress, many discussed active inductor designs in literature fail to simultaneously achieve desired key specifications, such as compactness, low power consumption, high Q-factor, tunability, and low noise. Thus, creating a robust active inductor that integrates these features into a single, efficient design remains a critical challenge.
This paper presents a novel DAI with a focus on its 2nd-order tunable BPF implementation for UHF band applications using 0.13 µm CMOS technology. By incorporating several design techniques, the DAI achieves enhanced performance metrics, including an improved Q-factor, broader frequency range, and greater tunability. The proposed differential structure distinguishes itself from existing designs in the literature. To validate its effectiveness, a reconfigurable BPF is designed and simulated. This filter achieves a wide center frequency range from 280 MHz to 2.426 GHz, enabling compatibility with various multi-standard, sensor-based communication modules such as LTE, GSM, Bluetooth, Wi-Fi, ZigBee, LoRa, and others. The transistors used in our design were sized and biased in the saturation region following the gm/ID methodology.
The remainder of this paper is structured as follows. Section 2 describes the basic concept of a gyrator-C-based active inductor and describes the design of the novel DAI. Section 3 deals with tunable second-order BPF design using the proposed DAI in Section 2. Section 4 focuses on the simulation results of the BPF, along with a performance comparison to other reported works, where process—voltage—temperature (PVT) analyses are also presented. Section 5 concludes and discusses the proposed work.

2. Differential Active Inductor Design

2.1. Basic Concept of Differential Active Inductors

To create a floating active inductor, the key principle is to design a circuit that satisfies the short-circuit admittance matrix equation for a two-port network,
I 1 I 2 = s L e q 1 1 1 1 V 1 + V 1 V 2 + V 2
where Leq is the equivalent inductance value of the floating active inductor.
A gyrator can be created by incorporating two operational transconductances amplifier (OTA) connected back to back in single loop. One OTA is positioned in the feedback-path, which provides a negative transconductance and another in the forward path, which provides a positive transconductance. When a capacitor is connected to the output of a gyrator, the resulting configuration is known as a gyrator-C network. This network leverages the intrinsic properties of the gyrator to transform the capacitive impedance into an inductive impedance, effectively emulating the behavior of a physical inductor. Figure 2 depicts the structure of a lossy differential gyrator-C network.
KCL equations at nodes 2+ and 2− show that the input admittance Yin can be given as follows:
Y i n = I i n V 2 + V 2 = s C 2 2 + G o 2 2 + 1 s C 1 2 G m 1 G m 2 + G o 1 2 G m 1 G m 2
where go1, and go2, are the total conductances at node 1 and node 2, respectively.
Since the input/output impedances of the transconductors in the gyrator-C network are inherently limited, the synthesized equivalent inductor exhibits a lossy behavior. Equation (2) is characterized by the presence of parasitic components, including resistance and capacitance, which degrade the ideal inductive response. The behavior of the lossy inductor can be modeled by the equivalent RLC network, as illustrated in Figure 2, and its equivalent circuit parameters are derived from the small-signal analysis of the gyrator-C network. These parameters are calculated as follows:
C p = C 2 2 , R p = 2 G o 2 , L e q = C 1 / 2 g m 1 g m 2 and R s = G o 1 / 2 g m 1 g m 2
where Rp is the parallel resistance, Cp is parallel capacitance, Rs is the series resistance, and the Leq equivalent inductance.
The corresponding resonance frequency and Q-factor are given by matching the real and imaginary parts equations
ω 0 = 1 L e q C p = 2 g m 1 g m 2 C 1 C 2
Q ω L e q R s R p R p + R s 1 + ω L e q R s 2 1 R s ² C p L e q ω ² L e q C p
Due to the presence of parasitic components, originating from the intrinsic characteristics of MOS transistors, the input admittance is not purely inductive behavior over all the entire frequency range. To achieve low ohmic losses, it is essential to maximize the parasitic parallel resistance Rp, thereby reducing energy dissipation through parallel leakage paths, and to minimize the parasitic series resistance Rs, which reduces resistive losses and enhances the overall performance.
To overcome the loss limitations, a feedback resistance is frequently employed to mitigate the parasitic series resistance, which adversely affects the Q-factor of the active inductor [39]. This approach introduces a parallel effect on both the series resistance and the inductance. Consequently, the feedback resistance simultaneously decreases the series loss resistance and enhances the emulated inductance Ls, theoretically driving Q toward infinity. Moreover, the feedback mechanism serves to fine-tune the phase of the input admittance and the self-resonant frequency (SRF), thereby optimizing the active inductor’s operating frequency at which Q achieves its maximum value.
A regulated cascode configuration is also used by many designers to take advantage of the benefits of series-connected transistors, as described in [40]. This approach effectively solved the main limitation of the active inductor, which arises from the conductance go1 at node 1. By adopting the cascode topology, the quality factor can be significantly improved and help reduce both the power consumption and supply voltage requirements of the active inductor. Furthermore, the transconductance of other transistors can be used to change the emulated inductive value.
Indeed, active inductors involve a tradeoff between SRF, equivalent inductance, power consumption, and Q-factor. To achieve low power consumption and a high SRF, a smaller input transistor is required. Conversely, a larger input transistor is essential for attaining a higher quality factor and greater inductance value, highlighting the design compromises necessary for optimizing performance [41].

2.2. Implementation of the Proposed Differential Active Inductor

The configuration of the proposed differential active inductor, computed using the design custom method, is shown in Figure 3. It comprises two grounded active inductors coupled through two cross-coupled transistor pairs that form negative resistances. Each inductor consists of three transistors M1, M2, M3 for AI-1, and M15, M16, M17 for AI-2, providing the basic inductance functionality.
Two controllable current mirror sources are employed for polarization, modeled by transistors M18 to M22 for active inductor (1) and M4 to M8 active inductor (2). To improve the Q-factor and the inductance value of each active inductor, cascode stages, implemented with transistors M3 and M17, are utilized as gain-boosting stages. These cascode configurations effectively reduce the output conductance gds of transistors M2 and M16, respectively, leading to enhanced performance. Furthermore, two controllable feedback resistors, Rf, are incorporated to optimize the Q-factor while introducing a negative conductance to compensate for the parasitics and losses inherent in the AIs. These resistors are implemented using PMOS transistors M9 and M23, biased in their ohmic regions. The negative resistance is generated through a double cross-coupled PMOS/NMOS differential pair M10, M24 and M11, M25, which delivers a negative resistance of −2/gm, where gm represents the transconductance of each cross-coupled transistor.
The operation of the tunable active inductor is as follows. When a differential input voltage is applied to the gates of the common-source transistors M1 and M15, the transconductances gm1 and gm15 convert the voltage into a drain current, charging the gate capacitances Cgs2 and Cgs17 of transistors M2 and M17, respectively. The voltages developed across Cgs3 and Cgs16 are subsequently converted into input currents by the transconductances of transistors M3 and M16, respectively, emulating the voltage-current characteristics of a shunt inductance.
Due to the symmetry of the circuit topology, the simplified small-signal equivalent half-circuit is presented in Figure 4.
By assuming Cdsi ≪ Cgsi, the input admittance Yin can be derived as:
Y i n = 1 Z i n = g m 9 2 g m 10 2 + R f + r d s 9 + r d s 10 2 + s C g s 1 + s C g s 9 + s C g s 10 2 + g m 2 1 + R f + r d s 9 + r d s 10 s C g s 1
where Cgsi is the gate-source capacitance, and gmi and gdsi are the transconductance and output conductance of the i-th MOS transistor, respectively. The equivalent RLC model parameters of the DAI are derived as follows:
L e q = 2 C g s 1 R f + 1 / g d s 2 + 1 / g d s 3 g m 1
C p = C g s 1 + C g s 9 + C g s 10 2
G p = g m 9 + g m 10 2 + R f + 1 / g d s 2 + 1 / g d s 3 2
R s = 2 g m 1
The inclusion of a feedback resistance Rf introduces a positive contribution to both the resistance and inductance in the RLC equivalent model of the proposed active inductance. By carefully selecting the value of Rf, the equivalent inductance Leq can be significantly enhanced, while the parallel resistance Rp can be maximized. Furthermore, maintaining a fixed transconductance for transistor M1 effectively reduces the equivalent series resistance Rs of the active inductor. Additionally, Leq can be independently tuned by adjusting Rf or modifying the conductance of other transistors, providing enhanced flexibility in circuit optimization.
The self-resonant frequency of the DAI, defined as the frequency at which the imaginary part of the input admittance becomes zero, can be calculated as:
ω 0 - D A I 1 L e q C p = 1 R f + r d s 2 + r d s 3 g m 1 C g s 1 2 + C g s 9 + C g s 10
The Q-factor of the DAI is defined as the ratio of the real part to the imaginary part of the input admittance Yin and is expressed as:
Q D A I = ( 1 / Y i n ) ( 1 / Y i n ) R p ω L e q = g m 1 ω C g s 1 R f + r d s 2 + r d s 3 R f + r d s 2 + r d s 3 g m 9 g m 10
where g e q = 1 / R f + r d s 2 + r d s 3 .
Based on Equation (12), by selecting Rf + rds2 + rds3 to be approximately equal to − gm9′ − gm10′, the parallel resistance Rp approaches infinity. Consequently, the Q-factor at frequencies below the self-resonant frequency can become significantly large.
According to the above equations, the characteristics of the proposed DAI can be independently adjusted with minimal mutual interference. Variations in bias currents provide an additional mechanism for controlling both the inductance Leq and the Q-factor QDAI. By varying the transconductance, an additional capacitance is introduced into the node described in Equation (6), offering further tuning capabilities for the proposed circuit.

3. Proposed Tunable Bandpass Filter

3.1. Bandpass Filter Design Methodology

In this section, the implementation of an RF bandpass filter using the proposed DAI is presented, with the intent of improving filter performance and minimizing chip area. The circuit diagram of the differential BPF including the compensation and tuning circuitry is illustrated in Figure 5, where a second order filter architecture was designed. It incorporates the required degrees of freedom to facilitate tuning the desired center frequency, bandwidth, noise figure, and power consumption.
The circuit comprises four RF stages, with the first stage being a differential input buffer. This buffer converts the differential input RF voltages, VRFin+ and VRFin−, into a current, which is then applied to the RLC network represented by the tunable DAI.
This block is implemented using two common source-follower NMOS transistors, M12 and M26, connected to two resistors R. The second stage, which forms the filter core, comprises the AIs connected via two cross-coupled negative resistances. The negative resistance reduces resistive losses, thereby increasing and controlling the Q-factor of the circuit. The third stage is a differential output buffer configured in a source-follower topology, consisting of transistors M13 and M27, along with two controlled gate transistors, M14 and M28. This stage is designed to drive the resistive load while preventing the load from significantly degrading the BPF’s performance. Device dimensions, i.e., width W, length L, as well as the number of fingers Nf, of all the devices (M1 through M26) for the simulation of the proposed DAI and BPF are reported in Figure 5. A multi-finger MOSFET layout technique [41,42] was employed to optimize these parameters, aiming to improve performance while minimizing parasitic effects, reducing power consumption and ensuring better matching and layout efficiency [43].
For optimal matching of input and output impedance, the transistors in the input buffer M12 and M26 were sized with a W of 70 μm, L of 0.13 μm, and Nf = 2, after optimization. To achieve input-impedance matching of 50 Ω, a resistor R with a value of 40 Ω is incorporated, eliminating the need for additional capacitors or inductors. In the output stage, a second impedance-matching circuit is crucial to ensure maximum transfer power and to mitigate the impact of the load on the filter’s frequency response and quality factor. To match the output impedance to 50 Ω, a 600 fF metal-insulator-metal (MIM) capacitor is integrated into the output buffer. Furthermore, the output stage transistors M13 and M27 are sized with a W of 60 μm and Nf = 2. The gate-to-source voltage Vgs of the cascode stages (M2, M3, and M16, M17) is set to ground (GND) to simplify the design.
The resonant frequency can be adjusted by varying the control voltages Vres and Vbias2. The resonant frequency of the BPF is expressed as:
ω B P F = 1 L C p + C p a r
where Cp ≈ Cgs1 + Cgs1′, and Cpar represents the parasitic capacitance contributed by the input Cgd26 and output buffers Cgs27. Considering the assumption Cgd ≪ Cgs, the parasitic capacitance is approximated as Cpar = Cgs27 + Cgs27′.

3.2. Theoretical Analysis

Figure 6 illustrates the configuration of the bandpass filter, including the RLC equivalent network of the DAI. To minimize the equivalent parasitic capacitance and resistance in the active inductor, a negative impedance transformation −RNEG is employed. Referring to Figure 6, the gain-bandwidth (GBW) of the bandpass filter, derived from the overall model, can be approximated as:
A 2 = g m G p G N E G
Therefore, the Q-factor of the BPF is obtained as:
Q B P F = ω 0 - D A I B W = 1 G p G N E G L e q ω 0 - D A I = Q D A I 1 / R p 1 / R p 1 / R N E G
In order to achieve optimal compensation condition, the parallel resistance Rp must equal the magnitude of the negative resistance ∣RNEG∣, whereas the stability condition of the BPF is satisfied when Gp > ∣RNEG∣.

4. Simulation Results and Discussion

The proposed DAI and resulting BPF were designed and implemented using 130 nm CMOS technology from STMicroelectronics. The circuit operation and performance was evaluated using the Virtuoso ADE© environment from Cadence IC.6.1.9©, with parasitic effects taken into account. A comprehensive set of simulations was performed to evaluate the performance parameters and their robustness under various operating conditions. These simulations encompassed DC analysis, S-parameter evaluation, noise characterization, periodic steady-state (PSS) analysis, process corner evaluation, and Monte Carlo tests, ensuring a thorough assessment of the design’s behavior across diverse environments. It is worth noting that all simulation results and graphical data were obtained using Cadence Virtuoso and subsequently processed and visualized using Origin Lab©. The simulated performance results of the proposed active BPF topology are summarized in Table 1.

4.1. Differential Active Inductor Performances

In the first phase of the analysis, the performance of the proposed differential active inductor was evaluated by analyzing its key characteristic parameters, including input impedance Zin, effective equivalent inductance Leq, Q-factor QDAI, and self-resonance frequency.
The input impedance Zin was composed of a resistive component {Zin} and an inductive reactance {Zin}, which were determined using a two-port S-parameter analysis. The frequency dependence of Zin, {Zin}, and {Zin} for the differential active inductor is depicted in Figure 7. Additionally, the inductive behavior of the proposed DAI was verified by analyzing the phase response under varying Rf. According to [44], the emulated inductance value Leq and Q-factor QDAI can be derived from the input impedance Zin as follows:
Z i n = Z 11 Z 21 Z 12 + Z 22
Q D A I = Z i n Z i n L e q = Z i n ω
The variation in Rf is achieved by adjusting Vres from −0.5 V to −1.2 V. Under these conditions, Leq was found to range between approximately 33 nH and 98 nH. Additionally, the circuit exhibited self-resonance frequencies of up to 3.3 GHz.
As illustrated in Figure 8, QDAI was optimized to exceed 20 across the entire operating frequency band. Notably, a peak quality factor of approximately 388 was achieved at 2.31 GHz when Vres was set to −0.7 V. This high degree of tunability in the Q-factor is a significant advantage of the proposed design.
To further validate the inductive behavior of the circuit, a detailed analysis of its phase characteristics was conducted. As shown in Figure 9, when Vres was set to −0.7 V, the phase approached 89.85° across the entire operating frequency range, confirming that the circuit maintained its inductive behavior within this band. For frequencies beyond 3.3 GHz, the reactance component of Zin becomes negative, indicating a transition to capacitive behavior.
The linearity of the differential AI has been investigated in prior studies [45,46]. Intermodulation tests were conducted using a two-tone input signal at 1.997 GHz and 2 GHz, yielding an input third-order intercept point (IIP3) of −9.71 dBm. The 1 dB compression point (P1dB) for the proposed DAI was determined to be 2.25 dBm, corresponding to a voltage swing of 0.82 V at the inductor’s differential input with a 1.2 V supply voltage.

4.2. Bandpass Filter Performances

To achieve a wide frequency tuning range, two control voltages were utilized. The first control voltage was adjusted by regulating Vres, which controls the feedback transistors M9 and M23. The second control voltage was implemented through controllable current sources transistors M12 and M26 via Vbias2. These control mechanisms were calibrated to minimize the noise figure while achieving high gain and low power consumption. Note that Vres was employed for coarse frequency tuning, whereas Vbias2 enabled fine frequency adjustments. The supply voltage remained fixed at 1.2 V, with the tuning range of Vres and Vbias2 varying from −1.2 V to 0.2 V, and from 0.5 V to 0.8 V, respectively. Meanwhile, the bias voltages Vbias and Vbias3 were fixed at 0.5 V, providing an optimal compromise between power consumption and compensation.
In Figure 10, the simulated S-parameters of the filter are presented under its fourth tuning state (see Table 1), providing valuable insight into the circuit’s performance. This piece of detail is usually missing in the state-of-the-art of active BPF. However, their behaviors were imported for evaluating the stability characteristics and the input and output impedance matching of the design. The proposed BPF is centered at 1.228 GHz and exhibits an input reflection coefficient (S11) and an output reflection coefficient (S22) of −33 dB and −31 dB, respectively, indicating excellent impedance matching and optimal power transfer across the filter’s passband.
At this center frequency, the filter reached a 3-dB BW of approximately 400 MHz, which aligns with the expected behavior of a 2nd-order active BPF in the literature. In contrast, various previously reported 2nd-order BPFs utilizing active inductors achieve narrow 3-dB BW by incorporating a high value of negative series resistance. Although this technique effectively reduces the bandwidth, it often introduces practical stability concerns and significantly increases the risk of oscillation. Therefore, to ensure reliable operation, an active inductor must always maintain unconditional stability throughout its intended operating range.
Figure 11 illustrates the variation in the transmission coefficient S21 response. The results demonstrate that the center frequency fc can be effectively tuned from 0.28 GHz, to 2.426 GHz, by adjusting the values of Vres and Vbias2 according to the configuration settings listed in Table 1, spanning Config 1 to Config 9. A maximum forward transmission coefficient (S21) of 16.54 dB was achieved at the center frequency of 0.28 MHz, i.e., Config 1, indicating optimal signal transmission.
The resulting −3 dB bandwidth (BW) of the proposed design ranged from 287 MHz to 406 MHz. Additionally, the medium gain of the filter enabled it to achieve an effective Q-factor of 6.4 at a center frequency of 1.83 GHz.
A comprehensive noise analysis of the proposed topology was performed, including its dynamic range, evaluated in terms of the noise figure (NF). Figure 12 depicts the NF as a function of frequency, revealing that the NF varied from 18.41 dB at 1.57 GHz to 24.6 dB at 280 MHz. In comparison to other active bandpass filters reported in the literature [47,48], these NF values remain well-suited for practical applications.
The results obtained for various combinations of control voltages Vres and Vbias2 are summarized in Table 1.
The linearity of the BPF was evaluated using the 1 dB compression point (P1dB) and the third-order intercept point (IIP3). To assess its performance, two-tone signals with a spacing of ±2.5 MHz were applied. As shown in Figure 13a,b, the P1dB distortion point was achieved at an input power of −3.78 dBm, while the IIP3 was measured at approximately −0.897 dBm. These results demonstrate that the proposed active filter exhibits high linearity, making it well-suited for multiband applications in integrated circuit design.
In this part, the analysis results of the proposed design regarding alternative stability factor B1f and Rollet stability factor Kf responses are depicted in Figure 14.
As defined by Equations (11) and (12), a two-port circuit is considered unconditionally stable if B1f > 0 and Kf > 1.
B 1 f = 1 Δ 2 + S 11 2 S 22 2 > 0
K f = 1 Δ 2 + S 11 2 S 22 2 2 S 11 S 21 > 1
From the plot shown in Figure 14, the stability analysis of the proposed filter confirms a sufficient stability margin across the operational frequencies and beyond, ensuring an unconditionally stable design. Specifically, B1f was approximately 0.966 (greater than 0), while Kf remained consistently greater than 1 over the entire frequency band.

4.3. Monte Carlo Analysis

In the nanoscale area, several factors contribute to the degradation of MOS device performance [49,50], including oxide thickness variation (OTV), metal-gate work-function fluctuations (WKF), random dopant fluctuations (RDF), and line-width roughness (LWR).
To investigate these phenomena and analyze their impact on circuit performance, process-sensitive device parameters such as doping concentration, channel width W, channel length L, and gate oxide thickness tox were incorporated into a Monte Carlo simulation analysis. The variability in the bandpass filter’s forward transmission coefficient S21, noise figure NF, Rollet stability coefficient Kf, and alternative stability factor B1f was estimated using 1000 Monte Carlo simulation runs. This approach ensured a reduction in the standard deviation (σ) of the estimated values to below 5%, enhancing the accuracy of the variability analysis.
Table 2 summarizes the sensitivity analysis under mismatch and process variation obtained from statistical histograms. The Gaussian distributions at ±3δ level obtained from the Monte Carlo analysis indicate that approximately 99.7% of the simulated values for the forward gain S21 fell within the range of 13.67 to 15.94 dB, in distribution of the mean value (µ), i.e., 14.9 dB. Similarly, NF maintained a stable mean of 19.94 dB with a majority of samples concentrated between 19.81 and 20.04 dB, reflecting excellent consistency. Furthermore, Kf and B1f parameters exhibited average values of 9.82 and 0.998, respectively, with corresponding variations of 4.93% and 0.09%. Both metrics consistently remained well above the critical stability thresholds throughout all Monte Carlo runs, indicating that the circuit maintained unconditional stability even under worst-case mismatch and process variation scenarios. The Monte Carlo analysis demonstrates that the proposed circuit exhibited tightly bounded variation under mismatch and process fluctuations, confirming its robustness, high reliability, and suitability for stable, high-yield integration in practical CMOS-based RF systems.

4.4. Process Corner Analysis

Figure 15 illustrates the forward transmission coefficient S21, which maintained a value of 14.9 dB under three process variation scenarios: typical NMOS-typical PMOS (TT), fast NMOS-fast PMOS (FF), and slow NMOS-slow PMOS (SS). The best-case scenario occurred under the FF corner at 1.1 V and a temperature of −40 °C, while the worst-case scenario was observed under the SS corner at 0.9 V and a temperature of +125 °C.
The forward transmission coefficient S21 of the filter maintained an acceptable value (>12 dB) even under varying conditions, ensuring it remained sufficient for the target BPF’s gain. However, the center frequency fc demonstrated significant sensitivity to voltage and temperature variations, exhibiting a shift of approximately 34% across all process corners. This degradation is primarily attributed to fluctuations in the reference current. These variations can be effectively mitigated by adjusting the bias current generator or incorporating a temperature-compensated current reference [51]. This underscores the robustness of the proposed design for integration into wireless communication devices and systems.

4.5. Layout and Performance Comparison

The layout of the proposed BPF is depicted in Figure 16. It occupies a total area of 2046 µm2 (30 µm × 68.2 µm), including the output-matching impedance capacitor. The circuit’s DC power consumption ranges between 14.1 mW and 20.3 mW, operating at a nominal supply voltage of 1.2 V.
The design metrics of the proposed BPF are summarized in Table 3 and compared against the state-of-the-art active BPFs reported in the literature. The dynamic range (DR) of the bandpass filter, as defined in [48], is expressed as:
D R = P 1 d B P n = P 1 d B k T F
where P1dB represents the input 1-dB compression point in dBm, and Pn is the filter’s noise power relative to a 1-Hz bandwidth, given by Pn = kT. Here, kT (thermal noise floor) is −174 dBm/Hz and F is the noise factor.
To evaluate and compare the performance of bandpass filters, a figure-of-merit (FoM) was utilized, where a higher FoM indicates superior filter performance. Two widely adopted definitions of FoM, as detailed in [52,53,54,55], are given by:
F o M 1 = 1 P D C P 1 d B P n
and
F o M 2 = P 1 d B W 𝑓 B P F Q B P F N P D C N F
where P1dB and P1dBw denote the 1-dB compression points in dBm and watts, respectively, PDC is the DC power consumption in mW, Pn is the noise power corresponding to a 1-Hz bandwidth in dBm, QBPF is the quality factor, fBPF is the resonant frequency of the filter, N is the number of poles, and NF is the noise figure (expressed as a unit less ratio rather than in decibels).
Table 3 highlights that the proposed bandpass filter (BPF) achieved a first figure of merit (FoM1) of 140.3 dB-Hz/mW, positioning it among the top-performing designs. Although slightly lower than those reported in [39,56], the difference is primarily attributed to the lower power consumption achieved in those works. Nonetheless, the proposed architecture demonstrated outstanding performance, including the widest tuning range and smallest chip area among the compared works, at an expense of a moderate noise figure. Furthermore, the design maintained reasonable power consumption, reinforcing its overall efficiency and competitiveness. The second figure of merit (FoM2), which evaluated frequency selectivity, power efficiency, linearity, and noise performance, reached 73 dB, making it comparable to existing implementations in the literature.
Table 3. Performance comparison inductor-based BPF.
Table 3. Performance comparison inductor-based BPF.
Reference[39][55][56][57][58][59]Our Work
Parameter
Filter order (N)2222212
Inductor TopologyGyrator-C active inductor
SRF (GHz)1.16–3.275.15–5.351.6–2.91–23.02–3.760.1–1.30.28–2.426
−3dB BW (MHz)6515801000-300294–406
VDD (V)1.01.21.81.220-1.2
Inductor Q300–794964--12-388
Gain S21 (dB)26.62–33.45155.750.5−12.2–13.708.89–16.54
PDC (mW)4.04–6.447.7163050612014.1–20.3
P1dB (dBm)2.72−2.8−7−21−11.6553.78
NF (dB)14.48–16.5621.7619N/A1118.41–24.60
DR (dB-Hz)161.7 a149.5 a161 a 134 a-82151.8
FoM1 (dB-Hz/mW)153.9 a140.6 a153.1 a119.2 a-61.2 a140.3
FoM2 (dB)92.7 a92.2 91 a41.8 a-64 a73
Area (mm2)0.017-0.0230.180.062-0.002 b
Sim./Meas.Sim.Sim.Sim.Meas.Meas.Meas.Sim.
Process
Technology
130 nm CMOS40 nm CMOS180 nm CMOS130 nm CMOS0.5 µm pHEMT GaNN/A130 nm CMOS
a Calculated based on data available in article. b Including impedance matching circuit.

5. Conclusions

In this paper, a reconfigurable and a fully differential active inductor has been introduced for UHF band applications, enabling the implementation of a robust tunable second-order bandpass filter using 130 nm CMOS technology with a 1.2 V supply voltage. Implementing feedback resistor and cascode transistor elements enhanced the quality factor of the active inductor, ensuring superior performance and reliability. The post-layout simulation results demonstrate that the proposed BPF design achieved a wide frequency tuning range from 280 MHz to 2.246 GHz, enabled by two control voltages. The design further exhibited high linearity, good selectivity, low power consumption, and a compact layout, making it a highly efficient and practical solution for advanced RF and sensor-oriented applications.
Future advancements in active inductor design present substantial opportunities for enhancing filter performance in sensor interface circuits. While the use of a cascode structure and series feedback resistor successfully minimizes losses and improves the quality factor, these features may introduce stability issues and limit the tunability of inductance, especially in high-frequency scenarios. Adjusting the current bias magnitude could address these challenges and optimize performance across a wide range of conditions. However, increasing the bias current leads to higher power consumption, while lowering it may result in greater losses due to elevated series resistance RS. In future sensor-based systems, a focus on managing distortion under high input power conditions will be crucial. The nonlinearity of the inductance, arising from the limited transconductance in the feedback path, will need to be addressed. This nonlinearity could also restrict performance in low-voltage, high-swing environments, where stacked transistors must operate in strong inversion to meet high-frequency demands. Employing level shifters in such cases could provide an effective solution, ensuring that stacked devices in the feedback path function reliably. Overcoming these challenges and balancing the associated trade-offs will be critical in satisfying the demands of next-generation sensor technologies.

Author Contributions

Conceptualization, S.S. and A.B.H.; methodology, A.B.H.; software, S.S.; validation, S.S, F.H., and A.B.H.; formal analysis, S.S.; investigation, F.H.; resources, S.S. and F.H.; data curation, S.S.; writing—original draft preparation, S.S. and F.H.; writing—review and editing, S.S. and F.H.; visualization, S.S. and A.B.H.; supervision, F.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Acknowledgments

The authors would like to think Fayrouz Haddad for their extraordinary support in this research project and for providing the essential resources that facilitated the preparation of this manuscript.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
BPFBandpass Filter
DAIDifferential Active Inductor
UHFUltra-High Frequency
CSCommon Source
CMOSComplementary metal-oxide-semiconductor
IIP3Third-Order Input Intercept Point (IIP3)
DCDirect Current
RFRadio Frequency
IoTInternet of Things
MTCMachine-Type Communication
ISMIndustrial, Scientific, and Medical
Wi-FiWireless Fidelity
BANBody Area Network
HARTHighway Addressable Remote Transducer
ZigBeeZonal Intercommunication Global Standard
LPWANLow-Power Wide Area Networks
WSNWireless Sensor Network
GNSSGlobal Navigation Satellite System
ICIntegrated Circuit
LNALow-Noise Amplifier
LPFLow Pass Filter
VGAVariable Gain Amplifier
DACDigital Analog Converter
ADCAnalog Digital Converter
PDPhase Detector
CPCharge Pump
VCOVoltage Controlled Oscillator
FDFrequency Divider
MOSFETMetal-Oxide-Semiconductor Field-Effect Transistor
GBWPGain-Bandwidth Product
Op-AmpsOperational Amplifiers
DTTDigital Terrestrial Television
LTELong-Term Evolution
GSMGlobal System for Mobile communication
LoRaLong Range Wide Area
PVTProcess Voltage Temperature
OTAOperational Transconductance Amplifier
SRFSelf-Resonant Frequency
MIMMetal Insulator Metal
GBWGain Bandwidth
PSSPeriodic Steady State
IIP3Input Third-Order Intercept Point
BWBandwidth
NFNoise Figure
OTVOxide Thickness Variation
WKFMetal Gate Work Function Fluctuations
RDFRandom Dopant Fluctuations
LWRLine-Width Roughness
DRDynamic Range
FoMFigure of Merit

References

  1. ITU. Utility Radio Communications Operating in the Land-Mobile Service; ITU: Geneva, Switzerland, 2018. [Google Scholar]
  2. Aragón-Zavala, A.; Angueira, P.; Montalban, J.; Vargas-Rosales, C. Radio propagation in terrestrial broadcasting television systems: A comprehensive survey. IEEE Access 2021, 9, 34789–34817. [Google Scholar] [CrossRef]
  3. Więcek, D.; Mora, M.; Michalski, I. Coexistence of 5G/6G and Digital Terrestrial Television Networks. Appl. Sci. 2024, 14, 5756. [Google Scholar] [CrossRef]
  4. Graham, A.; Kirkman, N.C.; Paul, P.M. Mobile Radio Network Design in the VHF and UHF Bands: A Practical Approach, 1st ed.; John Wiley & Sons: Hoboken, NJ, USA, 2007. [Google Scholar]
  5. Quintana-Diaz, G.; Nodar-Lopez, D.; Muíño, A.G.; Agelet, F.A.; Cappelletti, C.; Ekman, T. Detection of radio interference in the UHF amateur radio band with the Serpens satellite. Adv. Space Res. 2022, 69, 1159–1169. [Google Scholar] [CrossRef]
  6. Xu, Z.; Liu, J.; Zhang, W.; Zhou, S.; Waug, W. Capacity analysis and comparison for cellular systems over different frequency bands. In Proceedings of the IEEE International Conference on Network Infrastructure and Digital Content (IC-NIDC), Beijing, China, 23–25 September 2016; pp. 135–138. [Google Scholar]
  7. Mekki, K.; Bajic, E.; Chaxel, F.; Meyer, F. Overview of cellular LPWAN technologies for IoT deployment: Sigfox, LoRaWAN, and NB-IoT. In Proceedings of the IEEE International Conference on Pervasive Computing and Communications Workshops (Percom Workshops), Athens, Greece, 19–23 March 2018; pp. 197–202. [Google Scholar]
  8. Al-Yasir, Y.I.A.; Ojaroudi Parchin, N.; Abd-Alhameed, R.A.; Abdulkhaleq, A.M.; Noras, J.M. Recent Progress in the Design of 4G/5G Reconfigurable Filters. Electronics 2019, 8, 114. [Google Scholar] [CrossRef]
  9. Banu, M.; Tsividis, Y. The MOSFET-C technique. In Design of High Frequency Integrated Analogue Filters, 1st ed.; Sun, Y., Ed.; Series 14; The Institution of Engineering and Technology: London, UK, 2007; pp. 35–56. [Google Scholar]
  10. Banu, M.; Tsividis, Y. Fully integrated active RC filters in MOS technology. IEEE J. Solid-State Circuits 1983, 18, 644–651. [Google Scholar] [CrossRef]
  11. Chen, W.-K. Passive, Active, and Digital Filters, 1st ed.; CRC Press: Boca Raton, FL, USA, 2005; Chapter 17. [Google Scholar]
  12. Mohan, P.A. VLSI Analog Filters: Active RC, OTA-C, and SC, 2nd ed.; Springer Science & Business Media: New York, NY, USA, 2012; pp. 251–432. [Google Scholar]
  13. Sun, Y. (Ed.) Architectures and design of OTA/gm-C filters. In Design of High Frequency Integrated Analogue Filters, 1st ed.; Series 14; The Institution of Engineering and Technology: London, UK, 2007; pp. 1–34. [Google Scholar]
  14. Li, D.; Tsividis, Y. Active LC filters on silicon. IEE Proc.-Circuits Devices Syst. 2000, 147, 49–56. [Google Scholar] [CrossRef]
  15. Dandan, L.; Yannis, T. Active filters using integrated inductors. In Design of High Frequency Integrated Analogue Filters, 1st ed.; Sun, Y., Ed.; Series 14; The Institution of Engineering and Technology: London, UK, 2007; pp. 57–80. [Google Scholar]
  16. Bahl, I.J. High-performance inductors. IEEE Trans. Microw. Theory Tech. 2001, 49, 654–664. [Google Scholar] [CrossRef]
  17. Mathúna, C.Ó.; Wang, N.; Kulkarni, S.; Roy, S. Review of integrated magnetics for power supply on chip (PwrSoC). IEEE Trans. Power Electron. 2012, 27, 4799–4816. [Google Scholar] [CrossRef]
  18. Burghartz, J.; Edelstein, D.; Jenkins, K.; Jahnes, C.; Uzoh, C.; O’Sullivan, E.; Chan, K.; Soyuer, M.; Roper, P.; Cordes, S. Monolithic spiral inductors fabricated using a VLSI Cu-damascene interconnect technology and low-loss substrates. In Proceedings of the International Electron Devices Meeting, Technical Digest. San Francisco, CA, USA, 8–11 December 1996; pp. 99–102. [Google Scholar]
  19. Yeşil, A.; Kaçar, F.; Gürkan, K. Lossless grounded inductance simulator employing single VDBA and its experimental band-pass filter application. AEU Int. J. Electron. Commun. 2014, 68, 143–150. [Google Scholar] [CrossRef]
  20. Kacar, F.; Kuntman, H. CFOA-based lossless and lossy inductance simulators. Radioengineering 2011, 20, 627–631. [Google Scholar]
  21. Kumar, N.; Vista, J.; Ranjan, A. A tuneable active inductor employing DXCCTA: Grounded and floating operation. Microelectron. J. 2019, 90, 1–11. [Google Scholar] [CrossRef]
  22. Senani, R.; Bhaskar, D.R.; Singh, A.K. Current Conveyors: Variants, Applications and Hardware Implementations; Springer: Cham, Switzerland, 2015; pp. 469–497. [Google Scholar]
  23. Yuan, F. CMOS Active Inductors and Transformers: Principle, Implementation and Applications, 1st ed.; Springer: New York, NY, USA, 2008. [Google Scholar]
  24. Wang, X.; Zhang, W.; Na, W.; Jin, D.; Xie, H.; Ren, Y.; Li, N.; Gao, W.; Wu, Y. A novel CMOS active inductor with high quality factor, high linearity and mutually independent tuning of inductance and quality factor. Microelectron. J. 2023, 141, 105984. [Google Scholar] [CrossRef]
  25. Momen, H.G.; Yazgi, M.; Kopru, R. A low loss, low voltage and high Q active inductor with multi-regulated cascade stage for RF applications. In Proceedings of the IEEE International Conference on Electronics, Circuits, and Systems (ICECS), Cairo, Egypt, 6–9 December 2015; pp. 149–152. [Google Scholar]
  26. Gardeshkhah, R.; Saatlo, A.N. A New Tunable Gyrator-C-Based Active Inductor Circuit for Low Power Applications. J. Circuits Syst. Comput. 2023, 32, 2350235. [Google Scholar] [CrossRef]
  27. Lai, Q.-T.; Mao, J.-F. A new floating active inductor using resistive Feedback Technique. In Proceedings of the IEEE MTT-S International Microwave Symposium, Anaheim, CA, USA, 23–28 May 2010; pp. 1748–1751. [Google Scholar]
  28. Faruqe, O.; Amin, M.T. Active inductor with feedback resistor based voltage controlled oscillator design for wireless applications. Int. J. Electron. Telecommun. 2019, 65, 57–64. [Google Scholar] [CrossRef]
  29. Saberkari, A.; Ziabakhsh, S.; Martinez, H.; Alarcon, E. Design and comparison of flipped active inductors with high quality factors. Electron. Lett. 2014, 50, 925–927. [Google Scholar] [CrossRef]
  30. Nediyara Suresh, L.; Manickam, B. Multiple cascode flipped active inductor-based tunable bandpass filter for fully integrated RF front-end. IET Circuits Devices Syst. 2020, 14, 93–99. [Google Scholar] [CrossRef]
  31. Farahani, M.M.; Mazloum, J.; Fouladian, M. An ultra-wideband low noise amplifier with cascaded flipped-active inductor for cognitive radio applications. Integration 2023, 93, 102046. [Google Scholar] [CrossRef]
  32. Jeong, Y.J.; Kim, Y.M.; Chang, H.J. Low-power CMOS VCO with a low-current, high-Q active inductor. IET Microw. Antennas Propag. 2012, 6, 788–792. [Google Scholar] [CrossRef]
  33. Sackinger, E.; Fischer, W.C. A 3-GHz 32-dB CMOS limiting amplifier for SONET OC-48 receivers. IEEE J. Solid-State Circuits 2000, 35, 1884–1888. [Google Scholar] [CrossRef]
  34. Yodprasit, U.; Ngarmnil, J. Q-enhancing technique for rf CMOS active inductor. In Proceedings of the IEEE International Symposium on Circuits and Systems (ISCAS), Geneva, Switzerland, 28–31 May 2000; pp. 589–592. [Google Scholar]
  35. Sato, T.; Ito, T. Design of low distortion active inductor and its applications. Analog Integr. Circuits Signal Process. 2013, 75, 245–255. [Google Scholar] [CrossRef]
  36. Pantoli, L.; Stornelli, V.; Leuzzi, G. Class AB gyrator-based active inductor. In Proceedings of the Integrated Nonlinear Microwave and Millimetre-Wave Circuits Workshop (INMMiC), Taormina, Italy, 1–2 October 2015; pp. 1–3. [Google Scholar]
  37. Thanachayanont, A.; Sae Ngow, S. Class AB VHF CMOS active inductor. In Proceedings of the 45th Midwest Symposium on Circuits and Systems (MWSCAS), Tulsa, OK, USA, 4–7 August 2002; p. I-64. [Google Scholar]
  38. Gorjizad, Z.; Shojaee, A.; Abrishamifar, A. A compact and tunable active inductor-based bandpass filter with high quality factor for Wireless LAN applications. AEU Int. J. Electron. Commun. 2024, 187, 155540. [Google Scholar] [CrossRef]
  39. Ben Hammadi, A.; Haddad, F.; Mhiri, M.; Saad, S.; Besbes, K. RF and microwave reconfigurable bandpass filter design using optimized active inductor circuit. Int. J. RF Microw. Comput.-Aided Eng. 2018, 28, e21550. [Google Scholar] [CrossRef]
  40. Saad, S.; Mhiri, M.; Hammadi, A.B.; Besbes, K. A new low-power, high-Q, wide tunable CMOS active inductor for RF applications. IETE J. Res. 2016, 62, 265–273. [Google Scholar] [CrossRef]
  41. Saad, S.; Haddad, F.; Hammadi, A.B. Impact of multi-finger MOSFET geometry on the electrical performance of RF circuits. Microelectron. Reliab. 2022, 129, 114445. [Google Scholar] [CrossRef]
  42. Ho, C.C.; Kuo, C.W.; Chan, Y.J.; Lien, W.Y.; Guo, J.C. 0.13-µm RF CMOS and varactors performance optimization by multiple gate layouts. IEEE Trans. Electron Devices 2004, 51, 2181–2185. [Google Scholar] [CrossRef]
  43. Yeh, K.L.; Guo, J.C. Narrow-width effect on high-frequency performance and RF noise of sub-40-nm multifinger nMOSFETs and pMOSFETs. IEEE Trans. Electron Devices 2012, 60, 109–116. [Google Scholar]
  44. Ler, C.L.; bin A’ain, A.K.; Kordesch, A.V. Compact, High-Q, and Low-Current Dissipation CMOS Differential Active Inductor. IEEE Microw. Wirel. Compon. Lett. 2008, 18, 683–685. [Google Scholar] [CrossRef]
  45. Ler, C.L.; A’ain, A.K.B.; Kordesch, A.V. CMOS active inductor linearity improvement using feed-forward current source technique. IEEE Trans. Microw. Theory Tech. 2009, 57, 1915–1924. [Google Scholar] [CrossRef]
  46. Li, C.; Gong, F.; Wang, P. Analysis and design of a high-Q differential active inductor with wide tuning range. IET Circuits Devices Syst. 2010, 4, 486–495. [Google Scholar] [CrossRef]
  47. Mehra, R.; Kumar, V.; Islam, A. Reliable and Q-Enhanced Floating Active Inductors and Their Application in RF Bandpass Filters. IEEE Access 2018, 6, 48181–48194. [Google Scholar] [CrossRef]
  48. Kumar, V.; Mehra, R.; Islam, A. A 2.5 GHz Low Power, High-Q, Reliable Design of Active Bandpass Filter. IEEE Trans. Device Mater. Reliab. 2017, 17, 229–244. [Google Scholar] [CrossRef]
  49. Li, Y.; Hwang, C.H.; Li, T.Y.; Han, M.H. Process-variation effect, metal-gate work-function fluctuation, and random-dopant fluctuation in emerging CMOS technologies. IEEE Trans. Electron Devices 2009, 57, 437–447. [Google Scholar] [CrossRef]
  50. Saha, S.K. Compact MOSFET modeling for process variability-aware VLSI circuit design. IEEE Access 2014, 2, 104–115. [Google Scholar] [CrossRef]
  51. Sundarasaradula, Y.; Thanachayanont, A. A 1-V, 6-nW programmable 4th-order bandpass filter for biomedical applications. Analog Integr. Circuits Signal Process. 2016, 89, 89–98. [Google Scholar] [CrossRef]
  52. He, X.; Kuhn, W.B. A fully integrated Q-enhanced LC filter with 6 dB noise figure at 2.5 GHz in SOI. In Proceedings of the IEE Radio Frequency Integrated Circuits (RFIC) Systems, Digest of Papers. Fort Worth, TX, USA, 6–8 June 2004; pp. 643–646. [Google Scholar]
  53. Christensen, K.T.; Lee, T.H.; Bruun, E. A high dynamic range programmable CMOS front-end filter with a tuning range from 1850 to 2400 MHz. Analog Integr. Circuits Signal Process. 2005, 42, 55–64. [Google Scholar] [CrossRef]
  54. Gao, Z.; Ma, J.; Yu, M.; Ye, Y. A fully integrated CMOS active bandpass filter for multiband RF front-ends. IEEE Trans. Circuits Syst. II Express Br. 2008, 55, 718–722. [Google Scholar]
  55. Yadav, P.K.; Misra, P.K. Tunable bandpass filter using double resistive feedback floating active inductor for 5 GHz wireless LAN Applications. Analog Integr. Circuits Signal Process. 2021, 108, 205–213. [Google Scholar] [CrossRef]
  56. Hu, F.; Mouthaan, K. L-band bandpass filter with high out-of-band rejection based on CMOS active series and shunt inductors. In Proceedings of the IEEE MTT-S International Microwave Symposium, Tampa, FL, USA, 1–6 June 2014; pp. 1–3. [Google Scholar]
  57. Suresh, L.N.; Manickam, B. Design and application of CMOS active inductor in bandpass filter and VCO for reconfigurable RF front-end. Integration 2022, 82, 115–126. [Google Scholar] [CrossRef]
  58. Herbert, T.B.; Hyland, J.S.; Abdullah, S.; Wight, J.; Amaya, R.E. An active bandpass filter for LTE/WLAN applications using robust active inductors in gallium nitride. IEEE Trans. Circuits Syst. II Express Br. 2021, 68, 2252–2256. [Google Scholar] [CrossRef]
  59. Leuzzi, G.; Stornelli, V.; Del Re, S. A tuneable active inductor with high dynamic range for band-pass filter applications. IEEE Trans. Circuits Syst. II Express Br. 2011, 58, 647–651. [Google Scholar] [CrossRef]
Figure 1. Transceiver front-end circuit.
Figure 1. Transceiver front-end circuit.
Sensors 25 03089 g001
Figure 2. Schematic of a lossy DAI based on gyrator-C networks and its equivalent RLC circuit.
Figure 2. Schematic of a lossy DAI based on gyrator-C networks and its equivalent RLC circuit.
Sensors 25 03089 g002
Figure 3. Proposed differential active inductor circuit.
Figure 3. Proposed differential active inductor circuit.
Sensors 25 03089 g003
Figure 4. Equivalent small signal model of the proposed differential active inductor.
Figure 4. Equivalent small signal model of the proposed differential active inductor.
Sensors 25 03089 g004
Figure 5. Proposed DAI based bandpass filter.
Figure 5. Proposed DAI based bandpass filter.
Sensors 25 03089 g005
Figure 6. Configuration block of a DAI-based BPF.
Figure 6. Configuration block of a DAI-based BPF.
Sensors 25 03089 g006
Figure 7. The real and imaginary parts of the proposed active inductor’s tuning for different Vres.
Figure 7. The real and imaginary parts of the proposed active inductor’s tuning for different Vres.
Sensors 25 03089 g007
Figure 8. Q-factor value versus frequency over Vres and input-referred noise voltage of the proposed DAI.
Figure 8. Q-factor value versus frequency over Vres and input-referred noise voltage of the proposed DAI.
Sensors 25 03089 g008
Figure 9. Phase response of the proposed DAI over Vres.
Figure 9. Phase response of the proposed DAI over Vres.
Sensors 25 03089 g009
Figure 10. Simulation response of S-parameters of the proposed BPF.
Figure 10. Simulation response of S-parameters of the proposed BPF.
Sensors 25 03089 g010
Figure 11. Transmission coefficient S21 of the proposed BPF for different combinations of [Vres, Vbias2].
Figure 11. Transmission coefficient S21 of the proposed BPF for different combinations of [Vres, Vbias2].
Sensors 25 03089 g011
Figure 12. Noise figure of the proposed BPF.
Figure 12. Noise figure of the proposed BPF.
Sensors 25 03089 g012
Figure 13. (a) 1-dB compression point, (b) input referred third-order of the proposed BPF.
Figure 13. (a) 1-dB compression point, (b) input referred third-order of the proposed BPF.
Sensors 25 03089 g013
Figure 14. Stability properties of the proposed BPF (a) alternative stability factor B1f, (b) Rollet stability factor Kf.
Figure 14. Stability properties of the proposed BPF (a) alternative stability factor B1f, (b) Rollet stability factor Kf.
Sensors 25 03089 g014
Figure 15. Performance variation of transmission coefficient S21 of the proposed BPF with various corner conditions.
Figure 15. Performance variation of transmission coefficient S21 of the proposed BPF with various corner conditions.
Sensors 25 03089 g015
Figure 16. Layout of the proposed BPF.
Figure 16. Layout of the proposed BPF.
Sensors 25 03089 g016
Table 1. Simulated design metrics of the proposed BPF.
Table 1. Simulated design metrics of the proposed BPF.
ConfigVres (mV)Vbias2 (mV)fc (GHz)S21 (dB)BW (MHz)QBPFNF (dB)PDC (mW)
12006000.2816.543210.924.6015.2
21306000.5915.843571.721.8615.2
31005800.8715.523702.3519.6314.9
41005001.2214.92406320.1914.1
5−2008001.5713.422975.318.4120.3
6−2508001.8314.022876.418.4820.3
7−300800213.383186.2918.5520.3
8−4008002.1911.783975.5218.6320.3
9−12008002.4268.895934.118.7320.3
Table 2. Monto Carlo analysis: sensitivity of the proposed BPF to mismatch and process variations.
Table 2. Monto Carlo analysis: sensitivity of the proposed BPF to mismatch and process variations.
TypeS21 (dB)NF (dB)KfB1f
Mean (µ)14.919.949.820.998
Standard deviation (σ)0.3450.3790.4840.000875
Variability (σ/Mean)2.32%1.9%4.93%0.09%
Minimum (−3σ)13.6719.818.60.997
Maximum (+3σ)15.9420.0411.61.002
Total number of iterations/hits = 1000, evaluated at Vres = 0.1V and Vbias2 = 0.5V.
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

Saad, S.; Haddad, F.; Ben Hammadi, A. A Compact and Tunable Active Inductor-Based Bandpass Filter with High Dynamic Range for UHF Band Applications. Sensors 2025, 25, 3089. https://doi.org/10.3390/s25103089

AMA Style

Saad S, Haddad F, Ben Hammadi A. A Compact and Tunable Active Inductor-Based Bandpass Filter with High Dynamic Range for UHF Band Applications. Sensors. 2025; 25(10):3089. https://doi.org/10.3390/s25103089

Chicago/Turabian Style

Saad, Sehmi, Fayrouz Haddad, and Aymen Ben Hammadi. 2025. "A Compact and Tunable Active Inductor-Based Bandpass Filter with High Dynamic Range for UHF Band Applications" Sensors 25, no. 10: 3089. https://doi.org/10.3390/s25103089

APA Style

Saad, S., Haddad, F., & Ben Hammadi, A. (2025). A Compact and Tunable Active Inductor-Based Bandpass Filter with High Dynamic Range for UHF Band Applications. Sensors, 25(10), 3089. https://doi.org/10.3390/s25103089

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