Next Article in Journal
Investigating a Low-Cost Dryer Designed for Low-Cost PM Sensors Measuring Ambient Air Quality
Previous Article in Journal
Discrimination of Ignitable Liquid Residues in Burned Petroleum-Derived Substrates by Using HS-MS eNose and Chemometrics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

BPF-Based Thermal Sensor Circuit for On-Chip Testing of RF Circuits

1
Electronic Engineering Department, Universitat Politècnica de Catalunya–BarcelonaTech, 08034 Barcelona, Spain
2
Institute of Microelectronics of Barcelona (IMB-CNM)—CSIC, 08193 Cerdanyola del Vallés, Spain
*
Author to whom correspondence should be addressed.
Sensors 2021, 21(3), 805; https://doi.org/10.3390/s21030805
Submission received: 17 December 2020 / Revised: 19 January 2021 / Accepted: 20 January 2021 / Published: 26 January 2021
(This article belongs to the Section Physical Sensors)

Abstract

:
A new sensor topology meant to extract figures of merit of radio-frequency analog integrated circuits (RF-ICs) was experimentally validated. Implemented in a standard 0.35 μm complementary metal-oxide-semiconductor (CMOS) technology, it comprised two blocks: a single metal-oxide-semiconductor (MOS) transistor acting as temperature transducer, which was placed near the circuit to monitor, and an active band-pass filter amplifier. For validation purposes, the temperature sensor was integrated with a tuned radio-frequency power amplifier (420 MHz) and MOS transistors acting as controllable dissipating devices. First, using the MOS dissipating devices, the performance and limitations of the different blocks that constitute the temperature sensor were characterized. Second, by using the heterodyne technique (applying two nearby tones) to the power amplifier (PA) and connecting the sensor output voltage to a low-cost AC voltmeter, the PA’s output power and its central frequency were monitored. As a result, this topology resulted in a low-cost approach, with high linearity and sensitivity, for RF-IC testing and variability monitoring.

1. Introduction

The temperature in a surface point of an integrated circuit (IC) depends on the power dissipated by the devices placed nearby (so-called self-heating and thermal coupling), the structure and materials that constitute the packaging (which determine the thermal impedances of the different devices) and the ambient temperature [1]. Traditionally, off-chip temperature measurement set-ups have been used to detect unexpected hot spots within digital circuits [2,3,4,5]. Hot spots might appear either due to the presence of a defect in the circuit structure [6,7,8,9] or by a nonuniform power dissipation on the die surface, which is a common situation in microprocessors [4,10]. In complex digital systems, such as microprocessors, temperature sensors are built-in within the same silicon die in order to ensure reliable system performance, i.e., they perform power-temperature monitoring to control the activation of cooling systems, to modulate microprocessor supply voltage or clock frequency, or to assert if the workload of a specific microprocessor block can be increased or should be reduced to avoid nonuniform power distributions [11,12,13,14,15,16,17,18]. Nevertheless, thermal monitoring is not restricted to digital circuits, but used as well in analog circuits. Most commonly, thermal measurements of analog circuits are usually performed to extract the thermal resistance of devices [19,20], especially in power devices. More recently, temperature measurements are done in high frequency analog circuits to perform testing applications. The test of high frequency analog circuits is a challenging task, as the presence of a defect or the effects of process-voltage-temperature variations and device aging, does not usually produce a catastrophic circuit failure, but a degradation of circuit performance and specifications [21,22,23,24,25], which compromises yield. A strategy to compensate these performance degradations is to build in monitor circuits with the analog circuit (called hereafter the circuit under test (CUT)) to track variations in their performance [26,27,28,29,30] and, in the eventual case of detecting a circuit degradation, to activate a feedback in the CUT bias to compensate for them. Recently, several studies have proved that by measuring the temperature in a surface point near the CUT, it is feasible to monitor the performances of radio frequency (RF) and millimeter wave (mmW) circuits [30,31,32,33,34,35,36,37] or the presence of structural defects [38,39]. The use of built-in temperature sensors to monitor high-frequency analog circuits is attractive. On the one hand, the sensor does not load any node of the CUT, avoiding the need of a CUT-sensor codesign. On the other hand, thanks to the Joule effect, there is a frequency down-conversion of the high-frequency information in the electrical domain to low frequency in the thermal domain in such a way that the same temperature sensor can be used to monitor different CUTs working at different frequency bands.
In this scenario, the goal of temperature measurements is to get solely a signature of the power dissipated by the CUT, as this power carries information of the high frequency electrical signals while rejecting any temperature variation due to changes in the ambient temperature or the case-to-ambient thermal resistance (e.g., the activation of a cooling system). To this end, one strategy is the use of differential temperature sensors embedded in the same silicon die [37,38,39,40,41]. Such sensors possess two sensing devices (temperature transducers T1 and T2). Whereas T1 is placed close to the CUT, T2 is placed far from it. This placement ensures that only T1 measures the temperature changes caused by the power dissipation of the CUT, but both T1 and T2 detect common-mode temperature changes, being eventually rejected. Generally, the two sensing devices form the differential pair of an operational transconductance amplifier (OTA), which operates in open-loop to have a high sensitivity [38,39,40]. Although these sensors have been proved useful to perform the test of analog high-frequency CUTs, they have several limitations. Firstly, the OTA output can be easily saturated by mismatches produced because of manufacturing variability and the DC temperature gradients generated by the DC CUT bias [36,38]. Secondly, since the OTA operates in open loop, the gain (and hence the sensitivity to temperature) is quite sensitive to integrated circuit (IC) manufacturing process variations [37]. Thirdly, mismatches, circuit topology and device limitations reduce the rejection to common mode temperature changes.
These limitations have motivated the design of a novel circuit topology that we presented in [42] (Figure 1). One sensing device (T1 placed near the CUT) is connected to an active band-pass filter (BPF). The filter’s band-pass is centered at Δf so that the bandwidth (and hence the noise) is limited around the frequency of interest. The BPF’s zero at the origin rejects slow temperature variations (provoked by either ambient temperature changes or the DC CUT power dissipation). To achieve a CUT thermal signature within the frequency band of the sensor, we propose to excite the CUT with a heterodyne technique [41,43]. The technique consists of applying two tones, whose frequencies are f1 and f2 = f1 + Δf, to the CUT input. This driving strategy generates a spectral component of power dissipated by the CUT devices (and hence of temperature) at Δf, whose magnitude depends on the CUT figures of merit at f1. For Δf values higher than a certain threshold (fixed by both the die thickness and semiconductor thermal properties [40]), the amplitude of this temperature component is no longer dependent on the package materials and package mounting thermal properties, i.e., the silicon die is seen by the CUT, from a thermal point of view, as a semi-infinite medium, independent of the thermal boundary conditions (including ambient temperature). This heterodyne technique is not new: it has been already used with differential temperature sensors connecting the sensor’s output node to a lock-in amplifier (LIA) locked at Δf [33,35]. To reduce costs, the BPF output signal is proposed to be measured using a low-cost digital multimeter (DMM), off-chip in this work, but that allows an easier integration than the LIA if a complete built-in test approach is required.
The main novelty and first goal of this paper is to present the first experimental characterization of this sensor topology, focusing on its sensitivity, noise and linearity. To this end, we implemented a realization of this sensor topology in a standard 0.35 μm CMOS technology (VDD = 3.3 V) together with MOS transistors acting as controllable dissipating devices. The second goal is to assess the temperature sensor for RF CUT monitoring. For that purpose, a tuned (420 MHz) class-A radio-frequency power amplifier (PA) is built-in together with the sensor, which shows its capability to monitor the PA central frequency and the output power delivered to the load (antenna).
Taking this into account, the paper is organized as follows: the sensor topology and design are described in Section 2. The sensor’s block characterization and full sensor validation are carried out in Section 3. Section 4 presents the PA used as CUT, the placement of the transducer within the PA layout and two application cases for this temperature sensor as a built-in RF monitor. Finally, Section 5 draws the main conclusions.

2. Sensor Description and Design

Figure 2 shows the proposed sensor schematic. It was made of two blocks: (a) the temperature transducer and (b) the active BPF. The goal of the temperature transducer is to generate a voltage at the node Vot proportional to the working temperature of the transistor T1, which is placed in the silicon die at the proximity of the CUT. On the other hand, the aim of the active BPF is to provide signal amplification at the output node Vo if the input signal Vif has its frequency within the passing band. The main figures of merit of the BPF frequency response are represented in Figure 2c: the low-frequency (fp1) and high-frequency (fp2) poles, which determine the pass-band, and the band-pass gain Av. Three different circuits were implemented in the IC: only (a) with Vot connected to an output pad; only (b) with Vif connected to an input pad and (a) and (b) with Vot internally shorted to Vif. A detailed description and theoretical analysis of each block is in the following two subsections, with emphasis on sensitivity, noise and linearity.

2.1. Temperature Transducer Description

The temperature transducer T1 was an nMOS transistor (dimensions: W = 24 μm, L = 1.5 μm) connected in diode configuration and biased with a DC constant current (IB in Figure 2). A small bias current placed the transistor in the weak inversion region, having an expected sensitivity of −1.5 mV/K at IB = 20 nA. With other dimensions and bias, transducer sensitivities in the range [−1.6 mV/K, 5 mV/K] can be achieved, as reported in [44,45]. To create such a small bias current, a current mirror with a ratio IB/Iext = 1/1000 was implemented. The internal current mirror in conjunction with an operational amplifier (OA) in voltage follower configuration ensures a low parasitic capacitance at node N1 and a low output impedance, enhancing the dynamic transducer behavior.

2.2. Band Pass Filter Amplifier: Noise and Linearity Analysis

The output of the temperature transducer was connected to an active BPF. The transducer signal is AC-coupled, amplified and filtered, and appeared at the output superposed on a DC level of VDD/2. Capacitors C1 = 50 pF and C2 = 100 fF set the band-pass gain Av which was designed to be:
| A v | C 1 C 2 = 500   ( 54   dB ) .  
Having a zero at the origin, the pass-band is determined by the low (fp1) and high (fp2) poles (cut-off frequencies), which are respectively:
f p 1 = 1 2 · π · R 2 C 2
f p 2 = G B W | A v |
Since the OA gain-bandwidth (GBW) product is 2.4 MHz, this results in a theoretical value for fp2 of 4.8 kHz. To have a low fp1, resistor values in the order of GΩ are required, thus R2 was implemented with two subthreshold biased pMOS, as shown in Figure 3 [46,47,48,49]. These MOS had W = 2 μm and L = 10 μm, and the external voltage VBIAS applied to their gate allows tuning to its equivalent resistance. From Equations (1) and (3), in a first order analysis, fp2 and Av should be independent from VBIAS.
Table 1 shows the values, obtained from simulations, of the R2 incremental resistance as a function of VBIAS (assuming both R2 terminals, N1 and N2, at VDD/2) as well as the BPF’s low and high cut-off frequencies. Note how fp1 is highly sensitive to VBIAS, which opens a discussion on its optimum value. As explained above, transducer signal consisted of a tone at a frequency Δf, which should fall within the filter passband. From the values in Table 1, Δf around 1–2 kHz are reasonable choices. The filter output would be measured with a DMM that evaluates the total RMS voltage. Then, a narrow band-pass (high fp1) seems desirable in order to filter out as much noise as possible and reduce the measurement noise level. But, on the other hand, the amplified signal at the output node modulates the R2 value, thus producing distortion. The effect of this R2 nonlinearity on the overall sensor nonlinearity is relevant whenever R2 has a significant contribution to the signal path, i.e., when Δf is around fp1 or lower. On the contrary, at Δf values well above fp1, the effects of R2, including its nonlinearity, become negligible on the output linearity and gain. From this point of view, a low fp1 is desirable.
In order to assess the effect of the different fp1 choices on the sensor noise and linearity, Table 1 shows simulation results of the filter alone (no transducer included). Noise was evaluated after integrating the power spectral density at the filter output up to 100 kHz, while the maximum harmonic distortion (THDMAX) was also evaluated at the filter output obtained for a 1 kHz input sinusoid just before output clipping. Simulations show how linearity was dramatically degraded as fp1 was set near the signal frequency, while could be improved as fp1 is moved away from the signal. The price to pay was a moderate noise degradation. In view of these simulation results, VBIAS = 1.2 V (fp1 = 248 Hz) was selected as a good default setting, which can be increased (fp1 decreased) to produce a more linear response.

3. Sensor Implementation and Validation

The sensor was designed and manufactured in the CMOS 0.35 µm process. This microelectronic process provides four metal layers and two polysilicon layers. Capacitances C1 and C2 were implemented with polysilicon capacitors. The supply voltage (VDD) was 3.3 V (unipolar), and the IC was packaged in a QFN56. Its validation was performed first by each block standing alone and finally both connected, as detailed below.

3.1. Temperature Transducer

To characterize its frequency response, the circuit in Figure 2a was implemented as stand-alone, with the node Vot connected to an output pad. Besides, a large diode-connected nMOS transistor acting as a heater (W = 450 μm, L = 1 μm and 15 fingers) was placed at a 5 μm distance from the transducer (Figure 4). The heater was biased with a gate (drain) voltage, vin(t):
V i n ( t ) = V D C h + A · cos ( 2 π · f i · t ) ,  
where VDCh is the heater bias voltage, A is the AC amplitude, and fi is its frequency. Assuming a linear response, the current flowing through the heater is:
I i n ( t ) = I D C h + g m A · cos ( 2 π · f i · t ) ,  
where gm is the heater AC transconductance at Vin = VDCh, and IDCh is the DC drain to source current when Vin = VDCh.
Multiplying Equations (4) and (5), the power dissipated by the heater in the frequency domain can be described as the sum of different spectral components. If we focus on the spectral component at the frequency fi (called pfi), it can be written as:
p f i ( t ) = P A · cos ( 2 π · f i · t ) ,  
where PA is the power dissipation amplitude at the frequency fi. The above-mentioned heater dimensions enabled us to easily have PA in the mW range while applying a low input voltage. pfi(t) causes a temperature oscillation tS(t) at the temperature transducer with the same frequency fi:
t S ( t ) = T A · cos ( 2 π · f i · t + Φ A ) ,  
where TA is the temperature amplitude and φA is the phase shift between ts and pfi. Both thermal amplitude and phase shift depend on the silicon physical properties (thermal conductivity and specific heat), the input frequency fi and the distance between the heater and the transducer [40]. The close proximity between the heater and the transducer ensures a good thermal coupling even in the MHz frequency range [32,40,50,51]. Then, the transducer output voltage can be written as:
V o t ( t ) = V G S T 1 + V A · cos ( 2 π · f i · t + Φ A ) ,  
where VGST1 is the DC gate-to-source voltage in T1, needed to sustain the drain current IB, and VA is the voltage amplitude generated by the oscillating temperature of amplitude TA. Here we assume that there is no phase shift between the transducer temperature and the output voltage, i.e., the pole induced by the node N1 in Figure 2b is at a frequency much higher than fi. Being a temperature-to-voltage transducer, its sensitivity is defined as VA/TA. However, as the goal of the sensor was to monitor the CUT characteristics through its power dissipation, the sensitivity expressed as VA/PA had a greater interest, and will be evaluated in this work.
In order to characterize the transducer response, the heater has was excited with the different input voltage levels (VL) reported in Table 2. PA values in this table were calculated from the experimental I-V characteristics of the heater. Figure 5 shows the measured transducer sensitivity as a function of the frequency (transducer biased with IB = 20 nA). Here, the amplitude VA was measured with an LIA (Signal Recovery 7265DSP) locked at fi with a constant time of 1 s.
The frequency response in Figure 5 shows a low-pass behavior corresponding to the thermal coupling from the heater to the transducer, which agrees with the theoretical, simulations and experimental data previously reported in [52]. In this particular set-up, the coupling attenuation increased for frequencies higher than 3 kHz. Besides that, the independence of the transducer’s sensitivity on the amplitude PA proves the transducer linear behavior.

3.2. Band Pass Filter Amplifier

The circuit in Figure 2b was also implemented as stand-alone, with input (Vif) and output nodes (Vo) connected, respectively, to input and output pads.
Figure 6 shows the measured frequency response of the BPF voltage gain. Input Vif(t) is a sinusoidal voltage of 20 mVRMS attenuated by a factor of 100 with an off-chip resistive voltage divider. Filter characterizations were done for the three VBIAS values reported in Table 1. For each VBIAS, the output voltage amplitude was measured using either the LIA or a DMM (HP 33401A) AC-coupled. The LIA was set with a time constant of 1 s, which implied that the measure was integrating the signal (and noise) in a bandwidth of 1 Hz. On the other hand, the DMM was integrating the signal and noise along all the filter bandwidth (i.e., the output harmonics were integrated as well).
Table 3 summarizes the BPF characteristics extracted from the responses in Figure 6, considering LIA measurements. Comparing Table 1 and Table 3, measured characteristics show a reasonable agreement to the simulation predictions.
More importantly, measurements done with a simple DMM had very good agreement with those made with the LIA for frequencies equal and higher than the BPF central frequency. However, DMM results were higher than the LIA ones when the input signal frequency was equal or smaller than fp1. As predicted by simulations, output-voltage nonlinearities became relevant when the input frequency was equal or smaller than fp1, due to the R2 modulation. Those nonlinearities produced significant harmonics, which were captured with the DMM, thus producing a slightly higher gain measurement. Finally, DMM results showed an equivalent noise at the filter output of 4 mVRMS, which, according to Section 2, translated into an equivalent noise at the temperature sensor’s input of 7.54 mK (@ IB = 20 nA). As predicted by simulations, the noise level at the output was almost independent from VBIAS: fp1 may change over one order of magnitude, but its values were in the range of tens-hundreds of Hz, with small effect after integrating over the whole bandwidth.
The effects of the manufacturing variability on R2 were also evaluated. Figure 7 shows the effect of R2 sample-to-sample variability on the measured fp1 values, as a function of VBIAS in three different IC samples. In all these situations, the maximum gain measured was 54.5 ± 0.43 dB, in agreement with the expected gain independence with VBIAS.

3.3. Overall Sensor Characterization

Besides the stand-alone temperature transducer (Figure 2a) and the stand-alone BPF (Figure 2b), the complete circuit in Figure 2 was also implemented with the transducer’s output connected to the input of the BPF, and with a nMOS transistor such as that depicted in Figure 4 as a controllable heater. Then, when the dissipating device is biased with vin(t) as in Equation (4), the sensor’s output voltage can be written as:
V O ( t ) = V D D 2 + A O · cos ( 2 π · f i · t + Φ O ) ,  
where AO is the amplitude of the sinusoidal component of VO(t) at the frequency fi. The filter introduces a phase shift (φO − φA) − φA is defined in Equations (7) and (8).
Figure 8 shows AO,RMS (RMS value of AO from Equation (9)) as a function of the frequency when measured with the DMM (@ IB = 20 nA, VBIAS = 1.2 V), using the different VLs in Table 2 for the heater excitation. The frequency response reported a central frequency, fc, of about 2 kHz for all the VL. The overall frequency response is the concatenation of the transducers’ and BPF frequency responses. For frequencies smaller than 100 Hz or higher than 10 kHz (the exact frequency depends on the particular bias), the sensor’s output voltage reached the noise level, which is 40 mVRMS. This noise level is higher than that observed in Figure 6, meaning that it was dominated by the heater-transducer circuits. The noise level sets the PA threshold that the sensor is able to detect. To calculate this threshold, Figure 9 shows the sensor’s sensitivity (AO/PA) as a function of the frequency, when measured with the LIA. The maximum sensitivity was about 140 mV/mW. Therefore, 40 mVRMS corresponded to an equivalent noise floor in PA of 285 μW.
If we now consider the sensor’s linearity, the overlapping points in Figure 9 show that the sensor sensitivity was not affected by the PA level when fi was equal or higher than fc. On the other hand, the sensitivity depended on the amplitude PA dissipated by the heater when fi was smaller than fc: The R2 nonlinear behavior affected the sensor linearity. These results have implications in the way the CUT must be driven when a test is done using a heterodyne excitation (Figure 1): Δf should be higher than fp1. The exact value of Δf is determined by the expected range of values of the amplitude PA dissipated by the CUT, which determines the required sensor sensitivity and the noise level. For instance, from Figure 8 and Figure 9, if Δf = 2.5 kHz the sensor’s sensitivity is 134 mW/mV. This sensitivity allows the sensor to detect the PA dissipated by the heater when driven by the VL 1. On the other hand, if Δf = 10 kHz the sensor’s sensitivity is 54 mW/mV, and the amplitude PA dissipated by the heater when driven by the VL 1 is below the noise level, whereas the sensor can barely detect the amplitude PA when driven by the VL 2. This indicates that Δf higher than the sensor’s central frequency can be a good choice to sense large PA levels. On the other hand, if Δf = fc, high values of PA might drive the sensor to saturation.

4. Use Case: RF Tuned Class-A Power Amplifier Monitoring

4.1. Circuit and Experimental Set Up Description

The BPF temperature sensor was integrated with a narrowband RF power amplifier (PA) used as a CUT. The schematic of the RF PA is shown in Figure 10. It is a class-A cascode amplifier with off-chip load, fully described in [32]. The cascode transistor M2 was made of three transistors in parallel connection (inset in Figure 10). The overall M2 dimensions were W = 1173 μm (implemented with 51 fingers), and L = 0.5 μm. The MOS temperature transducer was placed in the free space existing between two of these transistors. When the DC bias of the PA was VDD = 3.3 V and Vcnt = 3 V, it had a current consumption of 22 mA. Experimental characterization of the PA reported a central frequency of 420 MHz, a maximum gain of 12 dB and a 1 dB compression point referred to the input of −4 dBm. To minimize parasitics in the RF measurements, the chip was directly soldered to the board (chip on board).
In order to demonstrate the sensor monitoring capability, the PA was AC driven with a heterodyne approach: two sinusoidal signals of input power Pi and frequencies (fRF Δf/2) and (fRF + Δf/2). This driving generated power dissipation in transistor M2, with a spectral component at Δf and an amplitude PΔf. Previous work [32,33] indicated that PΔf depended on both the voltage gain of the amplifier at fRF and on the input-output matching at the same frequency. As elaborated in Section 3, this power dissipation generated an AC sensor output voltage superimposed to VDD/2, whose content at Δf can be written as:
v O ( t ) = A O · cos ( 2 π · f i · t + Φ O ) .  

4.2. Output Power Monitoring

Figure 11 shows the sensor’s output voltage amplitude AO as a function of the total output power delivered by the PA to a 50 Ω load. The temperature transducer was biased with a current IB = 20 nA and the sensor’s filter with VBIAS = 1.2 V. The PA was driven with two tones with fRF = 420 MHz (its central frequency) and Δf = 1 kHz. Pi was swept to get a total output power delivered to the load ranging from −40 dBm to 0 dBm. This output power range was below the 1 dB compression point, ensuring constant PA gain for all the cases. The PA output power was measured with the Agilent E4443A spectrum analyzer. The sensor output voltage was measured with both the DMM in AC mode, and the LIA locked at Δf.
Focusing on DMM measurements, the sensor tracked the power delivered to the load when it was in the range [−25 dBm, −6 dBm] Below this range (−27 dBm, as labelled in Figure 11), DMM readings reached the noise level (which was 12 mVRMS,). At the other end of the sensor’s linear range above −6 dBm, the sensor output signal became clipped. LIA measurements show that the sensor was able to track the output power delivered to the load for values lower than −6 dBm, with very good agreement with DMM measurements in the range [−19 dBm, −6 dBm]. When the sensor output voltage became clipped, DMM measurements were slightly higher than LIA ones as DMM AC measurements take into account both the sensor’s output fundamental and the harmonics generated by the clipping.
If the sensor must track the PA output power for values higher than −6 dBm, the sensor sensitivity (AO/PA) should be reduced. For example, reducing AO/PA by a factor of ten would enable us to measure up to 4 dBm instead of −6 dBm. This would allow, for example, extending the sensor linear response up to the PA compression, and thus be able to monitor its 1-dB compression point.
Several strategies can be followed to reduce the sensor sensitivity: (i) increasing the distance between the CUT and the temperature transducer T1 in the IC layout [40]; (ii) decreasing the BPF gain below the current Av = 500, which opens the possibility to implement BPF amplifiers with tunable gain for dynamic range extension and (iii) changing the transducer’s T1 dimensions or bias [43,44]. All these strategies require either a custom layout depending on the target measurements or the design of additional complex circuits, such as tunable gain networks or a programmable transducer bias. Nevertheless, there is another strategy that can be used without redesigning the sensor or the sensor placement. As pointed out by Figure 9 and discussed in Section 3, the sensor sensitivity can be reduced by choosing a Δf value higher than fc. To illustrate this sensitivity reduction, Figure 12 shows the AO,RMS measured with the DMM when increasing Δf above 1 kHz, for three constant PA output power values. Focusing on Pout = −6 dBm, with Δf = 1 kHz the sensor output had already reached the saturation level observed in Figure 11. As Δf was increased, the measured output amplitude decreased (i.e., the sensor enters in the linear range), thus enabling the possibility to monitor higher output powers. Finally, when Δf = 200 kHz, the DMM output reached the noise level, and Pout = −6 dBm became the PoutNOISE_LEVEL. From Figure 12, PoutNOISE_LEVEL was −21 dBm when Δf = 10 kHz, and −11 dBm when Δf = 100 kHz. Therefore, Δf selection allowed us to easily adjust the linear response of the sensor to different ranges of dissipated power.

4.3. Central Frequency Monitoring

Figure 13 shows the RF power delivered to the load at the frequency (fRF + Δf/2), measured with the RF spectrum analyzer connected to the PA’s output; and the sensor’s output voltage measured with the DMM in AC mode; both as a function of fRF.
Sensor and PA had the same bias than the one reported in the previous section. In this experiment, the only swept input variable was the frequency fRF (from 120 to 920 MHz). In all the measurements, Δf = 1 kHz, and the total PA input power was −20 dBm (−23 dBm each tone), i.e., the PA had a linear behavior. The spectrum analyzer measurements indicated that the PA central frequency was 420 MHz, which agreed with the frequency fRF where the sensor’s output amplitude at Δf was maximum [31].

5. Conclusions

A single MOS transistor used as a temperature transducer connected to a BPF amplifier was presented, characterized and assessed for IC testing applications. The overall sensor circuit was implemented in a standard 0.35 µm CMOS technology, and was built-in with devices acting as controlled heat sources, and an RF power amplifier was used as a CUT. As strong points, heterodyne measurements could be done with a simple DMM, allowing a simplification of the measurement set-up. As the LIA was not required, there was no need for locking the frequencies of all the generators involved in the set-up measurements. Moreover, the task of the DMM could be easily integrated with the CUT, allowing a complete built-in self-test (BIST) solution. As weak point, as DMM integrated noise on a wider bandwidth, DMM readings did not reach the sensitivity levels achieved with an LIA. The sensor characterization showed that the nonlinear R2 behavior did not affect the sensor’s linearity as long as Δf was higher than the first BPF’s cut-off frequency. Moreover, the sensor sensitivity could be reduced by selecting a Δf higher than the BPF’s central frequency, allowing extension of the sensor’s dynamic range. As a proof of concept, we showed the feasibility of the circuit to track the output power delivered to the load and the central frequency of a RF class-A power amplifier.
Future directions of our research are the usage of temperature sensors as monitors in circuits used to compensate the effects of time-variability (e.g., aging) in RF circuits.

6. Patents

PCT/ES2013/070095: Sensor Circuit for Obtaining small-signal temperature measurements in integrated circuits.

Author Contributions

Conceptualization, F.R., E.B. and J.A.; methodology, F.R., D.M., X.P., and J.A.; circuit simulation, X.A.; circuit design, F.R., E.B., J.A.; board design, E.B. and F.R.; experimental validation, A.B., X.P., D.M., E.B. and J.A.; writing—original draft preparation, J.A., F.R., X.P., X.A. and D.M.; funding acquisition, D.M. and X.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Spanish AEI–Agencia Estatal de Investigación–grant number PID2019-103869RB-C33. (X.P.) has also received founds from the Spanish Ministry of Science, Innovation and Universities through Agencia Estatal de Investigación (AEI) (projects: HIPERCELLS, RTI2018-098392B-I00, and “Fiabilidad Inteligente”, PCI2020-112028).

Data Availability Statement

Data sharing is not applicable to this article.

Acknowledgments

Authors would like to thank Anant Rungta for the useful discussion and preliminary experimental work.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Altet, J.; Claeys, W.; Dilhaire, S.; Rubio, A. Dynamic surface temperature measurements in ICs. Proc. IEEE 2006, 94, 1519–1533. [Google Scholar] [CrossRef]
  2. Soden, J.J.; Anderson, R.E. IC failure analysis: Techniques and tools for quality reliability improvement. Proc. IEEE 1993, 81, 703–715. [Google Scholar] [CrossRef]
  3. Perpiñà, X.; Vellvehi, M.; Jordà, X. Chapter 13: Thermal issues in microelectronics. In Thermometry at the Nanoscale: Techniques and Selected Applications, 1st ed.; Nanoscience & Nanotechnology Series; Carlos, L., Palacio, F., Eds.; The Royal Society of Chemistry: Cambridge, UK, 2016. [Google Scholar]
  4. Hamann, H.F.; Weger, A.; Lacey, J.A.; Hu, Z.; Bose, P.; Cohen, E.; Wakil, J. Hotspot-limited microprocessors: Direct temperature and power distribution measurements. J. Solid State-Circuits 2007, 42, 56–65. [Google Scholar] [CrossRef]
  5. Perpiñà, X.; Jordà, X.; Vellvehi, M.; Altet, J. Study of heat sources interacting in integrated circuits by laser mirage effect. Appl. Phys. Lett. 2014, 105, 084101. [Google Scholar] [CrossRef]
  6. Arumi, D.; Rodriguez-Montanes, R.; Figueras, J.; Eichenberger, S.; Hora, C.; Kruseman, B.; Lousberg, M.; Majhi, A.K. Diagnosis of bridging defects based on current signatures at low power supply voltages. In Proceedings of the 25th IEEE VLSI Test Symposium (VTS’07), Berkeley, CA, USA, 6–10 May 2007; pp. 145–150. [Google Scholar] [CrossRef]
  7. Segura, J.; Rubio, A. A detailed analysis of CMOS SRAMs with gate oxide short defects. J. Solid State-Circuits 1997, 21, 1543–1550. [Google Scholar] [CrossRef]
  8. Lai, J.; Chandrachood, M.; Majumda, A.; Carrejo, J.P. Thermal detection of device failure by atomic force microscopy. IEEE Electron. Device Lett. 1995, 16, 312–315. [Google Scholar] [CrossRef]
  9. Bianic, S.; Allemand, S.; Kerrosa, G.; Scafidi, P.; Renard, D. Advanced backside failure analysis in 65 nm CMOS technology. Microelectron. Reliab. 2007, 47, 1550–1554. [Google Scholar] [CrossRef]
  10. Mahajan, R.; Chiu, C.; Chrysler, G. Cooling a microprocessor chip. Proc. IEEE 2006, 94, 1476–1486. [Google Scholar] [CrossRef]
  11. Chouhan, S.S.; Halonen, K. A low power temperature to frequency converter for the on-chip temperature measurement. IEEE Sens. J. 2015, 15, 4234–4240. [Google Scholar] [CrossRef] [Green Version]
  12. 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]
  13. Chen, P.-Y.; Huang, G.-Y.; Shyu, Y.-T.; Chang, S.-J. A primary-auxiliary temperature sensing scheme for multiple hotspots in system-on-a-chips. IEEE Sens. J. 2014, 14, 2633–2642. [Google Scholar] [CrossRef]
  14. Yang, T.; Kim, S.; Kinget, P.R.; Seok, M. Compact and supply-voltage-scalable temperature sensors for dense on-chip thermal monitoring. J. Solid State-Circuits 2015, 50, 2773–2785. [Google Scholar] [CrossRef]
  15. An, Y.J.; Ryu, K.; Jung, D.-H.; Woo, S.-H.; Jung, S.-O. An energy efficient time-domain temperature sensor for low-power on-chip thermal management. IEEE Sens. J. 2014, 14, 104–110. [Google Scholar] [CrossRef]
  16. Shor, J. Compact thermal sensors for dense cpu thermal monitoring and regulation: A review. IEEE Sens. J. 2020. [Google Scholar] [CrossRef]
  17. Eberlein, M.; Pretl, H. A No-Trim, Scaling-Friendly Thermal Sensor in 16 nm FinFET using bulk diodes as sensing elements. IEEE Solid-State Circuits Lett. 2019, 2, 63–66. [Google Scholar] [CrossRef]
  18. Bass, O.; Shor, J. A miniaturized 0.003 mm2 PNP-based thermal sensor for dense CPU thermal monitoring. IEEE Trans. Circuits Syst. I Reg. Pap. 2020, 67, 2984–2992. [Google Scholar] [CrossRef]
  19. Perpiñà, X.; Werkhoven, R.J.; Vellvehi, M.; Jakovenko, J.; Jordà, X.; Kunen, J.M.G.; Bancken, P.; Bolt, P.J. Thermal analysis of LED lamps for optimal driver integration. IEEE Trans. Power Electron. 2015, 30, 3876–3891. [Google Scholar] [CrossRef]
  20. Vellvehi, M.; Perpinya, X.; Leon, J.; Avino-Salvado, O.; Ferrer, C.; Jordà, X. Local thermal resistance extraction in monolithic microwave integrated circuits. IEEE Trans. Ind. Electron 2020. [Google Scholar] [CrossRef]
  21. Rodriguez, R.; Crespo-Yepes, A.; Martin-Martinez, J.; Nafria, M.; Aragones, X.; Mateo, D.; Barajas, E. Experimental monitoring of aging in CMOS RF linear power amplifiers: Correlation between device and circuit degradation. In Proceedings of the 2020 IEEE International Reliability Physics Symposium (IRPS), Dallas, TX, USA, 28 April–30 May 2020; pp. 1–7. [Google Scholar] [CrossRef]
  22. Crespo-Yepes, A.; Barajas, E.; Martin-Martinez, J.; Mateo, D.; Aragones, X.; Rodriguez, R.; Nafria, M. MOSFET degradation dependence on input signal power in a RF power amplifier. Microelectron. Eng. 2017, 178, 289–292. [Google Scholar] [CrossRef] [Green Version]
  23. Yuan, J.S.; Xu, Y.; Yen, S.D.; Bi, Y.; Hwang, G.W. Hot carrier injection stress effect on a 65 nm LNA at 70 GHz. IEEE Trans. Device Mater. Rel. 2014, 14, 931–934. [Google Scholar] [CrossRef]
  24. Simicic, M.; Weckx, P.; Parvais, B.; Roussel, P.; Kaczer, B.; Gielen, G. Understanding the impact of time-dependent random variability on analog ICs: From single transistor measurements to circuit simulations. IEEE Trans. Very Large Scale Integr. (VLSI) Syst. 2019, 27, 601–610. [Google Scholar] [CrossRef]
  25. Yen, H.D.; Yuan, J.S.; Huang, G.W.; Yeh, W.K.; Huang, F.S. Reliability performance of a 70-GHz mixer in 65-nm technology. IEEE Trans. Device Mater. Rel. 2016, 16, 101–104. [Google Scholar] [CrossRef]
  26. Vinaya, M.M.; Paily, R.; Mahanta, A. A new PVT compensation technique based on current comparison for low-voltage, near sub-threshold LNA. IEEE Trans. Circuits Syst. I Reg. Pap. 2015, 62, 2908–2919. [Google Scholar] [CrossRef]
  27. Barabino, N.; Silveira, F. Digitally assisted CMOS RF detectors with self-calibration for variability compensation. IEEE Trans. Microw. Theory Techn. 2015, 63, 1676–1682. [Google Scholar] [CrossRef]
  28. Wursthorn, J.; Knapp, H.; Al-Eryani, J.; Aufinger, K.; Maurer, L. Absolute mm-wave power sensor using a switching quad output stage. In Proceedings of the 2017 IEEE 17th Topical Meeting on Silicon Monolithic Integrated Circuits in RF Systems (SiRF), Phoenix, AZ, USA, 15–18 January 2017; pp. 40–42. [Google Scholar] [CrossRef]
  29. Onabajo, M.; Silva-Martinez, J. Analog Circuit Design for Process. Variation-Resilient System-on-a-Chip, 1st ed.; Springer: New York, NY, USA, 2012. [Google Scholar] [CrossRef]
  30. Bowers, S.M.; Sengupta, K.; Dasgupta, K.; Parker, B.D.; Hajimiri, A. Integrated self-healing for mm-wave power amplifiers. IEEE Trans. Microw. Theory Tech. 2013, 61, 1301–1315. [Google Scholar] [CrossRef]
  31. Perpiñà, X.; León, J.; Altet, J.; Vellvehi, M.; Reverter, F.; Barajas, E.; Jordà, X. Thermal phase lag heterodyne infrared imaging for current tracking in radio frequency integrated circuits. Appl. Phys. Lett. 2017, 110, 094101. [Google Scholar] [CrossRef] [Green Version]
  32. Perpiñà, X.; Reverter, F.; León, J.; Barajas, E.; Vellvehi, M.; Jordà, X.; Altet, J. Output power and gain monitoring in RF CMOS class A power amplifiers by thermal imaging. IEEE Trans. Instrum. Meas. 2019, 68, 2861–2870. [Google Scholar] [CrossRef] [Green Version]
  33. Aldrete, E.; Mateo, D.; Altet, J.; Salhi, M.A.; Grauby, S.; Dilhaire, S.; Onabajo, M.; Silva-Martinez, J. Strategies for built-in characterization testing and performance monitoring of analog RF circuits with temperature measurements. Meas. Sci. Technol. 2010, 21, 075104. [Google Scholar] [CrossRef]
  34. Altet, J.; Gómez, D.; Perpiñà, X.; Mateo, D.; González, J.L.; Vellvehi, M.; Jordà, X. Efficiency determination of RF linear power amplifiers by steady-state temperature monitoring using built-in sensors. Sens. Act. A Phys. 2013, 192, 49–57. [Google Scholar] [CrossRef]
  35. Altet, J.; Mateo, D.; Gómez, D.; González, J.L.; Martineau, B.; Siligaris, A.; Aragones, X. Temperature sensors to measure the central frequency and 3 dB bandwidth in mmW power amplifiers. IEEE Microw. Wirel. Compon. Lett. 2014, 24, 272–274. [Google Scholar] [CrossRef]
  36. Onabajo, M.; Gómez, D.; Aldrete, E.; Altet, J.; Mateo, D.; Silva-Martinez, J. Survey of robustness enhancement techniques for wireless systems-on-a-chip and study of temperature as observable for process variations. J. Electron. Test. 2011, 27, 225–240. [Google Scholar] [CrossRef] [Green Version]
  37. Onabajo, M.; Altet, J.; Aldrete, E.; Mateo, D.; Silva-Martinez, J. Electrothermal design procedure to observe RF circuit power and linearity characteristics with homodyne differential temperature sensor. IEEE Trans. TCAS I 2011, 58, 458–469. [Google Scholar] [CrossRef]
  38. Abdallah, L.; Stratigopoulos, H.; Mir, S.; Altet, J. Defect-oriented non-intrusive RF test using on-chip temperature sensors. In Proceedings of the 31st IEEE VTS, Berkeley, CA, USA, 29 April–2 May 2013. [Google Scholar] [CrossRef]
  39. Altet, J.; Rubio, A.; Rossello, J.L.; Segura, J. Structural RFIC device testing through built-in thermal monitoring. IEEE Commun. Mag. 2003, 41, 98–104. [Google Scholar] [CrossRef]
  40. Altet, J.; Rubio, A.; Schaub, E.; Dilhaire, S.; Claeys, W. Thermal coupling in integrated circuits: Application to thermal testing. J. Solid State-Circuits 2001, 36, 81–91. [Google Scholar] [CrossRef]
  41. Barajas, E.; Aragones, X.; Mateo, D.; Altet, J. Differential temperature sensors: Review of applications in the test and characterization of circuits, usage and design methodology. Sensors 2019, 19, 4815. [Google Scholar] [CrossRef] [Green Version]
  42. Reverter, F.; Gómez, D.; Altet, J. On-chip MOSFET temperature sensor for electrical characterization of RF circuits. IEEE Sensors J. 2013, 13, 3343–3344. [Google Scholar] [CrossRef]
  43. Altet, J.; Aldrete, E.; Mateo, D.; Perpiñà, X. A heterodyne method for the thermal observation of the electrical behavior of high-frequency integrated circuits. Meas. Sci. Technol. 2008, 19, 115704. [Google Scholar] [CrossRef]
  44. Reverter, F.; Altet, J. On-chip thermal testing using MOSFETS in weak inversion. IEEE Trans. Instrum. Meas. 2015, 64, 524–532. [Google Scholar] [CrossRef] [Green Version]
  45. Reverter, F.; Altet, J. MOSFET temperature sensors of on-chip thermal testing. Sens. Act. A Phys. 2013, 203, 234–240. [Google Scholar] [CrossRef]
  46. Olsson, R.H.; Buhl, D.L.; Sirota, A.M.; Buzsaki, G.; Wise, K.D. Band-tunable and multiplexed integrated circuits for simultaneous recording and stimulation with microelectrode arrays. IEEE Trans. Biomed. Eng. 2005, 52, 1303–1311. [Google Scholar] [CrossRef]
  47. Olsson, R.H.; Gulari, A.N.; Wise, K.D. A fully integrated bandpass amplifier for extracellular neural recording. In Proceedings of the 1st International IEEE EMBS Conference on Neural Engineering, Capri, Italy, 20–22 March 2003; pp. 165–168. [Google Scholar]
  48. Rezaee-Dehsorkh, H.; Ravanshad, N.; Lotfi, R.; Mafinezhad, K.; Sodagar, A.M. Analysis and design of tunable amplifiers for implantable neural recording applications. IEEE J. Emerg. Sel. Top. Circuits Syst. 2011, 1, 546–556. [Google Scholar] [CrossRef]
  49. Perlin, G.E.; Wise, K.D. An ultra compact integrated front end for wireless neural recording microsystems. J. Microelectromech. Syst. 2010, 19, 1409–1421. [Google Scholar] [CrossRef] [Green Version]
  50. Altet, J.; González, J.L.; Gómez, D.; Perpiñà, X.; Claeys, W.; Grauby, S.; Dufis, C.; Vellvehi, M.; Mateo, D.; Reverter, F.; et al. Electro-thermal characterization of a differential temperature sensor in a 65 nm CMOS IC: Applications to gain monitoring in RF amplifiers. Microelectron. J. 2012, 45, 484–490. [Google Scholar] [CrossRef] [Green Version]
  51. Nenadovic, N.; Mijalkovic, S.; Nanver, L.K.; Vandamme, L.K.J.; d’Alessandro, V.; Schellevis, H.; Slotboom, J.W. Extraction and modeling of self-heating and mutual thermal coupling impedance of bipolar transistors. J. Solid State-Circuits 2004, 39, 1764–1772. [Google Scholar] [CrossRef]
  52. Reverter, F.; Perpiñà, X.; Barajas, E.; León, J.; Vellvehi, M.; Jordà, X.; Altet, J. MOSFET dynamic thermal sensor for IC testing applications. Sens. Act. A Phys. 2016, 242, 195–202. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Circuit under test (CUT) with heterodyne driving. Temperature sensor made of a single metal -oxide semiconductor (MOS), transistor temperature transducer (T1), an active Band Pass Filter (BPF), and a Digital Multimeter (DMM). Temperature variations at Δf carry information about the CUT figures of merit at f1.
Figure 1. Circuit under test (CUT) with heterodyne driving. Temperature sensor made of a single metal -oxide semiconductor (MOS), transistor temperature transducer (T1), an active Band Pass Filter (BPF), and a Digital Multimeter (DMM). Temperature variations at Δf carry information about the CUT figures of merit at f1.
Sensors 21 00805 g001
Figure 2. Temperature sensor schematic. (a) Temperature transducer based on an nMOS transistor (T1). (b) Active band pass filter (BPF). In our design: Iext is off-chip. (c) Typical BPF frequency response of the circuit in (b), characterized by the low-frequency (fp1) pole, the high-frequency (fp2) pole and the band-pass gain Av.
Figure 2. Temperature sensor schematic. (a) Temperature transducer based on an nMOS transistor (T1). (b) Active band pass filter (BPF). In our design: Iext is off-chip. (c) Typical BPF frequency response of the circuit in (b), characterized by the low-frequency (fp1) pole, the high-frequency (fp2) pole and the band-pass gain Av.
Sensors 21 00805 g002
Figure 3. R2 implemented with two pMOS transistors biased in subthreshold.
Figure 3. R2 implemented with two pMOS transistors biased in subthreshold.
Sensors 21 00805 g003
Figure 4. (a) Schematic of the nMOS heater and the temperature transducer T1. (b) Layout floorplan.
Figure 4. (a) Schematic of the nMOS heater and the temperature transducer T1. (b) Layout floorplan.
Sensors 21 00805 g004
Figure 5. Measurements of transducer sensitivity as a function of the frequency. Three different voltage levels (VL) for the heating device are considered.
Figure 5. Measurements of transducer sensitivity as a function of the frequency. Three different voltage levels (VL) for the heating device are considered.
Sensors 21 00805 g005
Figure 6. Frequency response of the amplifier BPF’s voltage gain, plot for three different VBIAS values: 1.15, 1.2 and 1.25 V. For each VBIAS, measurements are taken either with a digital multimeter (DMM) or with a lock-in amplifier (LIA).
Figure 6. Frequency response of the amplifier BPF’s voltage gain, plot for three different VBIAS values: 1.15, 1.2 and 1.25 V. For each VBIAS, measurements are taken either with a digital multimeter (DMM) or with a lock-in amplifier (LIA).
Sensors 21 00805 g006
Figure 7. fp1 as a function of VBIAS. Measurements performed in three different samples.
Figure 7. fp1 as a function of VBIAS. Measurements performed in three different samples.
Sensors 21 00805 g007
Figure 8. Sensor’s output voltage as a function of the frequency for four different heater bias. Measurements done with the DMM.
Figure 8. Sensor’s output voltage as a function of the frequency for four different heater bias. Measurements done with the DMM.
Sensors 21 00805 g008
Figure 9. Sensor’s sensitivity as a function of the frequency. See Table 2 for the VL description and the PA values. Measurements taken with the LIA.
Figure 9. Sensor’s sensitivity as a function of the frequency. See Table 2 for the VL description and the PA values. Measurements taken with the LIA.
Sensors 21 00805 g009
Figure 10. Schematic of the RF power amplifier used as CUT. The temperature transducer is placed close the cascode transistor M2. Inset shows a photo of the IC layout, with the placement of the active transistor M1, cascode transistor M2 (formed with three equal transistors connected in parallel), and the MOS acting as temperature transducer.
Figure 10. Schematic of the RF power amplifier used as CUT. The temperature transducer is placed close the cascode transistor M2. Inset shows a photo of the IC layout, with the placement of the active transistor M1, cascode transistor M2 (formed with three equal transistors connected in parallel), and the MOS acting as temperature transducer.
Sensors 21 00805 g010
Figure 11. Sensors output voltage amplitude as a function of the output power delivered by the CUT to the load. Measurements obtained with a LIA and a DMM in AC mode. Label indicates at which POUT the DMM readings reach the noise level.
Figure 11. Sensors output voltage amplitude as a function of the output power delivered by the CUT to the load. Measurements obtained with a LIA and a DMM in AC mode. Label indicates at which POUT the DMM readings reach the noise level.
Sensors 21 00805 g011
Figure 12. Sensor output voltage amplitude as a function of Δf for three different output power levels. DMM measurements.
Figure 12. Sensor output voltage amplitude as a function of Δf for three different output power levels. DMM measurements.
Sensors 21 00805 g012
Figure 13. Sensor’s output measured with the DMM, superposed to power delivered to the load at the frequency (fRF + Δf/2) as a function of fRF. Δf = 1 kHz, and total PA input power = −20 dBm (−23 dBm/tone). Sensor output voltage represented as in [33].
Figure 13. Sensor’s output measured with the DMM, superposed to power delivered to the load at the frequency (fRF + Δf/2) as a function of fRF. Δf = 1 kHz, and total PA input power = −20 dBm (−23 dBm/tone). Sensor output voltage represented as in [33].
Sensors 21 00805 g013
Table 1. Active band pass filter (BPF) simulated characteristics.
Table 1. Active band pass filter (BPF) simulated characteristics.
VBIAS
(V)
R2
(GΩ)
fp1
(Hz)
fp2
(kHz)
Av_max
(dB)
Noise
(μV2RMS)
THDmax
(dB)
1.151.479735.554.82.04−15.7
1.26.422484.654.72.22−29.9
1.2528.5554.654.52.46−52.7
Table 2. Excitation applied to the heater.
Table 2. Excitation applied to the heater.
VLVDCh (V)A (mVRMS)PA (mW)
11.677.80.44
21.67150.84
31.67231.29
41.67311.75
Table 3. Active BPF measured characteristics.
Table 3. Active BPF measured characteristics.
VBIAS (V)fp1 (Hz)fp2 (kHz)Av_max (dB)
1.1512585.154.1
1.23503.954.9
1.25793.954.6
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Altet, J.; Barajas, E.; Mateo, D.; Billong, A.; Aragones, X.; Perpiñà, X.; Reverter, F. BPF-Based Thermal Sensor Circuit for On-Chip Testing of RF Circuits. Sensors 2021, 21, 805. https://doi.org/10.3390/s21030805

AMA Style

Altet J, Barajas E, Mateo D, Billong A, Aragones X, Perpiñà X, Reverter F. BPF-Based Thermal Sensor Circuit for On-Chip Testing of RF Circuits. Sensors. 2021; 21(3):805. https://doi.org/10.3390/s21030805

Chicago/Turabian Style

Altet, Josep, Enrique Barajas, Diego Mateo, Alexandre Billong, Xavier Aragones, Xavier Perpiñà, and Ferran Reverter. 2021. "BPF-Based Thermal Sensor Circuit for On-Chip Testing of RF Circuits" Sensors 21, no. 3: 805. https://doi.org/10.3390/s21030805

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