Next Article in Journal
Investigations on Field Distribution along the Earth’s Surface of a Submerged Line Current Source Working at Extremely Low Frequency Band
Next Article in Special Issue
Microfabrication, Characterization, and Cold-Test Study of the Slow-Wave Structure of a Millimeter-Band Backward-Wave Oscillator with a Sheet Electron Beam
Previous Article in Journal
Using Enhanced Test Systems Based on Digital IC Test Model for the Improvement of Test Yield
Previous Article in Special Issue
Experimental Study of AM and PM Noise in Cascaded Amplifiers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Advances in Microwave Large-Signal Metrology: From Vector-Receiver Load-Pull to Vector Signal Network Analyzer and Time-Domain Load-Pull Implementations (Invited Paper)

by
J. Apolinar Reynoso-Hernández
1,*,
Manuel Alejandro Pulido-Gaytan
1,
Thaimí Niubó-Alemán
1,2 and
Marlon Molina-Ceseña
1
1
Centro de Investigación Científica y de Educación Superior de Ensenada, Baja California, División de Física Aplicada, Departamento de Electrónica y Telecomunicaciones, Ensenada CP 22860, Mexico
2
Department of Electrical and Computer Engineering, The Ohio State University, Columbus, OH 43210, USA
*
Author to whom correspondence should be addressed.
Electronics 2022, 11(7), 1114; https://doi.org/10.3390/electronics11071114
Submission received: 21 February 2022 / Revised: 13 March 2022 / Accepted: 14 March 2022 / Published: 31 March 2022
(This article belongs to the Special Issue Analysis and Test of Microwave Circuits and Subsystems)

Abstract

:
Radiofrequency (RF) power amplifiers (PAs) are important elements of modern communication systems. The most important components in PAs are the transistors, which are operated under large-signal regimes in such applications. Designing and optimizing PAs are challenging tasks that demand the highest accuracy on large-signal measurements at both device and circuit levels. Large-signal power transistor characterization demands the development and utilization of high-frequency and low-frequency load-pull systems. Furthermore, the static and dynamic characterization of two-input PAs requires the development of new measurement systems that consist of an nonlinear vector network analyzer (NVNA) and an arbitrary waveform generators (AWG). This paper reviews the research activities, achievements, and current research goals at CICESE Research Center in Mexico in the large-signal microwave metrology field.

1. Introduction

Although advances in wireless communications are usually driven by developments in digital coding, modulation techniques, and the liberation of new fragments of the electromagnetic spectrum, the elements that remain present in every communication system are the power amplifiers (PAs) [1,2,3]. The most important elements in a PA are the transistors, which are operated under large-signal regimes in such applications [4]. Hence, characterizing the nonlinear behavior of power devices, either transistors or amplifiers, is of paramount importance for PA design and optimization [5,6].
Characterization of the large-signal behavior of power devices using only S-parameters is insufficient since the representation of their performance by means of S-parameters is only valid when describing their small-signal behavior [7]. In this paper, a review of several solutions aimed to assess the large-signal behavior of power devices is presented, from vector-receiver load-pull systems to the implementation of measurement systems dealing with wideband modulated signals, and time-domain load-pull solutions.
The paper is organized as follows. Section 2 presents the foundations of vector-receiver load-pull (VRLP) systems; a description of their calibration procedure is presented and the impact of the calibration techniques on devices’ characterization is studied. Section 3 shows a measurement system that uses a vector receiver jointly with arbitrary waveform generators to characterize devices’ performance under modulated signal excitations, hence transforming a conventional vector receiver into a vector signal analyzer. In Section 4, a low-frequency time-domain harmonic load-pull system is built by substituting the mixer-based vector receiver used in VRLP systems by a four-channel oscilloscope to capture voltage and current waveforms at the device’s intrinsic plane. Such a system is used to evaluate and compare the RF capabilities of power transistors fabricated in different technologies. Section 5 is devoted to presenting a final summary along with the current research goals being pursued at CICESE Research Center in Mexico in the field of microwave large-signal metrology.

2. Vector-Receiver Load-Pull Systems: Description and Calibration Procedure

The demand for enhanced performance from microwave power devices requires operating them in their large-signal regime. The characterization of nonlinear devices requires complex measurement systems, where the most utilized are the load-pull systems [8,9,10]. In legacy power-based load-pull systems, measurement accuracy relied on pre-calibrated impedance tuners and power receivers [11,12]. By comparison, VRLP systems (Figure 1) measure the impedances presented to the device under test (DUT) and the power at its ports in real-time [13,14] using a calibrated vector receiver [15].

2.1. System Calibration

Characterizing devices using a VRLP system requires removal of the effects of the system’s elements using calibration methods [16,17,18,19]. The calibration of a system such as the one depicted in Figure 1 comprises two parts: relative calibration, used to determine ratios of parameters, and power calibration, used to determine the power levels at the DUT ports. Relative calibration may be carried out using two-port calibration techniques, whereas power calibration is performed using a power reference device (e.g., power meter) connected at a coaxial plane located close to the calibration plane. In this section, a complete mathematical formulation of the calibration procedure, developed using the ABCD-parameters matrix formalism [18,19], is presented.

2.1.1. Relative Calibration

At present, vector network analyzers (VNAs) allowing access to the instrument receivers are used as the vector receiver (VR) in load-pull systems (Figure 2a). Thus, VNA calibration techniques, such as the thru-reflect-line (TRL) [20,21,22], thru-reflect-match (TRM) [23,24], and short-open-load with unknown thru (SOLT) [25] may be used to carry out the relative calibration of VRLP systems. This calibration relates ratioed quantities measured at the VR (U1-U2 planes), to ratioed quantities at the calibration plane (D1-D2 planes), as depicted in Figure 2b. The voltage and current quantities at the vector receiver ports, V k U , and I k U ; k = 1 , 2 , are defined as a function of the measured traveling waves, A k U , and B k U , as [7]:
V k U = Z 0 1 + B k U A k U A k U = Z 0 A k U B k U + 1 B k U
I k U = 1 Z 0 1 B k U A k U A k U = 1 Z 0 A k U B k U 1 B k U
where Z0 represents the measurement system impedance. Using the eight-term error model, V k U and I k U are defined as a function of the voltage and current quantities at the calibration reference plane, V k and I k , as:
V 1 I 1 = T A 1 V 1 U I 1 U
V 2 I 2 = T B V 2 U I 2 U
T A and T B are ABCD-parameters matrices denoting the networks connecting U1-U2 planes to D1-D2 planes. As shown in Appendix A, these matrices are defined as:
T A = D X A X ¯ B X ¯ C X ¯ 1 Z B Z A 1 1 1
T B = D Y Z B Z A 1 1 A Y ¯ B Y ¯ C Y ¯ 1
where ZA and ZB represent the impedance of a calibration structure used in a particular calibration technique. In the TRL technique, ZA = ZB = ZL, where ZL represents the characteristic impedance of the line used as calibration structure [18,22]. In the TRM technique, ZA equals the impedance of the load used as the match standard at port one, and ZB equals the impedance of the load used at port two [19,24]. Seven out of the eight error terms in Equations (5) and (6), namely A X ¯ , B X ¯ , C X ¯ , A Y ¯ , B Y ¯ ,   C Y ¯ , D X D Y , may be determined from measurements of either TRL or TRM calibration structures; Table A1 in Appendix A summarizes the measurements necessary to determine them on both TRL and TRM techniques.
Let R z be the ratio of Z B to Z A ( R z = Z B / Z A ). By substituting Equations (5) and (6) in Equations (3) and (4), and developing the resultant equations, the following expressions for the voltages and currents at the calibration plane may be obtained as:
V 1 = Z A D X Δ X R Z + C X ¯ B X ¯ R Z + A X ¯ Y 1 U V 1 U
I 1 = 1 D X Δ X 1 C X ¯ B X ¯ A X ¯ Y 1 U V 1 U
V 2 = Z A D Y C Y ¯ A Y ¯ R Z 1 B Y ¯ R Z Y 2 U V 2 U
I 2 = D Y C Y ¯ + A Y ¯ 1 + B Y ¯ Y 2 U V 2 U
where Δ X = A X ¯ B X ¯   C X ¯ and Y k U = I k U / V k U ; k = 1,2. From Equations (7)–(10), it may be noticed that any ratioed quantity at the calibration plane (e.g., V 2 / I 2 depends only on the seven terms calculated from the relative calibration, the value of ZA and ZB, and ratioed quantities measured at the VR (e.g., I 2 U / V 2 U , V 2 U / V 1 U ). Let us consider three quantities commonly used in large-signal characterization: the input impedance ( Z I N = V 1 / I 1 ), the load impedance ( Z L D = V 2 / I 2 ), and the voltage gain ( G V = V 2 / V 1 ), which may be expressed as follows:
Z I N = Z A R Z + C X ¯ B X ¯ R Z + A X ¯ Y 1 U 1 C X ¯ B X ¯ A X ¯ Y 1 U
Z L D = Z A C Y ¯ A Y ¯ R Z 1 B Y ¯ R Z Y 2 U C Y ¯ + A Y ¯ 1 + B Y ¯ Y 2 U
G V = D X D Y Δ X C Y ¯ A Y ¯ R Z 1 B Y ¯ R Z Y 2 U R Z + C X ¯ B X ¯ R Z + A X ¯ Y 1 U V 2 U V 1 U
Although ZLD is not a power-dependent quantity, ZIN and GV, along with other gain metrics (e.g., power gain), are power dependent. Nevertheless, such dependency cannot be captured using information provided by the relative calibration process.

2.1.2. Power Calibration

To complete the calibration, we must determine the power at the calibration plane for which ratioed quantities obtained from the relative calibration are defined. Expressions for the power at the D1 and D2 planes, PIN and POUT, are obtained from the analysis of the structures depicted in Figure 2b as:
P I N = V 1 2 Re Y I N ,
P O U T = V 2 2 Re Y L D .
YIN = 1/ZIN and YLD = 1/ZLD are determined from the relative calibration. Thus, |V1|and |V2| must be determined. As shown in Equations (7) and (9), this requires calculating the value of |DX| and |DY|; the procedure presented next describes how to determine these terms.
In most practical cases, power devices are characterized in non-connectorized environments (e.g., in-fixture, on-wafer). In these conditions, a power reference cannot be connected at the calibration plane. A more general procedure consists in connecting the power reference at a coaxial plane near the calibration plane, while a thru connection is performed at the D1-D2 planes. (For the sake of clarity, power calibration is described for the case in which a zero-length thru connection is connected at the calibration plane. A similar procedure may be derived for cases in which the test ports are connected through transmission lines of arbitrary length [16,17], with no major modifications.) (Figure 3a) [13,16,17]. As shown in Figure 3b, there is a two-port network connecting the D1-D2 plane to a coaxial plane (C0). Let TC, defined as:
T C = D C A C ¯ B C ¯ C C ¯ 1 ,
be the ABCD-parameters matrix representing such two-port network, such that:
V 1 I 1 = V 2 I 2 = T C V C I C .
By substituting Equation (16) in (17), Equation (18) may be derived:
A C ¯ Z C + B C ¯ C C ¯ Z C Z I C = Z I C ,
Z 1 C = V 1 / I 1 = V 2 / I 2 represents the impedance at the calibration plane’s input impedance when a load of known impedance Z C = V C / I C is attached at the plane (C0) as shown in Figure 3. Then, by using the connection of three loads at such a plane and using Equation (18), one can form a system of equations, which, after being solved, allows determination of the values of A C , B C and C C .
The next step in the power calibration is to connect a power reference (power meter) at the plane C0. First, the impedance of the power meter ( Z P R ) may be determined by solving Equation (18) for the term Z C as:
Z P R = Z I C , P R B C ¯ / A C ¯ C C ¯ Z I C , P R ,
where Z 1 C , P R denotes the input impedance at plane D1-D2, when the input impedance at the calibration plane and a power meter of impedance ZPR is connected at the plane C0. By analyzing the structure presented in Figure 3b, and using Equation (16), one can derive the following expression:
V 1 = V 2 = D C A C ¯   Z P R + B C ¯ V P R ,
where VPR represents the voltage at the power reference. Let P P R = V P R 2 Re Y P R , Y P R = 1 / Z P R , be the power measured at the power reference; from Equation (20) the following expression is obtained:
V 1 = V 2 = D C A C ¯ Z P R + B C ¯ P P R / Re Y P R .
Since the two-port network represented by TC is assumed to be reciprocal [16,17], the determinant of TC should be identical to the unity, and the value of D C = 1 / A C ¯ B C ¯ C C ¯ .
Finally, the following expressions for |DX| and |DY| may be determined by solving Equations (7) and (9) for DX and DY, respectively, and using Equation (21) in the obtained equations:
D X = Z A D C R Z + C X ¯ B X ¯ R Z + A X ¯ Y 1 U Δ X A C ¯ Z P M + B C ¯ P P M / Re Y P M V 1 U ,
D Y = D C Z A A C ¯ Z P M + B C ¯ 2 P P M / Re Y P M C Y ¯ A Y ¯ R Z 1 B Y ¯ R Z Y 2 U 1 V 2 U .
From Equations (7), (14) and (22), it is noted that power calibration relates PIN to an absolute quantity measured at the VR, V 1 U . Similarly, Equations (9), (15) and (23), establish a relationship between POUT and an absolute quantity measured at the VR, V 2 U .

2.2. Large-Signal Characterization of Power Devices

In this section, two cases of devices’ large-signal characterization are presented. The first shows the usefulness of the TRL calibration implemented using lines of arbitrary impedance. In the second case, a transistor is measured in a system calibrated using the TRM calibration procedure, implemented using nonsymmetrical loads. In both cases, a Keysight PNA-X N5245A is used as a vector receiver, and a mechanical tuner is used to vary the load impedance.

2.2.1. Characterization of Devices in Systems Calibrated Using the TRL Technique

With the aim of measurement, packaged transistors are accommodated in low-impedance transmission lines; the higher the device’s periphery, the lower the line’s impedance. Thus, the test fixtures used to accommodate these devices use impedance transformers to adapt transmission lines of low impedance to Z 0 (Figure 4a).
A 45 W transistor from Wolfspeed (CGH40045) was mounted in a fixture that comprises Klopfestein impedance transformers, with the aim of adapting transmission lines of 10 Ω characteristic impedance to 50 Ω transmission lines (Figure 4a). The system was calibrated at 3 GHz using the TRL calibration structures shown in Figure 4b–d, and contours of constant P O U T at the DUT plane were measured. The actual value of ZL was assumed to be purely real (10 Ω).
Figure 5 shows contours of constant output power determined using ZL = 10 Ω in the system calibration, along with output power contours determined using several erroneous ZL values. As noted from Equation (12), since in the TRL technique RL = 1, errors in ZL represent proportional errors in ZLD. Hence, the impedance for maximum power (ZOPT) may be either underestimated or overestimated when the wrong value of ZL is used during the system calibration. Similar linear dependence of ZIN as a function of ZL was reported in [18].
The impact of errors in ZL in determining the large-signal gain is also addressed. From Equation (13), it may be noted that the calculation of the voltage gain (GV) is not dependent on ZL. Thus, the gain defined in terms of the incident (A1) and transmitted (B2) waves was evaluated as:
G D = B 2 A 1 = V 2 Z 0   I 2 V 1 + Z 0   I 1 = 1 + Z 0 / Z L D 1 + Z 0 / Z I N G V .
The GD calculation depends on ZL through ZIN and ZLD. The plots shown in Figure 6 show the amplitude and angle of GD determined using different ZL values in the calibration. It was observed that the calculation of the phase is more sensitive to ZL than the magnitude.

2.2.2. Characterization of Devices in Systems Calibrated Using the TRM Technique

This section shows the usefulness of the proposed calibration procedure, allowing nonsymmetrical loads of arbitrary impedance to be used as a standard in the TRM technique. A VRLP system was calibrated at the center frequency of 3.5 GHz using the calibration structures shown in Figure 7b–d. The load (match) standard is formed by a 50 Ω resistor at port one and two 50 Ω resistors (connected in parallel) at port two. The loads at ports one and two were characterized using a network analyzer calibrated using the TRL technique [21,22]. The impedance of these loads was determined at 3.5 GHz as 53 + j13 Ω and 26 + j10 Ω, respectively.
Contours of constant Pout obtained from the measurement of a packaged transistor from the Wolfspeed (CGH40010) were measured, and the importance of taking into account the electrical characteristics of the match standard on the calculation of output power contours was analyzed.
First, to evaluate the accuracy of the TRM procedure, the results were compared to those obtained using the TRL calibration. The TRL procedure was implemented using the thru and reflecting loads used in the TRM technique and an additional transmission line. Figure 8 shows output power contours determined using these two techniques. The contours calculated using both procedures present high correlation, thus showing the accuracy of the proposed procedure.
Then, the impact of not taking into account the electrical characteristics of the match structure on the determination of output power contours was experimentally studied by considering the following three cases [19]:
  • The value of ZM1 and ZM2 were considered as nonsymmetrical, but purely real and identical to their DC values (ZM1 = 50 Ω; ZM2 = 25 Ω). Figure 9a shows the calculated contours of constant P O U T . These contours considerably differ from those shown in Figure 8a,b.
  • Contours of constant output power determined by considering that the match structure is formed by two loads of identical impedance of 53 + j13 Ω are shown in Figure 9b.
  • Similarly, contours of constant P O U T , shown in Figure 9c, were determined considering that the match standard is formed by two loads of identical impedance. The impedance of these loads is assumed to be identical to ZM1 = ZM2 = 26 + j10 Ω.
For the case considered in this paper (ZM2 < ZM1), as shown in Figure 9b,c, ignoring the asymmetry of the match standard caused a contraction in the calculated impedance map. On the other hand, the load impedance map rotates when the imaginary part of the match standard impedance is ignored, as presented in Figure 9a. More results in which simultaneous rotation and contraction, caused by ignoring both asymmetry and frequency-dependence of ZM2 and ZM1, are shown in [19].

3. Large-Signal Characterization Using Modulated Signals

Modern communication systems are empowered by the use of complex modulation formats (e.g., Orthogonal Frequency-Division Multiplexing, OFDM) in order to meet data rate and spectral efficiency requirements. However, using these modulation formats results in the appearance of intermittent peaks that generate high Peak to Average Power Ratio (PAPR) levels. In order to obtain a desired performance (efficiency, linearity) at peak and average power levels, transmitter configurations using multiple amplifiers, e.g., Doherty, Chireix, have to be used. The principle of operation of these transmitter configurations is similar: the load of an amplifier, referred to as main, is modulated by an auxiliary amplifier, guaranteeing maximum swing in voltage and current independently of the signal input power.
Figure 10 shows the measurement system developed in [26,27], which aims to capture both static and dynamic behavior of the two amplifiers in a transmitter based on Doherty or Cheirex principles, as well as the behavior of the whole transmitter. In this system, the modulated baseband signals used to drive the inputs of the amplifiers are upconverted to RF frequency using a dual-channel AWG (Keysight M8190A). In the AWG, the IQ modulation is performed in the digital domain, thus avoiding signal distortion typically caused by analog IQ modulators. Highly linear amplifiers are used to boost the signals created by the AWG in order to properly drive the transmitter’s inputs. Bi-directional couplers along with Keysight PNA-X’s external receivers (including the H85 option) are used to sample the incident and reflected waveforms at both the inputs and the output of the transmitter, as depicted in Figure 10. In order to capture both CW and modulated signals, the PNA’s Spectrum Analyzer option was used. Independent DC supplies were utilized to bias the gate and drain of the two amplifiers, and two digital multimeters were used to measure the current at the output of the main and peaking amplifiers.

3.1. Calibration of the Measurement System

In order to accurately characterize dual-input devices using the testbed shown in Figure 10, the system has to be calibrated using the procedure introduced in [26,27]. Since the system comprises three ports (two inputs and one output), three-port relative calibration along with absolute power calibration are required. The calibration reference planes are denoted in Figure 10 with green lines. For the relative calibration, the well-known short-open-load-reciprocal (SOLR) calibration technique [25] was used (center frequency = 2.08 GHz, span = 500 MHz, tone spacing = 5 kHz). The phase of the different tones, comprising the signals at the input and output ports of the transmitter, are calculated by utilizing the PNA-X’s spectrum analyzer option. It is worth mentioning that, in this system, no phase reference is required for the measurement to be performed, as in the case of the well-known Nonlinear Vector Network Analyzer.
The calibrated PNA-X’s receivers realize a five-channel Vector Signal Analyzer. Then, the system proposed in [27] is referred to as Vector Signal Network Analyzer (VSNA). The amplitude- and phase-corrected signals at the PNA-X are captured by calculating ϕ r c a l i , the differences between the signal at the VSNA and the reference OFDM signal at the AWG’s input, for the i tones at the frequency ω i as:
ϕ r C A L i = ϕ r r a w i ϕ A W G i + ω i ω L O Δ τ r + Δ ϕ L O , r
where Δ τ r represents the delay in each PNA-X’s receiver and Δ ϕ L O , r represents the local oscillator phase. Since, in practice, ϕ r C A L i is nearly identical for all the receivers, the average phase correction ϕ a v g C A L i over the receivers can be used to simplify Equation (25) as:
ϕ r C A L i = ϕ a v g C A L i + δ ϕ r C A L i
with δ ϕ r C A L i defined as the residual phase correction for the channel r.
In Figure 11 an example of the amplitude and phase correction for three receivers is shown.
The signal under analysis is an OFDM signal, with 64 QAM, tone spacing = 5 kHz, bandwidth = 500 MHz, and a sampling frequency = 1 GHz, where the average phase correction for the five receivers of the VNA is the term ϕ a v g C A L i and the term δ ϕ r C A L i represent the residual phase correction for the channel r .
The amplitude and phase corrections obtained for each receiver are shown in Figure 11a,b. For their extraction, an OFDM signal was used (bandwidth 500 MHz, 64 QAM), with tone spacing of 5 kHz, sampling frequency of 1 GHz, an average power of 18.87 dBm and PAPR 10.78 dB.
Figure 12 shows the phases of the 100,000 subcarriers (original and received) along with the relative errors for each subcarrier. For the sake of clarity, the phase of each subcarrier after alignment and after residual phase correction are also presented. The nearly constant difference between the original and the aligned phases ( Δ ϕ L O ) is used to define a phase offset between local oscillators of the AWG and the developed VSNA.
Figure 13 shows a flowchart that summarizes the post-processing calibration performed by the VSNA on the raw data.
In order to assess the accuracy of the VSNA, a comparison of its performance against a commercial VSA was performed. Table 1 shows the SNR and EVM values calculated after correction for the case of several OFDM signals. The comparison was performed over different signal bandwidths. The results show that the EVM obtained using the VSNA is about 10 dB smaller than that obtained using the VSA, while the SNR is about 10 dB higher for all the cases analyzed.

3.2. HD-OPA Dynamic Characterization

To demonstrate the usefulness of the measurement testbed previously described, the static and dynamic behavior of a hybrid Doherty-Outphasing PA (HDOPA), shown in Figure 14, were extracted. Details about the design of the HDOPA were reported in [28]. However, for completeness, it is briefly described in this section. It consists of two power amplifiers, designed to operate at 2.1 GHz center frequency, the main PA working as class-F and the peaking PA as class-C, and an output power combiner. GaN-HEMT transistors model CGH40010, 10 W from Cree Inc. were used as active devices. The passive elements of the PA were designed and fabricated on Rogers substrate (RT Duroid 5880).
The Look-Up Table, shown in Table 2, was extracted using the measurement system reported in Figure 10. It provides the highest drain efficiency, which is applied to the original discrete time modulated signal x n = x n exp j ϕ x n . The input signals to the main and peaking PA are x 1 n and x 2 n , which are given by:
x 1 n = P i n , 1 P o u t n   e j ϕ x n + 1 2 θ P o u t n
x 2 n = P i n , 2 P o u t n   e j ϕ x n 1 2 θ P o u t n
where ϕ x = ϕ x φ O P A P o u t n and P i n , i P o u t n , θ P o u t n are the incident power at the DUT input ports and the outphasing angle between its inputs, respectively. Figure 15 shows the HD-OPA drain efficiencies versus its power obtained from simulation (black line) and measurement testbed (blue line), and the measured optimal drain efficiency is depicted with the red line. For the DUT simulation, the Harmonic Balance of the Advanced Design System (ADS) was used. The measurements were taken by sweeping the outphasing angles between 20 to 130 degrees for several incident power levels at the DUT inputs.
Table 2 shows the LUT extracted to obtain the optimal drain efficiency; it contains the input power levels P i n , 1 / 2 and the outphasing angle θ needed to drive the PA for each output power P o u t .
The static and the dynamic behavior of the DUT are shown in Figure 16, represented with red and blue dots, respectively. From Figure 16 it is observed that the dynamic curves agree fairly well with the static ones, despite the existence of memory effects.
The HD-OPA dynamic gain behavior, reported in Figure 16, is obtained by applying the LUT from Table II to an original 10 MHz OFDM signal. The slight upper shift of the dynamic gain is believed to be due to the dynamic bias effect [29].
The VSNA proposed in [27] allows obtaining the static and dynamic curves of multiple-input PAs using the same test setup, without modifying the system for different measurement needs.

4. Low-Frequency Active Harmonic Load-Pull: An Experimental Study of the RF Capabilities of Power Transistors

Efficiency, power, and linearity are the key metrics used to evaluate the performance of RF power amplifiers. These metrics are typically determined using high-frequency harmonic load-pull systems, either passive or active. However, regarding the assessment of the maximal performance related to transistor technology used in power amplifier applications under different amplifier operation modes, it is required to identify the voltage and current waveforms occurring at the transistor’s current generator plane.
Two methods are used to identify these waveforms: from simulations using CAD tools [30] or from large-signal measurements [31,32,33,34]. In order to study the behavior of a power amplifier using CAD tools, nonlinear transistor models are required; thus, the capabilities of the transistor model to represent the device large-signal behavior heavily impacts the accuracy of this method [35]. By comparison, waveform measurements using commercial high-frequency (HF) load-pull systems [31,32] require the use of state-of-the-art measurement equipment, such as a nonlinear vector network analyzer or a large-signal network analyzer [36]. Further, both linear and nonlinear de-embedding procedures are needed since the effect of the device parasitic and the nonlinear capacitive networks have to be removed in order to access the intrinsic current generator. At low frequencies, the effect of reactive parasitic elements can be neglected, and the transistor’s equivalent circuit may be simplified to that shown in Figure 17.

4.1. Time-Domain Low-Frequency Load-Pull System

In Figure 17, the schematic diagram and picture of the time-domain LF active harmonic load-pull system developed in [35] is shown. In [35] the system operation was extensively described; in this paper a brief review of the measurement system functionality is provided. The main elements of the system are a four-channel oscilloscope and directional couplers, which allow sensing of the waves at the input and output ports of the DUT, along with other elements that allow the system to properly operate, such as bias networks, DC power supplies, and driver power amplifiers to boost the signals at the DUT ports. The DUT voltage ( v g s t , v d s t ) and current ( i d s t ) waveforms may be expressed as a function of the incident ( v g s i t , v d s i t ) and reflected ( v g s r t , v d s r t ) sensed voltage waveforms as:
v g s t = v g s i t + v g s r t ,
v d s t = v d s i t + v d s r t ,
i d s t = v d s r t v d s i t / Z 0 ,
where Z0 is the measurement system impedance (typically 50 Ω).

4.2. Experimental Results

In this paper, three examples of the usefulness of the time-domain low-frequency active harmonic load-pull system, shown in Figure 18, are presented. In the first, the load lines of a GaN-HEMT transistor are measured at four different drain-source voltages, and for two different gate-source voltages, as reported in Figure 19. It can be noted that, as the quiescent bias point shifts to large drain-source voltages, the knee voltage increases, and the drain-source current decreases. These phenomena, reported by P. Tasker [31] and A. Raffo [33,34], are produced by traps and are known as knee walkout and current collapse, respectively. Figure 20 shows the load-lines measured for a SiC MESFET biased as Class-B mode at four different drain-source voltages. Notice that SiC MESFET is free of knee walkout and current collapse, but the knee voltage in SiC is larger than that observed in GaN. A large knee voltage typically limits the maximum achievable power and efficiency of a transistor in a power amplifier.
.
As a second example of the usefulness of the measurement system shown in Figure 18, the behavior of transistors under harmonically-tuned operation modes was studied. An example, Class-F, is presented. According to theory, the Class-F [37] operation mode should be biased as Class-AB mode but loading the intrinsic current generator with a short circuit and an open circuit at the second and third harmonic, respectively [38]. In Figure 21, the impedance design space of the Class-F modes for the fundamental, second, and third harmonics for a GaN-HEMT and SiC-MESFET are presented.
Figure 22 shows the load lines along with the current and voltage waveforms of a GaN-HEMT measured when loading the transistor with the impedances for maximum output power. Note that the half-rectified shape of the drain-source current waveform and the rectangular shape of the drain-source voltage waveform are consistent with the theory [38]. Figure 23 shows the same set of load lines along with the current and voltage waveforms of a SiC-MESFET. These intrinsic load-lines cannot be determined directly from time-domain HF load-pull measurements.
A recently published work [36] reports an experimental study about Class-J power amplifiers using the test bench shown in Figure 17. The Class-J amplifier [32,39] is generated by appropriately manipulating the fundamental (reactive load instead of resistive load) and second harmonic (reactance instead of short circuit) impedances. The reactive load at the fundamental is the main difference between Class-J and Class-F amplifiers. The third example is the Class-J amplifier. Figure 24 shows the impedance design space (fundamental and second harmonics), synthesized using the load-pull system shown in Figure 24, for the Class-J amplifier.
Figure 25 shows measurements of the current and voltage waveforms at the current generator plane, when the fundamental load is the optimal load and when the transistor is biased as Class-B, and varying the amplitude and phase of the second harmonic impedance.
Figure 26 shows the output power and drain efficiency of a Class-J amplifier. According to the theory of resistive reactive Class-J [32], the output power and drain efficiency must be independent of the α parameter but dependent on the β parameter. However, small variations in Pout versus α are observed. According to experimental results published in [36], when α is close to ±1, the current reduction seems to be the origin of the small variations in the output power with α.

5. Conclusions

The design of high-performance power amplifiers is greatly benefited by the use of advanced measurement systems, which need to be calibrated correctly to operate efficiently. In this paper, the complete process to achieve large-signal characterization of power transistors and amplifiers is presented, from calibration of the VRLP system to its use in the implementation of systems allowing the characterization of devices under modulated signal excitation, and systems allowing the implementation of time-domain active load-pull systems. The goal of CICESE’s RF and Microwave Research Group is to generate knowledge in the field of large-signal RF metrology with the aims of characterizing power transistors and designing power amplifiers. Current activities are focused on the study of continuous mode Class-J and Class-F power amplifiers, using the measurement systems described in previous sections. In the following order, our current goals are: (1) to develop high- and low-frequency active harmonic load-pull systems, (2) to design continuous mode Class-J and Class-F amplifiers using non-linear de-embedding, and (3) to study and characterize dual-input power amplifiers, stimulated with modulated signals using a vector signal analyzer and vector network analyzer.

Author Contributions

The research reviewed in this paper is based on the dissertation of three CICESE (M.A.P.-G., T.N.-A. and M.M.-C.) conducted under the direction of J.A.R.-H. The PhD work on the VSNA (T.N.-A.) was pursued in collaboration with the Ohio State University. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported in part by the Center for Scientific Research and Higher Education at Ensenada (CICESE), Baja California, and in part by the National Council for Science and Technology (CONACYT-Mexico) under Project INFR-2014-01 (228580), Project CB 2013-222949-Y, Project INFR-201601 (269927), and Project INFR-2019-01 (301058), and in part by the US National Science Foundation under grant 1740119.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

This appendix summarizes procedures for the TRL and TRM calibration techniques that are developed using the ABCD-parameters matrix formalism [22,24]. The TRL and TRM techniques use as calibration elements a through connection of the test ports (thru) and a pair of symmetrical reflecting loads (reflect), along with either a uniform transmission line (TRL) [20,21,22] or a pair of broadband loads (TRM) [23,24], as shown in Figure A1.
Figure A1. Calibration structures: (a) thru (TRL/TRM), (b) symmetrical reflecting loads (TRL/TRM), (c) uniform transmission line (TRL), (d) symmetrical/nonsymmetrical broadband loads (TRM).
Figure A1. Calibration structures: (a) thru (TRL/TRM), (b) symmetrical reflecting loads (TRL/TRM), (c) uniform transmission line (TRL), (d) symmetrical/nonsymmetrical broadband loads (TRM).
Electronics 11 01114 g0a1
Using the eight-term error model, the measurement of the thru structure shown in Figure A1a, may be expressed as:
M T = T A I T B
where TA and TB are ABCD-parameters matrices representing the errors in ports 1 and 2 of the VNA, and I is the identity matrix representing a zero-length thru. As demonstrated in [24], the measurement of the thru structure can also be expressed as:
M T = T A T Z T Z 1 T B = T X T Y ,
where:
T X = T A T Z = A X B X C X D X = D X A X ¯ B X ¯ C X ¯ 1 ,
T Y = T Z 1 T B = A Y B Y C Y D Y = D Y A Y ¯ B Y ¯ C Y ¯ 1 .
In Equations (A3) and (A4), the matrix TZ, denoted as:
T Z = Z B Z A 1 1 ,
contains information about one of the calibration structures used in a particular calibration technique. In the TRL technique [22], ZA = ZB = ZL, with ZL defined as the characteristic impedance of the line used as the calibration structure (Figure A1c). In the TRM technique [24], the value of ZA equals the impedance of the load used as the match standard at port one (ZA = ZM1), and ZB equals the impedance of the load used as the match standard at port two (ZB = ZM2) of the VNA (Figure A1d). This allows the loads used as the match standard to be either symmetrical or nonsymmetrical.
The problem of VNA calibration relies on calculating seven of the eight elements ( A X ¯ , B X ¯ , C X ¯ , A Y ¯ , B Y ¯ , C Y ¯ , D X D Y ) of the matrices TX and TY. Since detailed procedures to determine them have been previously published [22,24], here a concise summary is provided. Table A1 summarizes the measurements that are used to determine the seven error terms. The first step is to calculate the term B X ¯ and the ratio A X ¯ / C X ¯ . In the TRL technique, these terms are determined by using measurements of the thru-line combination [22]. On the other hand, in the TRM technique they are calculated from measurements of the thru and the broadband loads used as the match standard [24].
Table A1. Calculation of error terms in TRL and TRM calibration techniques. The sequence is intended to be organized in the table from left to right.
Table A1. Calculation of error terms in TRL and TRM calibration techniques. The sequence is intended to be organized in the table from left to right.
Calibration
Terms
B X ¯ A X ¯ / C X ¯ C X ¯ A Y ¯ , B Y ¯ , C Y ¯ , D X D Y
TRL [22]
(ZA = ZB = ZL)
Thru-LineThru-LineThru-Reflecting
loads
Thru
TRM [24]
(ZA = ZM1; ZB = ZM2)
Match P1Thru-Match P2Thru-Reflecting
loads
Thru
The second step is to determine the term C X ¯ , so that the variables forming the ratio A X ¯ / C X ¯ can be calculated separately; this term is determined using a pair of symmetrical reflecting loads. Finally, the remaining four terms, A Y ¯ , B Y ¯ , C Y ¯ , D X D Y , are calculated using the measurement of the thru along with the knowledge of A X ¯ , B X ¯ and C X ¯ [22,24].
The ABCD-parameters of a DUT (TD) may be calculated as a function of its measured ABCD-parameters (MD) as:
T D = T Z T X 1   M D   T Y 1 T Z 1 .
As denoted in Equation (A6), the elements of the matrix TZ must be known in order to perform the error correction process. Therefore, in order to accurately refer the TRL and TRM calibrations to the measurement system impedance, the impedance of either a transmission line (TRL) or a pair of broadband loads (TRM) must be known.

References

  1. Chireix, H. High power outphasing modulation. Proc. Inst. Radio Eng. 1935, 23, 1370–1392. [Google Scholar] [CrossRef]
  2. Doherty, W.H. A new high efficiency power amplifier for modulated waves. Proc. Inst. Radio Eng. 1936, 24, 1163–1182. [Google Scholar]
  3. Shakib, S.; Dunworth, J.; Aparin, V.; Entesari, K. mmWave CMOS Power Amplifiers for 5G Cellular Communication. IEEE Commun. Mag. 2019, 57, 98–105. [Google Scholar] [CrossRef]
  4. Aen, P.; Wood, J.; Pla, J. RF and Microwave Power Transistors. In Modeling and Characterization of RF and Microwave Power FETs; Cambridge University Press: Cambridge, MA, USA, 2007; pp. 1–50. [Google Scholar]
  5. De Groote, F.; Teyssier, J.; Gasseling, T.; Jardel, O.; Verspecht, J. Introduction to measurements for power transistor characterization. IEEE Microw. Mag. 2008, 9, 70–85. [Google Scholar] [CrossRef]
  6. Roblin, P.; Root, D.E.; Verspecht, J.; Ko, Y.; Teyssier, J.P. New Trends for the Nonlinear Measurement and Modeling of High-Power RF Transistors and Amplifiers with Memory Effects. IEEE Trans. Microw. Theory Tech. 2012, 60, 1964–1978. [Google Scholar] [CrossRef]
  7. Root, D.; Verspecht, J.; Horn, J.; Marcu, M. S-Parameters—A concise review. In X-Parameters: Characterization, Modeling, and Design of Nonlinear RF and Microwave Components; Cambridge University Press: Cambridge, UK, 2013; pp. 1–19. [Google Scholar]
  8. Ghannouchi, F.M.; Hashmi, M.S. Passive load-pull systems. In Load-Pull Techniques with Applications to Power Amplifier Design; Springer: Dordrecht, The Netherlands, 2012; pp. 55–84. [Google Scholar]
  9. Ferrero, A.; Pirola, M. Harmonic load-pull techniques: An overview of modern systems. IEEE Microw. Mag. 2008, 14, 116–123. [Google Scholar] [CrossRef]
  10. Hashmi, M.S.; Ghannouchi, F.M. Introduction to load-pull systems and their applications. IEEE Instrum. Meas. Mag. 2013, 16, 30–36. [Google Scholar] [CrossRef]
  11. Cusack, J.; Perlow, S.; Perlman, B. Automatic load contour mapping for microwave power transistors. In Proceedings of the IEEE Microwave Theory and Techniques International Symposium, Atlanta, GA, USA, 12–14 June 1974; pp. 269–327. [Google Scholar]
  12. Secci, F.; Perlow, M.; Perlman, B.S. A computer-controlled microwave tuner for automated load pull. RCA Rev. 1983, 44, 566–583. [Google Scholar]
  13. Tucker, R.S.; Bradley, P.D. Computer-Aided Error Correction of Large-Signal Load-Pull Measurements. IEEE Trans. Microw. Theory Tech. 1984, 32, 296–301. [Google Scholar] [CrossRef]
  14. Camarchia, V.; Teppati, V.; Corbellini, S.; Pirola, M. Microwave Measurements Part II: Non-linear Measurements. IEEE Instrum. Meas. Mag. 2007, 10, 34–39. [Google Scholar] [CrossRef] [Green Version]
  15. Teppati, V.; Bolognesi, C.R. Evaluation and reduction of calibration residual uncertainty in load-pull measurements at millimeter-wave frequencies. IEEE Trans. Instrum. Meas. 2012, 61, 817–822. [Google Scholar] [CrossRef]
  16. Ferrero, A.; Pisani, U. An improved calibration technique for on-wafer large-signal transistor characterization. IEEE Trans. Instrum. Meas. 1993, 42, 360–364. [Google Scholar] [CrossRef]
  17. Roblin, P. Large-signal vector measurement techniques with NVNAs. In Nonlinear RF Circuits and Nonlinear Vector Network Analyzers: Interactive Measurement and Design Techniques; The Cambridge RF and Microwave Engineering Series; Cambridge University Press: Cambridge, MA, USA, 2007; pp. 17–39. [Google Scholar]
  18. Pulido-Gaytán, M.A.; Reynoso-Hernández, J.A.; Maya-Sánchez, M.C.; Loo-Yau, J.R. The impact of knowing the impedance of the lines used in the TRL calibration on the load-pull characterization of power transistors. In Proceedings of the 87th ARFTG Microwave Measurement Conference, Atlanta, GA, USA, 3–4 December 2015; pp. 1–5. [Google Scholar]
  19. Pulido-Gaytán, M.A.; Reynoso-Hernández, J.A.; Maya-Sánchez, M.C.; Loo-Yau, J.R. Calibration of a real-time load-pull system using the generalized theory of the TRM technique. In Proceedings of the 87th ARFTG Microwave Measurement Conference, San Francisco, CA, USA, 27 May 2016; pp. 1–5. [Google Scholar]
  20. Engen, G.F.; Hoer, C.A. Thru-Reflect-Line: An Improved Technique for Calibrating the Dual Six-Port Automatic Network Analyzer. IEEE Trans. Microw. Theory Tech. 1979, 27, 987–993. [Google Scholar] [CrossRef] [Green Version]
  21. Reynoso-Hernandez, J.A.; Pulido-Gaytan, M.A.; Zuniga-Juarez, J.E.; Monjardin-Lopez, J.R.; Garcia-Osorio, A.; Orozco-Navarro, D.; Loo-Yau, J.R.; Maya-Sanchez, M.C. Using lines of arbitrary impedance as standards on the TRL calibration technique. In Proceedings of the 81st ARFTG Microwave Measurement Conference, Seattle, WA, USA, 1–4 June 2013; pp. 1–4. [Google Scholar]
  22. Reynoso-Hernandez, J.A.; Pulido-Gaytan, M.A.; Maya-Sanchez, M.C.; Loo-Yau, J.R. What can the ABCD parameters tell us about the TRL? In Proceedings of the 79th ARFTG Microwave Measurement Conference, Montreal, QC, Canada, 3–4 June 2012; pp. 1–4. [Google Scholar]
  23. Eul, H.J.; Schiek, B. A generalized theory and new calibration procedures for network analyzer self-calibration. IEEE Trans. Microw. Theory Tech. 1991, 39, 724–731. [Google Scholar] [CrossRef]
  24. Pulido-Gaytan, M.A.; Reynoso-Hernandez, J.A.; Loo-Yau, J.R.; Zarate-de Landa, A.; Maya-Sanchez, M.C. Generalized Theory of the Thru-Reflect-Match Calibration Technique. IEEE Trans. Microw. Theory Tech. 2015, 63, 1693–1699. [Google Scholar] [CrossRef]
  25. Ferrero, A.; Pisani, U. Two-Port Network Analyzer Calibration Using an Unknown ‘Thru’. IEEE Microw. Guid. Wave Lett. 1992, 2, 505–507. [Google Scholar] [CrossRef] [Green Version]
  26. Niubó-Alemán, T.; Hahn, Y.; Roblin, P.; Teyssier, J.P.; Reynoso-Hernández, J.A.; Chen, V.; Rajan, S. Calibrated Digital Predistortion Using a Vector Network Analyzer as the Receiver. In Proceedings of the 2019 93rd ARFTG Microwave Measurement Conference (ARFTG), Boston, MA, USA, 7 June 2019; pp. 1–4. [Google Scholar]
  27. Niubo-Aleman, T.; Liang, C.; Hahn, Y.; Reynoso-Hernandez, J.A.; Teyssier, J.P.; Roblin, P. Time-Domain Characterization and Linearization of a Dual-Input Power Amplifier Using a Vector Network Analyzer as the Receiver. IEEE Trans. Microw. Theory Tech. 2021, 69, 2386–2398. [Google Scholar] [CrossRef]
  28. Liang, C.; Niubo-Aleman, T.; Hahn, Y.; Roblin, P.; Reynoso-Hernandez, J.A. Optimal Two-Way Hybrid Doherty-Outphasing Power Amplifier. In Proceedings of the 2020 IEEE Topical Conference on RF/Microwave Power Amplifiers for Radio and Wireless Applications (PAWR), San Antonio, TX, USA, 26–29 January 2020; pp. 26–29. [Google Scholar] [CrossRef]
  29. Cabral, P.M.; Pedro, J.C.; Carvalho, N.B. Bias networks impact on the dynamic AM/AM contours in microwave power amplifiers. In Proceedings of the 2006 International Workshop on Integrated Nonlinear Microwave and Millimeter-Wave Circuits, Aveiro, Portugal, 30–31 January 2006; pp. 38–41. [Google Scholar] [CrossRef]
  30. Jang, H.; Roblin, P.; Xie, Z. Model-based nonlinear embedding for power-amplifier design. IEEE Trans. Microw. Theory Tech. 2014, 62, 1986–2002. [Google Scholar] [CrossRef]
  31. Tasker, P.J. Practical Waveform Engineering. IEEE Microw. Mag. 2009, 10, 65–76. [Google Scholar] [CrossRef]
  32. Friesicke, C.; Quay, R.; Jacob, A.F. The resistive-reactive class-J power amplifier mode. IEEE Microw. Wirel. Compon. Lett. 2015, 25, 666–668. [Google Scholar] [CrossRef]
  33. Raffo, A.; Scappaviva, F.; Vannini, G. A new approach to microwave power amplifier design based on the experimental characterization of the intrinsic electron-device load line. IEEE Trans. Microw. Theory Tech. 2009, 57, 1743–1752. [Google Scholar] [CrossRef]
  34. Raffo, A.; Vadalà, V.; Vannini, G. Nonlinear Embedding and De-Embedding: Theory and Applications; Crupi, G., Schreurs, M.M.-P.D., De-embedding, M., Eds.; Academic Press: Cambridge, MA, USA, 2014; Chapter 9. [Google Scholar]
  35. Molina-Ceseña, M.; Reynoso-Hernández, J.A.; Pulido-Gaytán, M.A.; Loo-Yau, J.R.; Maya-Sánchez, M.C. Experimental Investigation of Resistive-Reactive Class-J Mode Using Time-Domain Low-Frequency Active Harmonic Load-Pull Measurements. IEEE Microw. Wirel. Compon. Lett. 2022, 32, 96–99. [Google Scholar] [CrossRef]
  36. Van Moer, W.; Gomme, L. NVNA versus LSNA: Enemies or friends? IEEE Microw. Mag. 2010, 11, 97–103. [Google Scholar] [CrossRef]
  37. Carrubba, V.; Clarke, A.L.; Akmal, M.; Lees, J.; Benedikt, J.; Tasker, P.; Cripps, S.C. On the Extension of the Continuous Class-F Mode Power Amplifier. IEEE Trans. Microw. Theory Tech. 2011, 59, 1294–1303. [Google Scholar] [CrossRef]
  38. Colantonio, P.; Giannini, F.; Leuzzi, G.; Limiti, E. On the class-F power amplifier design. Int. J. RF Microw. Comput.-Aided Eng. 1999, 9, 129–149. [Google Scholar] [CrossRef]
  39. Cripps, S.C.; Tasker, P.; Clarke, A.L.; Lees, J.; Benedikt, J. On the continuity of high efficiency modes in linear RF power amplifiers. IEEE Microw. Wirel. Compon. Lett. 2009, 19, 665–667. [Google Scholar] [CrossRef]
Figure 1. Typical vector-receiver load-pull system. Red dotted lines denote the calibration plane.
Figure 1. Typical vector-receiver load-pull system. Red dotted lines denote the calibration plane.
Electronics 11 01114 g001
Figure 2. (a) System configuration for relative calibration; (b) measurement of calibration structures. Red dotted lines denote the calibration plane; blue dotted lines denote the vector receiver plane.
Figure 2. (a) System configuration for relative calibration; (b) measurement of calibration structures. Red dotted lines denote the calibration plane; blue dotted lines denote the vector receiver plane.
Electronics 11 01114 g002
Figure 3. (a) Measurement system for power calibration; (b) measurement of coaxial short-open-load (SOL) structures and power reference. Red dotted lines denote the calibration plane; blue dotted lines denote a coaxial plane where the power reference and SOL structures are connected.
Figure 3. (a) Measurement system for power calibration; (b) measurement of coaxial short-open-load (SOL) structures and power reference. Red dotted lines denote the calibration plane; blue dotted lines denote a coaxial plane where the power reference and SOL structures are connected.
Electronics 11 01114 g003
Figure 4. Structures used to implement the TRL calibration: (a) DUT in test fixture, (b) thru, (c) line and (d) reflecting load (open circuit).
Figure 4. Structures used to implement the TRL calibration: (a) DUT in test fixture, (b) thru, (c) line and (d) reflecting load (open circuit).
Electronics 11 01114 g004
Figure 5. Contours of constant output power of a 45 W transistor (VDS = 20 V, VGS = −2.5 V) using: (a) ZL = 5 Ω, (b) ZL = 10 (actual value) Ω, (c) ZL = 20 Ω, (d) ZL = 50 Ω in the VRLP calibration. (From [18] with permission, © 2022 IEEE).
Figure 5. Contours of constant output power of a 45 W transistor (VDS = 20 V, VGS = −2.5 V) using: (a) ZL = 5 Ω, (b) ZL = 10 (actual value) Ω, (c) ZL = 20 Ω, (d) ZL = 50 Ω in the VRLP calibration. (From [18] with permission, © 2022 IEEE).
Electronics 11 01114 g005
Figure 6. (a) Magnitude and (b) angle of the calculated DUT’s large-signal gain using different ZL values in the calibration. (From [18] with permission, © 2022 IEEE).
Figure 6. (a) Magnitude and (b) angle of the calculated DUT’s large-signal gain using different ZL values in the calibration. (From [18] with permission, © 2022 IEEE).
Electronics 11 01114 g006
Figure 7. (a) DUT in the test fixture and TRM calibration structures: (b) thru, (c) symmetrical reflecting load (short circuit) and, (d) nonsymmetrical loads (match). (From [19] with permission, © 2022 IEEE).
Figure 7. (a) DUT in the test fixture and TRM calibration structures: (b) thru, (c) symmetrical reflecting load (short circuit) and, (d) nonsymmetrical loads (match). (From [19] with permission, © 2022 IEEE).
Electronics 11 01114 g007
Figure 8. Contours of constant output power calculated using in the VRLP calibration: (a) TRL and (b) TRM using nonsymmetrical loads as the match standard. (From [19] with permission, © 2022 IEEE).
Figure 8. Contours of constant output power calculated using in the VRLP calibration: (a) TRL and (b) TRM using nonsymmetrical loads as the match standard. (From [19] with permission, © 2022 IEEE).
Electronics 11 01114 g008
Figure 9. Output power load-pull contours calculated using (a) ZM1 = 50 Ω and ZM2 = 25 Ω, (b) ZM1 = ZM2 = 53.0 + j13.0 Ω, and (c) ZM1 = ZM2 = 26.0 + j10 Ω in the calibration. (From [19] with permission, © 2022 IEEE).
Figure 9. Output power load-pull contours calculated using (a) ZM1 = 50 Ω and ZM2 = 25 Ω, (b) ZM1 = ZM2 = 53.0 + j13.0 Ω, and (c) ZM1 = ZM2 = 26.0 + j10 Ω in the calibration. (From [19] with permission, © 2022 IEEE).
Electronics 11 01114 g009
Figure 10. Testbed for the characterization of multiple-input devices. (From [27] with permission, © 2022 IEEE).
Figure 10. Testbed for the characterization of multiple-input devices. (From [27] with permission, © 2022 IEEE).
Electronics 11 01114 g010
Figure 11. Amplitude (a) and phase corrections (b) for R1 (a1), R2 (a2), and C (b3) (VNA receivers). (From [27] with permission, © 2022 IEEE).
Figure 11. Amplitude (a) and phase corrections (b) for R1 (a1), R2 (a2), and C (b3) (VNA receivers). (From [27] with permission, © 2022 IEEE).
Electronics 11 01114 g011
Figure 12. Subcarrier phases vs. the actual ones in the frequency domain. (From [27] with permission, © 2022 IEEE).
Figure 12. Subcarrier phases vs. the actual ones in the frequency domain. (From [27] with permission, © 2022 IEEE).
Electronics 11 01114 g012
Figure 13. Flowchart of the post-processing applied to the raw data of each channel by the VSNA to correct the multi-channel data. (From [27] with permission, © 2022 IEEE).
Figure 13. Flowchart of the post-processing applied to the raw data of each channel by the VSNA to correct the multi-channel data. (From [27] with permission, © 2022 IEEE).
Electronics 11 01114 g013
Figure 14. Device under test: hybrid Doherty-Outphasing PA. (From [27] with permission, © 2022 IEEE).
Figure 14. Device under test: hybrid Doherty-Outphasing PA. (From [27] with permission, © 2022 IEEE).
Electronics 11 01114 g014
Figure 15. Drain efficiency (%) versus Pout (dBm) and outphasing angle (°). (From [27] with permission, © 2022 IEEE).
Figure 15. Drain efficiency (%) versus Pout (dBm) and outphasing angle (°). (From [27] with permission, © 2022 IEEE).
Electronics 11 01114 g015
Figure 16. (a) HD-OPA characteristics: AM-AM and (b) AM-PM. (From [27] with permission, © 2022 IEEE).
Figure 16. (a) HD-OPA characteristics: AM-AM and (b) AM-PM. (From [27] with permission, © 2022 IEEE).
Electronics 11 01114 g016
Figure 17. Equivalent electrical circuit of a field-effect transistor at low frequency.
Figure 17. Equivalent electrical circuit of a field-effect transistor at low frequency.
Electronics 11 01114 g017
Figure 18. Time-domain LF active harmonic load-pull system: (a) measurement bench and (b) schematic diagram.
Figure 18. Time-domain LF active harmonic load-pull system: (a) measurement bench and (b) schematic diagram.
Electronics 11 01114 g018
Figure 19. Load-lines measured in a GaN-HEMT transistor at four different drain-source voltages (VDS = 10, 15, 20, 25, 30 V; VGS = −3 V) and two different power levels: (a) allowing maximum v g s t = 0.4   V and (b) allowing maximum v g s t = 0   V .
Figure 19. Load-lines measured in a GaN-HEMT transistor at four different drain-source voltages (VDS = 10, 15, 20, 25, 30 V; VGS = −3 V) and two different power levels: (a) allowing maximum v g s t = 0.4   V and (b) allowing maximum v g s t = 0   V .
Electronics 11 01114 g019
Figure 20. Load-lines measured in a SiC-MESFET at four different drain-source voltages (VDS = 10, 15, 20, 25, 30 V; VGS = −9 V) and two different power levels: (a) allowing maximum v g s t = 1.5   V and (b) allowing maximum v g s t = 0   V .
Figure 20. Load-lines measured in a SiC-MESFET at four different drain-source voltages (VDS = 10, 15, 20, 25, 30 V; VGS = −9 V) and two different power levels: (a) allowing maximum v g s t = 1.5   V and (b) allowing maximum v g s t = 0   V .
Electronics 11 01114 g020
Figure 21. Impedance design space of the Class-F modes for the fundamental, second, and third harmonics for (a) GaN HEMT and (b) SiC MESFET.
Figure 21. Impedance design space of the Class-F modes for the fundamental, second, and third harmonics for (a) GaN HEMT and (b) SiC MESFET.
Electronics 11 01114 g021
Figure 22. (a) Load lines and (b) current and voltage waveforms of a GaN-HEMT measured when loading the transistor with the impedances for maximum output power.
Figure 22. (a) Load lines and (b) current and voltage waveforms of a GaN-HEMT measured when loading the transistor with the impedances for maximum output power.
Electronics 11 01114 g022
Figure 23. (a) Load lines and (b) current and voltage waveforms of a SiC-MESFET measured when loading the transistor with the impedances for maximum output power.
Figure 23. (a) Load lines and (b) current and voltage waveforms of a SiC-MESFET measured when loading the transistor with the impedances for maximum output power.
Electronics 11 01114 g023
Figure 24. The measured impedance at the fundamental frequency Zfo (-o-) and the impedance at second harmonic frequency Z2fo (-*-) depending α y β for optimal resistance RL1 = 37 Ω.
Figure 24. The measured impedance at the fundamental frequency Zfo (-o-) and the impedance at second harmonic frequency Z2fo (-*-) depending α y β for optimal resistance RL1 = 37 Ω.
Electronics 11 01114 g024
Figure 25. Measured voltage and current waveforms of the Class-J power amplifier at the current generator plane for the on-wafer GaN-HEMT (Z = 2 mm, Lg = 0.25 μm) transistor biased at VDS = 20 V, VGS = −2.3 V with v g s t maximum value equal to −0.2 V: (a) β = 0 and α = [−1, 1], (b) β = [0, 0.4] and α = 0, (c) β = 0.3 and α = [−1, 1], (d) β = 0.4 and α = [−1, 1].
Figure 25. Measured voltage and current waveforms of the Class-J power amplifier at the current generator plane for the on-wafer GaN-HEMT (Z = 2 mm, Lg = 0.25 μm) transistor biased at VDS = 20 V, VGS = −2.3 V with v g s t maximum value equal to −0.2 V: (a) β = 0 and α = [−1, 1], (b) β = [0, 0.4] and α = 0, (c) β = 0.3 and α = [−1, 1], (d) β = 0.4 and α = [−1, 1].
Electronics 11 01114 g025
Figure 26. Measured drain efficiency and output power versus α and β for the on-wafer GaN-HEMT (Z = 2 mm, Lg = 0.25 μm) transistor biased at VDS = 20 V, VGS = −2.3 V with v g s t maximum value equal to −0.2 V.
Figure 26. Measured drain efficiency and output power versus α and β for the on-wafer GaN-HEMT (Z = 2 mm, Lg = 0.25 μm) transistor biased at VDS = 20 V, VGS = −2.3 V with v g s t maximum value equal to −0.2 V.
Electronics 11 01114 g026
Table 1. Average SNR and EVM captured using the VSNA and VSA. (From [27] with permission, © 2022 IEEE).
Table 1. Average SNR and EVM captured using the VSNA and VSA. (From [27] with permission, © 2022 IEEE).
VSNA/VSA
BW10 MHz20 MHz100 MHz120 MHz
avg SNR (dB)43.2929.7643.0229.4040.9428.5241.0827.50
avg EVM (dB)−45.15−35.91−44.68−34.38−40.81−29.12−40.61−27.37
avg EVM (%)0.551.600.581.900.913.490.934.28
Table 2. LUT of the hybrid Doherty-Outphasing PA for maximum efficiency. (From [27] with permission, © 2022 IEEE).
Table 2. LUT of the hybrid Doherty-Outphasing PA for maximum efficiency. (From [27] with permission, © 2022 IEEE).
P o u t ( dBm ) P i n , 1 ( dBm ) P i n , 2 ( dBm ) θ
(degree)
ϕ O P A ( degree )
27.3813.7111.3621.04−51.83
29.5116.2113.8722.50−51.22
31.0317.9515.6919.15−52.68
32.2119.4117.0322.33−50.81
34.2921.9119.8332.82−43.14
34.8322.5620.7332.91−41.59
36.3623.8322.2842.88−35.39
38.6825.6024.9352.72−28.54
40.2526.7326.9462.23−24.56
42.7628.4029.82100.69−16.62
43.2628.3929.81110.69−16.34
43.6928.3929.81120.69−16.10
43.9928.3829.80130.69−16.10
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Reynoso-Hernández, J.A.; Pulido-Gaytan, M.A.; Niubó-Alemán, T.; Molina-Ceseña, M. Advances in Microwave Large-Signal Metrology: From Vector-Receiver Load-Pull to Vector Signal Network Analyzer and Time-Domain Load-Pull Implementations (Invited Paper). Electronics 2022, 11, 1114. https://doi.org/10.3390/electronics11071114

AMA Style

Reynoso-Hernández JA, Pulido-Gaytan MA, Niubó-Alemán T, Molina-Ceseña M. Advances in Microwave Large-Signal Metrology: From Vector-Receiver Load-Pull to Vector Signal Network Analyzer and Time-Domain Load-Pull Implementations (Invited Paper). Electronics. 2022; 11(7):1114. https://doi.org/10.3390/electronics11071114

Chicago/Turabian Style

Reynoso-Hernández, J. Apolinar, Manuel Alejandro Pulido-Gaytan, Thaimí Niubó-Alemán, and Marlon Molina-Ceseña. 2022. "Advances in Microwave Large-Signal Metrology: From Vector-Receiver Load-Pull to Vector Signal Network Analyzer and Time-Domain Load-Pull Implementations (Invited Paper)" Electronics 11, no. 7: 1114. https://doi.org/10.3390/electronics11071114

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