5a 059 PDF
5a 059 PDF
5a 059 PDF
Mauro Marchetti
Mixed-Signal Instrumentation for
Large-Signal Device Characterization and
Modelling
PROEFSCHRIFT
door
Mauro MARCHETTI
Dottore in Ingegneria Elettronica
van Università degli Studi di Napoli “Federico II”, Italië
geboren te Napoli, Italië
Dit proefschrift is goedgekeurd door de promotor:
Prof. dr. J. R. Long
Samenstelling promotiecommissie:
Mauro Marchetti,
Mixed-Signal Instrumentation for Large-Signal Device Characterization and
Modelling,
Ph.D. Thesis Delft University of Technology,
with summary in Dutch.
ISBN: 978-94-6203-471-6
1 Introduction 1
1.1 Trends in wireless communication . . . . . . . . . . . . . . . . . 1
1.2 Requirements on the power amplifier . . . . . . . . . . . . . . . 4
1.3 The need for advanced measurement tools . . . . . . . . . . . . 6
1.4 Thesis objectives . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.5 Thesis outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
i
ii CONTENTS
Bibliography 133
Summary 145
Samenvatting 149
Acknowledgments 157
Introduction
The appearance of modern smart phones and their quick penetration to con-
sumer markets has drastically changed the way we communicate as a society.
The manner in which social networks, for example, are influencing every day
life has been magnified by the capabilities of truly portable, always connected
devices. While just a few years ago the cellular phone was only used for voice
calls and SMS, the newer generation smart phones are reaching more and more
the capabilities of pocket size personal computers which are always connected
to the network. Consequently, applications ranging from internet browsing
to video chatting, from high-quality video streaming to internet calls and so-
cial networks, which were previously tied to personal computers alone, have
become accessible anytime and from almost anywhere to the user.
As a result, it is becoming increasingly more difficult to imagine our life
without a mobile internet connection. This trend is reflected in the 2012 Cisco
World Technology Report [1], which indicates that more than 60 % of “Gen
Y” (18 to 30 year olds) do not even wait to get out of bed in the morning
before checking their smart phones, while more than 40 % of them (including
the author of this thesis) would feel anxious, almost lost, if they couldn’t check
their smart phones constantly.
1
2 Introduction
Figure 1.1: Mobile data traffic forecast vs. time in exabytes per month.
Source Cisco VNI Mobile Forecast [2].
10000
802.11ac
LTE
1000 Advanced
802.11n
Peak Data Rate [Mb/s]
LTE
100 WiMAX
HSPA+
802.11ag
10 802.11b HSPA Mobile
HSDPA WLAN
802.11
1
UMTS R99
EDGE
0.1
GPRS
0.01 GSM
1990 1995 2000 2005 2010 2015
Year
Figure 1.2: Development of data rates over time in wireless networks [3].
• The design and implementation of novel and more energy efficient hard-
ware.
0.08
0.07
0.06
0.05
Density
0.04
0.03
0.02
0.01
0
−30 −25 −20 −15 −10 −5 0
Normalized Signal Power [dB]
−50
Spectral Power Density [dBm/Hz]
−60
−70
−80
−90
−100
−110
2392 2402 2412 2422 2432 2442 2452 2462 2472
Frequency [MHz]
To complicate matters, the demands for high efficiency and high linearity
are normally in conflict with each other. The linearity requirement, in fact,
could very simplistically be met by operating the PA well below its saturation
point, in its linear region of operation. However, since the peak efficiency of a
PA is normally achieved at peak output power, this would drastically reduce
PA efficiency [11]. On the other hand, driving the PA closer to its saturation
point, in its nonlinear region of operation, would certainly increase efficiency,
but would result in nonlinear distortion effects.
For this reason, higher complexity amplifier concepts like Doherty [12],
outphasing [13] and envelope tracking [14], which make use of dynamic varia-
tions of the load and voltage conditions for the output stage transistor(s), are
becoming increasingly more popular, since they can relax the compromise be-
tween efficiency and linearity. It is clear that all these constraints on efficiency
and linearity, in combination with these dynamic changes to the operating
conditions for the active devices, make designing a power amplifier capable of
meeting today’s communication industry standards a challenging task, which
requires accurate, fast and versatile measurement tools.
as output power, efficiency, linearity, and many others, while tuning the various
impedances offered to the active device.
Due to the aforementioned need to operate the transistor close to its non-
linear region, the models used for PA design need to correctly represent the
large-signal and distortion behaviors of the active device. Moreover, to prop-
erly predict the transistor performance when employing wideband complex
modulated signals, effects such as bias or temperature-induced memory [16]
must be included in the model. Finally, the transistor model needs to be
verified.
Therefore, to enable the extraction and the verification of accurate large-
signal transistor models, there is a great need for advanced large-signal RF
and microwave measurement tools. In fact, while linear device characterization
(e.g., small-signal S-parameters) is mainly useful to extract the small-signal
behavior of the transistor, pulsed DC, RF and large-signal measurements are
mandatory to extract the nonlinear behavior of the active device and to verify
the model. As an example, pulsed-DC and S-parameters measurements are re-
quired to characterize the active device under isothermal conditions (maintain-
ing the core of the transistor at a constant temperature), in order to develop
device models which include self-heating or trapping effects [17–20]. Moreover,
load-pull measurements are necessary to obtain realistic large-signal data (e.g.,
gain and phase distortion data), for the extraction of large-signal behavioral
models [21], or for model validation.
Additionally, since the model creation process can be a lengthy and cum-
bersome task, it is in many cases common practice to directly use load-pull
measurement data for all activities related to PA design, from the technology
development to the actual power amplifier design. Therefore it is of fundamen-
tal importance that the measurement data extracted from these characteriza-
tion systems is able to predict with the highest level of accuracy the active
device behavior in the final application environment (i.e., circuit termination
and signal stimulus).
However, in spite of the significant progress made over the last decade in
the field of RF measurements, there is still a great need for more accurate and
advanced isothermal and large-signal characterization systems. In particular,
pulsed-DC and S-parameter measurement systems still suffer from limitations
in the minimum pulse widths that can be used and in measurement accu-
racy [22–26]. These impairments limit the usability of these systems for the
characterization of new device technologies under truly isothermal conditions.
A comparable situation exists in load-pull measurement techniques. Al-
though significant advances were made during past years [27], bringing conven-
tional load-pull setups employing passive mechanical tuners and power meters
8 Introduction
signal excitations.
In Chapter 6, the basic theory behind the measurement of the high fre-
quency time-domain voltage and current waveforms at the device reference
planes is discussed. An extension of the mixed-signal load-pull system de-
scribed in the previous chapters is presented, with particular attention on the
requirements of the calibration device used for the system calibration. Finally,
an approach to time-domain waveform analysis of multi-tone signals which are
closely spaced in frequency is introduced.
In Chapter 7, several examples of applications that highlight the most
unique capabilities of the system described are reported. In particular an out-
of-band linearity optimization of an HBT device, the characterization of a GaN
device for high efficiency PA design, and some high power device measurements
for base-station applications are described.
Finally, in Chapter 8 some conclusions are drawn, with recommendations
for future work.
Chapter 2
Isothermal Measurement
Systems
11
12 Isothermal Measurement Systems
Carrier frequency fC
PRF
1/PW
fC
narrowband detection, the analyzer samples are not synchronized with the
incoming pulses, therefore no trigger pulse is required. The advantage of this
detection mode is that there is no lower pulse-width limit, since no matter
how broad the pulse spectrum is, most of it is filtered away, leaving only
the central spectral component. The disadvantage of narrowband detection is
that measurement dynamic range decreases as the duty cycle increases. This
phenomenon is also known as pulse desensitization and the degradation in
dynamic range can be expressed as 20 · log(DutyCycle).
At the time of commencing this project two different network analyzer
products from Agilent Technologies were available for performing pulsed-RF
measurements: the 8510C (with pulsed option 008) and the PNA. The 8510C
[30] uses the wideband detection technique. After a first superheterodyne
down-conversion to 20 MHz, the pulsed signal is down-converted to baseband
IQ pulses and then digitized (Fig. 2.4a). Each I and Q output has a bandwidth
of 1.5 MHz, for a total bandwidth of 3 MHz, yielding pulses with 300 ns rise and
fall times (tr /tf = 1/BW ). Since the pulse width must be larger than several
tr /ts in order for the detector to properly acuire the pulses, this results in a
minimum specified measurable pulse width of 1 µs [29]. The down-conversion
chain of the PNA network analyzer [29] is shown in Fig. 2.4b. After a first
down-conversion to 8.33 MHz, the pulsed signal is down-converted again to an
IF of 41.7 kHz. An anti-alias filter is placed just in front of the analog-to-digital
converter. The PNA has the possibility to operate both with a wideband and
narrowband detection technique. However, due to the low IF bandwidth of
35 kHz for the PNA and of 250 kHz for the PNA-L version, the minimum
pulse durations that can be measured with the wideband technique are 50 µs
and 10 µs respectively, therefore this mode is not suitable for performing
2.1 Pulsed measurements fundamentals 15
(a)
(b)
Figure 2.4: (a) 8510C down-conversion chain using the wideband detection
technique. (b) PNA down-conversion chain. The PNA has the possibility
to operate both with a wideband and narrowband detection technique.
Source Agilent Technologies [28].
Figure 2.5: Dynamic range vs. duty cycle for different pulse widths for
the Agilent 8510 and the PNA pulsed network analyzers [28].
that the device will remain isothermal. For example, a pulse width of 100 ns
with a duty cycle of 0.1 % can be considered a satisfactory condition for most
devices. However, at these settings the performance of current pulsed network
analyzers is strongly reduced, as the dynamic range becomes lower than 60 dB,
as shown in Fig. 2.5.
It is important to note that further developments to the PNA have im-
proved the dynamic range of the new PNA-X for low duty cycles. In particular,
for this instrument the pulse desensitization problem, although still present, it
has now been reduced to a value of 10 · log(DutyCycle) [31], enabling the use
of this instrument for isothermal measurements. These later developments,
which occurred after this thesis work, are not discussed here.
DC Pulse
Triggers
RF Pulse
Trigger
DUT
Bias Bias
Tee Tee
Agilent 8110A
Pulse Generator V1 V2
I1 I2
b1
b2
Synthesizer CH. 0 CH. 0 CH. 0 CH. 0
HP 85110A
a2 TRIG TRIG TRIG TRIG
Test-Set
PXI Based
Synthesizer
DAQ
RF RF signals IF signals
GPIB Bus
Digital filtering
Figure 2.6: Block diagram of the custom pulsed RF - pulsed I-V setup.
TUDelft Synthetic Network Analyzer
Isothermal Measurement Systems
2.2 System configuration 19
The dynamic range (DR) of an ADC is the ratio of the largest to the smallest
signal that the converter can represent. The largest signal is usually taken to
be a sine wave that covers the full voltage range of the ADC, while the smallest
signal is usually taken to be the total noise level of the ADC. Consequently
the dynamic range expressed in decibels (dB) is
S
DR = 20 · log , (2.6)
R
where S is the rms amplitude of the largest signal, and R is the rms ampli-
tude of the smallest signal. In an ideal ADC, the noise is a direct result of
the quantization noise of the ADC [34], therefore the dynamic range can be
calculated directly from the number of bits N as
In a real ADC, though, the presence of noise other than from quantization
is responsible for a decrease in dynamic range. In our particular situation,
the A/D converters employed are 60 MS/s NI PXI-5105 12-bit digitizers, and
their theoretical dynamic range can be calculated from the number of bits and
the input range. For an input range of 50 mV peak-to-peak, the maximum
20 Isothermal Measurement Systems
(V maxRM S )2
Pmax (dBm) = 10 · log 1000 · =
50 Ω
" √ #
(V maxP eak / 2)2
= 10 · log 1000 · =
50 Ω
" √ #
(25 mV / 2)2
= 10 · log 1000 · ≈ −22 dBm .
50 Ω
(2.8)
The lowest power that can be measured due to the total noise of the digitizer
as specified in the data sheet is
(V noiseRM S )2
Pmin (dBm) = 10 · log 1000 · =
50 Ω
(19 µV )2
= 10 · log 1000 · ≈ −81.4 dBm ,
50 Ω
(2.9)
If we assume that the source of noise in our ADC appears as white noise, we
can use digital filtering to filter out noise components outside the bandwidth
of interest. In this case a correction factor, called process gain [34], must be
added to equation 2.7 to account for the related increase in the signal-to-noise
ratio,
fS
DR = 6.02 · N + 1.76 + 10 log . (2.11)
2 · BW
The process of sampling a signal at a rate which is greater than twice the
bandwidth is referred to as oversampling. Performing an M-point FFT over
the acquired waveform to extract information about a particular frequency
component, is equivalent to digitally filtering the signal with a bandwidth
equal to the frequency resolution of the FFT, that is fS /M . Therefore the
dynamic range due to the discrete Fourier transform is
M
DR = 6.02 · N + 1.76 + 10 log . (2.12)
2
2.2 System configuration 21
Output Voltage
From equation 2.12 it’s clear that to improve the dynamic range it is suf-
ficient to acquire a higher number of samples. In particular, doubling the
number of FFT points drops the magnitude of the asynchronous noise compo-
nents by a factor of 2, or 3 dB on a log scale. This expected improvement drops
off somewhat for large record lengths due to round-off errors in computing the
FFT.
Time-domain averaging is another technique to improve dynamic range as
it attenuates asynchronous noise sources by averaging time-domain waveforms
from multiple triggers. Averaging in the time domain not only reduces the
white noise present on the signal acquired, but also permits to increase the
number of effective bits of the digitizer, and thus increases the dynamic range.
This technique, often referred to as “oversampling and decimation” [35], can
be achieved by oversampling the signal and then taking the average of the
subsequent samples. For each additional bit of resolution, n, the signal must
be oversampled four times,
140
Theoretical
Measured
Dynamic Range, dB
120
100
80
60 0 2 4 6
10 10 10 10
Number of Acquired Samples
Vmeas
ADC
VDS
Imeas
VGS ADC
R
R
Figure 2.9: Current sense configuration with series resistor in the DUT
ground path.
are 200 V and 10 A. The pulsed currents and voltages at the device ports
are measured using 4 channels of the same PXI NI-5105 A/D converter board
used for the RF measurements. Time-domain averaging is also used on the
DC waveforms to increase the measurement accuracy.
Two different methods are available for sensing the current. The DC
pulsers employed provide a current monitor port with an internal 0.1 Ω sense
resistor to measure high currents. This resistor is inserted in the ground path
going from the DUT ground to the negative port of the DC supply as illus-
trated in Fig. 2.9. To sense lower currents with high resolution, a 1 Ω or 10 Ω
resistor can be placed in the external ground path just before the bias-tee.
Before the actual measurement can take place, the sense resistor is calibrated
by applying a range of voltages over a set of known resistors, while measuring
the current. For the highest accuracy, the calibration resistors and voltages
have to be selected according to the intended measurement range.
One advantage of using this method for sensing the current is certainly
the high bandwidth, which will result in shorter transient times, allowing
measurement of shorter pulses. On the other end, the sense resistor will cause
a voltage drop at the device terminals. For this reason the voltage needs to be
constantly monitored and iterations are necessary to reach the desired voltage
at the DUT ports. Furthermore, due to the position of the sense resistor
in the ground path, the source providing biasing to the pulser should have
a floating ground with respect to the common ground of the measurement
setup. Failing to do so would result in a parasitic current flow which is not
directed through the current sense resistor, reducing the current measurement
accuracy. The use of a completely floating source, like an Agilent 6629A, can
solve this problem.
24 Isothermal Measurement Systems
(a) (b)
Figure 2.10: (a) S21 (dB) of a band-pass filter for different pulse widths
with a PRP of 1 ms, (b) S21 (dB) of a band-pass filter for different duty
cycles with a PW of 1 µs.
range of the setup is more than 80 dB. The use of the wideband detection
technique, which employs a trigger to synchronize the data acquisition, com-
bined with digital processing of the acquired signal, ensures a wide dynamic
range that is totally independent from the pulse width. Moreover, the high
sampling rate of the A/D converters significantly reduces the constraints on
the minimum pulse width, which can be as short as 100 ns. In Fig. 2.10b
the results obtained for different duty cycle values, with a fixed pulse width of
1 µs, are presented. As expected, the duty cycle does not affect the dynamic
range of the setup. Fig. 2.10 also shows a comparison between the dynamic
range of the synthetic pulsed VNA with that of commercially-available instru-
ments such as the HP 8510C and the Agilent PNA. It is clear that the new
setup has a 20 dB improvement in dynamic range with respect to the 8510C.
Also, for low duty cycles the new setup has a higher dynamic range than the
PNA and the PNA-X.
1
10
Range0-1 V
Range0-10 V
0
10
Minimum current [A]
-1
10
0.1 Ohm
Imax. 10 A
-2
10 1 Ohm
Imax. 1 A
-3
10
10 Ohm
Imax. 0.1 A
-4
10 0 0.2 0.4 0.6 0.8 1 1.2
Pulse width [µs]
Figure 2.11: Minimum current vs. pulse width which can be measured
with 1 % accuracy using the current sense resistors for the indicated cur-
rent ranges (Imax). Due to ringing effects the current measurement accu-
racy is dependent on the voltage range (0-1 V) and (0-10 V) used in the
calibration.
different current sense resistors (0.1, 1.0 and 10 Ω). After calibration, indepen-
dent reference resistors, which were not included in the calibration, ranging
from 5 Ω to 220 Ω are measured and the deviation of the measured value from
their nominal resistance value is calculated to quantify the relative error in
the current measurement. The loading and reference resistors were previously
measured, in a temperature-controlled environment, by using a recently cali-
brated Agilent E5270B parameter analyzer, at small applied voltages to avoid
thermal effects on the resistors. Since the accuracy of the parameter analyzer
is higher than the A/D converter, the resistance values are accurately known.
Note that the gain and offset errors of the current measurement are corrected
by the current calibration. Therefore, the error calculated in such a way in-
cludes the inaccuracies of the A/D converters, the errors in the voltage setting
and the ringing of the pulse for smaller pulse widths.
The minimum current that can be measured with 1 % accuracy is plotted
as a function of pulse width for the different sense resistors in Fig. 2.11. In
all the experiments the averaging in the data acquisition was kept constant
at 1,024. From Fig. 2.11 it’s possible to conclude that currents above 1 mA
can be measured with an accuracy of 1 % on a resistive load, provided that
the pulse width is at least 500 ns. It appears that the low-voltage range has a
2.4 Measurements examples 27
slightly better accuracy for short pulse widths. This can be explained by the
lower ringing of the bias pulse in the low voltage range. This effect disappears
when the pulse width increases.
0.5 0.5
0.45 1
2 0.45 1
1 3 2
2 45
0.4 3 0.4 1 2 43
45 6 7 2 43 5 6
12 3 6 7 8 5 6 7
0.35 45 8 9 0.35 43 7 8
9 8 9
1
7 10 10 9 10
Ids , A
Ids , A
0.3 11 0.3 10
1
6 11 6
9 10 1 12 11
8
32 1
12
8
0.25 1 13 0.25 9 10 11 12
4312 2 1
7
13 12
9810 7 45 32 1
12
5
14 13 13
0.2 14 0.2 13
13
12 11
74 3612 312
15 14
15
0.15 0.15
53426 32
14
14
14
14
9 180
11
0.1 0.1
14 14 13
12
13 13 13 13 13 12
12 12
11 12
4 3521
0.05 11 11 12 0.05 11 11
3 65
8 10 8 10
956 4 10
9 10 10 10
4 7 3
956
24
7 31 2 7 31 8 56 6 8 75 9 6 8 75 9 6 8 75 9
0 0
0 5 10 15 20 0 5 10 15 20
Vds , V Vds , V
(a) (b)
0.5
0.45
0.4
1 1
0.35 2 2
23 4 35 4 35
4 5 6 6
Ids , A
0.3 7 8 7 7
9 8
9
7 5 321 1
0.25 8 9 10
10
6
11 10
11 11
8 93 12 4 1
0.2 12
12 12
13 13
0.15
1101
2413
0.1
6
13 13
13
13 12 12
12
34521
0.05 11 11
10
10 10
69 8 45 7 69 8 45 7 69 8 45
7
0
0 5 10 15 20
Vds , V
(c)
0.4 0.4
0.3 0.3
Ic , A
Ic , A
0.2 0.2
0.1 0.1
0 0
0 1 2 3 4 5 0 1 2 3 4 5
Vce , V Vce , V
(a) (b)
0.4 0.4
0.3 0.3
Ic , A
0.2 Ic , A 0.2
0.1 0.1
0 0
0 1 2 3 4 5 0 1 2 3 4 5
Vce , V Vce , V
(c) (d)
Figure 2.13: Measured output characteristic for a GaAs HBT device for a
pulse width of (a) 300 ns, (b) 500 ns, (c) 1 µs, and (d) 5 µs, with a pulse
period = 1 ms.
0.25 0.25
0.2 0.2
0.15 0.15
Ic
Ic
0.1 0.1
0.05 0.05
0 0
0 1 2 3 4 0 1 2 3 4
Vce Vce
(a) (b)
0.5
0.4
0.3
Ic
0.2
0.1
0
0 1 2 3 4
Vce
(c)
0.05 0.05
40
35 30
5
0.04 0.04
02
52
130
40
500
101
Ic , A
150
140
40
Ic , A
40
170
160
0.03 0.03
1000
180
30
200
35
25
0.02 35 35 0.02
1500
20
250
1015
30 30 30
0.01 0.01
300
25 25 25
2000
20 20 20
15
10 15
10 15
10
0 0
0.5 1 1.5 2 2.5 3 0.5 1 1.5 2 2.5 3
Vce , V Vce, V
(a) (b)
Figure 2.15: Measured contours of (a) constant fT (GHz) and (b) constant
Cbc (fF) for a QUBiC4plus SiGe HBT device plotted in the Ic(Vce) plane
using a pulse width of 1 µs (pulse period = 1 ms).
2.5 Conclusions
A new, low-cost, very flexible isothermal measurement setup has been pro-
posed which facilitates fast and accurate isothermal device characterization.
Measurements of different types of active devices have been performed, show-
ing that the setup is capable of measuring pulses as short as 200 ns for a wide
range of bias and power conditions, with good accuracy and measurement
speed.
The system is based on a custom pulsed network analyzer that achieves
a significantly higher dynamic range and measurement speed than currently
available commercial solutions for isothermal measurements, when a low duty
cycle is required. Moreover, the new setup does not suffer from the pulse
desensitization issue typical of a narrowband detection technique, allowing
the measurement of RF pulse widths down to 100 ns without any degradation
in the dynamic range, facilitating true isothermal measurements conditions.
32 Isothermal Measurement Systems
Chapter 3
33
34 Source and Load-pull Architectures
Bias-circuits (baseband)
IN OUT
(a)
Load-pull
Provide the same test
condition to the device
without building the circuit
b2
a2
a2
L
Reference plane
b2
(b)
The actual load impedance (ZL ) presented to the DUT in the microwave
domain is represented by the reflection coefficient related to this load, namely
ΓL . It represents the ratio between the reflected wave from the load (a2 )
and the forward traveling wave (b2 ). The generalized formula for ΓL at all
3.1 Passive load-pull 35
Figure 3.2: A conventional passive tuner, with moving probe and slab
line.
a2,n (fx )
ΓL,n (fx ) = , (3.1)
b2,n (fx )
in which a2,n is the wave reflected from the load, b2,n is the forward traveling
wave, n the harmonic index and fx a frequency in the band of interest. This
notation allows us to describe all the relevant circuit properties of the output
matching network and of the bias circuitry (Fig. 3.1a), while we can use a sim-
ilar formalism for the input matching. In a source or load-pull measurement
we control the reflection coefficients offered to the active device in an artificial
way and imitate the complete circuit without actually building it (Fig. 3.1b).
This yields major advantages in reducing the development time of new tran-
sistor technologies and their application in power amplifiers, while providing
a significantly better understanding of the behavior of the active device.
In the following sections the traditional load-pull architectures, along with
their strengths and weaknesses, will be described in detail.
RF
(a)
a1 b1 b2 a2
RF
(b)
RF
Signal
Generator RF
DUT
Reference Active Load
planes A fA
a1 a2
Amplifier
b1 b2
Figure 3.5: Open-loop active load-pull configuration.
content of the signal. This makes the closed-loop concept suitable for fast
device characterization. On the other hand the use of a feedback topology
can cause loop oscillations [47] when the forward gain of the feedback loop
is bigger than unity, or if |ΓL | · |ΓDU T | > 1 at any frequency. This is more
likely to occur when employing wideband loop amplifiers. To avoid the risk of
oscillation, most closed-loop systems include sophisticated filtering [48]. For
this purpose, a recent publication proposes to control the reflection coefficient
at baseband frequencies [49], where it is possible to implement better frequency
selectivity.
A further disadvantage of this technique is that high-power and very linear
injection amplifiers are needed in order to maintain the loop linearity, which
can increase the overall system cost tremendously.
The basic active tuning chain consists of a signal source, a variable phase
shifter and a variable gain stage. More advanced versions include in-phase /
quadrature (IQ) modulators to control the amplitude and phase of the injected
(a2 ) wave [54]. Since the a2 wave injected into the output of the DUT is no
longer a direct function of b2 , the reflection coefficients realized are power and
phase dependent. Therefore, iterations are needed during the measurements
to find the optimal injection signals to offer the desired reflection coefficients to
the DUT. This makes open-loop systems slower than the closed-loop ones. On
the other hand, there is no feedback path and therefore no chance of tuning-
loop oscillation compared to the closed-loop technique. Open-loop load-pull
also requires high-power amplifiers to achieve the desired reflection coefficients
when testing a high-power device. However, these amplifiers do not have to
be linear, since the user specified reflection coefficient is reached by successive
software iterations.
Furthermore, when extending the single-tone principle to wideband sig-
nals, determining the required content of the injection signals, at both the
fundamental and harmonic frequencies, becomes difficult. To tackle the prob-
lem, the use of multipliers in the signal path has been proposed [55]. However,
intermodulation distortion and memory effects will be present in practice,
making it impractical to obtain the required injection signals by a separate
analog multiplication process.
(a) (b)
Figure 3.6: (a) Phase delay caused by electrical lengths of the tuning
element plus cables, adapters and probe (b) Phase rotation of the reflection
coefficient for a 2.58 MHz wideband signal for a typical load-pull system.
probe in mechanical tuners), yields very large electrical delays, causing rapid
phase changes of the reflection coefficients versus frequency. Typical values
for these phase fluctuations start from about 3◦ /MHz for a fundamental pas-
sive mechanical tuner. If high-Q resonators or impedance tuners are used to
control the harmonic terminations, the phase change with frequency is even
greater. When considering active systems, the phase rotation can reach values
of 30◦ /MHz or more for conventional closed-loop systems (Fig. 3.6b). Note
that for a W-CDMA signal with adjacent and alternate channels (total band-
width of 25 MHz), the reflection coefficient offered by a passive tuner based on
these values would vary 75◦ in phase over this bandwidth, while in a conven-
tional active closed-loop the phase change of the offered reflection coefficient
would amount to 750◦ .
V2 Z 0 I 2
L
V2 Z 0 I 2
bDU T · (ZDU T + Z0 )
q
EDU T = √ bDU T = 2 · Pb2 · (1 − |ΓDU T |2 )
Z0
(3.2)
bSY S · (ZSY S + Z0 )
q
ESY S = √ bSY S = 2 · Pa2 · (1 − |ΓSY S |2 ) .
Z0
(3.3)
With reference to the schematic of Fig. 3.7, the injected power required
to achieve a certain ΓL , or in other words a certain impedance ZL = V2 /I2
3.4 Injection power and load amplifier linearity 43
Vdut
I_Probe P_1Tone
PORT2
I_dut
Num=2
S2P_Eqn
P_1Tone Z=50 Ohm
S2P1
PORT1 P=polar(pow er,phi)
S[1,1]=0
Num=1 Freq=1 GHz
S[1,2]=1
Z=2 Ohm
S[2,1]=1
P=200
S[2,2]=0
Freq=1 GHz
Z[1]=5
Z[2]=50
Figure 3.8: ADSTM schematic testbench for the evaluation of the required
injected power by active load-pull.
It is clear from equation (3.4) that the injected power needed not only
depends on the output power of the DUT and the desired ΓL , but also on
the output impedance of the device. When considering high-power devices,
with output impedances in the order of a few ohms, the required injection
power to cover the desired area of the Smith chart can be extremely high in
a 50 Ω system (e.g., 2 to 10 times higher than the maximum output power of
the DUT). To overcome this problem, pre-matching is typically used, which
converts the 50 Ω impedance of the system to a value that is much closer to
the output impedance of the DUT. This widely-used technique (also applied
in passive load-pull) not only reduces the losses, but also lowers the power
requirements of the load-injection amplifier [59]. As an example, consider a
DUT with an output impedance of 2 Ω and an available output power of
200 W. To synthesize a load impedance of 1 Ω in a 50 Ω system, the required
injection power would be more than 2 kW. Reducing the system impedance to
10 Ω with a lossless pre-match fixture lowers the required injection power for
the same load condition to 360 W, while with a pre-match to 5 Ω the required
injection power is only 142.2 W.
To provide a better understanding of the problem, a simulation using Ag-
ilent’s Advanced Design Simulator (ADS), was performed using the simple
schematic testbench illustrated in Fig. 3.8. The power source at port 1 repre-
sents a DUT with 200 W of output power and an output impedance of 2 Ω,
while the power source at port 2 represents the injection amplifier in a 50 Ω
44
GammaL Source and Load-pull Architectures
GammaL
(a) (b)
GammaL
(c)
− 2 · IP3,a2 , (3.5)
where Pb2 ,f und is the available power coming from the DUT at the fundamental
tones, and Pa2 ,f und and IP3,a2 are the power injected by the load amplifier at
the fundamental tones and its output third-order intercept point, respectively.
Another harmonic balance simulation with Agilent ADS is performed using
the simple schematic illustrated in Fig. 3.10. In this schematic, an amplifier
component based on a polynomial model is used to simulate the DUT and
the injection amplifier linearity. The same DUT is used as for the single-
tone considerations, with a maximum output power of 200 W and an output
impedance of 2 Ω. The output third-order intercept (OIP3 ) is set in this
simulation to 63 dBm. For this device, the output power is set equal to 50 W
per tone in order to achieve the same peak voltage as in the single-tone case.
These conditions yield an actual third-order intermodulation (IM3 ) of the
DUT of -30.35 dBc.
The results of the simulation are shown in Fig. 3.11, where the apparent
IM3 of the DUT is plotted as a function of the decreasing OIP3 of the injection
46 Source and Load-pull Architectures
vin vout
P_nT one P_nT one
I_Probe
PORT 1 PORT 2
I_dut2
Num=1 Amplifier2 Amplifier2 Num=2
S2P_Eqn
Z=50 Ohm DUT Inj_Amp Z=50 Ohm
S2P1
Freq[1]=2 GHz S21=38.456 S21=dbpolar(50,0) Freq[1]=2 GHz
S[1,1]=0
Freq[2]=2.1 GHz S11=polar(0,0) S11=polar(0,0) Freq[2]=2.1 GHz
S[1,2]=1
P[1]=dbmtow(7.7472) S22=polar(0.9231,180) S22=polar(0,180) P[1]=polar(dbmtow(power),180)
S[2,1]=1
P[2]=dbmtow(7.7472) S12=0 S12=0 P[2]=polar(dbmtow(power),180)
S[2,2]=0
Z[1]=50
Z[2]=5
Figure 3.10: ADSTM schematic testbench for the evaluation of the required
injection amplifier linearity by active load-pull.
3.5 Conclusions
When looking at traditional load-pull topologies, it is clear that all of them
have advantages and disadvantages. Mechanical tuners are simple, relatively
inexpensive and can handle high power, but they are slow and limited by losses.
Active load-pull systems can provide ΓL ≥ 1, are compact and therefore eas-
ily integrated for on-wafer measurements, but require expensive band-limited
amplifiers for high-power devices.
By combining passive tuners and active topologies in the same “hybrid”
system, it is possible to obtain many of the advantages of both systems while
reducing the disadvantages [27]. Traditional passive mechanical tuners can be
used to reflect high power at the fundamental frequency, thus pre-matching the
3.5 Conclusions 47
−20
Z =5 Ω
sys
Z sys =7 Ω
−25 Z sys =10 Ω
Passive Impedance
IM 3 level DUT [dBc]
Calculated
−30
−35
−40
−45
75 73 71 69 67 65 63 61 59
OIP Injection Amplifier [dBm]
3
Figure 3.11: Harmonic balance simulated IM3 level of the DUT vs. de-
creasing OIP3 of the injection amplifier for different impedance pre-match
values. The dotted line is the actual IM3 level as would be achieved with
passive matching techniques. The dot-dash line represents the IM3 level
only due to the Pa2 ,IM3 as approximated by equation 3.5. A polynomial
model was used for the amplifier linearity.
Mixed-Signal Active
Load-Pull with Realistic
Wideband Modulated Signals
The previous chapter has introduced the traditional load-pull topologies along
with their strengths and weaknesses, and has highlighted several areas where
major improvements are required in order to meet the demands of today’s
power amplifier design. The following chapters describe a novel active har-
monic load and source-pull system, which was developed at the Electronics
Research Laboratory of the Delft University of Technology during this the-
sis work, to overcome the limitations of conventional load-pull configurations
outlined previously. The proposed solution couples traditional analog and mi-
crowave techniques with low-frequency signal acquisition and generation and
the related digital signal processing, and will be referred to as “mixed-signal”
active load-pull.
This chapter describes the principles of this approach, the specific hard-
ware and software requirements and the ability of this load-pull setup to work
with realistic wideband communication signals. Measurement examples are
provided that illustrate the realized system functionality.
49
Mixed-Signal Active Load-Pull with Realistic Wideband
50 Modulated Signals
in which ax,n and bx,n are the incident and reflected waves at port x and
harmonic index n, while Γx,n represent the user-defined reflection coefficients
versus frequency for port x and harmonic index n.
As in the classical open-loop approach, only the content of the driving
waveform (as ) is known prior to the acquisition. All other injection signals
(a1inject,n and a2inject,n ) containing all the frequency components of the signal
4.1 The wideband, open-loop load-pull approach 51
NO
Figure 4.2: Flow diagram for the optimization of the reflection coefficients.
Mixed-Signal Active Load-Pull with Realistic Wideband
52 Modulated Signals
load-pull system.
Source Source
HPR
a1,f0 a1,2f0 a1,BB b1,f0 b1,2f0 b1,BB b2,f0 b2,2f0 b2,BB a2,f0 a2,2f0 a2,BB To
LO
4.2 System configuration
X2
a1,f0 a1,2f0 b1,f0 b1,2f0 b2,f0 b2,2f0 a2,f0 a2,2f0
To RF To RF
LO LO LO LO LO LO @ f0 @ 2f0
On-wafer configuration
RF @ f0 PA @ f0 RF @ f0
To DC To DC PA @ f0
DUT
Bias Bias
Tee Tee
Input section Output section
Reference
BaseBand I V Planes I V BaseBand Load
Reference
PA @ 2f0 Plane PA @ 2f0
RF @ 2f0 RF @ 2f0
To DC To DC
DUT
Reference Planes
BaseBand I V I V BaseBand
cients are synthesized by injecting signals into the DUT which are generated
by baseband arbitrary waveform generators (AWG) and up-converted using
in-phase/quadrature (IQ) modulators. All the AWGs and the A/D converters
are integrated in a PXI express platform. They share the same time-base and
are fully synchronized. Since data generation and data acquisition of both
RF signals and DC parameters are handled through the PXI-based D/A and
A/D instrumentation, no mechanical tuners, VNA or DC-parameter analyzer
is needed, yielding a cost-effective, high-end characterization solution. Custom
bias-Tees with low inductance are placed directly at the DUT reference planes
in order to minimize the electrical delay of the baseband (BB) impedance,
which is implemented for now as a passive impedance switch bank [41]. Low-
frequency couplers for the baseband impedance measurement are also imple-
mented on the baseband board. Note that in a more extended version, the
baseband impedance can also be made active and controlled by an additional
AWG (see Section 7.1.2).
f sAW G 1
∆fAW G = = , (4.2)
NAW G TM OD
where ∆fAW G represents the frequency bin size of the generated signals, and
f sAW G and NAW G are (respectively) the sampling frequency and the number
of samples used by the arbitrary waveform generators to construct the wave-
form. TM OD is the minimum period of the source signal in the time domain,
that is needed to comply with the specifications of the standard test model.
To provide the reader with an example, a W-CDMA signal has a chan-
nel bandwidth of 5 MHz, a chip rate of 3.84 Mcps, 2560 chips/slot and
15 slots/frame. When considering one frame, the complex waveform is 10 ms
long, or in other words, it will have a frequency resolution of 100 Hz. If we
then consider a single slot, the frequency resolution becomes 1.5 kHz. This
frequency representation allows us to analyze modulated communication sig-
nals like “classical” multi-tone signals, but now with a very large number of
frequency tones (e.g., more than 23,000 frequency tones when considering a
bandwidth of 35 MHz).
Mixed-Signal Active Load-Pull with Realistic Wideband
58 Modulated Signals
Amplitude [dBm]
Amplitude [dBm]
0 0
I Q
∆f
-40 -40
BW/2
-80 -80
0 2 4 6 8 10 12 14 0 2 4 6 8 10 12 14
Frequency [MHz] Frequency [MHz]
(a)
Amplitude [dBm]
-30
0 20 40 60 80 100
Frequency [MHz]
(b)
Amplitude [dBm]
Amplitude [dBm]
-30 -30
I Q
-60 -60 IM3
IM3 IM5
IM5
-90 -90
0 10 20 30 40 0 10 20 30 40
Frequency [MHz] Frequency [MHz]
(c)
Figure 4.4: Illustration of the generated and acquired signals in the pro-
posed load-pull system. Each circle corresponds to a frequency bin. (a)
Frequency-binned spectral content of the I and Q waveforms for generat-
ing the drive signal of the DUT. (b) Down-converted low IF representation
of the spectrum in the fundamental band at the output of the DUT. (c)
Spectral content of the I and Q waveforms for generating the active load
injection signal to achieve the user-defined reflection coefficient over the
fundamental band.
where ∆fA/D is the resulting frequency bin size of the acquired signals, f sA/D
and NA/D are (respectively) the sampling frequency and the number of samples
used by the A/D converters, and k is an integer.
The frequency bins of the acquisition and the generation should match for
a correct measurement, thus the frequency resolution of the A/D converter
should be set equal (k=1), or an integer factor larger (smaller frequency bin
size) than that of the generated signals.
−60
Receiver Attenuator = 0 dB
−65 Receiver Attenuator = 10 dB
−70
−75
IM 3 [dBc]
−80
−85
−90
−95
−100
−15 −10 −5 0 5 10 15 20 25 30
Output Power [dBm]
0.5
Magnitude
0.4
0.3
0.2
0.1
2.13 2.135 2.14 2.145 2.15
(a) Frequency [GHz]
-90
Phase [degrees]
-120
-150
-180
2.13 2.135 2.14 2.145 2.15
(b) Frequency [GHz]
Magnitude [dBm/Hz]
0 60 dB
-20
-40
-60 Control threshold
-80
2.13 2.135 2.14 2.145 2.15
(c) Frequency [GHz]
Figure 4.7: Measured load Γ in (a) magnitude and (b) phase, and (c) power
spectrum (resolution bandwidth of 6 kHz) at the output reference plane of
a HBT overdriven with a W-CDMA signal yielding spectrum broadening.
The user-specified reflection coefficient is ΓL = |0.283|∠−135o . The power
threshold level for the load reflection coefficient is set to -65 dBm.
the edge of the band corresponds to those frequencies where the signal power
falls below the user-specified threshold.
1 1
-1 -1
-1 1 -1 1
Figure 4.8: Source and load reflection coefficients at the device refer-
ence plane in the fundamental (2.1225 GHz - 2.1575 GHz) and harmonic
(4.245 GHz - 4.315 GHz) frequency ranges, with electrical delay (open
symbols) and without electrical delay (filled symbols).
namely: ΓL,f 1 = |0.6| ∠45o and ΓS,f 1 = |0.5| ∠90o . The input and output
baseband impedances enforce a short condition, and the input and output 2nd
harmonics are set to an open circuit condition (ΓL,f 2 = ΓS,f 2 = |0.95|) in or-
der to optimize the efficiency [67]. To highlight the excellent wideband control
and the delay-free electrical operation of the new measurement setup, a com-
parison is made with a previously developed state-of-the-art active harmonic
load-pull system [41], which was optimized especially for minimum electrical
delay. For this purpose, we use a two channel W-CDMA signal (centered at
2.135 GHz and 2.145 GHz) and set the input and output reflection coefficients
in the newly developed setup to the following two cases:
2. With an electrical delay of 4.85◦ /MHz for the fundamental source and
load and 4.6◦ /MHz for the 2nd harmonic source and load.
Mixed-Signal Active Load-Pull with Realistic Wideband
64 Modulated Signals
Fig. 4.8 illustrates the source and load matching conditions provided to
the active device under test for the two different cases. Note that the filled
markers represent the source and loading conditions for the two-carrier W-
CDMA signal without any electrical delay, yielding completely overlapping
points in the Smith chart. As shown in Fig. 4.8, the fundamental load trajec-
tory has been shifted for the case with electrical delay such that the optimum
matching condition is now centered at 2.135 GHz. This was required to avoid
the unstable region of the active device.
It is important to note that this is a comparison to the “best known case” of
a classical closed-loop active load-pull system, since closed-loops are subject to
amplitude variations within the control frequency bands in practice. Moreover,
oscillation conditions in closed-loop systems for these very large bandwidths
are difficult to avoid due to the use of wideband loop filters. Passive load-pull
systems with harmonic tuning will have a comparable or even worse phase
variation of the reflection coefficients versus frequency than the closed-loop
system used in the comparison.
The measurement results are summarized in Table 4.1. There is significant
performance degradation for the active device when measured with electrical
delay present in the reflection coefficients. This is also evident from Fig. 4.9a
and 4.9b which show the power spectral density at the device output reference
plane for the fundamental and 2nd harmonic frequency bands. Note that a
5 dB output power drop and close to 8 % degradation of the power-added
efficiency (PAE) can be observed when compared to the situation with no
electrical delay. Furthermore, due to the electrical delay, the output power
spectrum shows significant asimmetry both at fundamental and 2nd harmonic
frequency bands.
MEASUREMENT RESULTS
Without With
electrical delay electrical delay
PAE 24.2 % 16.3 %
POUT Ch. 1 20.3 dBm 20.5 dBm
POUT Ch. 2 20.6 dBm 15.4 dBm
ACLR1 Ch. 1 -43.9 dBc -43.0 dBc
ACLR2 Ch. 1 -42.2 dBc -41.6 dBc
ACLR1 Ch. 2 -42.1 dBc -41.8 dBc
ACLR2 Ch. 2 -39.6 dBc -39.2 dBc
-10 -20
Output Power Spectral Density [dBm/Hz]
-30 -40
-40 -50
-50 -60
-60 -70
Without delay Without delay
With delay With delay
-70 -80
2.125 2.13 2.135 2.14 2.145 2.15 2.155 4.25 4.26 4.27 4.28 4.29 4.3 4.31
Frequency [GHz] Frequency [GHz]
(a) (b)
Figure 4.9: Measured output power spectral density [dBm/Hz] vs. fre-
quency [GHz] of a NXP GEN6 LDMOS device (gate width 1.8 mm) in
the proposed load-pull setup (a) at the fundamental frequency band using
a 3 kHz resolution bandwidth. (b) at the 2nd harmonic frequency band
using a 6 kHz resolution bandwidth. The measurement is shown for the
two cases with (dashed line) and without electrical delay (drawn line).
The reflection coefficients offered to the device under test are given in
Fig. 4.8.
4.6 Conclusions
In this chapter, a novel active harmonic load-pull setup has been presented
which is suitable for large-signal device characterization under “real life” mod-
ulated signal stimulus. The new system is capable of synthesizing arbitrary
source and loading conditions at the fundamental and harmonic frequencies
over a frequency bandwidth of 120 MHz, and therefore permits testing of ac-
tive devices under realistic (circuit-like) conditions. Furthermore, even user-
defined reflection coefficients versus frequency can be downloaded to the mea-
surement system in order to perform active device testing.
The IQ signal up-conversion and low IF down-conversion provide a very
high dynamic range of about 80 dB in the signal detection and about 60 dB in
the modulated active load impedance control. This proves to be sufficient for
the communication standards used today in industry. Furthermore, this ap-
proach can be easily adjusted to any frequency band of interest (e.g., X-band)
since all of the parts required are commercially available. The IQ-based open-
loop approach in combination with the frequency binning technique described
resolves all of the conventional drawbacks of current load-pull techniques, while
Mixed-Signal Active Load-Pull with Realistic Wideband
66 Modulated Signals
High-Speed, High-Power,
Fully-Controlled,
Multi-dimensional Load-Pull
Parameter Sweeps
67
High-Speed, High-Power, Fully-Controlled, Multi-dimensional
68 Load-Pull Parameter Sweeps
RF Input Signal
Power 2
1 Power 1
Voltage [V]
−1
0 5 10 15 20 25 30
(a) Time [µs]
RF Load Injection Signal
Load 2 Load 3 Load 4
1 Load 1
Voltage [V]
Figure 5.1: Time-segmented RF waves for (a) multiple input power levels
and (b) load termination control. In this example four different loads over
a range of two power levels are presented to the DUT.
1
0.13
0.14 ΓLf0 1
0.12 0.13
0.14 ΓSf0
1.0
0.15 0.15
0.1
6
0.1
ΓLf0 Target 0.1
6
0.1
ΓSf0 Target
7 7
0.8 0.1
ΓL2f0 0.8 0.1
ΓS2f0
2.0
2.0
8 8
0.1
0.1
ΓL2f0 Target ΓS2f0 Target
9
9
0.6
0.2
0.6
0.2
0
0
0.2
0.2
1
1
0.4 5.0 0.4 5.0
1.0
1.0
0.22
0.22
0.23
0.23
0.2 0.4
10
0.2 0.4
10
0.24
0.24
2.0
5.0
1.0
2.0
5.0
10
20
50
10
20
50
0.25
0.25
0 0
(a) (b)
that can yield device failure (voltage or thermal breakdown); something that
is far from trivial when using analog load-modulation methods. An example
is given in Fig. 5.3, showing the constant PAE contours for a given PAV S as
measured with the proposed technique.
In this example, the output stability circle was first obtained from the (pre-
viously measured) device small-signal S parameters [37], then only the stable
region was addressed in the actual measurement by a fundamental load sweep
at various (much higher) power levels. In this measurement the output power
of the device has been obtained for 90 different load terminations, keeping ΓS
fixed to the previously specified value, while at the same time sweeping the
source and load harmonic terminations between open and short conditions,
and the power available from the source from 1 to 16 dBm. The total time for
this measurement, which consists of more than 5,000 controlled measurement
points, is below 5 minutes.
Given the importance of second harmonic source and load termination con-
trol in PA design, which has been addressed several times in literature [72,73],
the capability of the presented system to simultaneously sweep the 2nd har-
monic impedances presented to the DUT can prove to be very useful when in-
vestigating the optimal device terminations for high-efficiency / high-linearity
operation. The constant PAE contours plotted in Fig. 5.3 show that the
highest PAE values are obtained with the 2nd harmonic load impedance set
to an open, and the 2nd harmonic source impedance set to a short. For this
latter case, the measured PAE as a function of POU T and the measured trans-
ducer power gain (GT ) as a function of PAV S are plotted for all the different
fundamental loading conditions in Fig. 5.4a and Fig. 5.4b, respectively.
Finally, the constant PAE contours shown in Fig. 5.5a and 5.5b allow the
reader to compare the results obtained using the proposed “real-time” tech-
nique and a traditional open-loop load-pull technique under the same PAV S
and loading conditions. The excellent agreement between the two methods
indicates that the much higher measurement speed of the new approach does
not affect accuracy.
1.0
0.15 0.15
0.1 0.1
6 6
0.1 0.1
7 7
0.8 0.1
0.8 0.1
2.0
2.0
8 8
0.1
0.1
9
9
0.6
0.2
0.6
0.2
0
0
0.2
0.2
0.70
1
1
0.4 5.0 0.4 0.65 5.0
1.0
1.0
0.22
0.22
0.65
0.6
0.6
0.55
0.23
0.23
0.55
0.2 0.4 0.4 0.5 10
0.2 0.4
0.5 10
0.45 0.45
0.24
0.24
0.4
1.0
2.0
5.0
1.0
2.0
5.0
10
20
50
10
20
50
0.25
0.25
0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
(a) (b)
1.0
0.15 0.15
0.1 0.1
6 6
0.1 0.1
7 7
0.8 0.1
0.8 0.1
2.0
8 2.0 8
0.1
0.1
9
9
0.6
0.2
0.6
0.2
0
0
0.2
0.2
1
1.0
0.65
0.22
0.22
0.6
0.6
0.55
0.23
0.23
0.55
0.4
0.5
0.2 0.4 10
0.2 10
0.4
5 0.4
0.
4 0.5
0.24
0.24
0.45
1.0
2.0
5.0
1.0
2.0
5.0
10
20
50
10
20
50
0.25
0.25
0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
(c) (d)
Figure 5.3: Constant PAE load-pull contours for a NXP GEN6 LDMOS
device (PAV S = 14 dBm). In this measurement the power available from
the source is swept from 1 to 16 dBm for 90 different load terminations.
The source fundamental termination is set to ΓS = |0.5| ∠90o , while load
and source 2nd harmonic terminations are swept to: (a) open - short (b)
open - open (c) short - open (d) short - short using the proposed tech-
nique. Note that the unstable area for the DUT is avoided directly during
the measurement (i.e., no measurement points taken in the potentially
unstable region).
5.2 High-power, real-time pulsed-RF measurements 73
0.8 21
0.7
20
0.6
0.5 19
GT [dBm]
PAE
0.4 18
0.3
17
0.2
0.1 16
0 15
18 20 22 24 26 28 30 32 34 0 2 4 6 8 10 12 14 16 18
POUT [dBm] PAVS[dBm]
(a) (b)
1.0
0.15 0.15
0.1 0.1
6 6
0.1 0.1
7 7
0.8 0.1
0.8 0.1
2.0
2.0
8 8
0.1
0.1
9
9
0.2
0.2
0.6 0.6
0
0
0.2
0.2
1
0.6 0.6
0.4 5.0 0.4 5.0
0.22
0.22
1.0
1.0
0.5
0.5
0.23
0.23
0.4 0.4
0.2 0.4
10
0.2 0.4
10
0.24
0.24
0.3
0.3
1.0
2.0
5.0
1.0
2.0
5.0
10
20
50
10
20
50
0.25
0.25
0 0
(a) (b)
RF Input Signal
1
PW POWER 1 POWER 2
Voltage [V]
0.5
−0.5
PRP
−1
0 40 80 120 160 200 240
(a) Time [ µs]
RF Load Injection Signal
1
L1 L2 L3 L1 L2 L3
Voltage [V]
0.5
−0.5
−1
0 40 80 120 160 200 240
(b) Time [ µs]
Figure 5.6: Time-segmented pulsed-RF waves for (a) multiple input power
levels and (b) load termination control with pulsed-RF. In this simplified
example, three different loads for two input power levels are presented to
the DUT.
The results are shown in Fig. 5.7, where the maximum output power contour
and the PAE at 3 dB of gain compression are plotted. Also in Fig. 5.8, the
PAE versus output power calculated at a gain compression level of 3 dB is
shown. Note that the complete device characterization with 25 power levels
at each of the 50 load impedances takes less than 3 minutes.
0
Max PAE [%]
Pout @ compression [W]
−0.1
77.
−0.2 8
63.3
88.
.6
61.9
9
64
−0.3
99.
9
59.
11
1.0
3
−0.4
55 51.2
12
2
.2
13 .1
3.2
1
15 44.3
−0.5
45 16 .35
.9 6
−0.6 177 .4
.5
40.
−0.7 5
−0.8
−0.6 −0.5 −0.4 −0.3 −0.2 −0.1 0 0.1
70
65
Power Added Efficiency [%]
60
55
50
45
40
35
60 80 100 120 140 160 180 200
Output Power @ compression [W]
Figure 5.8: Measured PAE vs. output power at 3 dB power gain compres-
sion level for a NXP GEN7 LDMOS device at different load states and
input powers, using pulsed single-tone conditions (10 µs pulse width and
10 % duty cycle).
5.3 High-power measurements with modulated signals 77
0 0
Average PAE [%] @ 3 dB PAR reduction
Average PAE [%] for 30 W output power Average Pout [W] @ 3 dB PAR reduction
ACPR [dB] for 30 W output power
−0.1 −0.1
−0.2 −0.2
−0.3 −0.3
33.
−0.4 7 −0.4
32.
30. 3
22.5
32
23.7
9 .3
29
24.9
30
28 .5
26.1
−0.5 29 .9
27.3
−0.5
26 .1 2 .5
28.5
26 8.0
29.8
25 .8
30.9
32.2
2 .6
33.4
.
2 4 −34.9 2 5
22 3.8 .2
−0.6 22. 4 −0.6 2 .
19 1.0 3
21. 6 −34.3 18 .5
19. 2 −33.8 .1
8 −33.3
−0.7 −32.7 −0.7
−31−.232.2
−0.8 −0.8
−0.6 −0.5 −0.4 −0.3 −0.2 −0.1 0 0.1 −0.6 −0.5 −0.4 −0.3 −0.2 −0.1 0 0.1
(a) (b)
Figure 5.9: (a) Load-pull contours of average power added efficiency and
ACPR for an average output power of 30 W. (b) Load-pull contours of
average power added efficiency and average output power at 3 dB of peak-
to-average ratio reduction. The related peak-to-average power (PEP) is
as high as 150 W.
at high powers with wideband signals arises because the linearity of the injec-
tion amplifier needs to be taken into account. When the injection amplifier is
not sufficiently linear, it will introduce intermodulation products in a “conven-
tional” active load-pull system, that will cause significant measurement errors,
such as IM3 increase or cancellation effects. To have reliable linearity mea-
surements in a “conventional” active load-pull setup (even when pre-matching
is used), the injection amplifier linearity (and thus its peak power) needs to
be at least 10 times higher than that the of the DUT, as concluded from the
simulation shown in Fig. 3.11. When working with devices designed for base-
station applications where peak envelope powers are as high as 200 W, using
such amplifiers becomes extremely expensive and impractical.
Instead, an iteration process is performed in the mixed-signal active load-
pull system, to optimize the reflection coefficient of each individual frequency
component of the wideband signal (e.g., 23,362 tones with 1.5 kHz spacing
for a W-CDMA considering a total bandwidth of 35 MHz). Due to these
iterations, the injection amplifier is pre-distorted for its own non-linearities.
This allows the use of injection amplifiers with much lower linearity than what
is typically required in conventional active load-pull systems.
High-Speed, High-Power, Fully-Controlled, Multi-dimensional
78 Load-Pull Parameter Sweeps
As an example consider Fig. 5.9, which shows ACPR, average PAE and
output power for a single-channel W-CDMA signal at 2.14 GHz with a peak-
to-average ratio of 9.5 dB. It should be stressed that in these experiments the
maximum saturated power rating of the injection amplifier is only 200 W, with
an associated 60 dBm output IP3 . In these measurements, the non-linearity
of the injection amplifier does not affect the measurement results because ΓL
is controlled to the user-defined value in and out-of-band.
5.4 Conclusions
In this chapter, the measurement principles that enable the mixed-signal ac-
tive harmonic load-pull system to be compatible with the requirements of
high-speed, high-power, high-linearity base-station applications have been pre-
sented. The system provides ultra-fast large-signal device characterization in
both CW and pulsed conditions. For the latter, both the duty-cycle and the
gain compression of the DUT during the measurement can be controlled. All
of these features are crucial in guaranteeing the safe operating conditions for
high-power DUTs (> 100 W).
In addition, high-power device characterization with realistic W-CDMA
signals has been demonstrated. It was shown that the realized system can
compensate for nonlinearities of the injection amplifiers, which normally would
obscure the linearity and ACPR measurements. This property allows the use
of cheaper injection amplifiers, providing a lower PSAT than required in con-
ventional active load-pull systems. The ability to eliminate losses and electrical
delay while being completely free to define the source and load reflection coef-
ficients vs. frequency, allows perfect mimicking of in-circuit situations, making
the system an extremely valuable tool for the RF power amplifier developer.
Chapter 6
Measurement of
Time-Domain Waveforms
79
80 Measurement of Time-Domain Waveforms
1
ii (t) = √ · [ai (t) − bi (t)] , (6.2)
Z0
where i is the port number, and
N
X
ai (t) = ai,n · cos(nω0 t + φai ,n ) (6.3)
n=0
N
X
bi (t) = bi,n · cos(nω0 t + φbi ,n ) . (6.4)
n=0
Variables ai,n , bi,n and φai ,n , φbi ,n are the amplitude and phase coefficients
of the power waves measured at the fundamental and harmonic frequencies,
respectively.
The errors in a measurement system have to be removed by means of a
system calibration when working at microwave frequencies. In VNAs, all of
the error contributions of the measurement system are described by an error
model [62, 82]. As an example, consider the flow graph in Fig. 6.1 which
depicts the error model at the input port of the DUT. The four error terms
ed1 , es1 , i01 and i10 represent the error network between the DUT reference
plane and the measurement plane. By measuring several electrical standards,
the error terms of the model can be calculated and the measurement corrected
up to the DUT reference plane.
Conventional network analysis, however, allows only the measurement of
the ratio of power waves (e.g., b1 /a1 ). In fact, only the product term i01 · i10
is derived when referring to the classical 12 term error model used for VNA
calibration. This is better explained when considering the flow graph in Fig.
6.1. Equations 6.5 and 6.6 can be derived from the flow graph, which allow
the calculation of the calibrated incident and reflected waves at the input port
of the DUT from the measured quantities aM M
1 and b1 . It is clear that the
two terms i01 and i10 have to be known independently both in amplitude and
phase in order to calculate a1 and b1 according to
es1 · bM M
1 − ∆ · a1
a1 = (6.5)
i01
6.1 Time-domain waveform measurement fundamentals 81
a 1M i10 a1
b 1M i01 b1
bM M
1 − ed1 · a1
b1 = , (6.6)
i01
where ∆ = ed1 · es1 − i01 · i10 .
The addition of an extra calibration step with a power meter [63] allows
the calculation of the amplitude of the separate terms i01 and i10 from the
product term i01 · i10 , enabling the measurement of the absolute power of the
a and b waves. Finally, when the phase information of the input and output
signals are also required, an extra phase calibration step is needed to acquire
the phase information of the i01 and i10 terms.
Traditionally two techniques can be employed to generate a signal to be
used for a phase calibration:
aREF a1 b1
LO
Source
Comb
Generator
DUT
RF
Source
HPR
a1,f0 a1,2f0 a1,BB b1,f0 b1,2f0 b1,BB b2,f0 b2,2f0 b2,BB a2,f0 a2,2f0 a2,BB
Figure 6.3: Simplified block diagram of the phase reference channel needed
to enable waveform reconstruction in the mixed-signal load-pull system.
Detail from Fig. 4.3.
1
Phase Error [deg]
−1
−2
0 5 10 15
Frequency [GHz]
30
20
Output Power [dBm]
10
−10
−20
−30
0 2 4 6 8 10 12 14
Frequency [Hz]
50
0 f0
Phase Variation [deg]
2f0
−50 3f0
4f
0
−100 5f0
6f0
−150 7f0
−200
0 1 2 3 4 5
Input Power [dBm]
Figure 6.6: b2 phase variation for the different harmonics versus input
power for a Marki A-0030 amplifier.
88 Measurement of Time-Domain Waveforms
0.3 0.3
0.2 0.2 22
2 2
4 10
0.1 0.1 18
146
25 0
1
2 102
2 5
0 0 8
6
10 1 12 4
20 2
−0.1 −0.1 14
5 6
1 5 108
−0.2 15 0 −0.2
25 16
20 18
−0.3 −0.3
−0.4 −0.4
−0.4 −0.2 0 0.2 0.4 −0.4 −0.2 0 0.2 0.4
(a) (b)
th th
4 harmonic 5 harmonic
0.4 0.4
0.3 0.3
0.2 25 0.2 25
0.1 20 0.1 20
15 15
10 10
0 0
10 10 5
5
−0.1 −0.1 15
5 5
−0.2 15 −0.2
20 20
25 25
−0.3 −0.3
−0.4 −0.4
−0.4 −0.2 0 0.2 0.4 −0.4 −0.2 0 0.2 0.4
(c) (d)
0.3 0.3
0.2 5 35 0.2
30
0.1 25 0.1
5 20 5
0 15 0 4405
20 10 233505
10
5
25
−0.1 −0.1 125
5 15 230 50
20
1510 35
−0.2 30 −0.2
−0.3 −0.3
−0.4 −0.4
−0.4 −0.2 0 0.2 0.4 −0.4 −0.2 0 0.2 0.4
(e) (f)
0.3 0.3
0.2 0.2
0.1 0.1
4.5 4
0 0
2
1.5
3.5
0.5
3
2.5 2
3
1.5
−0.1 −0.1
3.5
1
2.5
4
2.53
0.5
1.5
1.52
0.5
1
2
1
0.5
−0.2 −0.2
−0.3 −0.3
−0.4 −0.4
−0.4 −0.2 0 0.2 0.4 −0.4 −0.2 0 0.2 0.4
(a) (b)
th th
4 harmonic 5 harmonic
0.4 0.4
0.3 0.3
65 3
4
0.2 0.2 1 32
1
2
2
1.5
0.5
2.5
0.5
1
1.5 1
0.1 0.1
1
2
2
3
0 0 3
6
3
2
2.5
−0.1 −0.1
1
5 2
4
3
−0.2 −0.2 1 1
−0.3 −0.3
−0.4 −0.4
−0.4 −0.2 0 0.2 0.4 −0.4 −0.2 0 0.2 0.4
(c) (d)
0.3 0.3 25
20
0.2 0.2
2.5 3
0.5 15
0.5 51
1. 2
1 10
3.5
3.5
0 0 5
5
4
−0.1 −0.1 10
15
−0.2 0.5 −0.2
20
−0.3 −0.3
−0.4 −0.4
−0.4 −0.2 0 0.2 0.4 −0.4 −0.2 0 0.2 0.4
(e) (f)
Figure 6.8: b2 phase variation of a Marki A-0030 amplifier vs. 2nd har-
monic output load Γ at: (a) 2nd harmonic, (b) 3rd harmonic, (c) 4th
harmonic, (d) 5th harmonic, (e) 6th harmonic, (f) 7th harmonic.
90 Measurement of Time-Domain Waveforms
−15
−20
Output Power [dBm]
−25
−30
−35
−40
−45
0 2 4 6 8 10 12 14
Frequency [GHz]
−5 f0
Phase Variation [deg]
2f0
−10 3f0
4f0
−15 5f0
6f0
−20 7f0
−25
0 2 4 6 8 10
Input Power [dBm]
Figure 6.10: b2 phase variation for the different harmonics versus power
for a HHFT MD-CG1 comb generator.
6.2 Waveform reconstruction on the mixed-signal load-pull
system 91
0.3 0.3
7 6 5 10 8
0.2 0.2
3
2 1
0.1 0.1 6
4 2
2 4
0 0
1
3 4
4 6
5 2 8
−0.1 7 6 −0.1
1
−0.2 −0.2 2 10
−0.3 −0.3
−0.4 −0.4
−0.4 −0.2 0 0.2 0.4 −0.4 −0.2 0 0.2 0.4
(a) (b)
th th
4 harmonic 5 harmonic
0.4 0.4
0.3 0.3
14 12 10 16 14 12 0
0.2 0.2 1
6 6
4 2
2 4
0.1 0.1
8
2 8 4
0 4 0 6
6 8
8 1102
−0.1 10 −0.1 2 14
2 12 2 16
−0.2 14 −0.2 18
−0.3 −0.3
−0.4 −0.4
−0.4 −0.2 0 0.2 0.4 −0.4 −0.2 0 0.2 0.4
(c) (d)
0.3 0.3
15 20
0.2 10 0.2 15
5
5
0.1 5 0.1
10 5
0 10 0
10
15 15
−0.1 20 −0.1
20
−0.2 25 −0.2 25
−0.3 −0.3
−0.4 −0.4
−0.4 −0.2 0 0.2 0.4 −0.4 −0.2 0 0.2 0.4
(e) (f)
0.3 0.3
1
0.2 0.2
1.8
1
1
211.61.4.2
6
0.
0.4
0.8
3
5
1
0.1 0.1 0.2
6
0.8
1.
0
0..4
2
1.4
6
6
0.2
0 0
0.6
6
7
0.
4
0.6
0.
−0.1 −0.1
00.
1
1
4
11. .86
3
0.
2
2
4
11..8.642
1
−0.2 −0.2 0.4
−0.3 −0.3
−0.4 −0.4
−0.4 −0.2 0 0.2 0.4 −0.4 −0.2 0 0.2 0.4
(a) (b)
th th
4 harmonic 5 harmonic
0.4 0.4
0.3 0.3
5
1
1.5
0.
0.2 0.2 2
2
1
1 1.5
2
1
5
1.
5
5
1
0.5
0 0
0.5
0.
1
1
1
0.5
2
1
0.5
1 2
−0.3 −0.3
−0.4 −0.4
−0.4 −0.2 0 0.2 0.4 −0.4 −0.2 0 0.2 0.4
(c) (d)
0.3 0.3
2 1
0.2 0.2
3
5.5 3.5
434.5 1
3
2.5
0.1 2 1 0.1 2
2
23.5 1 1
1.5
5
0 0
1
32
1.
1
−0.1 0.5 −0.1 1 1
2
1
32
1.52
0.5
1 2
−0.2 1.5 1 −0.2 4
1
5
2 6
−0.3 −0.3
−0.4 −0.4
−0.4 −0.2 0 0.2 0.4 −0.4 −0.2 0 0.2 0.4
(e) (f)
−5
Output Power [dBm]
−10
−15
−20
−25
−30
−35
0 2 4 6 8 10 12 14
Frequency [GHz]
Figure 6.13: Output Power Spectrum of the new comb generator proto-
type.
4 f0
Phase Variation [deg]
2f0
2 3f0
4f0
0 5f0
6f0
−2 7f0
−4
10 11 12 13 14 15 16 17
Input Power [dBm]
Figure 6.14: b2 phase variation for the different harmonics versus power
for the new comb generator prototype.
94 Measurement of Time-Domain Waveforms
0.3 0.3
0.2
0.2 0.2 0.4
1.4
11.2 1.21
0.
1
1
8
0.1 0.5 0.1 0.8
0.6 0.4
0.5 0.6
0.81
1.2
0 1 0 0.2
1 1.21
11.8.6
0.4 .6
0 .8
−0.1 −0.1 . 4 00.6
1 0.4
1.5
1.6
1
0.2
−0.2 −0.2 1.4 1.
1.5 2
1 2 1
−0.3 −0.3
−0.4 −0.4
−0.4 −0.2 0 0.2 0.4 −0.4 −0.2 0 0.2 0.4
(a) (b)
th th
4 harmonic 5 harmonic
0.4 0.4
0.3 0.3
1 1
1 2
0.2 0.2 1.5 2.5
0.5
1.
2.
5 1.51 .5 1.25 1
0.1 0.1 2 2 0.5
1
0.5 1.5
1.5
2
35 2
0 2 0
1
0.5
2.5
3
2
0.
2.
1 5
1.
1
−0.1 −0.1
1.5
−0.2 −0.2 2
2
−0.3 −0.3
−0.4 −0.4
−0.4 −0.2 0 0.2 0.4 −0.4 −0.2 0 0.2 0.4
(c) (d)
0.3 0.3 1
2
1 1.5 2
1
0.2 1 0.2 3
0.5 2
0.1 1 2 0.1
1.5
3
1
2
0.5
0 2 1.5 1
2.5 0 1
4
2
5 53
3.4
1. 3
2
2
−0.1 −0.1
4
2
1
−0.3 −0.3
−0.4 −0.4
−0.4 −0.2 0 0.2 0.4 −0.4 −0.2 0 0.2 0.4
(e) (f)
Figure 6.15: b2 phase variation of the new comb generator prototype vs.
fundamental output load Γ at: (a) 2nd harmonic, (b) 3rd harmonic, (c)
4th harmonic, (d) 5th harmonic, (e) 6th harmonic, (f) 7th harmonic.
6.2 Waveform reconstruction on the mixed-signal load-pull
system 95
0.3 0.3
0.6
0.2 0.4 2 0.2 0.24
1. 1 0.
0.6
0..86
0
0.81 1
0.4
0.1 8 0.1
0. 0.6 0.04.61
10.8
1.2
1
0.6
2
4
0 0 2 0.2
1.
0.2 11. 0.8
0.4 0.2
−0.1 0.6 −0.1 0.4 .60.81
0.4 0 111.4
0.2
1.2 .2
1.4 0.8 .6
−0.2 −0.2
1
−0.3 −0.3
−0.4 −0.4
−0.4 −0.2 0 0.2 0.4 −0.4 −0.2 0 0.2 0.4
(a) (b)
th th
4 harmonic 5 harmonic
0.4 0.4
0.3 0.3
0.5 1
0.2 0.2 .5
12
0.5
1
0.1 0.1
0.5
0.5 0.5
1 0 1.5
1 .5 0 1 1
1
.5 3
2 1.5
1.5
0 0 2.5
5
1.
1 2
1
−0.1 −0.1
11.5
1
1.5
2
1.5
−0.3 −0.3
−0.4 −0.4
−0.4 −0.2 0 0.2 0.4 −0.4 −0.2 0 0.2 0.4
(c) (d)
0.3 0.3
1 1.25
2
1
00.5.5
1
5
0.1 0.1
1.
1
1.5
5 5 1 1. 2
2 0. 1.2 5 5
4 1.
23.5
0 2 0 3 1
2
122.5
5
1
1.
−0.2 −0.2 2
4
1.5
23.52
2
2 1
−0.3 −0.3
−0.4 −0.4
−0.4 −0.2 0 0.2 0.4 −0.4 −0.2 0 0.2 0.4
(e) (f)
Figure 6.16: b2 phase variation of the new comb generator vs. 2nd har-
monic output load Γ at: (a) 2nd harmonic, (b) 3rd harmonic, (c) 4th
harmonic, (d) 5th harmonic, (e) 6th harmonic, (f) 7th harmonic.
96 Measurement of Time-Domain Waveforms
80
3
70 2.5
60 2
v2 [V]
1.5
50
i2 [A]
1
40
0.5
30
0
20 −0.5
0 0.2 0.4 0.6 0.8 1 20 30 40 50 60 70 80
t [s] −9
x 10 v2 [V]
(a) (b)
3.5
2.5
20 Ω 30 Ω 40 Ω
2
i2 [A]
15 Ω 25 Ω 35 Ω
1.5
0.5
0
0 0.2 0.4 0.6 0.8 1
t [s] −9
x 10
(c) (d)
120
100
80
v2 [V]
60
40
20
0
0 0.2 0.4 0.6 0.8 1
t [s] −9
x 10
(a)
3
i2 [A]
−1
0 0.2 0.4 0.6 0.8 1
t [s] −9
x 10
(b)
Figure 6.19: (a) Voltage (VDS ) and (b) current (ID ) time-domain wave-
forms measured as a function of power at the output current generator
of the DUT for a fundamental loading termination of 25 Ω. The source
available power (PAV S ) is swept from 15 dBm to 37 dBm in 1 dB steps.
Custom BiasTee
To LF coupler
PNA-X
Source 1 Source 2
OUT 1
OUT 1
Rear panel
R1 R3 R4 R2
A C D B
Figure 6.20: Simplified block diagram of the system, where HPR is the
harmonic phase reference (e.g., square-wave amplifier), and cHPR is the
calibration harmonic phase reference (e.g., Agilent U9391C).
6.3 Waveform reconstruction for closely–spaced multi-tone
signals 101
Figure 6.21: Measured S21 and S31 of the custom bias-Tee representing
the RF path and baseband path, respectively.
frequency phase relations at the desired reference plane of the device under
test (DUT).
The hardware of the PNA-X is designed to work in the frequency band
ranging from 10 MHz to 26.5 GHz. When exciting a non-linear device with a
multi-tone signal, intermodulation products will appear in the high frequency
band (e.g., 2f1 , 3f1 , etc.) as well as in the baseband (e.g., f2 − f1 , etc.).
When the stimulus is composed of closely-spaced multi-tones, the baseband
components will appear at frequencies below the PNA-X detection band. For
this reason, external hardware is required to properly sample these waves in
order to support correct waveform reconstruction for the overall signal. In the
proposed setup the a and b baseband waves (e.g., below 10 MHz) are sensed
using external low-frequency bridges (e.g., Minicircuits ZFDC-10-6).
The resulting coupled waves are then routed directly to the wideband A/D
converter (e.g., 16 MHz) of the PNA-X. Custom designed bias-Tees are em-
ployed to separate the low-frequency components (e.g., below 100 MHz) from
the high frequency components (e.g., above 1.5 GHz), because the low fre-
quency bridges would present too high loss for the high frequency components
above 10 MHz. Fig. 6.21 shows the measured transmission parameters for
both the RF and baseband path of the bias-Tees.
102 Measurement of Time-Domain Waveforms
f1 f2
DC f1+f2
Frequency
Figure 6.22: Frequency list selected by the user, based on harmonic order
(H) and intermodulation order (M). In this figure H=3 and M=3.
1. First step
A single-tone at baseband frequency fin = f2 − f1 is applied to the calibra-
tion harmonic phase reference (cHPR), generating a homogeneous grid at the
output of the cHPR (Fig. 6.23). The first step of the phase calibration (e.g.,
the extraction of the phase component of the i10 error term [63]) is performed
up to the highest tone in the fundamental band of the selected frequency grid
(e.g., 2f2 − f1 in Fig. 6.22). Note that the phase calibration will give cor-
rect results only for the frequency components which are also generated by
the measurement harmonic phase reference generator (e.g., spectral compo-
nents of Fig. 6.22). For all of the other frequency components, the data are
discarded.
6.3 Waveform reconstruction for closely–spaced multi-tone
signals 103
Frequency
Figure 6.23: Frequency grid generated by the calibration HPR during the
first step of the calibration procedure.
2f1-f2 f1 f2 2f2-f1
fin Frequency
Df*2 Df*n
angle i10 second step
f1 2f1 nf1
Frequency
Figure 6.24: Alignment of the error terms measured in the second step to
the correspondent frequency components measured during the first step
of the phase calibration procedure.
104 Measurement of Time-Domain Waveforms
2. Second step
During the second step, a single-tone is applied to the input of the cHPR at
frequency f1 in the fundamental frequency band. At this stage, the phase of
the error terms (e.g., i10 ) is acquired for all of the harmonic components of f1 ,
up to the highest frequency selected in the user-defined frequency grid (e.g.,
3rd harmonic in Fig. 6.22), with the upper limit being the bandwidth of the
instrument.
The phase of the error terms acquired during this second step will not be
consistent with the result of the first step (Fig. 6.24), since a new tone is
now applied to the phase reference that is not coherent with the stimulus of
the first step (e.g., f2 − f1 ). By comparing the phase of the i10 error term
obtained in the second step with the one measured in the first step, the phase
offset to be applied to all the error terms measured in the second step can be
calculated, as shown in Fig. 6.24.
This procedure is repeated when applying a new tone at f2 . In this way,
the phases of the error terms measured in the successive steps can be “aligned”
with the ones measured during the first calibration step. The result is a con-
sistent set of error terms for all the frequency components of interest. Note
that the phases of the error terms for the frequency components that cannot
be generated as an integer multiple of the tones present in the fundamental
frequency band (e.g., f1 +f2 ), are obtained via interpolation to the neighboring
components. This completes the calibration procedure, allowing the measure-
ment of corrected power waves at the input and output of the DUT for the
entire (user-defined) frequency grid, including the baseband components.
2. Lower phase noise and jitter of the spectral components generated, due
to the lower multiplication order.
(a)
(b)
sources which phase drift over time. In the homogeneous grid measurement
setup, the 10 MHz reference of the PNA-X is used to drive the Agilent U9391C
phase reference, while the two internal sources of the PNA-X drive the DUT.
Note that the 10 MHz reference is phase locked with the internal signal sources
and has very low phase noise, providing maximum output power for all of the
harmonics generated by the U9391C phase reference.
By measuring the phase of the input/output waves (e.g., b1 /b2 ) 100 times,
the standard deviation is computed and a Gaussian distribution (with nor-
malized mean and amplitude) is used to compare the data. Fig. 6.25 shows
the results of the two methods, assuming a ∆f of 10 MHz and the f1 tone at
2 GHz in both cases. In Fig. 6.25a, the measured phase variation is shown
at the 5th harmonic (10 GHz), while Fig 6.25b shows the measured results for
the 7th harmonic (14 GHz). As shown in these figures, the phase variation
of the homogeneous grid method rises with increasing harmonic order, while
the proposed method provides a constant phase deviation. Finally, the phase
deviation is shown in Fig. 6.26 for the new method when measuring the 5th
harmonic of a two-tone signal with f1 equal to 2 GHz when the applied tone
6.3 Waveform reconstruction for closely–spaced multi-tone
signals 107
0.3
0.2
Voltage, V
0.1
−0.1
−0.2
0 0.5 1 1.5 2
Time, s −7
x 10
(a)
0.3
0.2
Voltage, V
0.1
−0.1
−0.2
0 0.5 1 1.5 2
Time, s x 10
−7
(b)
Figure 6.27: time-domain waveform of the b2 wave measured for the square
wave LO amplifier, a) multi-step calibration procedure b) PNA-X hard-
ware and conventional calibration.
108 Measurement of Time-Domain Waveforms
spacing is 10 MHz, 1.25 MHz and 100 kHz. As expected, the accuracy of
the phase measurement does not reduce at narrower tone spacings, as would
happen for the homogeneous grid method.
6.4 Conclusions
In this chapter extensions to the mixed-signal active load-pull system that al-
low time-domain waveform measurements have been discussed. The require-
ments and the performance of commercially-available phase references and of
a newly-developed prototype have been presented. Finally, an approach for
time-domain waveform analysis dedicated to multi-tone signals closely spaced
in frequency has been introduced.
Chapter 7
Application Examples
In the previous chapters of this thesis, an innovative load-pull system has been
proposed. In this chapter, several examples of applications that highlight the
most unique capabilities of the system are presented.
109
110 Application Examples
BB: Baseband
B2nd: 2nd harmonic band
IML IMU
BB Ch B2nd
f
2CB
indirect mixing Linearized modulated
signal
Experimental proof of this theory is given in [105], where the active harmonic
load-pull system described in this thesis work is used to measure the linearity
of SiGe and GaAs heterojunction bipolar transistors (HBT) under two-tone
stimulus. The fundamental and 2nd harmonic source and load impedances are
actively controlled, and the baseband source and load impedance are controlled
passively using real impedances ranging from 0.5 to 2,048 Ω with a resistive
switch bank [64], as illustrated in Fig. 7.2.
Four different HBT devices are measured, as shown in table 7.1.
7.1 Out-of-band linearity optimization 111
I/Q 0 I/Q 0
I 1
RF @ 2f0 RF @ 2f0
From From
LO LO
RF @ f0 RF @ f0
On-wafer configuration
0 0
DUT
Bias Bias
Tee Tee
Input section Output section
0 0
Passive BB Passive BB
impedance DC impedance DC
board board
Figure 7.2: The mixed-signal based active load-pull setup with passive
baseband control extension.
Firstly, the fundamental source and load impedances are optimized for
PAE under single-tone excitation. From two-tone measurements, the baseband
impedance, second harmonic impedance and base-emitter voltage were swept,
and the optimal combination of Zs,bb , Zs,2f and collector current were found.
Finally, Zs,bb , and Zs,2f are fixed to their optimal values, and the input power
and Vbe are swept. The output third-order intercept point (OIP3 ) contours
are plotted in Fig. 7.3 on the output power and quiescent current plane. The
practical application of this method becomes apparent if the bias point for
achieving optimum linearity is found from these plots. The resulting IM3
level versus PAE is shown in Fig. 7.4. In particular, Fig. 7.4a shows the IM3
vs. the two-tone PAE for the collector quiescent currents that result in the
highest OIP3 value. Fig. 7.4b depicts the IM3 vs. the two-tone PAE for the
112 Application Examples
quiescent current that provides the highest two-tone linearity near the 1-dB
compression point, which was close to 3 mA for all devices shown.
Note that to guarantee low uncertainties in the linearity measurement,
the harmonic distortion products generated by the system should be at least
18 dB below the harmonic distortion level of the DUT. For this reason, the
measurements reported are only possible thanks to the high dynamic range of
the system, and the capability to control the impedance for a wideband signal
in and out-of-band.
8 P−1dB P−1dB
10
18
Collector Quiescent Current (mA)
18
21
20
20
22
9
7
optimal linearity 24
locus 24 8 optimal linearity 22
locus
6 26
21
7 24
28 26
20
18
18
30
26
22
20
22
5 28
24
26 26 24
24 5 22
21
24 20
22 18 21
4 20 4
18 16
20
3 18
20
14
18
20 16
22
20
3 12
22
16 18
−2 0 2 4 6 8 −4 −2 0 2 4 6 8 10
Output Power (dBm) Output Power (dBm)
(a) (b)
~P−1dB P−1dB
16
22
9
26 2
18
Collector Quiescent Current (mA)
10 optimal linearity
18
20
30
20
24
22
24
8
26
9 locus
34
32
28 32
8 7 30
30
22
26
16
7 28
32
18
26 6
26 28
24
18
36
20
20
34
22
26 28
24
6 24 30 30
32
22 5 2
5 22 22 4
20 20 26
18 20 4
4
16
16 18
18
20 22 24
20
22
18
18
24
3 14 16 3 16
−6 −4 −2 0 2 4 6 0 2 4 6 8
Output Power (dBm) Output Power (dBm)
(c) (d)
Figure 7.3: Measured OIP3 hi (dBm) contours on the output power and
quiescent current plane for: (a) device A, (b) device B, (c) device C and
(d) device D from Table 7.1. The 1-dB compression points and optimal
linearity loci are indicated. Courtesy of K. Buisman [105].
7.1 Out-of-band linearity optimization 113
0 0
−10 −10
−20 −20
IM3hi (dBc)
IM3hi (dBc)
−30 −30
−40 −40
(a) (b)
Figure 7.4: IM3 (dBc) versus two-tone PAE (%) for: (a) the bias current
for which the highest OIP3 was measured, and (b) the bias current for
which the highest two-tone linearity was measured near the 1 dB com-
pression point, which was 3 mA for all devices measured. Courtesy of K.
Buisman [105].
Trigger
and Clock
I Q I Q I Q I Q
2f0 f0 f0 2f0
I1 V1 V2 I2
a1 b1 b2 a2
LO LO LO LO
0 0
0 0
I1 V1 V2 I2
Baseband current and voltage sensing
1 Optimum f load Γ
0
0.6
0.4
0.2
−0.2
−0.4
−0.6
−0.8
−1
−1 −0.5 0 0.5 1
Figure 7.6: Optimum load fundamental and 2nd harmonic reflection coef-
ficients providing maximum PAE under pulsed-RF single-tone excitation.
correctly.
The measurement of the baseband impedance seen at the DUT reference
plane is obtained by acquiring the voltage and the current from DC to 48 MHz
using a current and voltage probe and analog-to-digital converters with a
sampling speed of 100 MS/s. To obtain a calibrated measurement of the
source and load baseband impedances seen by the DUT at the reference plane,
a conventional SOL calibration is performed. For this purpose, short, open
and load calibration standards compatible with the device test fixture have
been designed, implemented and modelled.
To illustrate the importance of controlling the baseband impedance prop-
erly when characterizing active devices with realistic communication signals,
a NXP GEN7 LDMOS device with Wg = 10 mm is measured with a 20 MHz
wide LTE signal using a peak-to-average ratio of 9.8 dB. Initially, the funda-
mental and 2nd harmonic load impedances yielding maximum PAE are found
by performing harmonic load-pull measurements under pulsed-RF, single-tone
excitation. The optimum values for the load reflection coefficients are depicted
in Fig. 7.6. Afterwards, the DUT is measured with an LTE input signal for the
(previously found) fundamental and harmonic conditions, and its ACPR and
EVM performance are evaluated as a function of the baseband load impedance.
Three different cases are evaluated in the following experiments, which
116 Application Examples
0.8 30 MHz
0.6
RF choke of
0.4 150 nH
0.2
0
10 kHz
−0.2
Passive baseband
−0.4 impedance
−0.8
−1
−1 −0.5 0 0.5 1
Figure 7.7: Measured impedance at the DUT reference plane for three
different cases.
• Baseband injection signal at load side set such that it mimics a feed
inductance of 150 nH in the DUT bias line.
The baseband impedance provided to the load plane at the DUT measured
up to the 30 MHz controlled bandwidth is illustrated in Fig. 7.7. The results
reported in Fig. 7.8 clearly highlight that proper termination of the baseband
load impedance to a short condition improves both the out-of-band (ACPR)
and the in-band (EVM) linearity of the active device. In particular an im-
provement of 10 dBc in the ACPR and 5 % in EVM is observed for an average
output power of 35 dBm. Note that the system is capable of providing any
arbitrary, user-defined impedance vs. frequency in the fundamental, 2nd har-
monic and baseband frequency bands. In these experiments, a bandwidth of
60 MHz is controlled at f0 to include the 20 MHz signal span and its adjacent
7.1 Out-of-band linearity optimization 117
−15 30
Passive BB Passive BB
BB short BB short
−20 BB 150 nH inductance
BB 150 nH inductance 25
−25
ACPR low [dBc]
20
−30 EVM [%]
−35 15
−40
10
−45
−50 5
25 30 35 40 25 30 35 40
P [dBm] P OUT [dBm]
OUT
(a) (b)
0
Passive BB
BB short
BB 150 nH inductance
Power Spectral Density [dBm/Hz]
−10
−20
−30
−40
−50
−60
1.97 1.98 1.99 2 2.01 2.02 2.03
Frequency [GHz]
(c)
Figure 7.8: Measured: (a) ACPR (dBc) low, (b) EVM (%), and (c) power
spectral density (dBm/Hz) at 30 dBm POU T for three different baseband
load impedance terminations.
118 Application Examples
1.1
Magnitude
0.9
0 5 10 15 20 25 30
200
Angle, deg
180
160
0 5 10 15 20 25 30
Frequency, MHz
Figure 7.9: Magnitude and phase (deg) of the controlled baseband load
reflection coefficient vs. frequency (MHz) at the DUT reference plane.
0.8
0.6
8.2
8.4
0.4
8.1
8.
3
0.2
8.
2
0
8
8.1
−0.2
−0.4 7.9
8
7.
8
−0.6 7.
7
−0.8
−1
−1 −0.5 0 0.5 1
channels. This will result in a 120 MHz bandwidth controlled at the 2nd har-
monic, while the baseband impedance is controlled over a 30 MHz frequency
span. An example is shown in Fig. 7.9, where the magnitude and angle of
the baseband load reflection coefficient vs. frequency is plotted. It can be
seen that an excellent control can be achieved throughout the measurement
bandwidth.
Finally, Fig. 7.10 shows contours of EVM for a 20 MHz LTE signal at
a constant average output power of 31 dBm as a function of the baseband
load impedance, plotted in a Smith chart with a reference impedance of 20 Ω.
It is clear that both the imaginary as well as the real parts of the baseband
impedance at the DUT reference plane affect the linearity of the active device,
and therefore require careful attention in both circuit implementation and
characterization.
65
60
60
50
55
40
PAE [%]
PAE [%]
50
30
45
20
10 40
0 35
0 50 100 150 200 250 300 350 100 150 200 250 300 350
PL [W] PL [W]
(a) (b)
PAE@GainCompression
0.1 PL_f0@GainCompression
−0.1
0 160
14
−0.2 180
55
50
200
−0.3 220
240
45
−0.4
60
−0.5
−0.6 300
280
40
35260
−0.7
−0.8
(c)
Figure 7.11: (a) Power-added efficiency (%) versus output power (W),
(b) power-added efficiency (%) versus output power (W) at 3 dB of gain
compression, and (c) POU T (W) and power added efficiency (%) contours
at 3 dB of gain compression for an NXP 7th generation LDMOS 200 W-
rated transistor.
7.2 High-power device measurements for base-station
applications 121
65
60
60
50
55
40
PAE [%]
PAE [%]
50
30
45
20
10 40
0 35
0 100 200 300 400 500 200 250 300 350 400 450 500
PL [W] PL [W]
(a) (b)
0.4
PAE@GainCompression
PL_f0@GainCompression
0.3
0.2
450
0.1
0
40
−0.1
−0.2
−0.3
60 400
45
25
55
50
−0.4
0
350
30
0
−0.5
(c)
Figure 7.12: (a) Power-added efficiency (%) versus output power (W),
(b) power-added efficiency (%) versus output power (W) at 3 dB of gain
compression, and (c) POU T (W) and power-added efficiency (%) contours
at 3 dB of gain compression for an NXP 8th generation LDMOS 360 W-
rated transistor.
122 Application Examples
(a) (b)
Figure 7.13: (a) Fundamental load Γ sweep, (b) second harmonic load Γ
sweep on a Smith chart (10 Ω reference impedance).
aged NXP GaN HEMT with a gate width of 12 mm, already used in Section
6.2.3. Also, the device in this example is measured at a frequency of 2 GHz
and is biased at a VDD of 50 V with a quiescent current IDQ of 150 mA. Once
again, the simple model shown in Fig. 6.17 has been used to de-embed the
voltage and current waveforms at the level of the internal drain node.
In order to find the optimum tuning impedances yielding the highest effi-
ciency, the fundamental and second harmonic load Γ are swept simultaneously
in the range depicted in Fig. 7.13.
The results are shown in Fig. 7.14, where the efficiency versus output
power is reported at a gain compression level of 3 dB. The gain is plotted
versus output power and the dynamic load line, measured at 35 dBm and at
45 dBm of output power, is depicted for every combination of fundamental
and 2nd harmonic load Γ.
It may be noted from Fig. 7.14c that the device is operating into class J
for the combination of Γ highlighted in yellow. This is more clearly visible
by looking at the voltage and current waveforms versus time as a function of
power, shown in Fig. 7.15a and 7.15c. Here, the peak voltage approaches the
theoretical value of 2.92 · VDD for class J very closely.
In Fig. 7.14a, however, it is easy to identify that the highest efficiency
reaches 80 % for the fundamental and 2nd harmonic Γ highlighted in red. In
124 Application Examples
80 24
22
70
20
60
Gain [dB]
18
Eff [%]
50
16
40
14
30 12
20 10
0 20 40 60 80 20 25 30 35 40 45 50
P [W] P [dBm]
OUT OUT
(a) (b)
6
2
5
1.5 4
3
1
i2 [A]
i2 [A]
0.5 1
0
0
−1
−0.5 −2
0 20 40 60 80 100 −50 0 50 100 150
v2 [V] v2 [V]
(c) (d)
Figure 7.14: (a) Efficiency (%) versus output power (W) at 3 dB of gain
compression, (b) power gain (dB) versus output power (dBm), (c) dy-
namic load line measured at POU T = 35 dBm and (d) dynamic load line
measured at POU T = 45 dBm plotted for every combination of funda-
mental and 2nd harmonic load Γ. The highighted points correspond to
the combination of fundamental and 2nd harmonic load Γ which are also
highlighted in Fig. 7.13.
7.3 Device characterization for high efficiency power amplifier
design 125
150 140
120
100 100
80
v2 [V]
v2 [V]
50 60
40
0 20
−50 −20
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
t [s] x 10
−9 t [s] x 10
−9
(a) (b)
5 5
4 4
3 3
2 2
i2 [A]
i2 [A]
1 1
0 0
−1 −1
−2 −2
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
t [s] −9 t [s] −9
x 10 x 10
(c) (d)
Figure 7.15: (a) Voltage (VDS ) and (c) current (ID ) time-domain wave-
form measured as a function of power at the output current generator of
the DUT for the combination of fundamental and 2nd harmonic load Γ
highlighted in yellow in Fig. 7.13. (b) Voltage (VDS ) and (d) current (ID )
time-domain waveform measured as a function of power at the output cur-
rent generator of the DUT for the combination of fundamental and 2nd
harmonic load Γ highlighted in red.
126 Application Examples
this case, it is not easy to identify what class the device is operating in from the
load-line, while from Fig. 7.15b and 7.15d it can be seen that the peak voltage
for this particular point is slightly lower than for pure class-J operation.
This example clearly highlights that the ability to measure time-domain
waveforms gives powerful insight into the device behavior, while providing
valuable information such as peak voltage and current which can be helpful for
considerations such as ruggedness. However, the high speed of the system that
allows to quickly search through a wide range of fundamental and harmonic
impedances proves to be the real added value when using such a large-signal
measurement system for power amplifier design.
7.4 Conclusions
Several examples have been presented which highlight the most important
features of the mixed-signal load-pull system developed during this thesis work.
Experimental results obtained using this setup show that active device lin-
earity, when operating with wideband-modulated signals (e.g., LTE) is severely
influenced by the baseband impedance offered to the device under test. The
ability to control these baseband impedances vs. frequency allows the user
to optimize the linearity of active devices for complex modulated signals, and
to troubleshoot memory effects due to the baseband impedance terminations
provided by a realistic matching network implementation as well.
Furthermore, the system is flexible enough to serve many different kinds of
applications, ranging from low-power, high-linearity measurements to the char-
acterization of extremely high-power devices for base-stations applications.
Finally, the high speed of the system coupled with the ability to measure
time-domain waveforms at the device intrinsic current generator plane, makes
the mixed-signal load-pull system a valuable tool for high-efficiency power
amplifier design.
Chapter 8
8.1 Conclusions
The focus of this thesis is the creation of new characterization techniques
that support technology development, (compact) model device extraction /
validation and design activities, that address the specific needs of advanced
wideband communication systems. Since the transmit path in these next-
generation wireless systems presents the largest challenges in terms of effi-
ciency, linearity and bandwidth, special attention is given to extending the
current state-of-the-art in pulsed and large-signal characterization techniques
in terms of duty-cycle, accuracy, power, impedance control, bandwidth, mea-
surement speed and functionality.
One of the key inventions in this work is a novel time gating and data
alignment approach that, when applied in an isothermal measurement system,
enables pulsed-DC and RF measurements down to 200 ns with an excellent
dynamic range which is independent from duty-cycle. As such, the realized
setup facilitates the characterization of RF / microwave devices under truly
isothermal conditions. This functionality gives significant advantages in defin-
ing, extracting and verifying (compact) models for various RF power devices
that typically suffer severely from bias and operation-dependent self-heating
and/or trapping effects.
The second and dominant part of this thesis relates to the development
and realization of a revolutionary active harmonic load-pull system. The sys-
tem capabilities are summarized and compared to state-of-the-art load-pull
systems in Fig. 8.1. This system provides, for the first time, the capability
to synthesize truly arbitrary source and loading conditions vs. frequency at
the fundamental and harmonic frequencies over a large bandwidth (currently
127
128 Conclusions and Future Work
120 MHz). This setup enables testing of active devices under realistic (circuit-
like) conditions with wideband modulated signals. The ability to control up
to three harmonics at high power levels with an extremely high speed (up to
1,000 measurement points per minute), dramatically enhances the process of
developing new transistor technologies and their application in very efficient
and linear power-amplifiers. Moreover, the option to measure time-domain
voltage and current waveforms can provide significant insight into the actual
device behavior, which benefits power amplifier design, ruggedness evaluation,
as well as (database) model extraction and validation. The usefulness of the
system has been demonstrated by applying this newly developed load-pull
characterization system to several relevant application examples.
All measurement setups and techniques introduced in this thesis are based
on the innovative mixed-signal measurement concept, which replaces tradi-
tional analog solutions with digital data acquisition and digital signal gener-
ation. This approach, sometimes also referred to as “synthetic” or “software-
defined” instruments [106], provides much higher flexibility, functionality, per-
formance and speed in many different applications when compared to tradi-
tional techniques. It is expected that mixed-signal systems will change the
current landscape of RF/microwave characterization.
curate models for circuit simulation. In a fast growing industry like telecom-
munication, where time-to-market is extremely important, enabling the rapid
and simple extraction of new device models directly from measurements has
received significant interest. For this reason, significant effort has been put
into the development of behavioral models in recent years, such as the poly-
harmonic distortion (PHD) model [108], which enable the extraction of device
models directly from large-signal measurements.
These models employ the measurement of time-domain voltage and cur-
rent waveforms under the different boundary conditions which affect the DUT
performance in order to create a database describing the behavior of the DUT.
For this reason, they require the device to be measured throughout the multidi-
mensional space of all of the different tunable parameters, such as fundamental
and harmonic source and load impedance, frequency, bias, etc.
Thanks to its very high measurement speed of up to 1,000 measurement
point per minute, the load-pull system described in this thesis can prove to
be an extremely valuable asset for the extraction of these type of models.
Furthermore, the development of additional measurement techniques to extend
their range of validity (e.g., to include memory effects) is also a topic of interest.
In terms of compact modeling, an issue is represented by the continuous
scaling in device size which is necessary to achieve higher cut-off frequencies
[109]. For this reason, the thermal resistance of some high-frequency devices
can reach values as high as 10,000 K/W [110,111]. To model these new device
generations, faster pulsed-DC measurements than what is currently available
are necessary to achieve truly isothermal conditions.
Conventional pulsed bias approaches employ analog pulsers to provide cur-
rent drive and to deliver the required voltage shape to the DUT. Nevertheless,
these methods cannot compensate for the distorting effect on the voltage wave-
form, arising from the interconnects and parasitic loading, due to the analog
nature of the driver. Therefore the minimum DC pulse widths achievable with
conventional pulsers are severely limited (to approximately 200 - 300 ns).
A more elaborate approach, based on a mixed-signal technique, can achieve
significantly shorter DC pulse widths (e.g., 10 ns or lower) by using a high-
speed AWG and a DC-coupled driver amplifier to generate the DC pulse, a
high-speed A/D converter to measure the calibrated DC pulse at the DUT,
and an iterative approach to optimize the DC pulse provided to the DUT [112].
However, in order for such an approach to be successful, the design of a high-
power DC-coupled amplifier, capable of providing high voltage and current
(e.g., 200 V and 30 A) with a wide bandwidth (e.g., 100 MHz), is required.
132 Conclusions and Future Work
[1] http://newsroom.cisco.com/release/1114955.
[2] http://www.cisco.com/en/US/solutions/collateral/ns341/ns525/
ns537/ns705/ns827/white paper c11-520862.html.
133
134 BIBLIOGRAPHY
[10] LAN/MAN standards Committee and others, “Part 11: Wireless LAN
medium access control (MAC) and physical layer (PHY) specifications,”
IEEE-SA Standards Board, 2003.
[46] G. Bava, U. Pisani, and V. Pozzolo, “Active load technique for load-pull
characterisation at microwave frequencies,” Electronics Letters, vol. 18,
no. 4, pp. 178 –180, Feb. 1982.
[54] B. Bunz and G. Kompa, “Active load pull with fourth harmonic tuning
based on an IQ modulator concept,” in Microwave Conference, 2003.
33rd European, Oct. 2003, pp. 359 –361.
[61] D. Poulin, J. Mahon, and J.-P. Lanteri, “A high power on-wafer pulsed
active load pull system,” Microwave Theory and Techniques, IEEE
Transactions on, vol. 40, no. 12, pp. 2412 –2417, Dec. 1992.
[67] D. Hartskeerl, I. Volokhine, and M. Spirito, “On the optimum 2nd har-
monic source and load impedances for the efficiency-linearity trade-off of
RF LDMOS power amplifiers,” in Radio Frequency integrated Circuits
(RFIC) Symposium, 2005. Digest of Papers. 2005 IEEE, June 2005, pp.
447 – 450.
[96] http://www.aeroflex.com/ams/metelics/micro-metelics-prods-comb.
cfm.
[102] J. Pedro and N. De Carvalho, “On the use of multitone techniques for
assessing RF components’ intermodulation distortion,” Microwave The-
ory and Techniques, IEEE Transactions on, vol. 47, no. 12, pp. 2393
–2402, dec 1999.
[104] M. P. van der Heijden, “RF amplifier design techniques for linearity
and dynamic range,” Ph.D. dissertation, Delft University of Technology,
2005.
[107] L. Galatro, M. Marchetti, and M. Spirito, “60 GHz mixed signal active
load-pull system for millimeter wave devices characterization,” in Mi-
crowave Measurement Symposium (ARFTG), 2012 80th ARFTG, 2012,
pp. 1–6.
145
146 Summary
pulses as short as 200 ns, while featuring a very high dynamic range (≈ 85 dB)
under pulsed-RF conditions, which is independent on the duty-cycle. The sys-
tem performance is discussed in detail through a set of benchmarks, and some
examples on isothermal active device characterization are provided.
The second and dominant part of this thesis introduces a revolutionary
active harmonic load-pull approach. Load-pull device characterization is fun-
damental to all activities related to PA design and PA development, from tech-
nology development, to model extraction and validation, to the actual power
amplifier design. For this reason Chapter 3 reviews conventional passive and
active source and load-pull architectures, and discusses their basic limitations,
with particular attention to the problems arising when characterizing devices
with wideband complex modulated signals. Moreover, the requirements of
active load-pull systems to perform high power measurements with complex
modulated signals are also explained.
To solve the problems of conventional load-pull systems when dealing with
wideband modulated signals, a novel active harmonic load-pull system based
on a mixed-signal approach is described in detail in Chapter 4. The system
developed during this thesis work enables the measurement of active devices
up to 120 MHz of modulation bandwidth, and allows arbitrary control of the
reflection coefficient in this band. Measurement data highlighting the system
performance, and measurement results on active devices are presented.
To enhance the process of developing new transistor technologies and
their application in very efficient and linear PAs, in Chapter 5, a new ap-
proach for enabling high-speed multidimensional source and load-pull param-
eter sweeps is introduced. The method described allows any combination of
multiple parameters (e.g., input power and/or fundamental and harmonic load
impedance) to be swept, at a very high speed, while maintaining all other pa-
rameters (e.g., second harmonic source impedance) accurately controlled to a
user-defined value. Moreover, several measurements are reported, with par-
ticular emphasis on the high-power capabilities of the system, both in CW as
well as under modulated signal excitations.
The option to measure time-domain voltage and current waveforms can
provide significant insight into the actual device behavior, which benefits to
power amplifier design, ruggedness evaluation, as well as to (database) model
extraction and validation. In Chapter 6 the basic theory behind the mea-
surement of high frequency time-domain voltage and current waveforms at the
device reference planes are discussed, and an extension to the mixed-signal
load-pull system described in the previous chapters is presented, with particu-
lar attention on the requirements of the calibration device used for the system
calibration. Furthermore an approach for time-domain waveform analysis of
Summary 147
149
150 Samenvatting
Journal Papers
M. Marchetti, “Mixed–signal active load pull: the fast track to 3G
and 4G amplifiers,” Microwave Journal, pp. 108–118, Sep. 2010.
C. Huang, P. J. Zampardi, K. Buisman, C. Cismaru, M. Sun, K. Stevens,
J. Fu, M. Marchetti and L. C. N. de Vreede, “A GaAs junction varac-
tor with a continuously tunable range of 9 : 1 and an OIP3 of 57 dBm,”
IEEE Electron Device Letters, vol. 31, no. 2, pp. 108–110, Feb. 2010.
153
154 List of Publications
Conference Papers
A. Kumar Manjanna, M. Marchetti, k. Buisman, M. Spirito, M. J.
Pelk and L. C. N. de Vreede, “Device characterization for LTE applica-
tions with wideband baseband, fundamental and harmonic impedance
control,” in Proc. 43rd European Microwave Conference, 2013, Nurem-
berg, Germany, Oct. 2013.
K. Buisman, M. Marchetti, M. P. van der Heijden, P. J. Zampardi
and L. C. N. de Vreede, “Evaluation of HBT device linearity using
advanced measurement techniques,” in Proc. 43rd European Microwave
Conference, 2013, Nuremberg, Germany, Oct. 2013.
Patents
M. Marchetti, M. J. Pelk, L. C. N. de Vreede, “Open loop load pull ar-
rangement with determination of injections signals,” World Intellectual
Property Organization Patent No. 2009131444, 30 Oct. 2009.
Book Chapters
M. Spirito and M. Marchetti, “Broadband large signal measurements
for linearity optimization,” in Modern RF and Microwave Measurement
Techniques, V. Teppati, A. Ferrero, and M. Sayed, Eds. Cambridge
University Press, 2013, ch. 14.
157
158 Acknowledgments
First and foremost I have to thank Michele Squillante for too many things
to mention. To name a few, thank you for developing big part of the current
load-pull software, for sharing most of the burden and many late evenings of
work without ever complaining, for the sausage with beans and the occasional
mozzarella, and most of all for being a good friend. I would also like to ac-
knowledge the other two members of the current Anteverta team for their hard
work, commitment and good spirit, Ajay Kumar Manjanna and Giampiero “’o
’mericano” Esposito (who has joined us and found the new continent). This
is the proper moment to also thank Marco Pelk for being able to solve most
of the practical technical issues we encountered and for knowing the answer
to some of the most difficult questions.
I also want to express my appreciation for Han Oey, Paul Althuis and
Ronald Gelderblom and all the other people of the Delft Valorisation Center
for their guidance during the start-up of Anteverta and for their continuous
support.
A special acknowledgment goes to all the people at NXP that have be-
lieved in us and in our system when we were still writing lines of code to
run a measurement. Particularly, I am grateful to Rob Heeres for sharing his
knowledge, for the continuous feedback, for his patience with the early ver-
sions of the software and for being one of the most sincere supporters of our
system. I would also like to thank Steven Theeuwen for generously providing
many test devices for us to write papers, test new software and sometimes to
destroy. A special acknowledgment also goes to Dr. Mark van der Heijden for
many discussions and for even bringing customers to us. Furthermore, I have
to thank Dr. Martino Lorenzini for the many measurement sessions and for
providing the GaN devices, Rob Bubeck, Angelo Andres, Petra Hammes and
Lex Harm for their feedback and for many useful discussions.
I would also like to thank all the people at Maury Microwave for helping
us in turning the research shown in this thesis into a commercial product. In
particular, I have to thank Steve Dudkiewicz for the many advices and almost
daily Skype calls.
Carrying out my Ph.D. first in the HiTec group, then in the ELCA group,
has proven to be a very enjoyable experience. I have had the pleasure to work
with some of the brightest and most interesting people I know. First of all I am
indebted to Dr. Koen Buisman for being always the first user and debugger
of my measurement systems, for answering my many questions about device
physics, for translating the summary and propositions of this thesis in Dutch,
and for a whole lot of other things. I have also had the pleasure to work
closely with Dr. Edmund Neo, Dr. Jawad Qureshi and Dr. Cong Huang to
whom goes my gratitude. Furthermore, for the enjoyable atmosphere and the
Acknowledgments 159
many fun times I would like to acknowledge my colleagues Morteza Alavi, Rui
Hou, David Calvillo Cortes, Luca Galatro, Gennaro Gentile, Yi Zhao, Akshay
Visweswaran. I would also like to thank Atef Akhnouk for his help in the
measurement room, for his phone calls to the Dutch tax office on my behalf,
and for continuously reminding me to be careful. I would also like to thank
Marysia Lagendijk, Marian Roozenburg, Bianca Knot and Marion de Vlieger
for their support.
Being far from home is never easy, therefore I would like to thank all
the friends that have made me feel at home both in Delft and in Eindhoven:
Luigi “van der Appels” Mele, Agata Ŝakić who shares my same chocolate
addiction, Riccardo “Materassi” Donatantonio, Luis Alberto Cusati nonos-
tante non sia il mio presidente, Roberto “Johnny” Amabile with whom I’ll
always be happy to go surfing, Francesco Vitale and all his wooden spoons,
Salvatore “rum e cotechino” Russo my MATLAB GUI teacher, Fabio San-
tagata, Elina Iervolino, Alessandro “the wall” Baiano, Maria De Biase, An-
drea Ingenito and his wonderful light blue Vespa, Daniel Tajari Mofrad, Ben-
jamin Mimoun, Luigi La Spina, Giovanna Razzano, Francesco Sarubbi, Gian-
paolo Lorito, Theodoros Zoumpoulidis, Olindo Isabella, Joelle Olivet, Ghaz-
aleh Nazarian, Vladimir “the professor” Jovanovic, Yann “purpette” Civale,
Giuseppe Fiorentino, Bruno Morana, the Andricciola’s, the De Maio’s, the
Tripodi’s, Fabio Sebastiano the master of beer, Elisa “Dr. House” Buonanno,
Salvo Drago, Lisa de Vries, Alberto Fazzi, Laura Pirani, Gerard Villar Pique,
Muhammed Bolatkale, the Van der Weide’s, the “6 o’ clock heroes” football
team.
Thanks to my friends back at home (who are now spread throughout
the world) because when I am back at home it feels like I never left: Gia-
como “o’ So’” Sozio, Alessandro “l’avvocato” Caprio, Ciro Pileggio, Marco
”o’ lion’” Leone, Michele “sapone” Esposito, Francesco Miglino, Francesco
Liquido, Claudio “culo” Esposito, Andrea “o’ So’ grande” Sozio, Luca Russo,
Alfonso “pivell” Losco, Roberta “Enzo” Esposito.
Being myself an Italian man from the South, you should know that for us
(and probably not only for us), family is the most important thing (and no,
not in “the Godfather” kind of way). Truly, life without family is like a babà
(typical Neapolitan dessert) without rum.
I would like to thank all my relatives, my aunts and uncles, my parents
in law Carmela e Antonio Spella, Claudio “maestro spigola” Spella, Simona
“a milanese” De Toni, Francesca Spella, Giovanni Colacurcio e la piccola Gio-
vanna. Thanks to them all for their sustained encouragement and support.
I am grateful to my sister for bringing some mess (and a lot more fun) into
my super organized world while growing up together, for being my fashion
160 Acknowledgments
advisor, and for coming to visit me in the Netherlands more than I have gone
to visit her in Gorizia. I would also like to thank the new addition to the
family Salvo Pennisi, who some day will hopefully catch some fresh fish in
Palinuro when I am there.
I will always be grateful to my parents for raising me with love, for their
continuous encouragements, unconditional support and for making me what I
am today. Although we are geographically so distant now, we will always be
close to each other.
To conclude, I want to express my love and profound appreciation to my
wife Maristella. Thanks for lighting up my existence and for facing life with
me while supporting me every step of the way: without you no achievement
would be possible. Finally, thanks to my wonderful little son Matteo, because
you are my purpose of life and whatever you will grow up to be, you will
always be my biggest achievement.
About the Author
Mauro Marchetti was born in Naples, Italy, in 1981. He received his B.S. and
M.Sc. degrees (cum laude) in electrical engineering from the “Università degli
studi di Napoli Federico II”, Naples, Italy, in 2004 and 2006 respectively.
In 2006 he joined the Electronic Research Laboratory (ELCA) of Delft
University of Technology, The Netherlands, where he carried out his Ph.D.
research on the development and implementation of mixed-signal instrumen-
tation for large signal device characterization and modelling.
In 2010 he co-founded and was appointed CEO of Anteverta-mw B.V., a
spin-off company of Delft University providing mixed-signal harmonic load-
pull measurement systems.
161