ArticlePDF Available

Comparison of a 100-pF Capacitor With a 12 906-Ω Resistor Using a Digital Impedance Bridge

Authors:

Abstract and Figures

We tested a digital impedance bridge in a hybrid structure for comparison of a capacitor with a resistor where the impedance ratio was measured in two separate parts. The modulus of the impedance ratio was matched arbitrarily close to the input-to-output ratio, in magnitude, of a two-stage inductive voltage divider by adjusting the operating frequency of the bridge; the residual deviation between the two together with the phase factor of the impedance ratio was measured using a custom detection system based on a four-channel 24-bit digitizer. The ratio of the inductive voltage divider was calibrated, in situ, using a conventional four-arm bridge with two known capacitors. Fluctuations of the source voltages were largely removed through postprocessing of the digitized data, and the measurement results were limited by the digitizer error. We have achieved an overall bridge resolution and stability of 0.02 μF/F in 2 hours for measuring a 100 pF capacitor relative to a 12906 Ω resistor at 1233 Hz. The relative combined standard uncertainty (k = 1) is 0.13 μF/F, dominated by the digitizer error.
Content may be subject to copyright.
This work is licensed under a Creative Commons Attribution 4.0 License. For more information, see https://creativecommons.org/licenses/by/4.0/
This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/TIM.2021.3139709, IEEE
Transactions on Instrumentation and Measurement
> REPLACE THIS LINE WITH YOUR PAPER IDENTIFICATION NUMBER (DOUBLE-CLICK HERE TO EDIT) <
1
AbstractWe tested a digital impedance bridge in a hybrid
structure for comparison of a capacitor with a resistor where the
impedance ratio was measured in two separate parts. The modulus
of the impedance ratio was matched arbitrarily close to the input-
to-output ratio, in magnitude, of a two-stage inductive voltage
divider by adjusting the operating frequency of the bridge; the
residual deviation between the two together with the phase factor
of the impedance ratio was measured using a custom detection
system based on a four-channel 24-bit digitizer. The ratio of the
inductive voltage divider was calibrated, in situ, using a
conventional four-arm bridge with two known capacitors.
Fluctuations of the source voltages were largely removed through
postprocessing of the digitized data, and the measurement results
were limited by the digitizer error. We have achieved an overall
bridge resolution and stability of 0.02 μF/F in 2 hours for
measuring a 100 pF capacitor relative to a 12906 Ω resistor at 1233
Hz. The relative combined standard uncertainty (k = 1) is 0.13
μF/F, dominated by the digitizer error.
Index TermsAC voltage ratio; digital bridge; impedance
standard; lock-in detector; noise cancellation.
I. INTRODUCTION
IGITAL techniques can be readily used to generate two
synchronized ac voltages with a phase difference of π/2.
The digital bridges, based on such ac sources, have the potential
to greatly simplify comparisons between a capacitor and a
resistor. Precise measurements of such impedance ratios are
critical to develop quantum-based impedance standards. The
present status of the digital bridges as compared with the
traditional transformer-based impedance bridges has been
recently reviewed [1]. The latter still provide measurements
with the highest accuracy for the most demanding applications,
including the realization of the capacitance unit from calculable
capacitors or the ac quantized Hall resistance (QHR) through a
quadrature bridge [24]. However, the digital bridges have been
noticeably improving for impedance comparisons, offering
many advantages through computer control and automation [5
11]. In particular, Josephson arbitrary waveform synthesizers
establish a quantum-based voltage ratio standard that can be
used for impedance comparisons at any phase angle [5,6].
Digital signal sources custom-designed for impedance bridges
have also shown great promise. A dual-channel ac voltage
source with amplitude ratio stability better than 0.01 μV/V and
a phase resolution of 0.2 µrad at 1 kHz has been reported [10].
A fully-digital four-terminal-pair (4TP) bridge, using such a
custom-designed voltage ratio source for reference, has been
reported for RC comparisons with a 1:1 magnitude ratio with a
combined uncertainty of 9.2×10−8, showing great promise for
the realization of the unit of capacitance from an ac QHR
standard [11]. Another interesting approach [12] is to use
commercial synthesizers that are then stabilized with a negative
feedback loop, minimizing the bridge error signal.
Fig. 1. Schematic of Quad bridge with combining network.
When the voltage ratio of two synthesized sources is used
directly as the reference for impedance ratio measurements, as
described in the literature [5-11], the stability of the voltage
ratio can become a major limiting factor for the overall bridge
performance. It appears that an underexplored research area is
to mimic in the digital domains some analog techniques that are
commonly used in the analog bridges to correlate and combine
detector voltages, enabling suppression of the effect of source
fluctuations. Let us consider the Quad bridge [2,3], shown in
Fig. 1, as an example. The complex impedance ratio of a resistor
and a capacitor, with a phase of π/2, cannot be measured with
high accuracy with a single quadrature bridge because the
required voltage ratio at a phase angle of π/2 cannot be
accurately produced in an analog bridge. However, two such
Comparison of a 100 pF Capacitor with a
12906 Ω Resistor Using a Digital Impedance
Bridge
Mona Feige, Stephan Schlamminger, Senior Member, IEEE, Andrew Koffman, Senior Member, IEEE,
Dean Jarrett, Senior Member, IEEE, Shamith Payagala, Alireza Panna, Bryan Waltrip, Senior
Member, IEEE, Michael Berilla, Frank Seifert, and Yicheng Wang, Fellow, IEEE
D
This work is licensed under a Creative Commons Attribution 4.0 License. For more information, see https://creativecommons.org/licenses/by/4.0/
This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/TIM.2021.3139709, IEEE
Transactions on Instrumentation and Measurement
> REPLACE THIS LINE WITH YOUR PAPER IDENTIFICATION NUMBER (DOUBLE-CLICK HERE TO EDIT) <
2
ratios in sequence, forming a double quadrature bridge with a
total phase shift of π, can be measured with high accuracy using
a transformer ratio as reference. It is important to observe that
although the overall accuracy of a Quad bridge can be very
high, the error voltages of the individual quadrature bridges at
points A and B (Fig. 1) fluctuate significantly due to the
inevitable fluctuation of the quadrature voltage represented by
δV. An elegant feature of the Quad bridge is to combine the
error voltages with a RC combining network such that it forms,
with the main bridge components, a twin-T network from the
quadrature voltage to the detector, D; the twin-T network is a
notch filter and can be adjusted so that D is immune to δV at
the fundamental frequency of the bridge excitation.
The Quad bridge can be simplified using digital techniques.
Specifically, if the detector voltages at points A and B (Fig. 1)
are digitized, their correlation can be analyzed in post-
processing and the function of the analog combining network
can be replaced by software algorithms. One can further argue
that if the detector voltage of a single quadrature bridge is
synchronously digitized with the source voltages, their
correlation can also be analyzed to suppress the source
fluctuations. This paper describes our research in this direction,
aiming to develop a simple digital bridge for RC comparisons.
Fig. 2. Schematic of digital impedance bridge for comparison of
 with   (C = 100 pF and RH = 12906 Ω) at a frequency near
1233 Hz. Z3 is the feedback resistor of the current amplifier. S1 (Amplitude =
10 V, Phase = 90°), S2 (Amplitude = 0.1 V, Phase = 0), and S3 are waveform
generators. V1, V2 and V3 are ac voltmeters. V1 (Amplitude = 0.1 V, Phase = -
90°) and V2 (nominally, Amplitude = 0.1 V, Phase = 0) are connected to the
high-potential ports (A and B) and are periodically switched to minimize the
effect of their gain drift. S3 is adjusted such that V3 is nominally 0. Coaxial
chokes (omitted for clarity) are placed in every unwanted loop in the bridge
circuit [13].
II. BRIDGE SETUP
The digital impedance bridge, shown in Fig. 2, is designed
for comparisons between a 4TP Vishay* resistor, with a
nominal value of RH = 12906 Ω, in an air bath at 23 °C, and a
two-terminal-pair (2TP) Andeen-Hagerling capacitor, with a
nominal value of C = 100 pF. The impedance of the capacitor
and the resistor are represented with
 and  ,
respectively, and the associated impedance ratio is represented
by a complex number, 
. The low port of the 2TP
capacitor was connected directly to the low-current port of the
resistor without a combining network by following a method
described by Small et al. [14] to compare 4TP resistors with
2TP capacitors. A current amplifier (Femto DLPCA-200) with
transimpedance of Z3, which is used to detect the bridge error
voltage, was connected to the low-potential port of the 4TP
resistor. Hence, the cable and the contact resistance between the
low-current port of the resistor and the low port of the 2TP
capacitor were then considered part of the capacitance standard.
As long as the defining planes are applied consistently in
calibrations, the inclusion of contact resistance only affects the
dissipation factor of the capacitor slightly, with a negligible
contribution to the uncertainty of the capacitance
measurements.
We used two phase-locked channels (S1 and S2) of a Keysight
33500B waveform generator as the main sources to excite the
bridge through a 2TP current loop connecting to the high-
current ports of Z1 and Z2, applying root mean square (rms)
voltages of 7.07 V and 70.7 mV, respectively, to the capacitor
and the resistor at a frequency near 1233 Hz. To overcome the
limited resolutions of the generator outputs, another
synchronized 33500B generator (S3) was used to inject a fine
adjustment signal through a 10000:1 injection transformer
inserted into the lower excitation arm. An external time base
was used for both generators with their 10 MHz reference signal
locked to the Global Positioning System.
The modulus of the nominal impedance ratio is 100. To avoid
the digitizer nonlinearity of sampling the excitation voltages
with different amplitudes, a two-stage inductive voltage divider
(IVD), with its input-to-output ratio, ko, having a nominal value
of -100, was added between the high port of the capacitor and
the voltage measurement system. The operating frequency of
the bridge was fine-tuned such that the modulus of the
impedance ratio, r, was arbitrarily close to ko in magnitude and
the sampled V1 and V2 were nominally equal in amplitude. The
IVD ratio may slightly depend on the loading condition and
therefore was calibrated, in situ, using a conventional four-arm
bridge with two capacitors, Ca and Cb, of nominal values of 1
pF and 100 pF, respectively. A small micrometer-controlled
trim capacitor added in parallel to Cb was used to null the in-
phase component of the bridge error.
The IVD output (A) and the low-potential port (B) of Z2 form
a 2TP potential loop of a digital bridge with two voltage
detectors through a custom coaxial switching fixture, which
This work is licensed under a Creative Commons Attribution 4.0 License. For more information, see https://creativecommons.org/licenses/by/4.0/
This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/TIM.2021.3139709, IEEE
Transactions on Instrumentation and Measurement
> REPLACE THIS LINE WITH YOUR PAPER IDENTIFICATION NUMBER (DOUBLE-CLICK HERE TO EDIT) <
3
was described previously [15]. The two detectors were
periodically interchanged to minimize the effect of their gain
drift. A small loading change at A and B is equivalent to a small
change of the excitation voltage ratio, which is suppressed in
the digital domain by correlation with the bridge error signal.
We used a Keysight DAQM909A, a four-channel 24-bit
digitizer module in a Keysight DAQ973A data acquisition
system, to simultaneously sample V1, V2, and V3, preserving the
relative phase difference of the three signals. The digitizer was
set with a differential input, a sampling rate of 800,000 samples
per second, and a record length of 2,400,000 samples for each
measurement. The analog bandwidth of the digitizer is
approximately 125 kHz. The amplitude and the phase of each
sampled voltage were determined using an algorithm of three-
parameter least squares fit as described in IEEE Standard 1057-
2017 [16].
The digital bridge (Fig. 2) relies on accurate measurements
of voltage ratios to determine the phase factor of the impedance
ratio,  . In the ideal case, the excitation sources would be
adjusted to balance the bridge, such that for any measured
voltage, V2, at the high-potential port of Z2, the measured
voltage, V1, which is scaled down by the IVD from the high-
potential port of Z1, would be equal to a perfect value V1p,
achieving the condition of equal current through the two
impedances under comparison. The balance equation is

In practice, the balance is never perfect, and the source drift
always exists. The combined effect can be represented by an
error voltage, δV, superimposed on the ideal voltage V1p, and
we have V1=V1p+δV. The effect of the error voltage is
automatically balanced through the feedback resistor Z3 of the
current amplifier. The common low-potential port is kept at
virtual ground, and the detected error voltage, V3, relates to δV
through:

The phase difference between V3 and δV is approximately 90°.
We define the gain factor  
. Hence,    .
The bridge dynamics can be understood as a superposition of
the two voltage-balancing actions governed by (1) and (2).
III. TEST RESULTS AND DISCUSSIONS
A. Equal Voltage Test
The measurement accuracy of the DAQM909A for ac
voltage ratios depends on not only the resolution of the digitizer
but also the gain stability of the input amplifiers. To determine
the limitations of the digitizer, we connected two input channels
of a DAQM909A in parallel to the same sinewave voltage, with
a rms value of 0.1 V at 1 kHz, similar to the tests described in
the previous paper [15] when the two SR860 lock-in detectors
were used to measure large ac signals. The best results were
obtained when the two input channels, set at the 0.3 V input
range, were periodically interchanged through the coaxial
switching fixture, creating two virtually identical digitizing
channels. The Allan deviation of the measured unity voltage
ratio as a function of the averaging time follows a straight line
in a log-log plot, with its slope consistent with averaging over
white noise. It reaches below 0.01 μV/V in approximately four
hours, about a factor of 10 lower than what was achieved using
the SR860s.
B. Digitized Bridge Voltage
A major advantage of the digital bridge is that the
excitation voltages and the error signal can be fully digitized,
and the bridge dynamics can be analyzed in postprocessing. All
the test results presented herein were acquired with the bridge
setup shown in Fig. 2. The gain of the transimpedance amplifier
was set at 107 V/A, and the corresponding Z3 was approximately
10 MΩ; the 3 dB bandwidth at this setting is 50 kHz. Figures 3
and 4 show the measured V1, V2, V3 values, and the bridge error
voltage, ε, as a function of time that were acquired with S1 and
S2 set at 1233.19734Hz, a phase of 90° and 0°, and an amplitude
of 10 V and 0.1 V, respectively. The complex amplitude of S3
set at the same frequency was automatically controlled through
a computer to minimize the mean bridge error (V3), using a
simple proportional-integral feedback algorithm.
The phase difference between V1 and V2 is approximately -
90°. The digitized voltages are phase normalized such that the
phase of V2 is 0. Their complex components are more
conveniently compared between jV1 and V2. The real parts of
jV1 and V2 (Fig. 3) fluctuated, on the order of 10 μV, exceeding
a factor of 10 more than the imaginary counterparts (Fig. 4).
This reflects that the digital sources have better phase stabilities
than amplitude stabilities. The real component of V2 closely
follows that of jV1, resulting from the feedback action that
minimizes the bridge error signal.
For better comparison, the error voltage V3 is shown after
being scaled with an estimated gain factor.  is
dominated by white noise, and its mean is effectively locked to
0 through the feedback (Fig. 3). The fluctuations of 
This work is licensed under a Creative Commons Attribution 4.0 License. For more information, see https://creativecommons.org/licenses/by/4.0/
This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/TIM.2021.3139709, IEEE
Transactions on Instrumentation and Measurement
> REPLACE THIS LINE WITH YOUR PAPER IDENTIFICATION NUMBER (DOUBLE-CLICK HERE TO EDIT) <
4
form a mirror image of Im(jV1), with its mean also locked to 0
(Fig. 4).
Fig. 3. Real components of recorded voltages as a function of time: (1) jV1, (2)
V2, (3) V3 scaled with the gain factor, (4) ε. jV1 and V2 are shifted by c = 105
uV.
We can qualitatively understand how the detected error
voltage V3 relates to the source fluctuation δV by considering
that the transimpedance amplifier together with Z1 and Z2 form
a summing amplifier. Since Z1/ko and Z2 are nominally equal in
magnitude and differ by 90° in phase, we have V1+jV2 -δV.
We define
   
The real and imaginary components of ε are shown in Fig. 3 and
Fig. 4, respectively. Both components follow a white noise
distribution, with a standard deviation of less than 0.05 μV,
indicating a strong correlation between δV and .
Fig. 4. Imaginary components of recorded voltages as function of time: (1)
jV1, (2) V3 scaled with the gain factor, (3) ε.
C. Correlation Analysis and Noise Cancellation
To analyze the dynamics of the bridge balancing more
rigorously, we applied Kirchhoff’s law to the bridge circuit:
 
(4)
Using conventional notations, we define α and β as the real and
imaginary part of the deviation, respectively, from the nominal
impedance ratio that is perfectly matched to the IVD ratio in
magnitude:
     
(5)
where   .
Combining (4) and (5), we rewrite:
  
   
(6)
We define:
   
(7)
 
(8)
Eq. (6) becomes:
    
(9)
Using a linear fitting between the complex variables u and v,
we can determine . We then have
This work is licensed under a Creative Commons Attribution 4.0 License. For more information, see https://creativecommons.org/licenses/by/4.0/
This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/TIM.2021.3139709, IEEE
Transactions on Instrumentation and Measurement
> REPLACE THIS LINE WITH YOUR PAPER IDENTIFICATION NUMBER (DOUBLE-CLICK HERE TO EDIT) <
5
    
(10)
    
(11)
To visualize the effectiveness of the linear fitting, we plot
the imaginary part versus the real part for u and in
Fig. 5(a). The natural fluctuation of u is mainly along the real
axis, covering a range of about 150 μV/V, reflecting that the
digital sources have better phase stabilities than amplitude
stabilities. The pattern of w is similar to that of u, except that it
is tilted due to a phase shift of the current amplifier. The
residuals of the linear fitting can be seen in Fig. 5(b), showing
α versus β. The residual data points distribute tightly in a circle
of radius about 0.5 μV/V, indicating that the fluctuations of u
and v largely cancel out in determining α and β.
Fig. 5. (a). Imaginary part versus real part: u in light blue and w in cyan. w
shifted lower by j20 × 10-6 for clarity. (b). β versus α. 24 data points of the
fitting residuals distributed close to the perimeter of a circle are colored
progressively in (b); the corresponding u and w points in (a) show their
correlation.
Figure 6 shows as function of time over a period of 24
hours. The distribution of the data points is consistent with a
constant that is buried in white noise. Each data point in the
lower panel takes about 36 s to acquire, and all the data points
stay within ±0.7 × 106. Averaging 256 points, or about 2 hours’
worth of data, produces a new set of averaged data that
fluctuates within ±0.02 × 106 about their mean. The
fluctuations can be attributed predominantly to the limited
resolution of the digitizer.
Fig. 6. Determined α as a function of time. The black dots were obtained by
averaging 256 points, or about 2 hours’ worth of data. The error bars in the
top graph denote the 1-σ standard deviation of the 256 points.
Figure 7 shows as a function of time over the same period.
The distribution of the data points of β are similar to α and also
consistent with a constant value over time. All the data points
stay within ±0.7 × 106. Averaging 256 points also produces a
new set of averaged data that fluctuates within ±0.02 × 106
about their mean.
Fig. 7. Determined β as a function of time. The black dots were obtained by
averaging 256 points, or about 2 hours’ worth of data. The error bars in top
graph denote the 1-σ standard deviation of the 256 points.
The Allan deviations of α and β are shown in Fig. 8. Both
decrease to 2×108 level in about 3 h and show a monotonic
downward trend over the test time window, demonstrating the
stability of the digital bridge.
Fig. 8. Squares and circles are the real and imaginary part of the deviation
from the nominal impedance ratio. Error bars are 1-σ standard deviation of the
Allan deviation.
D. Results and Uncertainty Analysis
The digital impedance bridge enables us to measure the
capacitance of C in reference to RH with a Type A uncertainty
(k = 1) of 0.02 μF/F. Repeated measurements show that the
results of C using the digital bridge are consistent, within 0.11
μF/F, with its capacitance measured against the Farad Bank,
which is used to maintain the capacitance unit at the National
Institute of Standards and Technology (NIST) and is traceable
to the calculable capacitor [17]. The difference can be partly
attributed to the frequency dependence of C because the digital
bridge functions near 1233 Hz to keep the impedance ratio close
This work is licensed under a Creative Commons Attribution 4.0 License. For more information, see https://creativecommons.org/licenses/by/4.0/
This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/TIM.2021.3139709, IEEE
Transactions on Instrumentation and Measurement
> REPLACE THIS LINE WITH YOUR PAPER IDENTIFICATION NUMBER (DOUBLE-CLICK HERE TO EDIT) <
6
to 100:1 in magnitude while the capacitance measurement
relative to the Farad Bank has been restricted to 1592 Hz.
However, the largest uncertainty source for the digital bridge is
the digitizer error as shown in Table 1.
The digital errors associated with the digitizer may arise from
aliasing and spectral leakage. Stray capacitances in the digitizer
may also cause crosstalk between the ADC channels and
leakage to the ground. These errors have been estimated
experimentally and numerically by varying the sampling rate
and the record length, combined with temporarily introducing
extra cross capacitances and changing from the differential
input mode to the single-ended mode. Sensitivity to harmonics
has been estimated experimentally and numerically by
including selected harmonic base functions in the sine fit,
adding simulated harmonic content to the digitized data record
before the sine fit, and physically injecting additional third
harmonic voltage into the bridge excitation. Possible offset
error in the detected V3 due to non-linearity of the current
amplifier and the ADC, causing intermodulation distortion, was
also accessed by changing the gain settings of the amplifier and
the ADC; no correlated change was detected within the limit of
the bridge resolution.
In the future, we plan to modify the front analog circuit of the
digitizer to reduce its error. The uncertainty for the frequency
dependence determination, which has currently been limited by
the stability of a reference 1 pF cross capacitor at NIST, can
also be significantly reduced [17, 18].
Table 1. Uncertainty Budget (k = 1)
Relative standard
uncertainty (×10-6)
Type A
0.02
Digitizer error
0.10
Frequency dependence
of Ca (1 pF) and Cb (100 pF)
0.07
Ca and Cb relative to Farad Bank
0.03
RH relative to dc QHR
0.01
Frequency dependence of RH
0.01
Relative combined standard uncertainty
0.13
IV. CONCLUSION
We evaluated a digital impedance bridge in a hybrid structure
for comparison of a capacitor with a resistor where the
impedance ratio was measured in two separate parts. The
modulus of the impedance ratio was matched arbitrarily close
to the input-to-output ratio, in magnitude, of a two-stage IVD
by adjusting the operating frequency of the bridge; the residual
deviation between the two together with the phase factor of the
impedance ratio was measured using a custom detection system
based on a four-channel 24-bit digitizer. The IVD was
calibrated, in situ, using a four-arm bridge with two known
capacitors. In contrast to the conventional approach of
emphasizing precision and stability of the voltage sources
driving the bridge, we adopted an approach that focused on the
resolution and stability of the detectors. Fluctuations of the
source voltages were largely removed through postprocessing
of the digitized data, and the measurement results were limited
by the digitizer error. While we have achieved a low Type A
uncertainty (k = 1) of 0.02 μF/F in 2 hours for determining the
capacitance of a 100 pF capacitor relative to a 12906 Ω resistor
at 1233 Hz, the combined relative standard uncertainty (k = 1)
is 0.13 μF/F. Even though the uncertainty is not as low as for a
conventional IVD-based double-quadrature bridge which has
the modulus of the nominal impedance ratio equal to one, the
digital bridge discussed here has a key advantage. The modulus
of the nominal impedance ratio of the digital bridge is 100. This
approach has the advantage of shortening the measurement
chain from a 12906 Ω resistor to a 100 pF capacitor by two 10:1
ratio steps. In the future, we plan to focus our research on
reducing the digitizer error for the digital impedance bridge to
serve as an alternative system at NIST for realizing the
capacitance unit.
The detection system based on the DAQM909A for
measuring ac voltage ratios compares favorably to the system
based on the SR860 lock-in detectors which we evaluated
previously [15]. We achieved a factor of 10 improvement in
terms of the Allan deviations for the impedance ratio
measurements over a comparable averaging window. We can
attribute the improvement to the higher resolution of the
modern data acquisition board and the customized
demodulation method in post-processing, which is not
accessible with the commercial lock-in detectors.
Acknowledgments
The authors would like to thank Dr. Jürgen Schurr of the
Physikalisch-Technische Bundesanstalt for providing the
frequency dependence measurements of a Vishay resistor, and
Dr. David Newell of NIST for his support and helpful
comments.
* Certain commercial equipment, instruments, or materials are
identified in this paper to foster understanding. Such
identification does not imply recommendation or endorsement
by the National Institute of Standards and Technology, nor
does it imply that the materials or equipment identified are
necessarily the best available for the purpose.
REFERENCES
[1] F. Overney and B. Jeanneret, “Impedance bridges: from Wheatstone
to Josephson,” Metrologia, vol. 55, pp. S119S134, Jul 2018.
[2] A. M. Thompson, “An absolute determination of resistance based on
a calculable standard of capacitance,” Metrologia, vol 4, pp.17, Jan
1968.
This work is licensed under a Creative Commons Attribution 4.0 License. For more information, see https://creativecommons.org/licenses/by/4.0/
This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/TIM.2021.3139709, IEEE
Transactions on Instrumentation and Measurement
> REPLACE THIS LINE WITH YOUR PAPER IDENTIFICATION NUMBER (DOUBLE-CLICK HERE TO EDIT) <
7
[3] R. D. Cutkosky, “Techniques for comparing four-terminal-pair
admittance standards,” Journal of Research of the National Bureau
of Standards, vol.74C, p. 63, 1970.
[4] J. Schurr, V. Bürkel, B. P. Kibble, “Realizing the farad from two ac
quantum Hall resistances,” Metrologia, vol. 46, pp. 619628, Oct
2009.
[5] F. Overney, N. E. Flowers-Jacobs, B. Jeanneret, A. Rüfenacht, A. E,
Fox, J. M. Underwood, A. D. Koffman, S. P. Benz, Josephson-based
full digital bridge for high-accuracy impedance comparisons,”
Metrologia, vol. 53, pp. 10451053, Jun 2016.
[6] S. Bauer et al., “A four-terminal-pair Josephson impedance bridge
combined with a graphene-quantized Hall resistance,” Measurement
Science and Technology, vol. 32, p.065007, Mar 2021.
[7] G. Ramm, H. Moser, New multifrequency method for the
determination of the dissipation factor of capacitors and of the time
constant of resistors,” IEEE Transactions on Instrumentation and
Measurement, vol. 54, no. 2, pp. 521524, 2005.
[8] J. Kucera, T. Funck, J. Melcher, “Automated capacitance bridge for
calibration of capacitors with nominal value from 10 nF up to 10 mF,”
in 2012 Conference on Precision electromagnetic Measurements
(IEEE), pp. 596597, 2012
[9] R. Rybski, J. Kaczmarek, K. Kontorski, “Impedance comparison
using unbalanced bridge with digital sine wave voltage sources,”
IEEE Transactions on Instrumentation and Measurement, vol. 64, pp.
33803386, 2015.
[10] J. Kucera, J. Kováč, L. Palafox L, R. Behr, L. Vojáčková,
Characterization of a precision modular sinewave generator,”
Measurement Science and Technology, vol. 31, p. 064002, 2020.
[11] M. Marzano, M. Ortolano, V. D’Elia1, A. Müller, L. Callegaro, “A
fully digital bridge towards the realization of the farad from the
quantum Hall effect,” Metrologia, vol. 58, p. 015002, 2021.
[12] W. G. Kürten Ihlenfeld, R. T. B. Vasconcellos, “A digital four
terminal-pair impedance bridge,” Conference on Precision
Electromagnetic Measurements (IEEE), pp. 12, 2016.
[13] D. N. Homan, “Applications of coaxial chokes to A-C bridge
circuits,” Journal of Research of the National Bureau of Standards,
vol. 72C, pp. 161165, 1968.
[14] G.W. Small, J.R. Fiander, P. C. Coogan, “A bridge for the comparison
of resistance with capacitance at frequencies from 200 Hz to 2 kHz,”
Metrologia, vol. 38, pp. 363368, 2001.
[15] M. Feige, S. Schlamminger, B. C. Waltrip, M. Berilla, Y. Wang,
Evaluations of a Detector-Limited Digital Impedance Bridge,”
Journal of Research of the National Bureau of Standards, vol. 126, p.
126006, 2021.
[16] IEEE Standard for Digitizing Waveform Recorders, IEEE
Standard 1057-2017, Jan. 2018.
[17] P. Gournay et al., “Comparison CCEM-K4.2017 of 10 pF and 100
pF capacitance standards,” Metrologia 56, 01001-01001, 2019.
[18] Y. Wang, S. Shields, “Improved capacitance measurements with
respect to a 1 pF cross-capacitor from 200 to 2000 Hz,” IEEE
Transactions on Instrumentation and Measurement 54, 542, 2005
... Currently, automatic fully digital bridges allow comparisons of impedance standards with an uncertainty of 10 −6 -10 −5 in the audio frequency range [15][16][17][18][19][20][21]. Achieving such measurement uncertainties requires the use of systems based on DAC and ADC converters with extremely high resolutions and appropriate, often advanced, digital signal processing algorithms. ...
... To determine the measurement errors of the real part E and the imaginary part F of the impedance ratio, the complex errors occurring in relation (15)-the impedance ratio (16), the ratio of the measured voltages reproduced by digital voltage sources (17)(18) and the unbalanced voltages (19)(20)-are presented in the form of the sum of their real and imaginary parts: ...
... Considering the relations (16)(17)(18)(19)(20) in (15), the following form of the expressions defining the measurement error, respectively, of the real and imaginary parts of the ratio of the compared impedances Z1 and Z2 are obtained: ...
Article
Full-text available
This article presents an automatic digital unbalanced impedance bridge for comparing two-port impedances: R-R, R-C and C-C. The bridge consists of a two-channel digital source of voltage sinusoidal waveforms supplying the bridge arms and a three-channel precise digitizer, by means of which the complex ratio of the source output voltages and the relative bridge unbalanced voltage are determined. The hardware implementation of the bridge is based on a relatively inexpensive universal data acquisition (DAQ) card by National Instruments, USB-6281, which contains, among others, a multi-channel 18-bit analog-to-digital converter (ADC) and two 16-bit digital-to-analog converters (DAC). Thanks to the new approach to the bridge operation algorithm, consisting in the use of the interpolation method in the comparison process, the influence of the accuracy of the measurement of the bridge unbalanced voltage and some bridge parasitic admittances on the uncertainty of the impedance comparison were minimized. This simplifies the measurement procedure and shortens its execution time. The obtained results confirmed that a relatively simple digital unbalanced bridge can compare impedance standards with an uncertainty of less than 10−5. By reducing the requirements regarding the uncertainty of the measurement, and thus its execution time, the presented bridge supplemented with impedance sensors can be used in many areas of industry.
... Understanding the weak frequency dependence of the air capacitors at very low frequencies is an important stepping stone for impedance metrology at audio frequencies. At this frequency range, the air capacitors can be used as stable transfer standards, see, for example, [22]. ...
Article
Full-text available
Four terminal-pair air capacitors are important transfer standards to calibrate LCR meters up to a frequency of 10 MHz. We report a simple and new method to obtain the frequency dependence of the four terminal-pair capacitance of these standards using a four-channel vector network analyzer (VNA). The frequency dependence of the capacitance of an air capacitor and its uncertainty can be obtained from a single set of measurements without changing connections between the standard and the VNA, as has been the case in previously published work. The calculation of the frequency dependency is straightforward and model-independent. Nevertheless, an elementary model is provided to explain the observed frequency dependence. This article allows every laboratory with a four-channel VNA to measure the frequency dependence of these capacitors. Hence, a significant shortening of the traceability chain is achieved.
... The voltages across Z 1 , Z 2 , and Z 3 , namely V 1 , V 2 , and V 3 , respectively, are digitized synchronously using a Keysight DAQM909A 13 and demodulated independently at each excitation frequency, with their sine-fitting complex amplitude denoted by U 1 , U 2 , and U 3 , respectively. The bridge is dynamically balanced, using a simple proportional-integral feedback algorithm, through the coaxial injection S 2 such that U 3 = 0, statistically. ...
Article
We report a new method for determining the time constant of ac resistors with values around 10 kΩ using a digital impedance bridge for the comparison of two nominally equal resistors. This method involves adding a probing capacitor in parallel to one of the resistors to induce a quadratic frequency dependence in the real component of the admittance ratio between the two resistors. The magnitude of this quadratic effect is proportional to the self-capacitance of the unperturbed resistor, enabling us to determine its value and the associated time constant with an estimated standard uncertainty ( k = 1) of 0.02 pF and 0.2 ns, respectively.
Article
The markers that help to predict th function of a cardiovascular system are hemodynamic parameters like blood pressure (BP), stroke volume, heart rate, and cardiac output. Continuous analysis of hemodynamic parameters such as BP can detect abnormalities earlier, preventing cardiovascular diseases (CVDs). However, sometimes due to motion artifacts, it becomes difficult to monitor the BP accurately and classify it. This work presents an optimized deep learning model having the capability to estimate the systolic blood pressure (SBP) and diastolic blood pressure (DBP) and classify the BP stages simultaneously from the same network using only a single channel photoplethysmography (PPG) signal. The proposed model is designed by exploiting the deep learning framework of a convolutional neural network (CNN), exhibiting the inherent ability to extract features automatically. Moreover, the proposed framework utilizes the superlet transform method to transform a 1-D PPG signal into a 2-D super-resolution time–frequency (TF) spectrogram. A superlet transform separates the peaks related to true PPG signal components and motion artifacts components. Thus, the superlet provides a robust realtime approach to accurately estimating and classifying BP using a single PPG sensor signal and does not require additional ECG and PPG sensor signals for reference. Using a super-resolution spectrogram and CNN model makes the method profitable in motion artifact removal, feature selection, and extraction. Hence the proposed framework becomes less complex for deployment on wearable devices having limited battery resources. The performance of the proposed framework is demonstrated on the publicly available larger dataset MIMIC-III. This work obtained a mean absolute error (MAE) of 2.71 mmHg and 2.42 mmHg for SBP and DBP, respectively. The classification accuracy for the SBP prediction is about 96.79%, whereas it is 98.94% for DBP. From a motion artifact-affected PPG signal, SBP and DBP are estimated. Then the estimated BP is classified into three categories: normotension, prehypertension, and hypertension, and is compared with the state of art methods to show the effectiveness of the proposed optimized framework.
Article
Since 2008, the capacitance unit, the farad, has been realized at the Physikalisch-Technische Bundesanstalt (PTB) successfully and routinely using a measurement chain starting at two quantum Hall resistance (QHR) devices operated at ac. After the first years of practice, a systematic scatter became apparent which indicated an increased measurement uncertainty. This was not due to systematic errors in bridge circuitries (which were fixed in 2014) but accidental incidents. Since then, valuable operational experience has been gained on how to avoid them. As a result, the current uncertainty of the calibration of a 10-pF capacitance standard is back to the initially published relative uncertainty of $6\times 10^{-9}$ ( $k = 1$ ), which is still unmatched. In this article, we report the findings from the operational experience gained during this time, in particular the effect of charge carriers trapped in a QHR device due to accidental discharging incidents. Besides helping practitioners to avoid the pitfalls, the low uncertainty provides a basis for improving the capacitance standards and for reducing the calibration period. Furthermore, we present a new and unique digital sinewave generator. Finally, we show that precision ac measurements of gallium arsenide (GaAs)-based QHRs do not require a <sup xmlns:mml="http://www.w3.org/1998/Math/MathML" xmlns:xlink="http://www.w3.org/1999/xlink">3</sup> He cryo-magnet system but can be carried out in a cheaper and easier-to-operate <sup xmlns:mml="http://www.w3.org/1998/Math/MathML" xmlns:xlink="http://www.w3.org/1999/xlink">4</sup> He cryo-magnet system at 2.1 K.
Article
Full-text available
We tested a simple digital impedance bridge using two nominally equal resistors to form a 1:1 ratio. We focused on resolution and stability of the detectors. Fluctuations of the source voltages were largely removed through postprocessing of the digitized data, and the measurement results were limited by the detector noise. This detector-limited operating condition was first demonstrated using three modified Keysight 3458A multimeters for measurements of the voltage ratios, achieving 0.01 μV/V type A uncertainty in less than 15 min at 1 kHz. In an effort to extend the applicable frequency range and develop a system with off-the-shelf components, we tested a system using three lock-in detectors for measuring small deviations from the perfect AC ratio of unity magnitude, achieving stabilities and resolutions of 0.1 μV/V in a few hours for each point from 1 kHz to 5 kHz.
Article
Full-text available
This paper introduces a four-terminal-pair impedance bridge based on pulse-driven Josephson junctions arrays which is designed to link any kind of impedance to the quantized Hall resistance. The unique features of the quantized Hall resistance in a multiple series connection allows to dispense a combining network and leads to a compact and simple design of the whole setup. Moreover, the low noise of a quantized Hall resistance reduces the measurement time compared to resistance standards at room temperature which is essential for the characterization of quantum Hall devices in the AC regime. A first measurement campaign confirmed the expected low noise of 1.82 n V / H z for a link to a 10 nF capacitance standard. The repeatability of the bridge was found to be few parts in 10 ⁸ . Capacitance and resistance standards were measured at 1233.15 Hz against graphene based quantum Hall resistance devices.
Article
Full-text available
This paper presents the implementation of an electronic fully-digital impedance bridge optimized for RC comparisons with equal impedance magnitudes, together with an evaluation of the uncertainty. This bridge has been designed with the goal of realizing the farad directly from the quantum Hall effect with a bridge uncertainty component at the 10 ⁻⁷ level. Thanks to its simple design, ease of operation and affordability, this bridge is suitable to be industrially manufactured. Together with the increasing availability of graphene quantum Hall resistance standards, this can provide an affordable quantum realization of the unit farad for metrology institutes and calibration centres. In this paper we present the uncertainty budget of an example measurement and the results of the validation of the bridge against a suitably modified version of the traceability chain of the Italian national standard of capacitance. The combined uncertainty of the bridge resulted from repeated measurements (overall measurement time of about 200 min) is 9.2 × 10 ⁻⁸ , suitable for the primary realization of the unit of capacitance from a quantized Hall resistance standard. The crosstalk among the channels of the electrical generator is the most significant uncertainty component, possibly reducible with internal shielding and filtering of the electronic generator.
Article
Full-text available
At the Czech Metrology Institute (CMI) we have developed a precision modular sinewave generator for impedance ratio bridges. The generator was developed to improve previously available designs regarding amplitude and phase accuracy, linearity, absolute stability, stability of the ratio between two outputs and harmonic distortion. It generates up to 7 Vrms in the frequency range from 1 mHz up to 20 kHz, extendable to 100 kHz with small changes to the filters. The amplitude resolution is better than 0.01 μV/V of full scale with an output voltage stability of 0.05 µV/V/30 minutes and a stability for the ratio between two outputs of 0.02×10^-6 over several hours. The generator can be powered from internal batteries and is controlled via optically isolated connections. The internal clock and voltage references can be replaced by external ones, optically coupled in the case of the clock. In this paper, we discuss experimental results obtained with the generator used as a signal source in digital impedance bridges with relative combined uncertainties from 10-5 down to 10-8. The generators have been used in a bridge to drive a quantum Hall resistor in the AC regime. The use of the generator with the AC quantum voltmeter will also be discussed. The generator is not only applicable in the field of AC impedance metrology but also for on-site comparisons of AC quantum voltage standards or, in general, where there is a need for precision voltage sources.
Article
Full-text available
This paper reviews the recent evolution of the measurement techniques in the area of impedance measurement from the Wheatstone bridge to the most advanced Josephson based digital impedance bridge. The progress in the development of high sampling rate, high accuracy digital-to-analog and analog-to-digital converters has led to the development of digital impedance bridges that have profoundly modified the landscape of impedance metrology. Although these new bridges do not yet outperform the traditional transformer based bridge in terms of accuracy, they significantly improve their measurement capabilities and flexibility by allowing a complete automation, a full coverage of the complex plane, arbitrary bridge ratios and extended frequency range. In addition, after the redefinition of the International System of Unit, they will contribute to the realization of the henry and the farad by establishing a direct link between the quantized Hall resistance and the capacitance or inductance standards.
Article
Full-text available
This paper describes a Josephson-based full digital impedance bridge capable of comparing any two impedances, regardless of type (R-C, R-L, or L-C), over a large frequency range (from 1 kHz to 20 kHz). At the heart of the bridge are two Josephson arbitrary waveform synthesizer systems that offer unprecedented flexibility in high-precision impedance calibration, that is, it can compare impedances with arbitrary ratios and phase angles. Thus this single bridge can fully cover the entire complex plane. In the near future, this type of instrument will considerably simplify the realization and maintenance of the various impedance scales in many National Metrology Institutes around the world.
Article
In 2017 the Consultative Committee for Electricity and Magnetism (CCEM) commissioned a key comparison of electrical capacitance standards, the second time this quantity has been compared since the implementation of the Mutual Recognition Agreement by the Comité International des Poids et Mesures (CIPM—MRA) in 1999. This comparison—CCEM-K4.2017—was piloted by the Bureau International des Poids et Mesures (BIPM) and included seven National Metrology Institutes (NMI) belonging to four Regional Metrology Organizations. The measuring scheme adopted for the comparison was that of a star comparison consisting of a set of bilateral comparisons between the participating NMIs and the BIPM, whose capacitance reference base served as a common reference. For each of the bilateral comparisons, the measurands were the capacitance values of 10 pF travelling standard capacitors belonging to the NMIs and, optionally, the values of 100 pF standards. All the participants have been chosen from those able to realize and maintain a representation of the farad at the best known level of accuracy. Four of them, including the BIPM, were taking their traceability from dc or ac quantum Hall effect standards and, the four others, from a calculable capacitor. The comparison results analysis have evidenced an agreement within about ±5 parts in 10⁸ for the mandatory 10 pF measurements and within about ±10 parts in 10⁸ for the optional 100 pF measurements. Also, excepted for one of the participants, a good agreement has been found for the ratio 100 pF:10 pF (within ±5 parts in 10⁸). In addition to the comparison, it has been possible to evaluate the difference between the value of RK (von Klitzing constant) measured by electrical means from calculable capacitors and its last CODATA recommended value (CODATA 2014 adjustment). A difference of (43 ± 23) parts in 10⁹ (k = 1) has been found which is consistent with the difference that can be computed from the experimental data used in the CODATA 2014 adjustment of fundamental constants. This report presents the details of the measurements and analysis having led to these results. Main text To reach the main text of this paper, click on Final Report. Note that this text is that which appears in Appendix B of the BIPM key comparison database kcdb.bipm.org/. The final report has been peer-reviewed and approved for publication by the CCEM, according to the provisions of the CIPM Mutual Recognition Arrangement (CIPM MRA).
Article
This paper presents a digital automatic unbalanced bridge designed for comparing the impedances of R-R, R-C, and C-C. The measurement principle uses initial balancing of the bridge using two digital sine wave sources, then measuring the ratio of voltage drops on the compared impedances as well as the finite bridge unbalance voltage. The present method is novel because it is not necessary to obtain a complete voltage nulling and also because it is readily implemented with a commercial data acquisition (DAQ) card. An NI USB6281 DAQ card was used to obtain the results; this card combines the two 16-bit digital-to-analog converters with one multiplexed 18-bit analog-to-digital converter. Such a measurement system, which is relatively simple and inexpensive, can compare impedances with an error below 10 ppm. Some considerations concerning the relationship between the measurement accuracy of unbalanced voltage and the level of the bridge unbalance are given.