Open Access
29 December 2023 Loaded waveguide measurements of plastic explosives at V-band
Zachary J. Landicini, Jeffrey Barber, James C. Weatherall, Duane C. Karns, Peter R. Smith, Joaquín Aparicio-Bolaño, Wendy Ruiz
Author Affiliations +
Abstract

Dielectric measurements of plastic explosives using a loaded waveguide technique via vector network analyzer and banded millimeter wave extender modules operating at V-band (50 to 75 GHz) are performed. A portion of an explosive sample is inserted into a waveguide shim 2 mm in length and trimmed flush with the faces of the shim. Two-port S-parameter measurements are conducted on the explosive; the empty shim is similarly characterized. Using standard waveguide equations and the measured length of the shim, the complex S-parameter data obtained with the filled shim is optimized to four free parameters—complex permittivity and distance offsets for the two sample faces relative to the calibration planes. Permittivity data obtained from measurements of the plastic explosives C-4, Primasheet 1000, Primasheet 2000 and Semtex 10 are presented. Results obtained for C-4 and Primasheet 1000 are comparable to other data in the literature, and the data on Primasheet 2000 and Semtex 10 are the first known published permittivity values in this range. Excellent agreement between the experiment and the fit is obtained using a constant permittivity across the waveguide band, indicating that dispersion is not significant for these materials.

1.

Introduction

The Transportation Security Laboratory, operated by the U.S. Department of Homeland Security’s Science and Technology Directorate, performs dielectric measurements on explosives for the purpose of designing simulants to support the test and evaluation of advanced imaging technology (AIT)15 systems deployed for explosives detection. The development of AIT systems, such as the Rohde and Schwarz QPS imaging systems,69 operating in the range of 70 to 80 GHz necessitates that the dielectric properties of explosives be measured at higher frequencies. Past studies have used techniques such as resonant cavities1 and free space methods10 to characterize these materials. Going to higher frequencies using resonant cavities has been challenging. As the frequency increases, the wavelength gets smaller, and the resonant cavities must decrease in size with the wavelength. As a result, repeatability of the polyethylene fixtures used for these measurements have been poor as the wavelength approaches machining tolerances. Free space methods, although simple, require significant sample sizes (4 inches in diameter) and a uniform thickness. Getting uniform samples of a sticky putty substance, such as C-4, is challenging, and the size significantly increases the safety hazards.

The loaded waveguide technique for permittivity measurements has been used to characterize solid materials, such as plastics,11 glass,12 and concrete,13 as well as liquid and granular materials.14,15 The use of waveguides for permittivity measurements has also been adopted as an ASTM international standard method.16 When coupled with the aforementioned challenges to using other techniques, the loaded waveguide technique is particularly attractive for semi-solid explosives. The loaded waveguide can be filled by hand, and the volume of the sample can be reduced from the order of 100 milliliters (free space method) to tens of microliters, dramatically increasing the safety of the measurements.

This report documents the dielectric characterization of the explosives C-4, Primasheet 1000, Primasheet 2000, and Semtex 10 using a V-band (50 to 75 GHz) loaded waveguide technique. The theory of this technique and adaptations made to account for the nature of the sample are presented. Data obtained on the empty waveguide fixture are used to both determine the length of the sample holder and ensure that the fixture is in agreement with waveguide theory. Permittivity results of the plastic explosives are presented and discussed relative to other values available in the literature.

2.

Methods and Techniques

Figure 1 displays a diagram of the loaded waveguide method used in this work. The theory of electromagnetic propagation in a rectangular waveguide and its interaction with a dielectric sample are well understood.11,17 Briefly, the scattering parameters (S-parameters) obtained from a measurement of a sample are given as

Eq. (1)

S11=R12Γ(1T2)1Γ2T2,

Eq. (2)

S22=R22Γ(1T2)1Γ2T2,

Eq. (3)

S21=S12=R1R2T(1Γ2)1Γ2T2,
where Γ is the reflection coefficient of an infinite sample, T is the propagation factor, and R1 and R2 are the calibration plane transformation factors. These are further defined as

Eq. (4)

Γ=γ0γγ0+γ,

Eq. (5)

T=exp(γLsample),

Eq. (6)

γ=κc2εrκ02,

Eq. (7)

γ0=κc2κ02,

Eq. (8)

R1=exp(γ0L1),

Eq. (9)

R2=exp(γ0L2),
where γ0 and γ represent the wave propagation constants of the empty and sample-filled waveguides, respectively, Lsample is the sample thickness, εr is the relative permittivity of the sample, L1 and L2 are the distances between the calibration planes and the sample surfaces, κ0 is the wavenumber in free space (κ0=ω/c), and κc is the cutoff wavenumber of the TE10 (dominant) mode, given by κc=π/a, where a is the width of the waveguide.

Fig. 1

Diagram for the measurement of a dielectric sample in a rectangular waveguide. L1 and L2 can be positive (sample interface farther from a port relative to the reference plane) or negative (closer to the port). S12 and S22 are not depicted for clarity.

JARS_18_1_014501_f001.png

The calibration planes are separated by Lshim, an arbitrary length of waveguide containing the sample of interest, as illustrated in Fig. 1. For a sample of dielectric material with surfaces that are exactly at the reference planes (Lsample=Lshim), L1 and L2 are zero, leading to R1=R2=1 and, subsequently, S11=S22. This allows Eqs. (1)–(3) to be fit with only the real and imaginary portions of permittivity as free parameters. As an example, Fig. 2 shows the measured and calculated S-parameters for an empty (εr=1) section of a waveguide shim, 4.153 mm in length. Both the real and imaginary S11/S22 curves are zero as there is no sample to cause reflections, whereas S21/S12 vary sinusoidally as expected from the propagation phase. The calculated S-parameters are in excellent agreement with the measured data. In the absence of dispersion and sample inhomogeneity, a single real and imaginary permittivity value can be applied to the entire bandwidth of the waveguide measurement system; the entire set of S-parameter data can then be simulated by only two free parameters with no frequency dependence.

Fig. 2

Measured (solid) and calculated (dashed) S-parameters for an empty 4.153 mm length V-band shim. S11 and S22 overlap at zero, whereas S21 and S12 are symmetric and vary sinusoidally.

JARS_18_1_014501_f002.png

In the event that one or both of the sample surfaces are not at the calibration plane, L1 and L2 will be non-zero, causing R1 and/or R2 to no longer be in unity. This change is observed as a splitting of the S11 and S22 data that is dependent on frequency. Hence, L1 and L2 must be accounted for to properly fit the data.18 Figure 1 illustrates that L1 and L2 are positive when the sample surface is farther from the port than the calibration plane. Conversely, when the sample surface sticks out of the shim, thus moving past the reference plane, L1 and L2 are negative in sign, and Lsample becomes larger. The materials of interest are soft putties not machined to fixed dimensions. Samples are prepared by fully packing the waveguide shim, but even careful preparation and handling can lead to sample surfaces not located at the calibration planes. Changes in these surface locations could also result in an incorrect sample thickness, causing an error in the measured permittivity values. Instead of relying on a fixed sample thickness, the distance Lshim between the calibration planes is used in conjunction with L1 and L2 to obtain the sample thickness for calculating permittivity using the relation

Eq. (10)

Lsample=LshimL1L2.

Figure 3 shows an example of the effect on S-parameters produced by a shift in the sample location. Plastic explosives are weakly absorbing and typically have low permittivity at microwave frequencies,19 so a value εr=2.70.01i is adopted a priori for this simulation and will be seen later to be representative of measurements at V-band. The simulated S-parameters (solid lines) are for a 2 mm thick sample contained in a 2 mm shim in which the sample has been shifted toward port 2 by 0.05 mm. L1 is now +0.05  mm, and L2 is now 0.05  mm. The shift in the sample with respect to the calibration planes causes a splitting of the S11 and S22 real and imaginary data that varies with frequency. The addition of L1 and L2 brings the total of free parameters to four but allows for the effects seen in the simulated S-parameter data to be accounted for.

Fig. 3

Illustration of the effect on the S-parameters caused by a 0.05 mm shift in the sample location, leading to a splitting of S11 and S22 data. Parameters for the simulation were εr=2.70.01i, Lsample=Lshim=2  mm, L1=+0.05  mm, L2=0.05  mm. Solid lines represent simulated data and dashed/dotted lines represent optimized fits of the data.

JARS_18_1_014501_f003.png

3.

Experimental Procedure

A Keysight Technologies N5245A PNA-X 50 GHz network analyzer was used in conjunction with two OML Inc. V-band (WR-15) waveguide transceivers operating in the frequency range of 50 to 75 GHz. Data were collected in increments of 10 MHz (2501 points) with an intermediate frequency bandwidth of 3 kHz. The transceivers were calibrated using OML’s TRL (thru, reflect, line) calibration kit with measurements of shorts on both ports, a null thru, plus a null +1/4λ thru. A (nominally) 2 mm length shim obtained from Pasternack Enterprises Inc. was used as the sample holder, with a sample volume of 14 microliters (3.76  mm×1.88  mm×2  mm). A series of 1- and 2-port measurements, 24 in total, were performed on the 2 mm shim to determine the value of Lshim. The shim thickness was optimized to the data using Eqs. (1)–(3), using length as the only free parameter.

Table 1 contains the explosives investigated in this work. Samples were manually packed into the waveguide shim, with care taken to pack as uniformly, and as completely, as possible. A plastic spatula was used to trim excess material from the faces of the shim, bringing the sample as close to the shim thickness as possible to minimize L1 and L2. Figure 4 contains representative images of samples packed into the waveguide shim. The sample shim was inserted between the OML transceivers, and a two-port measurement was performed. Five measurements were performed, with the waveguide shim emptied and repacked after each measurement from different parts of the sample to obtain statistical variability. Experimental data were imported into MATLAB and fit to Eqs. (1)–(3) using a least squares fitting algorithm with εr (real, ε, and imaginary, ε), L1, and L2 as free parameters.

Table 1

Explosives investigated in this work.

ExplosiveManufacturer
C-4Accurate energetics
Primasheet 1000Ensign-Bickford
Primasheet 2000Ensign-Bickford
Semtex 10Explosia

Fig. 4

Images of plastic explosives loaded into a WR-15 waveguide shim from top left clockwise; C-4, Primasheet 1000, Primasheet 2000, and Semtex 10.

JARS_18_1_014501_f004.png

4.

Quality Control of Waveguide SHIMS

Figure 5 shows a representative two-port S-parameter data (solid lines) of the empty Pasternack 2 mm shim along with calculated S-parameters (dashed lines). Because the shim is empty, S11=S22 and S12=S21. Excellent agreement of the calculated and measured data is obtained using an optimum value of 2.011 mm as the only free parameter. Figure 6 plots the optimum thickness values obtained from 24 total 2-port (transmission/reflection) and 1-port (short-backed reflection-only) measurements on the empty shim. The empty shim resulted in a length value of 2.010±0.012  mm for the shim, which is consistent both with the length 2.00±0.02  mm measured with a Mitutoyo 500-197-20 digital caliper and the ±0.02  mm accuracy specified by the shim manufacturer. The average value 2.010 mm was used for Lshim for all analysis samples. Scatter in the data is likely indicative of the repeatability of the connections at the waveguide interfaces. The results obtained from analysis of the separate two-port and one-port measurements show that they are not statistically different.

Fig. 5

Representative two-port S-parameter data obtained from the empty Pasternack shim nominally 2 mm in thickness. The dashed lines represent the calculated S-parameters using an optimum value of 2.011 mm for this single measurement.

JARS_18_1_014501_f005.png

Fig. 6

Results obtained on 24 measurements of the empty sample shim using two-port and one-port (short backed) measurements. Lshim was determined to be 2.01±0.01  mm.

JARS_18_1_014501_f006.png

Additional waveguide shims were examined as part of this effort. It was determined that some parts did not conform to expectations from standard waveguide equations. Figure 7 shows representative S-parameter data on a 5 mm V-band shim from an external supplier. Despite optimizing the thickness, poor agreement is obtained between the calculated and measured data. An empty shim should produce no reflections (S11=S22=0), yet signal is observed. Further investigation was performed. Figure 8 presents images obtained with an optical microscope on the 2 mm shim used for data collection in this work [Fig. 8(a) left] and a 5 mm shim from another supplier [Fig. 8(b) right]. The walls and corners of the shim on the right are poorly formed, leading to the data obtained in Fig. 7. As a result of this investigation, this shim was rejected for use in the laboratory. A general rule for the acceptability of the experimental setup can be defined from the reflection coefficient of the shim. The S11 reflection coefficient from the rejected shim was 0.1 or 20  dB. The rule requires that the reflection from the shim be less than 10% of the reflection coefficient being measured. For a reflection coefficient of 0.14, the reflection from the shim can be specified to be less than 0.014 or 37  dB, to avoid the problems evident in Fig. 7.

Fig. 7

Representative two-port S-parameter data obtained from the empty nominally 5 mm shim obtained from a waveguide manufacturer. Poor agreement is obtained between the calculated and measured data. S11 and S22 values are zero for an ideal empty waveguide.

JARS_18_1_014501_f007.png

Fig. 8

Microscope images of V-band shim used for measurements in this work (a) versus another manufacturer (b). Note the poor uniformity of the machining of the walls and corners. The shim on the right was rejected for use in dielectric measurements.

JARS_18_1_014501_f008.png

5.

Results of Dielectric Measurements

A representative set of two-port S-parameters obtained from a sample of C-4 is presented in Fig. 9. Measured data are presented as solid lines, and calculated data are shown in dashed/dotted lines. Optimizing the scattering functions with the four free parameters—ε, ε, L1, and L2—yields excellent agreement of the calculated S-parameters to the measured data. The optimized values from this measurement (measurement 1) and four others are presented in Table 2. From the five measurements, C-4 is determined to have an average permittivity value of εr=3.19(5)0.029(3)i; the uncertainties in the last digits are given in parentheses. Uncertainty values were calculated from the standard deviation of the measurements and uncertainty in the thickness of the sample shim. This procedure was repeated for Primasheet 1000, Primasheet 2000, and Semtex 10. Table 3 provides the results obtained for all four explosives studied in this work along with data on C-4 and Primasheet 1000 available in the literature. Results obtained for C-4 and Primasheet 1000 in this work are comparable to the literature values, providing confidence in the results. It is also noted that, as manufactured products, samples of plastic explosives can be subject to variations in formulation and uniformity that could account for the slight differences in measured permittivity values.

Fig. 9

2-port S-parameter measured and calculated data for a sample of C-4 packed in the 2 mm waveguide shim. Optimized values for εr, L1, and L2 for this data are listed in Table 2 under measurement 1. The calculated S-parameters are in excellent agreement with the measurement.

JARS_18_1_014501_f009.png

Table 2

Results of dielectric measurements for samples of C-4. Real and imaginary permittivity, L1 and L2, are optimized; Lsample is determined using Eq. (10), and Lshim=2.010  mm.

Measurementε′ε″L1 (mm)L2 (mm)Lsample (mm)
13.2400.0280.030−0.0232.003
23.2180.0300.1020.0681.840
33.1630.0330.195−0.0671.882
43.1570.0280.0450.0891.876
53.1970.027−0.0030.0621.952
Average3.190.029
Std. Dev.0.040.002

Table 3

Compilation of plastic explosive dielectric data as a function of frequency. The uncertainties in the last digits are given in parentheses.

MaterialFrequency (GHz)ε′ε″TechniqueSource
C-450 to 753.19(5)0.029(3)Loaded waveguideThis work
C-4753.280.008THz-TDS20
C-450 to 1003.10 to 0.2Quasioptical21
Primasheet 100050 to 752.83(4)0.061(2)Loaded waveguideThis work
Primasheet 100060 to 902.99(7)0.086(7)Free space reflectometry10
Primasheet 100018 to 402.86(2)0.105(7)Free space reflectometry22
Primasheet 200050 to 753.19(7)0.021(2)Loaded waveguideThis work
Semtex 1050 to 752.84(5)0.048(3)Loaded waveguideThis work

6.

Conclusions

Complex permittivity values for C-4, Primasheet 1000, Primasheet 2000, and Semtex 10 in the operating frequency of 50 to 75 GHz were obtained. Excellent agreement between this experiment and the fit was obtained using a constant permittivity across the waveguide band, indicating that dispersion was not significant for these materials. Permittivity values for C-4 and Primasheet 1000 were comparable to values reported in the literature in the millimeter wave (MW) range; the data on Primasheet 2000 and Semtex 10 are the first known published permittivity values in this range.

Disclosures

The authors affirm that there are no conflicts of interest with this work.

Code and Data Availability

The datasets used and analyzed in the current study are publicly available in the Zenodo repository at https://zenodo.org/records/10213766.23 The lsqcurvefit function in MATLAB version 2019a was used to fit the data using the theory presented in the paper.

Acknowledgments

This study was funded by the U.S. Department of Homeland Security (DHS) Science and Technology Directorate (S&T). Battelle/MAG Aerospace were supported under DHS (Grant No. 70RSAT19FR0000136). The views and conclusions contained in this document are those of the authors and should not be interpreted as representing the official policies, either expressed or implied, of the DHS S&T or the U.S. Government. The U.S. Government is authorized to reproduce and distribute reprints for Government purposes not withstanding any copyright notation herein. References or images herein to any specific commercial products, processes, equipment, or services do not constitute or imply its endorsement, recommendation, or favoring by the U.S. Government or DHS, or any of its employees or contractors. This paper was presented at a conference: Z. J. Landicini, J. Barber, J. C. Weatherall, D.C. Karns, P.R. Smith, J Aparicio-Bolano, and W. Ruiz, “Loaded waveguide measurements of plastic explosives at V-band,” Proc. SPIE, vol. 12535, pp. 156-166, 2023.

References

1. 

D. Karns et al., “Simulant development of a powder explosive material for a millimeter wave imaging system using a resonant cavity with ensemble averaging and the Houston criterion,” Proc. SPIE, 11745 117450K https://doi.org/10.1117/12.2585931 PSISDG 0277-786X (2021). Google Scholar

2. 

C. Richard et al., “A protocol for simulant validation using active millimeter-wave imaging systems,” Proc. SPIE, 11411 114110A https://doi.org/10.1117/12.2558856 PSISDG 0277-786X (2020). Google Scholar

3. 

D. Karns et al., “Simulant development for liquid explosive materials using millimeter wavelength characterization and ensemble averaging,” Proc. SPIE, 11411 1141109 https://doi.org/10.1117/12.2560190 PSISDG 0277-786X (2020). Google Scholar

4. 

J. C. Weatherall et al., “Suitability of explosive simulants for millimeter-wave imaging detection systems,” Proc. SPIE, 10994 109940G https://doi.org/10.1117/12.2521723 PSISDG 0277-786X (2019). Google Scholar

5. 

J. Barber et al., “Millimeter wave measurements of explosives and simulants,” Proc. SPIE, 7670 76700E https://doi.org/10.1117/12.848794 PSISDG 0277-786X (2010). Google Scholar

6. 

, “QPS quick personnel security scanners,” www.rohde-schwarz.com/securityscanner (2022). Google Scholar

7. 

S. S. Ahmed, A. Schiessl and L. P. Schmidt, “A novel fully electronic active real-time imager based on a planar multistatic sparse array,” IEEE Trans. Microwave Theory Tech., 59 3567 –3576 https://doi.org/10.1109/TMTT.2011.2172812 IETMAB 0018-9480 (2011). Google Scholar

8. 

S. S. Ahmed et al., “Advanced microwave imaging,” IEEE Microwave Mag., 13 26 –43 https://doi.org/10.1109/MMM.2012.2205772 1527-3342 (2012). Google Scholar

9. 

A. Schiessl, S. S. Ahmed and L.-P. Schmidt, “Motion effects in multistatic millimeter-wave imaging systems,” Proc. SPIE, 8900 890007 https://doi.org/10.1117/12.2029348 PSISDG 0277-786X (2013). Google Scholar

10. 

P. R. Smith et al., “Measurements of the dielectric properties of explosives and inert materials at millimeter wave frequencies (V-band and above) using free space reflection methods,” Proc. SPIE, 10189 1018908 https://doi.org/10.1117/12.2267195 PSISDG 0277-786X (2017). Google Scholar

11. 

N. Reyes et al., “Complex dielectric permittivity of engineering and 3D-printing polymers at Q-band,” J. Infrared, Millimeter, Terahertz Waves, 39 1140 –1147 https://doi.org/10.1007/s10762-018-0528-9 (2018). Google Scholar

12. 

J. Baker-Jarvis, E. J. Vanzura and W. A. Kissick, “Improved technique for determining complex permittivity with the transmission/reflection method,” IEEE Trans. Microwave Theory Tech., 38 1096 –1103 https://doi.org/10.1109/22.57336 IETMAB 0018-9480 (1990). Google Scholar

13. 

S. Kim, J. Surek and J. Baker-Jarvis, “Electromagnetic metrology on concrete and corrosion,” J. Res. Natl. Inst. Stand. and Technol., 116 655 https://doi.org/10.6028/jres.116.011 JRITEF 1044-677X (2011). Google Scholar

14. 

K. J. Bois et al., “Dielectric plug-loaded two-port transmission line measurement technique for dielectric property characterization of granular and liquid materials,” IEEE Trans. Instrum. Meas., 48 1141 –1148 https://doi.org/10.1109/19.816128 IEIMAO 0018-9456 (1999). Google Scholar

15. 

J. Hunger et al., “Precision waveguide system for measurement of complex permittivity of liquids at frequencies from 60 to 90 GHz,” Rev. Sci. Instrum., 82 104703 https://doi.org/10.1063/1.3647982 RSINAK 0034-6748 (2011). Google Scholar

16. 

ASTM D 5568-22, Standard Test Method for Measuring Relative Complex Permittivity and Relative Magnetic Permeability of Solid Materials at Microwave Frequencies, 10 ASTM International, West Conshohocken, Pennsylvania (2022). Google Scholar

17. 

U. C. Hasar and C. R. Westgate, “A broadband and stable method for unique complex permittivity determination of low-loss materials,” IEEE Trans. Microwave Theory Tech., 57 471 –477 https://doi.org/10.1109/TMTT.2008.2011242 IETMAB 0018-9480 (2009). Google Scholar

18. 

C. Yang and H. Huang, “Determination of complex permittivity of low-loss materials from reference-plane invariant transmission/reflection measurements,” IEEE Access, 7 131865 –131872 https://doi.org/10.1109/ACCESS.2019.2940723 (2019). Google Scholar

19. 

A. L. Higginbotham Duque, W. L. Perry and C. M. Anderson-Cook, “Complex microwave permittivity of secondary high explosives,” Propell. Explos. Pyrotech., 39 275 –283 https://doi.org/10.1002/prep.201300032 PEPYD5 0721-3115 (2014). Google Scholar

20. 

K. Yamamoto et al., “Noninvasive Inspection of C-4 Explosive in Mails by Terahertz Time-Domain Spectroscopy,” Jpn. J. Appl. Phys., Part 2, 43 L414 –L417 https://doi.org/10.1143/JJAP.43.L414 JAPLD8 0021-4922 (2004). Google Scholar

21. 

E. Sáenz et al., “Sub-millimetre wave material characterization,” in Proc. 5th Eur. Conf. Antennas and Propag. (EUCAP), 3183 –3187 (2011). Google Scholar

22. 

J. C. Weatherall, J. Barber and B. T. Smith, “Spectral signatures for identifying explosives with wideband millimeter-wave illumination,” IEEE Trans. Microwave Theory Tech., 64 999 –1005 https://doi.org/10.1109/TMTT.2016.2518159 IETMAB 0018-9480 (2016). Google Scholar

23. 

EMXLAB DHS/S&T/TSL, “Loaded waveguide measurements of plastic explosives at v-band,” (2023). Google Scholar

Biography

Zachary J. Landicini is a scientist for the Battelle Memorial Institute, which supports the U.S. Department of Homeland Security Transportation Security Laboratory (TSL) in Egg Harbor Township, New Jersey, United States. He received his BS degree in biochemistry/molecular biology from Stockton University in 2015. Presently, he is pursuing a master of science degree from Rowan University in electrical and computer engineering.

Jeffrey Barber received his BS degree in chemistry from Allegheny College in 1998 and his PhD in chemistry from Oregon State University in 2003. His work has included dielectric measurements of explosives, developing inert simulants and image quality tools for active millimeter wave imaging systems, and the quality control of synthesized explosives via infrared and Raman spectroscopy. In addition to co-authoring numerous publications and conference proceedings, he has been awarded 18 US patents.

James C. Weatherall supports explosives detection science as a physical scientist in the Applied Research Division of the TSL. He received his BS degree from Caltech and his PhD in plasma physics from the University of Colorado. He is actively developing methods to improve the airport screening process using the response of potential threats to millimeter-waves during screening. Before coming to the TSL, he worked in high-power microwaves and astrophysics.

Duane C. Karns is a physical scientist within the ARD division of the TSL. He received his BS degree from Pennsylvania State University and his PhD in electrical and computer engineering from Carnegie Mellon University, where he specialized in near-field optical systems. His work at the TSL involves MMW characterization of material dielectric properties, including explosives and their simulants and the development of synthetic data for MMW and X-ray imaging systems.

Peter R. Smith received his BSEE degree from Northeastern University, Boston, Massachusetts, United States, and his dual MSEE degrees from Columbia University, New York, New York, United States. He is an RF research engineer investigating the interaction of millimeter electromagnetic waves and materials for MAG Aerospace/Battelle Memorial Institute as a contractor for the DHS Science and Technology Directorate at the TSL. He enjoys working on knotty technical problems.

Joaquín Aparicio-Bolaño received his MS degree in physics from the University of Puerto Rico at Mayagüez and his PhD in chemical physics from the University of Puerto Rico in Río Piedras in 2011. In the same year, he joined ALERT DHS-COE at UPRM. He is currently a staff physicist at the Battelle Memorial Institute. His current research interests at TSL focus on the physical and chemical characterization of high energetic materials.

Wendy Ruiz is a test engineer for the Battelle Memorial Institute working in the Electromagnetic Signatures of Explosives Laboratory at the TSL. She received her bachelor of science degree in biochemistry and biology with a pre-med concentration from Rowan University in 2010. She also has a degree in Spanish literature with a concentration in international studies.

CC BY: © The Authors. Published by SPIE under a Creative Commons Attribution 4.0 International License. Distribution or reproduction of this work in whole or in part requires full attribution of the original publication, including its DOI.
Zachary J. Landicini, Jeffrey Barber, James C. Weatherall, Duane C. Karns, Peter R. Smith, Joaquín Aparicio-Bolaño, and Wendy Ruiz "Loaded waveguide measurements of plastic explosives at V-band," Journal of Applied Remote Sensing 18(1), 014501 (29 December 2023). https://doi.org/10.1117/1.JRS.18.014501
Received: 16 October 2023; Accepted: 6 December 2023; Published: 29 December 2023
Advertisement
Advertisement
KEYWORDS
Waveguides

Explosives

Plastics

V band

Dielectrics

Calibration

Reflection

Back to Top