Sie sind auf Seite 1von 4

VOL. 3, NO.

10, OCTOBER 2019 2000504

Sensor materials

Orthogonal Sensors for the Trace Detection of Explosives


Peter P. Ricci1,2 , Andrew S. Rossi1,2 , and Otto J. Gregory1.2
1 Sensors and Surface Technology Partnership, University of Rhode Island, Kingston, RI 02881 USA
2 Department of Chemical Engineering, University of Rhode Island, Kingston, RI 02881 USA

Manuscript received June 11, 2019; revised July 9, 2019 and August 2, 2019; accepted September 21, 2019. Date of publication October 2, 2019; date of
current version October 15, 2019.

Abstract—There is a growing need for an electronic trace detection system that can continuously monitor explosives at
trace levels in the vapor phase. Triacetone triperoxide (TATP) is a common explosive used by terrorists as the initiator
or the energetic material in improvised explosive devices. TATP still goes largely undetected in many densely populated
public venues. Currently, no detection system exists that is capable of continuously monitoring TATP, as well as nitrogen-
based explosives such as 2,4-dinitrotoluene. We have demonstrated a thermodynamic sensor platform employing thin
film microheaters and metal oxide catalysts that can detect TATP and 2,4-DNT at the parts per billion (ppb) level.
The microheaters and catalysts were deposited onto ultrathin alumina ceramic substrates, which are responsible for
extraordinary sensitivity. Recently, we have added a conductometric sensor platform to this thermodynamic platform to
form an orthogonal sensor capable of interrogating the same catalyst using two different modalities. The orthogonal sensor
probes the same catalyst using a thermodynamic protocol, as well as a conductometric protocol where electrical resistivity
changes in the catalyst are measured as a function of analyte exposure. The advantages of this added orthogonality in
terms of mitigating false-positives and negatives are described within.

Index Terms—Sensor materials, catalysts, explosive vapor detection, orthogonality, resistivity.

I. INTRODUCTION [3]; however, because of the temperature dependence on separation


time, this process is too time consuming to be implemented in fast
Improvised explosive devices (IEDs) are commonly used by ter- moving security venues.
rorist organizations to threaten and/or cause harm to the general pub- Ion mobility spectrometry (IMS) is a detection method that is ex-
lic. The IEDs frequently employ triacetone triperoxide (TATP) as tremely fast (about 3 s to positively identify explosives); however, the
either the detonator or the energetic material itself. This is largely IMS is marginal for detecting explosives in the vapor [4]–[6]. Liquid
due to the ease of fabrication and availability of its precursors. In chromatography [7], as well as dielectric barrier discharge ionization-
just the past three years, terrorist attacks on airports, concerts, and mass spectrometry [8], require solids or solutionized explosives, and
even churches have employed IEDs using TATP. Both TATP and trini- therefore, would be ineffective for detection unless swabbed for
trotoluene (TNT) decompose and sublime from solid to vapor. The explosives.
presence of these explosives in the vapor phase provides the ability to Other researchers have been able to successfully detect explosive
establish a detection protocol for vapors. Due to the low vapor pressure vapors using conductometric techniques whereby electrical resistance
of TNT, detection relies on the decomposition product of TNT; i.e., changes in a metal oxide are measured as the explosive vapor interacts
2,4-DNT, which has a significantly higher vapor pressure, making it with it. However, conductometric sensors typically exhibit poor selec-
easier to detect. tivity, which has stimulated researchers to look for different strategies
Many platforms exist that are capable of detecting either solid ex- to overcome this drawback. To our knowledge, a sensor that probes
plosive or explosive vapor at very low concentrations. But each of the same metal oxide catalyst using a thermodynamic protocol as well
these platforms has their own drawbacks in terms of detecting ex- as conductometric protocol has never been demonstrated. The results
plosives quickly and effectively. For example, Mullen et al. [2] used presented within suggest that our conductometric platform provides a
single photon laser ionization techniques in conjunction with time of built-in redundancy that insures a much lower rate of false positives
flight mass spectrometry to detect certain explosives at the parts per than other techniques previously reported [9], [10]. Our orthogonal
billion (ppb) level in the vapor phase. However, this technique requires sensor platform first described by Mallin [11] is unique in that both
significant power and complex equipment that would not be feasible sensing modalities (conductometric and thermodynamic) are function-
from a maintenance and portability viewpoint. alized on the same substrate and the same catalyst is simultaneously
Fluorescence and gas chromatography techniques have been inves- interrogated using these two independent modalities.
tigated in an attempt to identify TATP in the vapor. However, the TATP As previously described by Rossi et al. [12], our thermodynamic
does not exhibit fluorescence [1], and gas chromatography is difficult sensor relies on two separate microheaters: one coated with a metal
to perform due to the tendency for TATP to decompose under moderate oxide catalyst and an uncoated reference microheater. The reference
heating [1]. It is possible to detect TATP by gas chromatography/mass is used to distinguish between heat effects that arise from catalyst-
spectrometry and mass chromatography/tandem mass spectrometry analyte interactions and those due to hydrodynamic effects such as
sensable heat effects. By subtracting the reference signal from the
catalyst-coated signal, the resulting difference is the heat effect due to
Corresponding author: Peter P. Ricci (e-mail: pricci3@my.uri.edu).
catalytic decomposition. The thermodynamic platform relies on oxi-
Associate Editor: J. Gardner.
Digital Object Identifier 10.1109/LSENS.2019.2944587 dation/reduction reactions on the catalyst surface after decomposition

1949-307X C 2019 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission.

See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.


2000504 VOL. 3, NO. 10, OCTOBER 2019

Fig. 2. Schematic of an apparatus used for explosives testing of an


orthogonal sensor.
Fig. 1. Photograph of an orthogonal sensor (a) and top view of an
orthogonal sensor (b). A cutaway view (c) of an orthogonal sensor
shows the various layers comprising the sensor: A metal oxide (catalyst) Analyte was introduced to the system using a 2-chamber apparatus
layer (1), nickel interdigitated electrodes (2), alumina coatings (3) and (see Fig. 2). Initially, dry air from a pressurized gas cylinder was
(4), nickel microheater (5), and alumina substrate (6). delivered to an empty cylinder where the flow was split evenly such that
identical streams were delivered to the catalyst-coated and reference
of the explosive molecule. These reactions can be exothermic or en- microheaters. Constant volumetric flow was maintained using mass
dothermic depending on the temperature of the sensor, the analyte, flow controllers (Allicat Scientific Mass Flow Controllers with Flow
and the metal oxide catalyst. When the explosive molecule interacts vision software). At this point, the microheater sensors were heated to
with the heated catalyst, the molecule decomposes and either oxidizes a predetermined set point temperature, and a power/resistance baseline
or reduces the metal oxide (catalyst), which in turn changes the oxi- was established using this carrier gas (air).
dation state in the metal oxide from a higher state to a lower state. The At the start of a test, the carrier gas (air) was diverted from the empty
aforementioned conductometric modality works simultaneously with chamber to an analyte chamber containing either solid 2,4-DNT or
the thermodynamic modality in that as the explosive molecule inter- TATP. Flow was diverted using mass flow controllers to control the
acts with the metal oxide catalyst, electrical resistance of the metal volumetric flow rate. Explosive vapors were picked up by the dry air
oxide is monitored and recorded. Resistivity of the metal oxide can carrier gas and delivered downstream to the active sensor elements. A
increase or decrease depending on the oxidation/reduction potential schematic of this apparatus is shown in Fig. 2.
of the decomposition products. When an explosive molecule interacts with the metal oxide cata-
lyst, the heat generated or absorbed by the catalyst in the presence of
the explosive molecule results in a change in electrical power. As a
II. EXPERIMENTAL result of this change in power, the software maintains a constant tem-
The orthogonal sensor is constructed on ultrathin alumina ceramic perature set point by either adding or subtracting electrical power to
substrates measuring 1.6 × 0.7 cm. Photolithography and radio fre- the catalyst-coated microheater. The electrical power change required
quency (RF) sputtering techniques were used to pattern nickel mi- to maintain a constant temperature is the thermodynamic heat effect
croheaters and interdigitated electrodes on these substrates. After the associated with the analyte-catalyst interaction. This power difference
nickel films were deposited, a 20 μm thick layer of alumina cement was monitored and recorded by proprietary software for signal pro-
was applied over the microheater to form a passivation layer between cessing and compilation of the data. During a typical experimental
the microheater and metal oxide catalyst. After curing the alumina ce- run, the microheater was cycled to a temperature of 500 °C and a
ment, the metal oxide catalyst was sputter-deposited onto the cement. 15 mA current was supplied from the constant current source to the
In order to measure the resistance changes across the metal oxide interdigitated fingers deposited on the metal oxide. Each experiment
catalyst after introduction of the analyte, interdigitated “fingers” or required several minutes to complete since the sensors were allowed
nickel electrodes were patterned onto the catalyst surface. Thus, the to reach peak response, i.e., the power at which the sensor reaches
orthogonal sensor was comprised of six leads for attachment to the equilibrium depends on the response to a specific analyte.
connector: four leads for the nickel microheaters and two leads for
electrical resistance measurement. The four leads to each microheater III. RESULTS AND DISCUSSION
permitted communication between the microheaters and the digital
control system. A photograph and schematic of the orthogonal sensor Figs. 3 and 4 show two orthogonal sensor responses to TATP and
used for trace detection of explosives is shown in Fig. 1. 2,4-DNT in the vapor phase using a SnO2 catalyst. As shown in Figs.
The metal oxide catalysts investigated in this study include SnO2 , 3 and 4, both TATP and 2,4-DNT have an endothermic thermody-
Cu2 O, ZnO, FeO, and WO. However, SnO2 , Cu2 O, and ZnO catalysts namic response when using a SnO2 catalyst. Here, the SnO2 catalyst
will be the primary focus of this article. A digital control system sup- is reduced upon analyte interaction (decomposition products derived
plies electrical power to the microheaters such that resistance heating from the TATP and 2,4-DNT explosives), going from a higher ox-
can be controlled using proprietary software. The exact temperature idation state to a lower oxidation state. This reduction reaction is
of the microheater can be determined from the temperature coefficient endothermic in nature, and thus, the power to the sensor is increased
of resistance or TCR and resting resistance of the nickel microheater in order to maintain a constant temperature. For our purposes, sensi-
at room temperature. tivity is defined as a ratio of the thermodynamic, or conductometric,
A constant current source supplied a small current to the interdigi- response to the concentration of the target molecule. The thermody-
tated fingers such that electrical resistance of the metal oxide could be namic responses shown in Figs. 3 and 4 translate into sensitivities of
measured. The current was varied from 1 to 15 mA, and the nominal 0.5 mW/ppm for TATP and 28 mW/ppm for 2,4-DNT, respectively.
electrical resistance was on the order of 1000 . The conductometric response corresponds to the change in charge
VOL. 3, NO. 10, OCTOBER 2019 2000504

Fig. 3. Orthogonal sensor responses to 20 ppm TATP utilizing a SnO2


catalyst. Thermodynamic response is highlighted in blue, while conduc-
Fig. 5. Orthogonal sensor responses to 180 ppb 2,4-DNT utilizing a
tometric response is highlighted in orange.
ZnO catalyst. Thermodynamic response is highlighted in blue, while
conductometric response is highlighted in red.

Fig. 4. Orthogonal sensor responses to 180 ppb 2,4-DNT utilizing a Fig. 6. Orthogonal sensor responses to 180 ppb 2,4-DNT utilizing a
SnO2 catalyst. Thermodynamic response is highlighted in blue, while Cu2 O catalyst. Thermodynamic response is highlighted in blue, while
conductometric response is highlighted in orange. conductometric response is highlighted in orange.

carrier concentration in the metal oxide as a result of the correspond- A major drawback of current conductometric platforms is an in-
ing redox reaction. In Figs. 3 and 4, the SnO2 undergoes a reduction herent limit on selectivity [14]. As shown in the responses above,
reaction in which the number of oxygen vacancies is increased. The sensors employing post-transition metal oxide catalysts display sim-
negative response of the conductometric platform is consistent with ilar responses to various explosive analytes. To address this, certain
this reduction reaction in that the resistance of the metal oxide de- semiconducting metal oxides, such as CuO, have been used to im-
creases as it interacts with the explosive molecule due to the increase prove selectivity [14]. Here, the CuO possess a filled valence electron
in oxygen vacancies. These conductometric responses translate into shell and a small band gap (1.2 eV), allowing for further exchange of
sensitivities of 1.5%/ppm for TATP and 94%/ppm for 2,4-DNT. electrons and, thus, better sensitivity. Fig. 6 shows the response of an
It has been shown that post-transition metal oxides such as SnO2 orthogonal sensor to 2,4-DNT in the vapor phase using a Cu2 O cata-
and ZnO possess d0 and d10 electron configurations, which can result lyst. Again, the sensors were heated to 500 °C and a 15 mA current was
in improved sensitivity for conductometric sensor applications [13]. supplied to the sensor using a constant current source. Here, the Cu2 O
These oxides possess relatively small band gaps (3–4 eV) and, thus, catalyst was oxidized upon decomposition of the TATP and 2,4-DNT
permit greater exchange of electrons between the catalyst and analyte molecules, resulting in an exothermic response. This response is a
[14]. In order to test the effectiveness of these post-transition metal result of the transition from a lower oxidation state to a higher oxida-
oxides, orthogonal sensors employing a ZnO catalyst were fabricated. tion state, and thus, a decrease in oxygen vacancies and corresponding
As shown in Fig. 5, the ZnO orthogonal sensors exhibited endothermic decrease in charge carrier concentration. This oxidation reaction, as
responses to 2,4-DNT. Here, the ZnO catalyst undergoes a reduction opposed to a reduction reaction, makes the Cu2 O catalyst uniquely se-
reaction, going from a higher oxidation state to a lower oxidation lective to the 2,4-DNT molecule. Here, the thermodynamic response
state, which requires more energy to maintain a constant temperature. displayed a sensitivity of 140 mW/ppm, which represents a significant
This orthogonal sensor exhibited a similar response to 2,4-DNT to increase over previous catalysts. The conductometric response here is
a sensor employing an SnO2 catalyst as shown in Fig. 4. A similar positive (electrical resistance of the Cu2 O increases) upon interaction
negative conductometric response also coincided with the change in with the decomposition products of 2,4-DNT, and an increase in sen-
charge carrier concentration based on the redox reaction. The ZnO sitivity of 1250%/ppm was observed. Compared to orthogonal sensors
catalyst displayed sensitivities of 33 mW/ppm for the thermodynamic employing SnO2 or ZnO catalysts, sensors employing Cu2 O catalysts
platform and 39%/ppm for the conductometric platform. displayed an extraordinary increase in sensitivity to 2,4-DNT.
2000504 VOL. 3, NO. 10, OCTOBER 2019

IV. CONCLUSION [3] M. E. Sigman et al., “Analysis of triacetone triperoxide by gas chromatogra-
phy/mass spectrometry and gas chromatography/tandem mass spectrometry by elec-
Overall, integration of a conductometric platform with the thermo- tron/chemical ionization,” Rapid Commun. Mass Spectr., vol. 20, no. 19, pp. 2851–
dynamic platform lead to unparalleled orthogonality for trace detec- 2857, Aug. 2006.
[4] G. A. Buttigieg, A. K. Knight, S. Denson, C. Pommier, and M. B. Denton, “Char-
tion. Here, metal oxide catalysts are utilized to measure the heat effect
acterization of the explosive triacetone triperoxide and detection by ion mobility
associated with selective redox reactions that occur on the catalyst sur- spectrometry,” Forensic Sci. Int., vol. 135, pp. 53–59, 2003.
face. Simultaneously, electrical resistance changes in the oxide that [5] R. Rasanen et al., “Determination of gas phase triacetone triperoxide with aspira-
occur as a result of the catalytic redox reactions are measured. The tion ion mobility spectrometry and gas chromatography–mass spectrometry,” Anal.
experimental results suggest that by combining both sensor platforms, Chim. Acta., vol. 623, pp. 59–65, 2008.
[6] R. G. Ewing, D. A. Atkinson, G. Eiceman, and G. Ewing, “A critical review of
highly selective responses to TATP and 2,4-DNT resulted in unique ion mobility spectrometry for the detection of explosives and explosive related
signal-temperature signatures for each explosive so that false positives compounds,” Talanta, vol. 54, pp. 515–529, 2001.
could be mitigated. Additionally, by implementing Cu2 O as the cata- [7] R. Schulte-Ladbeck et al., “Determination of peroxide-based explosives using
lyst for our orthogonal sensor, improved sensitivity and selectivity to liquid chromatography on-line infrared detection,” Anal. Chem., vol. 78, no. 23,
pp. 8150–8155, Dec. 2006.
2,4-DNT was realized.
[8] S. Hagenhoff et al., “Determination of peroxide explosive TATP and related com-
pounds by dielectric barrier discharge ionization-mass spectrometry (DBDI-MS),”
ACKNOWLEDGMENT Anal. Chem., vol. 89, no. 7, pp. 4210–4215, Mar. 2017.
[9] J. Li et al., “Carbon nanotube sensors for gas and organic vapor detection,” Nano
This work was supported by the U.S. Department of Homeland Security, Science and Lett., vol. 3, no. 7, pp. 929–933, Jun. 2003.
Technology Directorate, Office of University Programs, under Grant 50503578059. The [10] H. Lin and S. S. Kenneth, “A colorimetric sensor array for detection of triacetone
views and conclusions contained in this document are those of the authors and should not triperoxide vapor,” J. Amer. Chem. Soc., vol. 132, no. 44, pp. 15519–15521, Oct.
be interpreted as necessarily representing the official policies, either expressed or implied, 2010.
of the U.S. Department of Homeland Security. [11] D. Mallin, “Increasing the selectivity and sensitivity of gas sensors for the detection
of explosives,” M.S. Thesis, Univ. Rhode Island, Kingston, RI, USA, 2014.
[12] A. Rossi, P. Ricci, and O. J. Gregory, “Trace detection of explosives using metal
oxide catalysts,” IEEE Sens. J., vol. 19, no. 13, pp. 4773–4780, Jul. 2019.
REFERENCES [13] C. Wang et al., “Metal oxide gas sensors: Sensitivity and influencing factors,”
Sensors, vol. 10, no. 3, pp. 2088–2106, Mar. 2010.
[1] R. S. Ladbeck, P. Kolla, and U. Karst, “Trace analysis of peroxide-based explosives,” [14] S. Kanan et al., “Semiconducting metal oxide based sensors for selective gas
Anal. Chem., vol. 75, pp. 731–735, 2003. pollutant detection,” Sensors, vol. 9, no. 10, pp. 8158–8196, Oct. 2009.
[2] C. Mullen, A. Irwin, B. V. Pond, D. L. Huestis, M. J. Coggiola, and H. Oser,
“Detection of explosives and explosives-related compounds by single photon laser
ionization time-of-flight mass spectrometry,” Anal. Chem., vol. 78, pp. 3807–3814,
2006.

Das könnte Ihnen auch gefallen