Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

International Journal of Hydrogen Energy 55 (2024) 875–881

Contents lists available at ScienceDirect

International Journal of Hydrogen Energy


journal homepage: www.elsevier.com/locate/he

Hydrogen sensor with a thick catalyst layer anchored on soda-lime glass


Gustavo Panama a, Munhyung Jo a, Soon Seop Shim b, Min Su Kim b, Seung S. Lee a, *
a
Department of Mechanical Engineering, Korea Advanced Institute of Science and Technology, Daejeon, 34141, Republic of Korea
b
Catalyst Team, Hydrochem Co., Ltd., Gangneung, 25440, Republic of Korea

A R T I C L E I N F O A B S T R A C T

Handling editor: Umit Demirci Hydrogen is a green energy source with zero pollutant emissions and increasing demand. Hydrogen sensors are
required due to the flammable nature of hydrogen in air. Current catalytic combustion hydrogen sensors based
Keywords: on membrane platforms strike a balance between sensitivity and mechanical strength. A thick catalyst layer
Hydrogen sensor allows enhanced sensor sensitivity, producing an acceptable sensor signal change. A hydrogen sensor with a thick
Thick catalyst
catalyst layer is introduced on soda-lime glass, which has low thermal conductivity and prefabricated catalyst
Catalyst anchored
anchoring channels. The hydrogen sensor is composed of a 44 Ω platinum sensing electrode and a 100 Ω gold
Soda-lime glass
Catalytic combustion microheater, both of which are coated with a thin layer of alumina. The sensing electrode is buried in the catalyst
anchoring channels with a 10 μm depth and 78 μm width using a hydrogen combustible catalyst. The response
for 1 % H2 in air is 9.43 % in a test chamber at 80 ◦ C with a response time of 48 s and recovery time of 75 s. The
calculated temperature increment due to catalytic combustion is 84 ◦ C. A hydrogen sensor with a thick catalyst
layer on soda-lime glass shows enhanced heat generation, resulting in an improved thermoresistive response, and
consistent mechanical support.

1. Introduction combustion of hydrogen [15] in which catalyst design allows lower


working temperatures (25–100 ◦ C) [16–18].
Hydrogen is a strategic energy source for reducing the use of fossil A catalyst is required to enable hydrogen spillover reaction in the
fuels in industrial applications, including petrochemical refining, catalytic hydrogen combustion [19,20]. Pt and Pd nanoclusters show
transportation, and energy storage [1]. Interest in hydrogen energy is high activity for the hydrogen spillover, which strongly relies on the
due to its zero pollutant emissions and ability to mitigate climate change support type such as TiO2 [21,22]. Catalytic hydrogen combustion in­
[2]. Further developments in hydrogen infrastructure and vehicles could volves absorption and adsorption of hydrogen atoms on Pd and Pt sur­
enable widespread uses of hydrogen by 2050 [3]. Hydrogen technolo­ faces, respectively [23,24]. Pt containing catalysts are preferred due to
gies have been adopted with roadmaps including production of fuel cell enhanced catalytic activity. Pt based bimetallic catalyst shows improved
electric vehicles and power generation [4–6]. However, hydrogen in air catalytic activity and good reliability [25]. Pt/TiO2 catalyst results in
is readily flammable and demands high safety standards [7,8]. Thus, enhanced catalytic activity due to highly reactive Pt nanoparticles
hydrogen leak detection from production to utilization requires reliable dispersed and stabilized on the TiO2 support, which could enable sensor
hydrogen sensors. application at reduced temperatures.
Various types of hydrogen sensors have been investigated [9,10]. Previous catalytic combustion hydrogen sensors were fabricated
Among them, semiconductive metal oxide and catalytic using silicon and consisted of a sensing electrode and microheater on a
combustion-type hydrogen sensors have shown promise due to their membrane support, as shown in Fig. 1 [26,27]. Membrane platforms
simple sensing mechanisms based on resistance change [11]. Semi­ ensured rapid response and low parasitic heat transfer to silicon bulk
conductive metal oxide sensors directly react with hydrogen gas to substrate [16,28]. However, membrane platforms strike a balance be­
modulate charge carrier concentration, but they have the disadvantage tween sensitivity and mechanical strength, as the deposition of a thick
of elevated operating temperatures (150–400 ◦ C) [12–14]. Catalytic catalyst layer to enhance signal sensitivity compromises structural sta­
combustion sensors use a thin platinum layer acting as a resistance bility of the membrane support [29–31]. Additionally, catalyst deposi­
thermometer to detect the generated heat due to the catalytic tion is challenging on membrane supports due to its reduced mechanical

* Corresponding author.
E-mail address: sslee97@kaist.ac.kr (S.S. Lee).

https://doi.org/10.1016/j.ijhydene.2023.11.271
Received 2 August 2023; Received in revised form 20 November 2023; Accepted 23 November 2023
Available online 29 November 2023
0360-3199/© 2023 The Authors. Published by Elsevier Ltd on behalf of Hydrogen Energy Publications LLC. This is an open access article under the CC BY license
(http://creativecommons.org/licenses/by/4.0/).
G. Panama et al. International Journal of Hydrogen Energy 55 (2024) 875–881

Fig. 1. Catalytic combustion hydrogen sensor on membrane support. (a) Front


view. (b) Cross-sectional view.

strength and limited wettability of the catalyst precursor solutions,


which can be improved by solution chemistry, but delamination may
occur due to thermo-mechanical stress [29,32,33]. An alternative
method involves the direct growth of the catalytic layer by local Joule
heating which provides good adhesion but produces a thin layer of
catalyst with low productivity [34–37]. Nanowires show rapid response
without the requirement for a membrane platform, yet they present
challenges arising from poor repeatability and low readout [38,39].
In this work, a catalytic combustion hydrogen sensor with a thick
catalyst layer is introduced on soda-lime glass, which has low thermal
conductivity and prefabricated catalyst anchoring channels. Catalyst
anchoring channels can hold a thick catalyst layer in close contact with
the sensing electrode, and the catalyst can be deposited by blade
coating.

2. Methods

2.1. Device
Fig. 3. Structure of a hydrogen sensor with a thick catalyst layer anchored on
The catalytic combustion of hydrogen is depicted in Fig. 2. Catalytic soda-lime glass.
combustion occurs when hydrogen and oxygen are adsorbed on the
active sites of the catalyst particles, leading to a surface reaction that piranha solution using a sulfuric acid to hydrogen peroxide ratio of 3:1
releases heat and water as byproducts [15]. The resulting exothermic for 10 min and dried with a nitrogen gun. Catalyst anchoring channels
reaction increases the temperature of the catalyst-coated region. with a 10 μm depth are wet-etched in buffered oxide etchant (BOE) for
This study proposes a hydrogen sensor with a thick catalyst layer 12 min while stirring magnetically at 30 ◦ C. The masking layer is an 8-
anchored on soda-lime glass, as shown in Fig. 3. A soda-lime glass μm-thick AZ10XT photoresist (+PR) prepared by photolithography. The
substrate is employed for good thermal isolation (κ = 1.12 W/m•◦ C, resulting sample is shown in Fig. 4(c), step 2.
100 ◦ C) compared to silicon (κ = 115 W/m•◦ C, 100 ◦ C) [40,41]. Catalyst A platinum sensing electrode and gold microheater are assembled
anchoring channels of 10 μm depth and 78 μm width are prepared by the inside the catalyst anchoring channels which are coated with a thin layer
wet etching of glass. A sensing electrode and microheater are patterned of alumina. Each composing layer (platinum, gold, and alumina) is
inside the catalyst anchoring channels and coated by a dielectric layer. patterned by the lift-off process in Fig. 4(b) and deposited by an e-beam
Hydrogen selective catalytic paste is blade coated to fill in the catalyst evaporator at 3 × 10− 6 Torr, as shown in Fig. 4(c) from step 3 to 5. The
anchoring channels. masking layer is a 5-μm-thick AZ-nLOF3070 photoresist (-PR).
Catalytic powder (Pt cluster based H2 oxidation catalyst, HC-H2-P01,
2.2. Fabrication process Hydrochem Co., Ltd.) and deionized water (DI) are mixed in a 1:1 wt to
volume ratio to make a uniform catalytic paste. Catalytic paste is
Catalyst anchoring channels are developed by wet etching, as shown deposited by doctor blade filling in the catalyst anchoring channels, as
in Fig. 4(a). The glass substrate, which is 500 μm thick, is cleaned in a depicted in Fig. 4(c), step 6. Thermal annealing is performed at 250 ◦ C
for 5 min with the included microheater.

2.3. Heating test setup

The heating test setup consists of a probe station with an infrared


camera (Seek Thermal CompactPRO, sensitivity: 1 ◦ C), as shown in
Fig. 5. Data are monitored in a computer interface at 10 Hz. Infrared
camera video is recorded at 15 fps, and temperature is derived by image
processing in MATLAB. A fine thermocouple wire is employed to cali­
brate the emissivity and separation distance of the infrared camera.
Sensor temperature using a calibrated infrared camera results in negli­
gible accuracy error, as shown in Fig. 6.

Fig. 2. Catalytic combustion of hydrogen.

876
G. Panama et al. International Journal of Hydrogen Energy 55 (2024) 875–881

Fig. 6. Sensor temperature with calibrated infrared camera.

Fig. 4. Fabrication process. Schematic diagram for (a) wet etching and (b) lift-
off processes. (c) Fabrication steps of the proposed hydrogen sensor.

Fig. 7. Sensing test setup. (a) Schematic diagram. (b) Test chamber.

3. Results and analysis

3.1. Fabrication results

Fig. 5. Heating test setup. The catalyst anchoring channels have a depth of 10 μm, with a width
of 78 μm for the sensing electrode and 55 μm for the microheater, as
2.4. Sensing test setup shown in Fig. 8(a). The electrical resistances of the platinum sensing
electrode and gold microheater are 44 Ω and 100 Ω, respectively. Fig. 8
A hydrogen gas sensing test is performed by flowing hydrogen gas in (a) and (b) display the reference sensor without catalyst coating and
the concentration range of 0.01–1 % to the hydrogen sensor in the test catalytic sensor with a catalyst layer in the catalyst anchoring channels,
chamber, as shown in Fig. 7(a). The test chamber is composed of a Pyrex respectively.
cylinder (diameter: 50 mm, length: 230 mm, thickness: 2.5 mm) sealed Fig. 9 shows the material analysis of the Pt/TiO2 catalyst. Fig. 9(a)
tightly with Teflon caps, as shown in Fig. 7(b). The oxygen concentra­ depicts the X-ray diffraction peaks of the catalyst powder showing TiO2
tion is maintained at 21 %. Nitrogen is the balancing gas to produce air. anatase domains. Diffraction peaks of Pt are not detected which suggests
Each gas supply is controlled by a mass flow controller (MFC), and the Pt nanoparticles are highly dispersed on the TiO2 support and do not
total flow rate is 1000 sccm. The electrical resistance of the hydrogen exhibit crystallinity [42,43]. Fig. 9(b) shows the scanning electron mi­
sensor and test chamber temperature are retrieved at 2 Hz. croscopy image of the catalyst coated in the anchoring channels. The
coated catalyst shows a porous structure for the hydrogen infiltration.
Fig. 9(c) displays the elemental mapping of Pt/TiO2 catalyst by energy
dispersive X-ray spectroscopy. The elemental analysis shows 1 % atomic

877
G. Panama et al. International Journal of Hydrogen Energy 55 (2024) 875–881

concentration for Pt which is dispersed on TiO2 support.

3.2. Heating response

Heating response and recovery times are measured at 90 % of the


total resistance change (ΔR) of the sensing electrode with the micro­
heater in the ON and OFF states, as shown in Fig. 10(a) [44]. The initial
resistance in air at room temperature (R0) is 44 Ω. The response and
recovery times with the microheater at 10 V are 28 s and 67 s, respec­
tively, which are substantially dependent on the thermal isolation of the
substrate. The inset in Fig. 10(a) shows linear fitting (y = a+bx) for the
thermoresistive response with the microheater at 0, 2, 4, 6, 8, and 10 V.
Fig. 10(b) shows the temperature distribution on a sensor with the
microheater at 6 V, and the temperature is monitored at the mid edge of
the sensing electrode towards the microheater.

3.3. Anchoring channels for a thick catalyst layer

A thick catalyst layer on a bare substrate is more likely to delaminate


due to the weak adhesion force. The catalyst anchoring channels offers
several advantages: close contact with the sensing layout, mechanical
interlocking in the micro roughened channels, and shelter within glass
walls. Therefore, heat generated in the catalytic combustion can rapidly
Fig. 8. Fabrication results. (a) Reference sensor. (b) Catalytic sensor. The
magnified images in each rectangle are shown at the right. transfer to the sensing electrode. Additionally, a moderately thick
catalyst layer is retained inside glass walls, leading to higher sensitivity
and mechanical stability compared to membrane-type sensors on silicon
[35]. Thus, catalyst anchoring channels ensure a consistent thermor­
esistive response.
Effective anchoring channels should have enhanced contact area

Fig. 10. Heating response results. (a) Total resistance change (ΔR) of the
Fig. 9. Catalyst characterization. (a) X-ray diffraction patterns of Pt/TiO2
sensing electrode during Joule heating, and the calculated response and re­
catalyst. (c) Elemental mapping of Pt/TiO2 catalyst by energy dispersive X-ray
covery times. Linear fitting of temperature difference vs. heating response
spectroscopy.
(inset). (b) Temperature distribution on the sensor with the microheater at 6 V.

878
G. Panama et al. International Journal of Hydrogen Energy 55 (2024) 875–881

with the sensing electrode and hold sufficient catalyst. Good catalyst
anchoring is attained with channels of aspect ratios (depth/width)
higher than 0.1 which is assessed from the amount of retained catalyst
after a doctor blade coating. Fig. 11 shows the sensing response (ΔR/R0)
with a catalyst layer thickness of 5 and 10 μm at various temperatures of
the test chamber atmosphere for 0.6 % H2 (response parameters are
described in the next section). Enhanced sensing response is attained
with a 10 μm catalyst layer showing a slight dependence on tempera­
tures in the range of 25–120 ◦ C, whereas a 5 μm catalyst layer shows a
lower sensing response which notably increases for temperatures higher
than 80 ◦ C. This catalytic activity behavior is attributed to the auto­
thermal mechanism of a relatively thick catalyst layer which produces
sufficient heat from the catalytic hydrogen combustion to sustain the
reaction [45]. Thick catalyst layer enables a good sensing response be­
tween room temperature and 80 ◦ C which is attractive for sensors
operating at room temperature.
Fig. 12. Resistance change of the reference and catalytic sensors for several
hydrogen concentrations.
3.4. Sensing response

The catalytic combustion of hydrogen gas generates heat, which Table 1


causes a thermoresistive response of the sensing electrode. Resistance Calculated sensor temperature at steady state.
changes (ΔR) vs. time of the reference and catalytic sensors are depicted
Hydrogen [%] Reference Catalytic
in Fig. 12. Several concentrations of hydrogen flow into the test chamber
while the chamber atmosphere is maintained at 80 ◦ C under atmo­ ΔR [Ω] T [◦ C] ΔR [Ω] T [◦ C]
spheric pressure. The microheater is not used during the sensing test 0 0 80 0 80
because the set temperature of the microheater can be biased due to the 0.01 0.04 80 0.37 85
0.2 0.08 80 2.00 117
resistance change caused by the heat generated due to catalytic
1.0 0.18 81 4.44 164
hydrogen combustion. The catalytic sensor shows a significant resis­
tance increment compared to the reference sensor. The initial resistance
in air at 80 ◦ C (R0’) is 47.06 Ω. The resistance for 1 % H2 in air reaches the sensing response (ΔR/R0’) dependence on hydrogen concentration,
51.5 Ω and the calculated sensor temperature is 164 ◦ C. Hence, the as shown in Fig. 13(b). The sensing response exhibits a linear relation­
sensor response (ΔR/R0’) for 1 % H2 in air is 9.43 % in a test chamber at ship with hydrogen concentration from 0.01 to 0.2 %, but its slope de­
80 ◦ C with a response time of 48 s and recovery time of 75 s. The creases to one-third for hydrogen concentrations beyond 0.25 %. The
resistance change and calculated sensor temperature are shown in decrease in the sensing response above 0.25 % H2 is likely attributed to
Table 1Table I. Temperature increments are derived by linear fitting of parasitic heat transfer to the glass substrate in a seal tight test chamber.
the heating response described in Section 3.2. A linear response was obtained with a sensor on a polyimide film, which
Temperature changes are calculated from the resistance change ob­ implies that parasitic heat loss is an important factor leading to reduced
tained in the sensing test for various concentrations of hydrogen which response.
are related to the initial resistance in air at room temperature (ΔR/R0),
as shown in Fig. 13(a). The calculated temperature change due to cat­ 4. Conclusions
alytic combustion is noted for each response curve in the steady state
condition. Consequently, a temperature change of 84 ◦ C arises from 1 % This study on a catalytic combustion hydrogen sensor seeks to
H2 exposure which is an order of magnitude higher compared to a sensor enhance sensor sensitivity and show an acceptable response signal by
on a SiO2 beam operated at a similar temperature (72 ◦ C) [35]. A sig­ using a thick catalyst layer. A thick catalyst layer on conventional
nificant temperature change can be attributed to a thick layer of cata­ membrane platforms has limitations regarding adhesion and structural
lytic particles in the catalyst anchoring channels. stability. Thus, this work presents the following key contributions.
The heat loss behavior into the glass bulk substrate is derived from
a. A thick catalyst layer is deposited and retained in the catalyst
anchoring channels prepared by the wet etching of soda-lime glass. A
thick catalyst layer increases sensitivity due to enhanced heat gen­
eration in the catalytic combustion of hydrogen.
b. A sensing platform on soda-lime glass substrate, which has low
thermal conductivity, shows good mechanical strength compared to
membrane platforms. Sensor reliability is ensured as long as the
catalyst is active.
c. A catalytic combustion hydrogen sensor with a thick catalyst layer
anchored on soda-lime glass is composed of a 44 Ω platinum sensing
electrode and a 100 Ω gold microheater. Both components are
assembled in the catalyst anchoring channels with a depth of 10 μm
and a width of 78 μm for the sensing electrode and 55 μm for the
microheater, coated with a thin layer of alumina, and buried with a
Pt based commercial catalyst. The sensor response (ΔR/R0’) for 1 %
H2 in air is 9.43 % in a test chamber at 80 ◦ C with a response and
recovery times of 48 s and 75 s, respectively.
Fig. 11. Sensing response dependence on the thickness of the catalyst layer and
temperature of the test chamber atmosphere for 0.6 % H2.

879
G. Panama et al. International Journal of Hydrogen Energy 55 (2024) 875–881

References

[1] Toyota Motor Corporation. Sustainability data book. 2019.


[2] Slowik P, Hall D, Lutsey N, Nicholas M, Wappelhorst S. Funding the transition to all
zero-emission vehicles. The International Council on Clean Transportation; 2019.
White Paper.
[3] Rasul MG, Hazrat MA, Sattar MA, Jahirul MI, Shearer MJ. The future of hydrogen:
challenges on production, storage and applications. Energy Convers Manag 2022;
272:116326. https://doi.org/10.1016/J.ENCONMAN.2022.116326.
[4] Stangarone T. South Korean efforts to transition to a hydrogen economy. Clean
Technol Environ Policy 2021;23:509–16. https://doi.org/10.1007/S10098-020-
01936-6/METRICS.
[5] Iida S, Sakata K. Hydrogen technologies and developments in Japan. Clean Energy
2019;3:105–13. https://doi.org/10.1093/CE/ZKZ003.
[6] Wappler M, Unguder D, Lu X, Ohlmeyer H, Teschke H, Lueke W. Building the green
hydrogen market – current state and outlook on green hydrogen demand and
electrolyzer manufacturing. Int J Hydrogen Energy 2022;47:33551–70. https://
doi.org/10.1016/J.IJHYDENE.2022.07.253.
[7] Foorginezhad S, Mohseni-Dargah M, Falahati Z, Abbassi R, Razmjou A, Asadnia M.
Sensing advancement towards safety assessment of hydrogen fuel cell vehicles.
J Power Sources 2021;489:229450. https://doi.org/10.1016/j.
jpowsour.2021.229450.
[8] Dagdougui H, Sacile R, Bersani C, Ouammi A. Hydrogen logistics: safety and risks
issues. Hydrogen Infrastructure for Energy Applications 2018;127–48. https://doi.
org/10.1016/B978-0-12-812036-1.00007-X.
[9] Hübert T, Boon-Brett L, Black G, Banach U. Hydrogen sensors – a review. Sens
Actuators B Chem 2011;157:329–52. https://doi.org/10.1016/J.SNB.2011.04.070.
[10] Sisman O, Erkovan M, Kilinc N. Hydrogen sensors for safety applications. Towards
Hydrogen Infrastructure 2023:275–314. https://doi.org/10.1016/B978-0-323-
95553-9.00061-3.
[11] Boon-Brett L, Bousek J, Black G, Moretto P, Castello P, Hübert T, et al. Identifying
performance gaps in hydrogen safety sensor technology for automotive and
stationary applications. Int J Hydrogen Energy 2010;35:373–84. https://doi.org/
10.1016/J.IJHYDENE.2009.10.064.
[12] Aroutiounian V. Metal oxide hydrogen, oxygen, and carbon monoxide sensors for
hydrogen setups and cells. Int J Hydrogen Energy 2007;32:1145–58. https://doi.
org/10.1016/J.IJHYDENE.2007.01.004.
[13] Koo WT, Cho HJ, Kim DH, Kim YH, Shin H, Penner RM, et al. Chemiresistive
hydrogen sensors: fundamentals, recent advances, and challenges. ACS Nano 2020;
14:14284–322. https://doi.org/10.1021/ACSNANO.0C05307/ASSET/IMAGES/
LARGE/NN0C05307_0027 [JPEG].
[14] Sharma B, Sharma A, Kim JS. Recent advances on H2 sensor technologies based on
MOX and FET devices: a review. Sens Actuators B Chem 2018;262:758–70. https://
doi.org/10.1016/J.SNB.2018.01.212.
[15] Singh SA, Vishwanath K, Madras G. Role of hydrogen and oxygen activation over
Pt and Pd-doped composites for catalytic hydrogen combustion. ACS Appl Mater
Interfaces 2017;9:19380. https://doi.org/10.1021/ACSAMI.6B08019/ASSET/
Fig. 13. Sensing response results. (a) Calculated temperature changes based on IMAGES/LARGE/AM-2016-080196_0008. 8, [JPEG].
the resistance change with respect to the initial resistance in air at room tem­ [16] Houlet LF, Tajima K, Shin W, Itoh T, Izu N, Matsubara I. Platinum micro-hotplates
perature (ΔR/R0). (b) Sensing response (ΔR/R0’) dependence on hydrogen on thermal insulated structure for micro-thermoelectric gas sensor. IEEJ Trans
Sensors Micromachines 2006;126:568–72. https://doi.org/10.1541/
concentration.
IEEJSMAS.126.568.
[17] Lee EB, Hwang IS, Cha JH, Lee HJ, Lee WB, Pak JJ, et al. Micromachined catalytic
d. An extension of this work is ongoing using polyimide films with combustible hydrogen gas sensor. Sens Actuators B Chem 2011;153:392–7.
https://doi.org/10.1016/J.SNB.2010.11.004.
polymer-based catalyst anchoring channels, which have higher [18] Ivanov II, Baranov AM, Talipov VA, Mironov SM, Akbari S, Kolesnik IV, et al.
thermal resistance to enhance sensing performance and linearity. Investigation of catalytic hydrogen sensors with platinum group catalysts. Sens
Actuators B Chem 2021;346:130515. https://doi.org/10.1016/J.
SNB.2021.130515.
CRediT authorship contribution statement [19] Karim W, Spreafico C, Kleibert A, Gobrecht J, Vandevondele J, Ekinci Y, et al.
Catalyst support effects on hydrogen spillover. Nature 2017;541:68–71. https://
doi.org/10.1038/nature20782. 2017 541:7635.
Gustavo Panama: Methodology, Software, Validation, Formal Anal­
[20] Beck A, Kazazis D, Ekinci Y, Li X, Müller Gubler EA, Kleibert A, et al. The extent of
ysis, Investigation, Writing, Visualization. Munhyung Jo: Methodology, platinum-induced hydrogen spillover on cerium dioxide. ACS Nano 2022. https://
Investigation. Soon Seop Shim: Funding Acquisition, Investigation. Min doi.org/10.1021/ACSNANO.2C08152/ASSET/IMAGES/LARGE/NN2C08152_0005
Su Kim: Methodology, Validation, Formal Analysis, Investigation. Seung [JPEG].
[21] Shun K, Mori K, Masuda S, Hashimoto N, Hinuma Y, Kobayashi H, et al. Revealing
S. Lee: Writing – Review & Editing, Supervision, Project Administration, hydrogen spillover pathways in reducible metal oxides. Chem Sci 2022;13:
Funding Acquisition. 8137–47. https://doi.org/10.1039/D2SC00871H.
[22] Zhu J, Hu L, Zhao P, Lee LYS, Wong KY. Recent advances in electrocatalytic
hydrogen evolution using nanoparticles. Chem Rev 2020;120:851–918. https://
doi.org/10.1021/ACS.CHEMREV.9B00248/ASSET/IMAGES/MEDIUM/
Declaration of competing interest CR9B00248_0041 [GIF].
[23] Mao S, Zhou H, Wu S, Yang J, Li Z, Wei X, et al. High performance hydrogen sensor
The authors declare that they have no known competing financial based on Pd/TiO2 composite film. Int J Hydrogen Energy 2018;43:22727–32.
https://doi.org/10.1016/J.IJHYDENE.2018.10.094.
interests or personal relationships that could have appeared to influence [24] Patil LA, Suryawanshi DN, Pathan IG, Patil DG. Nanocrystalline Pt-doped TiO2 thin
the work reported in this paper. films prepared by spray pyrolysis for hydrogen gas detection. Bull Mater Sci 2014;
37:425–32. https://doi.org/10.1007/S12034-014-0705-Y/METRICS.
[25] Kozhukhova AE, du Preez SP, Bessarabov DG. Development of Pt–Co/Al2O3
Acknowledgments bimetallic catalyst and its evaluation in catalytic hydrogen combustion reaction.
Int J Hydrogen Energy 2023. https://doi.org/10.1016/J.IJHYDENE.2023.09.119.
[26] Simon I, Bârsan N, Bauer M, Weimar U. Micromachined metal oxide gas sensors:
This research was funded by the National Research Foundation of
opportunities to improve sensor performance. Sens Actuators B Chem 2001;73:
Korea, grant number NRF-N01210648. We would like to acknowledge 1–26. https://doi.org/10.1016/S0925-4005(00)00639-0.
the technical support from NI Korea.

880
G. Panama et al. International Journal of Hydrogen Energy 55 (2024) 875–881

[27] Briand D, Courbat J. Micromachined semiconductor gas sensors. Semiconductor devices. Adv Mater 2015;27:1207–15. https://doi.org/10.1002/
Gas Sensors 2020:413–64. https://doi.org/10.1016/B978-0-08-102559-8.00013-6. ADMA.201404192.
[28] Adedokun G, Geng L, Xie D, Xu L. Low power perforated membrane microheater. [37] Sosnowchik BD, Lin L, Englander O. Localized heating induced chemical vapor
Sens Actuators A Phys 2021;322:112607. https://doi.org/10.1016/J. deposition for one-dimensional nanostructure synthesis. J Appl Phys 2010;107:
SNA.2021.112607. 51101. https://doi.org/10.1063/1.3304835/905905.
[29] Bársony I, Ádám M, Fürjes P, Lucklum R, Hirschfelder M, Kulinyi S, et al. Efficient [38] Lee H, Manorotkul W, Lee J, Kwon J, Suh YD, Paeng D, et al. Nanowire-on-
catalytic combustion in integrated micropellistors. Meas Sci Technol 2009;20: Nanowire: all-nanowire electronics by on-demand selective integration of
124009. https://doi.org/10.1088/0957-0233/20/12/124009. hierarchical heterogeneous nanowires. ACS Nano 2017;11:12311–7. https://doi.
[30] Mehmood Z, Haneef I, Udrea F. Material selection for optimum design of MEMS org/10.1021/ACSNANO.7B06098/ASSET/IMAGES/LARGE/NN-2017-06098Z_
pressure sensors. Microsyst Technol 2020;26:2751–66. https://doi.org/10.1007/ 0004 [JPEG].
S00542-019-04601-1/FIGURES/5. [39] Yun J, Ahn JH, Moon D Il, Choi YK, Park I. Joule-heated and suspended silicon
[31] Dücso C, Ádám M, Fürjes P, Hirschfelder M, Kulinyi S, Bársony I. Explosion-proof nanowire based sensor for low-power and stable hydrogen detection. ACS Appl
monitoring of hydrocarbons by mechanically stabilised, integrable calorimetric Mater Interfaces 2019;11:42349–57. https://doi.org/10.1021/ACSAMI.9B15111/
microsensors. Sens Actuators B Chem 2003;95:189–94. https://doi.org/10.1016/ ASSET/IMAGES/LARGE/AM9B15111_0002 [JPEG].
S0925-4005(03)00415-5. [40] Prakash C. Thermal conductivity variation of silicon with temperature.
[32] Jiang Y, Shi K, Tang H, Wang Y. Enhanced wettability and wear resistance on Microelectron Reliab 1978;18:333. https://doi.org/10.1016/0026-2714(78)
TiO2/PDA thin films prepared by sol-gel dip coating. Surf Coat Technol 2019;375: 90573-5.
334–40. https://doi.org/10.1016/J.SURFCOAT.2019.07.051. [41] Bansal NP, Doremus RH. Handbook of glass properties. Handbook of Glass
[33] Puigcorbé J, Vilà A, Cerdà J, Cirera A, Gràcia I, Cané C, et al. Thermo-mechanical Properties 2013;1–680. https://doi.org/10.1016/C2009-0-21785-5.
analysis of micro-drop coated gas sensors. Sens Actuators A Phys 2002;97–98: [42] Zhang C, He H. A comparative study of TiO2 supported noble metal catalysts for
379–85. https://doi.org/10.1016/S0924-4247(01)00858-5. the oxidation of formaldehyde at room temperature. Catal Today 2007;126:
[34] Chen CC, Lin YS, Sang CH, Sheu JT. Localized joule heating as a mask-free 345–50. https://doi.org/10.1016/J.CATTOD.2007.06.010.
technique for the local synthesis of ZnO nanowires on silicon nanodevices. Nano [43] Kim S Bin, Shin JH, Kim GJ, Hong SC. Promoting metal-support interaction on Pt/
Lett 2011;11:4736–41. https://doi.org/10.1021/NL202539M/SUPPL_FILE/ TiO2Catalyst by antimony for enhanced carbon monoxide oxidation activity at
NL202539M_SI_001.PDF. room temperature. Ind Eng Chem Res 2022;61:14793–803. https://doi.org/
[35] Del Orbe Henriquez D, Cho I, Yang H, Choi J, Kang M, Chang KS, et al. Pt 10.1021/ACS.IECR.2C01518/ASSET/IMAGES/LARGE/IE2C01518_0010 [JPEG].
nanostructures fabricated by local hydrothermal synthesis for low-power catalytic- [44] Boon-Brett L, Black G, Moretto P, Bousek J. A comparison of test methods for the
combustion hydrogen sensors. ACS Appl Nano Mater 2021;4:7–12. https://doi.org/ measurement of hydrogen sensor response and recovery times. Int J Hydrogen
10.1021/ACSANM.0C02794/ASSET/IMAGES/LARGE/AN0C02794_0004 [JPEG]. Energy 2010;35:7652–63. https://doi.org/10.1016/J.IJHYDENE.2010.04.139.
[36] Yang D, Kim D, Ko SH, Pisano AP, Li Z, Park I. Focused energy field method for the [45] Fernandes NE, Park YK, Vlachos DG. The autothermal behavior of platinum
localized synthesis and direct integration of 1D nanomaterials on microelectronic catalyzed hydrogen oxidation: experiments and modeling. Combust Flame 1999;
118:164–78. https://doi.org/10.1016/S0010-2180(98)00162-X.

881

You might also like