Pd-Functionalized, Suspended Graphene Nano-Sheet for Fast, Low

appropriate voltage bias, the graphene temperature was increased to 180 °C with ... ACS Paragon Plus Environment. ACS Applied Nano Materials. 1. 2. 3...
0 downloads 0 Views 2MB Size
Subscriber access provided by NEW MEXICO STATE UNIV

Article

Pd-Functionalized, Suspended Graphene NanoSheet for Fast, Low-Energy Multi-Molecular Sensors Takamune Yokoyama, Takahisa Tanaka, Yoshihiko Shimokawa, Ryosuke Yamachi, Yuta Saito, and Ken Uchida ACS Appl. Nano Mater., Just Accepted Manuscript • DOI: 10.1021/acsanm.8b00667 • Publication Date (Web): 03 Jul 2018 Downloaded from http://pubs.acs.org on July 6, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Nano Materials

Pd-Functionalized, Suspended Graphene Nano-Sheet for Fast, Low-Energy Multi-Molecular Sensors Takamune Yokoyama, Takahisa Tanaka, Yoshihiko Shimokawa, Ryosuke Yamachi, Yuta Saito, Ken Uchida* Electrical and Electronics Engineering, Keio University 3-14-1 Hiyoshi, Kohoku-ku, Yokohama, Kanagawa 223-8522, Japan

ABSTRACT To demonstrate the feasibility of a fast, low-energy breath diagnosis method, hydrogen sensors utilizing Joule self-heating were developed. The sensors consisted of suspended graphene films functionalized with Pd nanoparticles as sensing layers and utilized self-heating to achieve fast responses and humidity robustness with low energy consumption. Thanks to nanoscale point contacts between the graphene and Au electrodes, heat dissipation to the electrodes was greatly suppressed. By applying an appropriate voltage bias, the graphene temperature was increased to 180 °C with a low power consumption per unit graphene width of 0.93 mW/µm. Because of the temperature increase caused by the Joule self-heating of the graphene, the sensors responded to ppm-level H2, and a response time of 15 s was achieved at a H2 concentration of 100 ppm. At temperatures over 100 °C, the sensor response realized by self-heating was lower than that by heating using an external heater. The response reduction was due to suppressed charged-carrier scatterings with H-induced potentials under high electric fields in the self-heated graphene. Finally, we demonstrated voltage-controlled multi-molecular detection by self-heating. At a two-terminal voltage difference of 0.1 V, Pd-functionalized graphene responded not to H2 but to humidity. Meanwhile, at 0.9 V, the sensor responded to 10-ppm H2 despite great humidity variations in the background. Temperature change by self-heating is much faster and requires less energy than that by heating using an external heater. Fast, low-energy multi-molecule detection realized by self-heating paves the way for mobile, low-power, real-time molecular sensors for use in health diagnosis applications. Keywords: graphene, hydrogen sensor, self-heating, low energy, palladium

1

ACS Paragon Plus Environment

ACS Applied Nano Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Analysis of molecules in breath is a promising diagnostic technique because a number of compounds in human breath are related to various kinds of diseases.1–3 Breath diagnosis has gained increasing attention due to its safety, simplicity, and swiftness. The H2 concentration of breath is known to be a good indicator of disorders in small intestine including bacterial overgrowth, colonic fermentation, abnormal fermentation, and carbohydrate intolerance.1,4–6 The typical H2 concentration ranges from a few ppm in a fasting state to several hundred ppm after a meal. However, breath contains many disturbing substances, such as a high concentration of water.4,7 When the ambient around the sensor is changed from the atmosphere to expired air, the water concentration is changed much more greatly than H2 concentration is. Thus, to develop a H2-based breath diagnosis method that can be used anywhere at any time, H2 sensors are required to be capable of detecting low and wide ranges of H2 concentrations, to be small, and to exhibit humidity robustness and low power consumption. Therefore, numerous studies of H2 sensors have been reported on, and the corresponding sensors can be categorized into two groups: 1) metal-oxide-based sensors5,8–11 and 2) metal-based sensors, namely, metal-catalyst-decorated sensors in most cases and metal thin-film sensors in some cases. Metal oxides have been the most widely used materials owing to their ability to detect low-concentration H2. However, they react with various components, or almost all organic compounds, and require high operation temperatures of 300–600 °C to maintain stable responses and to eliminate the effect of humidity.12 The utilization of external heaters prevents sensor-size and power-consumption reductions.13 Noble metals such as Au, Pt, and Pd are also used for gas sensors as catalytic metals to enhance the sensor response and selectivity.14 Among them, Pd has been used the most frequently for H2 sensing because of its high sensitivity to and selectivity of H2. The following reaction mechanisms between H2 and Pd have been reported.15–18 Hydrogen molecules dissociate on the Pd surface, and the dissociated H atoms diffuse into the Pd around the surface region to form Pd hydride. The change from Pd to Pd hydride affects material properties with respect to work function, grain size, and electrical resistance. Therefore, there have been various studies on Pd-based H2 sensors, including Pd-gated metal– oxide–semiconductor transistors, Pd thin film sensors, and Pd nanowire sensors to detect low-concentration H2.19,20 However, Pd-based sensors often exhibit a serious issue of vulnerability to humidity.21 This issue can be eliminated by raising the temperature of Pd. Therefore, high-temperature operation utilizing external heaters as well as Joule self-heating without external heaters, which is called self-heating, has been studied using various combinations with numerous materials. However, most related 2

ACS Paragon Plus Environment

Page 2 of 25

Page 3 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Nano Materials

studies have focused on sensor response time and recovery time improvements.13,22–24 No attention has been paid to self-heating-induced changes in the physical properties and device parameters except for temperature. In electrical gas sensors, metal catalysts are usually combined with conductive materials, which transduce chemical reactions to electrical signals. Among the conductive materials decorated with metal catalysts, graphene has attracted growing interest, since it has unique electric properties suitable for electrical signal transducers: a high electrical conductivity with a low carrier concentration and a high surface-to-volume ratio. Furthermore, its thermal properties are outstanding and are expected to improve the sensor characteristics effectively. Since graphene has a small heat capacity, self-heating is an effective means of increasing its temperature with low power consumption. Therefore, graphene is a promising sensing material,25,26 and graphene-based

sensors

have

been

intensively

studied.25,27–29

Pristine, or unfunctionalized, graphene exhibits responses to polar molecules such as NO2 and NH3

and no response to nonpolar molecules such as H2. However, graphene modified with Pd can detect nonpolar H2.30–32 Self-heating with graphene under constant humidity was already demonstrated for a NO2 sensor at a relatively high power consumption of 32 mW.33 However, no study is available on H2 sensors in self-heated graphene. In addition, it has been reported that in Pd-composite graphene sensors, the response to humid air is the same as that to 1% H234, suggesting that detection of hydrogen as low as 10 ppm under humidity variations is quite difficult. Thus, effectiveness of self-heating for eliminating the humidity effect even with substantial relative humidity (RH) variation should be explored. In particular, if a low power-consumption suitable for health diagnosis applications anywhere at any time is required, H2 detection becomes even more difficult. Furthermore, to design low-power and highly sensitive sensors, accurate understanding of the self-heating effects on the sensor properties is required. However, the effects of high electric fields in self-heated materials on the properties of sensors remain unexplored. In this work, we fabricated Pd-functionalized suspended graphene sensors to avoid heat transfer from the self-heated graphene to the substrate. The graphene was suspended on the electrodes using a soft lithography technique called the polydimethylsiloxane (PDMS) stamp method.35–39 The sensor responses to H2 were evaluated by changing H2 concentrations, substrate temperature, and applied bias conditions. The self-heating of the graphene was verified by comparing the response time in self-heated graphene with that in graphene heated by an external heater. Finally, we also explored the possibility of selective sensing of hydrogen/humidity by applying 3

ACS Paragon Plus Environment

ACS Applied Nano Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

an appropriate bias voltage in realistic gas environments, in which small changes in the H2 concentration and large variations in the RH were induced. EXPERIMENTAL SECTION Low energy consumption, small size, and humidity robustness are required for a breath H2 sensor accessible anywhere at any time. To meet these requirements, we fabricated a suspended graphene sensor functionalized by Pd nanoparticles (NPs). To reduce the sensor size and power consumption, a self-heating technique was utilized. In addition, we studied the difference between heating using an external heater and self-heating by comparing the response time of self-heated graphene with that of graphene heated by an external heater. Device fabrication. The 300-nm-thick SiO2 thermally grown on the p-type Si substrate was cleaned in both acetone and 2-propanol for 5 min. To fabricate the patterned electrodes, lift-off resist (MicroChem LOR-3B) and photo resist (AZ Electronic Materials AZ5206E) were spin-coated onto the substrate at 4000 rpm for 25 s. The substrate was then exposed in a mask aligner and developed in AZ-300MIF for 20 s. Ti, Pt, and Au metals with thicknesses of 10 nm, 100 nm, and 40 nm were subsequently deposited by electron beam deposition at acceleration voltages of 6 kV, 6 kV, and 8 kV, respectively. Ti works as the adhesion layer to SiO2. Pt acts as the physically hard layer to stand against the needle probes for electrical measurement, since Mohs hardness of Pt (4.0) is greater than that of Au (2.5).40 However, Pt is a catalytic metal and may react with hydrogen. Therefore, Au layer is deposited to prevent a possible reaction of Pt with H2. Moreover, Au electrode is easier to transfer graphene onto the electrode using PDMS technique than Pt electrode is. The metal-deposited substrate was cleaned in acetone, 2-propanol, TMAH (Tetramethylammonium hydroxide), and water with ultrasonication. Buffered HF was used to etch SiO2 in the areas without electrodes. The substrate was then cleaned in acetone and 2-propanol before graphene transfer. HOPG (430HP-AB Grade ZYA, GE Advance Ceramics) was slowly exfoliated several times using Nitto tape (SPV224-R, Nitto). Then graphene on Nitto tape was placed on PDMS (Gel-Film® WF-20X4 6mil, Gel-Pak). To increase the number of graphene flakes on PDMS, Nitto tape was exfoliated quickly from PDMS. The PDMS was used to transfer the graphene on the electrodes. The graphene was transferred by PDMS stamping (also called soft lithography) with a three-axis stage. The transferred graphene was cleaned in acetone and 2-propanol and baked in a uniform temperature heat-treatment system (THERMO RIKO GFA430VN) for 1 h at 300 °C in N2 atmosphere; then, 0.3-nm-thick Pd (KOJUNDO CHEMICAL LABORATORY CO., LTD) was deposited by electron 4

ACS Paragon Plus Environment

Page 4 of 25

Page 5 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Nano Materials

beam deposition and agglomerated in the same heat-treatment system for 30 min at 400 °C in N2 atmosphere. As a result, Pd NPs were formed on the graphene. A schematic of the fabricated device is provided in Figure 1a. HOPG, Nitto tape, PDMS and Pd used in this study are all commercially available. Characterization. Transferred graphene was verified by optical microscopy. Scanning electron microscope was operated at an accelerating voltage of 20 kV. Raman spectrum of the graphene was obtained with 532 nm wavelength. Pd on graphene and Au, and the interface between graphene and Au electrodes were explored by transmission electron microscopy (TEM). Sensing measurement. The measurement setup is illustrated in Figure 1b. The resistance was obtained from the relationship between the drain voltage (VD) and drain current (ID). The temperature was controlled using the hot chuck of the probe station. The sensing performance was evaluated by exposing the reference and test gases alternately every 3 min. To evaluate the H2 sensing properties, dry or humid air was employed as the reference gas and H2 diluted with dry and humid air was used as the test gas. The flow rates of the 100-ppm H2 balanced with dry air, dry air, and humid air were controlled by mass flow controllers (MFCs). The concentration of H2 was changed from 2 ppm to 100 ppm by adjusting the flow rates of the gases. In our experiments, the device was exposed to gases with a total flow rate of 500 sccm (standard cubic cm). As shown in Figure 1c, the sensor response was defined as follows 14:

Sensor Response =



× 100% =



× 100%,

(1)

where R0 represents the initial resistance when the device was exposed to the test gas, and R represents the resistance during the sensor response measurements under the test gas. Figure 1c also shows the definition of the response time, which is the time necessary to reach half of the maximum sensor response. COMSOL simulation. COMSOL Multiphysics was used to evaluate the graphene channel temperature resulting from Joule self-heating. First, we obtained the sheet conductivity from the linear relationship of the ID–VD characteristics. The resulting sheet conductivity of 3.0 mS sq. was used in the COMSOL simulation. The electrical transport was obtained by applying the Poisson and current conservation equations, and the thermal transport was calculated by employing the Fourier law. The material parameters except for graphene we used in the simulations were the default parameters in the COMSOL library. The thermal conductivity of graphene modified by Pd was assumed to be 200 W/m·K.41 The interfacial thermal resistance was used as a fitting parameter to match the experimental data of an electrically broken device referring to 5

ACS Paragon Plus Environment

ACS Applied Nano Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

the previous investigations showing that graphene burns at around 600 °C in the atmosphere.42,43 Using these parameters, the graphene channel temperature was evaluated at various VD, and the channel temperature in the self-heated graphene was compared with that controlled by the hot chuck. The time dependence of the temperature during self-heating was also calculated. Monte Carlo simulation. To demonstrate the difference between the sensor responses in low and high electric fields, an ensemble Monte Carlo simulation was performed. The time step was 0.1 ps, and the transports of 1×105 holes were simulated. For the band structure of graphene, linear energy dispersion relation was assumed. Ek =ħvf |k k|, (2) where vf = 108 cm/s is the Fermi velocity and k is the 2D wave vector from the Dirac point. In the simulation, intravalley and intervalley acoustic phonons, intervalley optical phonon, and impurity scattering were taken into account.44 The deformation potential and phonon energy were cited.45 The detailed formula of scattering ratio in the previous work46 is utilized. Due to the suspension of graphene, optical phonon originated from substrate was neglected. In our calculation, initial hole concentration and initial ionized impurity density of 3×1012 cm-2 and 1×1012 cm-2 were assumed, respectively. The polarized H atoms adsorbed on the Pd NPs were treated as ionized impurities. Therefore, the adsorption of H atoms caused changes in the hole concentration and ionized impurity density of -∆p and ∆p, respectively. The adopted value of ∆p was 2×1012 cm-2. The simulated sensor response is defined by the ratio between the resistivity change by hydrogen and initial resistivity.

RESULTS AND DISCUSSION Figure 2a shows optical and scanning electron microscopy (SEM) micrographs of the fabricated device. We confirmed the successful suspension of the graphene film on the electrodes. Figure 2b depicts the Raman spectrum (532 nm wavelength) of the graphene, which was obtained by subtracting the Raman spectrum of the Au from that of the graphene on the Au. The G (~1580 cm-1) and 2D (~2670 cm-1) intensities of the Raman spectrum indicate that the graphene consisted of multiple layers, and the small D band (~1350 cm-1) demonstrates the high quality of the suspended graphene.47 We consider that the little D peak stems from weakening of C-C bonds, which is caused by Pd atoms adsorbed above the midpoint of C-C bonds.48 A transmission electron microscopy (TEM) image of the Pd on the graphene is shown in Figure 2c. The Pd was placed on graphene as NPs, and the size of the Pd NPs are about 8 nm. The number of graphene layers was eight, which was determined from the TEM image shown in 6

ACS Paragon Plus Environment

Page 6 of 25

Page 7 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Nano Materials

Figure 2c. One might consider that eight-layer graphene could be too thick to work as a sensor material. However, it is reported that the electrical conductivity of eight-layer graphene can be well modulated by external electric field.49 Therefore, eight-layer graphene has the ability to work as a transducer of hydrogen-induced changes in external electric field. Figure 2d presents an atomic force microscopy (AFM) image of the Au electrode without Pd deposition. Figure 2e shows an AFM image of Au electrode with transferred graphene and Pd deposition. The upper area of the Figure 2e shows the Au electrode and the lower area corresponds to the graphene on the Au electrode. In the Pd-deposited Au film, a number of additional undulations with short correlation lengths were appeared as shown in Figure 2e. By comparing these figures, we consider that undulations with short correlation lengths correspond to Pd NPs. Therefore, the uniform distribution of Pd NPs on graphene is confirmed from the AFM image. We measured the sensor resistance before and after Pd deposition. The resistance was obtained from the relationship between the drain voltage (VD) and drain current (ID), which is shown in Figure 3a. The formation of the Pd NPs increased the channel resistance. The resistance after Pd deposition was 520 Ω. The resistance change should originate from two effects: hole doping and mobility reduction by the Pd NPs. Graphene in an air environment is originally hole doped with environmental oxygen. The graphene was further hole-doped by the Pd NPs because of the difference between the work functions of graphene and Pd, which are 4.5 eV and 5.6 eV, respectively.50,51 Therefore, hole doping by Pd NPs should result in a decrease in resistance. However, the carrier mobility was greatly reduced by the Pd NPs. As a result, ID, which is determined by the carrier concentration and mobility, was reduced, and the resistance was increased. Figure 3b depicts the sensor response to 100-ppm H2 as a function of time at various operating temperatures from the room temperature (RT) of 27 °C to 180 °C. Resistance was increased (decreased) during H2 (dry air) exposure at any temperatures, which results from two phenomenon: a reduction in hole concentration and an increase in Coulomb scattering. H2 adsorbed on Pd dissociate into H atoms. H atoms diffuse through the Pd NPs and reach at the Pd/graphene interface, where ionized H atoms and image charge in Pd NPs form dipole directed toward graphene. As a result, carrier concentration was decreased by induced H+. In addition, holes are scattered by polarized H atoms on the Pd NPs. Therefore, resistance of the sensor was increased during H2 exposure. Whereas the resistance was recovered during dry air exposure because of hydrogen desorption. The sensor response increases as the operating temperature increases to 135 °C, due to the promoted dissociative adsorption of H2. 7

ACS Paragon Plus Environment

ACS Applied Nano Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 25

However, the sensor response is decreased at 180 °C, resulting from the impact of hydrogen desorption from Pd being greater than that of dissociation and adsorption at 180 °C. Figures 3c and 3d show the time dependence of the sensor response to 100-ppm H2, where the device was operated at various VD from 0.01 V to 1.1 V at RT. The sensor response increases as VD increases to 0.8 V, then gradually decreases as VD continues to increase from 0.9 V to 1.1 V. This tendency of the VD-dependent sensor response is almost the same as that of the temperature-dependent sensor response, which strongly suggests that VD-induced self-heating was successfully achieved. An improved sensor response and response time were confirmed in graphene with a high VD; a response time of 15 s was achieved with a VD of 0.7 V at a H2 concentration of 100 ppm. The sensor was more stable at a VD of 0.6 V than at 0.8 V because sensor response reduction occurred at a VD of around 0.8 V. Therefore, we explored the sensor response to H2 with various concentrations at a VD of 0.6 V, as shown in Figure 3e. The sensor response is linearly dependent on the square root of the H2 concentration. This relationship can be explained by Langmuir adsorption isotherm theory. Assuming that the sensor response is proportional to the coverage of H on the Pd surface, the relationship between the adsorption and desorption rates at equilibrium can be expressed as (3)   1 −   =    Or !



(4)

=   " ,  

where  is the fractional surface coverage of adsorbed hydrogen atoms. ka and kd are the adsorption and desorption constants, respectively, and K is an equilibrium constant defined as ka/kd. P represents the partial pressure of H2. When the H2 concentration is low, the resistance change is given by 

!

∝  ≈   " .

(5)

The reaction between the H2 and Pd NPs on the graphene may be related to the diffusion of the H in the Pd NPs and the H coverage at the interface between the Pd NPs and graphene. However, the linear relationship between the sensor response and the square root of the H2 concentration suggests that the dissociation of H2 and adsorption of H on the Pd surface are the dominant mechanisms determining the sensor response. In the present experiments, we demonstrated that our sensor detected H2 with concentrations ranging from 2 ppm to 100 ppm, which covers the important range of H2 concentrations in expired air. Although H2 concentrations of less than 2 ppm were not 8

ACS Paragon Plus Environment

Page 9 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Nano Materials

tested in our experimental setup, a detection limit of 60 ppb was derived. To calibrate the graphene channel temperature when the self-heating technique was utilized, the sensor responses to 100 ppm H2 using the self-heating technique and hot chuck were compared. The present Pd-functionalized graphene sensors react with hydrogens and the sensor response shows temperature dependence because of temperature dependent reaction of hydrogens with Pd. Therefore, the comparison of sensor reactions on hot plate with negligibly small self-heating and those with prominent self-heating effects at room temperature enables us to evaluate operating temperature under self-heating conditions. As shown in Figure 4a, the normalized sensor response at a VD of 0.9 V at RT matches the response at a VD of 0.01 V at 135 °C with hot chuck heating. Although the sensor responses shown in Figure 3b (T = 135 °C) and 3c (VD = 0.9 V) are different in magnitude, their shapes matched with each other by normalizing sensor responses with their maximum values. The reason will be described later. We compared the response time in self-heated graphene with that in graphene heated by the hot chuck. Since the reaction between H2 and Pd is controlled mainly by temperature, the temperature realized by self-heating should be equal to that controlled by the hot chuck when the response times in both cases are the same, as shown in Figure 4a. The response times as functions of the input power and temperature are compared in Figure 4b, which clearly demonstrates the close agreement between the self-heated and heater-heated characteristics. From this result, the relationship between the temperature and input power was obtained and is depicted in Figure 4c, which shows that a graphene temperature of 180 °C was realized with an input power of 2.3 mW. It is worth noting that the temperature increase due to self-heating was negligibly small during the measurements on the hot chuck. The applied drain voltage of 0.01 V results in the self-heating-induced temperature increase of 0.014 K. We also verified the relationship between the input power and temperature in the graphene sensor using numerical simulations. The simulation data agree well with the experimental data as shown in Figure 4c. In the simulation, the electrical conductivity obtained by fitting the ID–VD characteristics of the device was used.52 In a previous report, the thermal conductivity was measured as a function of the area covered with Au NPs in suspended tri-layer graphene.41 This information was utilized to determine the thermal conductivity of the present eight-layer suspended graphene covered with Pd NPs. The interfacial thermal resistance between the Au electrodes and graphene was determined as follows. We prepared another suspended Pd-functionalized graphene. The device was electrically broken by applying high drain voltage. Figure 4d shows ID versus VD characteristics of the broken device. The electrical breakdown was caused by 9

ACS Paragon Plus Environment

ACS Applied Nano Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

burning out of the graphene in air at approximately 900 K.43 Therefore, the interfacial thermal resistance was determined to be 2500 µm2·K/mW to reproduce channel temperature of 900 K at a drain voltage of 1.5 V as shown in Figure 4e. The obtained interfacial thermal resistance between the graphene and Au was 100 times larger than the theoretical value of 25 µm2·K/mW.53,54 We consider this difference was due to the fact that the interface between the graphene and Au was not uniform and was comprised of point contacts, as shown in Figure 4f. Thus, there were numerous large gaps between the graphene and Au. In addition, it has been reported that the graphene transferred by PDMS stamping is blistered,55 which may have produced additional gaps between the Au and graphene. Since gaps prohibit heat transfer between materials, a higher thermal resistance could result from these gaps. Moreover, the diameter of each point contact area is much less than the phonon mean free path of graphite. Therefore, phonons from graphene cannot enter Au electrodes easily, which could significantly increase the thermal contact resistance. Figure 4g shows the simulated temperature distribution of the graphene channel. The temperature at the center of the graphene is the highest because finite heat dissipation into the Au electrodes occurred near the electrodes. However, the channel temperature distribution realized by self-heating is almost constant and similar to that obtained using the hot chuck. Figure 5a shows the sensor response as a function of the temperature (using self-heating and the hot chuck). The sensor response with self-heating is smaller than that resulting from using the hot chuck at high temperatures. Although the temperature distribution obtained using self-heating differs from that resulting from using the hot chuck by 15 °C, as shown in Figure 4g, the sensor response obtained at 130 °C with self-heating is smaller than that at 115 °C on the hot chuck. Therefore, the sensor response reduction in self-heated graphene cannot be explained by the difference between the channel temperature distributions. To clarify the reason for the sensor response reduction, Monte Carlo simulations were performed. The self-heating effect, suspended structure, and electric field effects were considered in the simulations. H2 sensing was taken into account by introducing charged impurities. The sensor response was calculated from the conductivity changes. As shown in Figure 5b, the calculation data indicate that increasing the electric field decreased the sensor response. The VD of 1.1 V in self-heated graphene corresponds to a lateral electric field of approximately 400 kV/µm. Although the experimentally observed response reduction is greater than that calculated, the difference is basically due to the simplified model used for the potentials induced by hydrogens. To illustrate the different effects caused by high and low electric fields, schematic diagrams are provided in Figure 5c. Hydrogens are 10

ACS Paragon Plus Environment

Page 10 of 25

Page 11 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Nano Materials

dissolved and adsorbed on the Pd NPs during H2 sensing. In a low lateral electric field, the holes are scattered greatly by polarized H atoms on the Pd NPs. However, the holes are less scattered in a high lateral electric field because of their high kinetic energy obtained from the electric field. Thus, the response in self-heated graphene is smaller than that in graphene heated with a hot chuck in a high electric field. Finally,

we

demonstrated

voltage-controlled

multi-molecule

detection

by

self-heating. Figure 6a shows the time dependence of the sensor response to humidity at a VD of 0.01 V. In this measurement, the reference and test gases were humid airs with RHs of 50% and 80%, respectively. The sensor resistance decreased during exposure to the higher humidity of 80%, because water molecules were adsorbed on the oxidized Pd and acted as acceptors (hole donors) for the graphene. As a result, the graphene was hole doped, and the hole concentration increase led to a resistance decrease. Figure 6b shows the sensor responses to 10-ppm H2 with constant RH of 0% and 50% at a drain voltage of 0.9 V. As shown in Figure 6b, the sensor responses were almost the same even under different RH conditions, thanks to the successful elimination of humidity around the Pd NPs by self-heating. To confirm the humidity robustness, we also explored the sensor response to 10-ppm H2 under humidity variation. Figure 6c shows the sensor response to 10-ppm H2 in humid air with a RH of 80% as the test gas. Again, humid air with a RH of 50% was used as the reference gas. As demonstrated in Figure 6c, the sensor function can be switched by changing VD for the same set of test and reference gases. The upper figure shows that at a VD of 0.1 V, around RT, the sensor responded to humidity. On the other hand, the lower figure shows that at a VD of 0.9 V, at an increased temperature of approximately 135 °C, the sensor responded to 10 ppm H2 even when a large RH change from 50% to 80% occurred simultaneously. Therefore, self-heating successfully prevented the effect of humidity and resulted in the detection of a low concentration of H2. These results lead to the conclusion that the sensor function can be changed or controlled using applied voltages, thanks to the self-heating effects. Further research is required to ensure high selectivity of H2 if other gas is mixed in humid air. We consider that the utilization of different temperatures realized by self-heating could be one of the efficient methods to distinguish one gaseous molecule to another using specific catalytic reactions at each temperature. Sensing properties of our sensors were compared with previous reports. As shown in Table 1, our sensor responded to low concentration of H2 even under great humidity variations. The multi-functionality realized with self-heating is advantageous in terms of the time necessary to change between functions. If a conventional external micro heater is used, the time necessary to change the sensor temperature should be much longer. 11

ACS Paragon Plus Environment

ACS Applied Nano Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 6d shows the simulated temperature change as a function of time, revealing that a temperature increase to 460 K and a decrease to 300 K can each be realized within 1 μs. The short time necessary to change the functionality results from the small thermal capacitances of self-heated materials.

CONCLUSION In summary, the Pd-functionalized, suspended graphene sensor fabricated in this study demonstrates the low H2 detection limit and improved performance achieved using a self-heating technique with low power consumption. The response time to 100-ppm H2 was decreased to 15 s, and the sensor response was increased in the self-heated graphene sensor compared to that at RT. We revealed the effect of a high lateral electric field during self-heating on the sensor response; sensor response reduction was caused by less scattering with potentials induced by target molecules when a high lateral electric field was present in the graphene channel. It was demonstrated that 10-ppm H2 can be detected even with significant RH variation from 50% to 80%. We also showed the possibility of selective sensing of hydrogen/humidity by applying an appropriate bias voltage in realistic gas environments, in which small changes in the H2 concentration and large variations in the RH were induced. The sensor reacted to humidity at a lower bias and detected H2 at a higher bias. Furthermore, we showed rapid channel temperature change within 1 µs by simulation. These results pave the way for mobile, fast, low-power, and real-time molecular sensors for use in health diagnosis applications accessible anywhere at any time.

AUTHOR INFORMATION Corresponding Authors *E-mail: [email protected]

ACKNOWLEDGEMENT This work was supported by JST CREST Grant Number JPMJCR1331.

REFERENCES (1)

Norma K.C. Ramsay, M. D. Prospective Comparison of Indirect Methods for

(2)

Detecting Lactase Deficiency. N. Engl. J. Med. 1982, 306, 392–397. Fuchs, P.; Loeseken, C.; Schubert, J. K.; Miekisch, W. Breath Gas Aldehydes as

(3)

Biomarkers of Lung Cancer. Int. J. Cancer 2010, 126, 2663–2670. Filipiak, W.; Ruzsanyi, V.; Mochalski, P.; Filipiak, A.; Bajtarevic, A.; Ager, C.; 12

ACS Paragon Plus Environment

Page 12 of 25

Page 13 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Nano Materials

Denz, H.; Hilbe, W.; Jamnig, H.; Hackl, M.; et al. Dependence of Exhaled Breath Composition on Exogenous Factors, Smoking Habits and Exposure to Air (4) (5)

Pollutants. J. Breath Res. 2012, 6, 036008. Shin, W. Medical Applications of Breath Hydrogen Measurements. Anal. Bioanal. Chem. 2014, 406, 3931–3939. Wang, B.; Zhu, L. F.; Yang, Y. H.; Xu, N. S.; Yang, G. W. Fabrication of a SnO 2 Nanowire Gas Sensor and Sensor Performance for Hydrogen. J. Phys. Chem. C

(6)

2008, 112, 6643–6647. Ghoshal, U. C. How to Interpret Hydrogen Breath Tests. J. Neurogastroenterol.

(7)

Motil. 2011, 17, 312–317. Dewit, O.; Pochart, P.; Desjeux, J.-F. Breath Hydrogen Concentration and Plasma Glucose, Insulin and Free Fatty Acid Levels after Lactose, Milk, Fresh or Heated Yogurt Ingestion by Healthy Young Adults with or without Lactose

(8)

Malabsorption. Nutrition 1988, 4, 131–136. Lu, C.; Chen, Z. High-Temperature Resistive Hydrogen Sensor Based on Thin Nanoporous Rutile TiO2 Film on Anodic Aluminum Oxide. Sensors Actuators, B

(10)

Chem. 2009, 140, 109–115. Varghese, O. K.; Gong, D.; Paulose, M.; Ong, K. G.; Grimes, C. A. Hydrogen Sensing Using Titania Nanotubes. In Sensors and Actuators, B: Chemical; 2003; Vol. 93, pp. 338–344. Qurashi, A.; Yamazaki, T.; El-Maghraby, E. M.; Kikuta, T. Fabrication and Gas

(11)

Sensing Properties of In2 O3 Nanopushpins. Appl. Phys. Lett. 2009, 95, 1–4. Jun, Y. K.; Kim, H. S.; Lee, J. H.; Hong, S. H. High H2 Sensing Behavior of

(9)

TiO2 Films Formed by Thermal Oxidation. Sensors Actuators, B Chem. 2005, 107, 264–270. (12) Fine, G. F.; Cavanagh, L. M.; Afonja, A.; Binions, R. Metal Oxide

(13)

Semi-Conductor Gas Sensors in Environmental Monitoring. Sensors 2010, 10, 5469–5502. Meng, G.; Zhuge, F.; Nagashima, K.; Nakao, A.; Kanai, M.; He, Y.; Boudot, M.; Takahashi, T.; Uchida, K.; Yanagida, T. Nanoscale Thermal Management of Single SnO 2 Nanowire: Pico-Joule Energy Consumed Molecule Sensor. ACS

(14)

(15)

Sensors 2016, 1, 997–1002. Tanaka, T.; Hoshino, S.; Takahashi, T.; Uchida, K. Nanoscale Pt Thin Film Sensor for Accurate Detection of Ppm Level Hydrogen in Air at High Humidity. Sensors Actuators B Chem. 2018, 258, 913–919. Lewis, F. a. The Palladium-Hydrogen System. Platin. Met. Rev. 1982, 26, 20–27. 13

ACS Paragon Plus Environment

ACS Applied Nano Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(16)

Lundström, I. Hydrogen Sensitive Mos-Structures. Part 1: Principles and

(17)

Applications. Sensors and Actuators 1981, 1, 403–426. Lundström, I.; Söderberg, D. Hydrogen Sensitive Mos-Structures Part 2:

(18)

Characterization. Sensors and Actuators 1981, 2, 105–138. Manchester, F. D.; San-Martin, A.; Pitre, J. M. The H-Pd (Hydrogen-Palladium)

(19)

System. J. Phase Equilibria 1994, 15, 62–83. Lundström, I.; Shivaraman, S.; Svensson, C.; Lundkvist, L. A

(20) (21)

Hydrogen-Sensitive MOS Field-Effect Transistor. Appl. Phys. Lett. 1975, 26. Öztürk, S.; Kılınç, N. Pd Thin Films on Flexible Substrate for Hydrogen Sensor. Journal of Alloys and Compounds, 2016, 674, 179–184. Itoh, T.; Matsubara, I.; Kadosaki, M.; Sakai, Y.; Shin, W.; Izu, N.; Nishibori, M. Effects of High-Humidity Aging on Platinum, Palladium, and Gold Loaded Tin

(22)

Oxide-Volatile Organic Compound Sensors. Sensors 2010, 10, 6513–6521. Ahn, J.-H.; Yun, J.; Moon, D.-I.; Choi, Y.-K.; Park, I. Self-Heated Silicon Nanowires for High Performance Hydrogen Gas Detection. Nanotechnology

(23)

2015, 26, 095501. Lee, E.; Lee, J. M.; Koo, J. H.; Lee, W.; Lee, T. Hysteresis Behavior of Electrical Resistance in Pd Thin Films during The Process of Absorption and Desorption of

(24)

(25)

(26) (27)

(28)

Hydrogen Gas. Int. J. Hydrogen Energy 2010, 35, 6984–6991. Yang, F.; Taggart, D. K.; Penner, R. M. Joule Heating a Palladium Nanowire Sensor for Accelerated Response and Recovery to Hydrogen Gas. Small 2010, 6, 1422–1429. Schedin, F.; Geim, A. K.; Morozov, S. V.; Hill, E. W.; Blake, P.; Katsnelson, M. I.; Novoselov, K. S. Detection of Individual Gas Molecules Adsorbed on Graphene. Nat. Mater. 2007, 6, 652–655. Geim, A. K.; Novoselov, K. S. The Rise of Graphene. Nat. Mater. 2007, 6, 183– 191. Yoon, H. J.; Jun, D. H.; Yang, J. H.; Zhou, Z.; Yang, S. S.; Cheng, M. M. C. Carbon Dioxide Gas Sensor Using a Graphene Sheet. Sensors Actuators, B Chem. 2011, 157, 310–313. Choi, H.; Jeong, H. Y.; Lee, D.-S.; Choi, C.-G.; Choi, S.-Y. Flexible NO 2 Gas Sensor Using Multilayer Graphene Films by Chemical Vapor Deposition.

(29)

Carbon Lett. 2013, 14, 186–189. Dan, Y.; Lu, Y.; Kybert, N. J.; Luo, Z.; Johnson, A. T. C. Intrinsic Response of

(30)

Graphene Vapor Sensors. Nano Lett. 2009, 9, 1472–1475. Johnson, J. L.; Behnam, A.; Pearton, S. J.; Ural, A. Hydrogen Sensing Using 14

ACS Paragon Plus Environment

Page 14 of 25

Page 15 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Nano Materials

Pd-Functionalized Multi-Layer Graphene Nanoribbon Networks. Adv. Mater. (31)

2010, 22, 4877–4880. Goto, A.; Takeuchi, G.; Yamachi, R.; Tanaka, T.; Takahashi, T.; Uchida, K. Impact of Hydrogen on Carrier Mobility and Concentration in Graphene

(32)

Decorated with Pd Nanoparticle. ECS Trans. 2016, 72, 7–12. Chung, M. G.; Kim, D. H.; Seo, D. K.; Kim, T.; Im, H. U.; Lee, H. M.; Yoo, J. B.; Hong, S. H.; Kang, T. J.; Kim, Y. H. Flexible Hydrogen Sensors Using Graphene with Palladium Nanoparticle Decoration. Sensors Actuators, B Chem.

2012, 169, 387–392. (33) Kim, Y. H.; Kim, S. J.; Kim, Y. J.; Shim, Y. S.; Kim, S. Y.; Hong, B. H.; Jang, H. W. Self-Activated Transparent All-Graphene Gas Sensor with Endurance to (34)

Humidity and Mechanical Bending. ACS Nano 2015, 9. Lange, U.; Hirsch, T.; Mirsky, V. M.; Wolfbeis, O. S. Hydrogen Sensor Based on

(36)

a Graphene-Palladium Nanocomposite. Electrochim. Acta 2011, 56, 3707–3712. Xia, Y. N.; Whitesides, G. M. Soft Lithography. Annu. Rev. Mater. Sci. 1998, 37, 551–575. Whitesides, G. M.; Ostuni, E.; Jiang, X.; Ingber, D. E. Soft Lithography in

(37)

Biology and Biochemistry. Annu. Rev. Biomed. Eng. 2001, 3, 335–373. Huang, Y. Y.; Zhou, W.; Hsia, K. J.; Menard, E.; Park, J. U.; Rogers, J. A.;

(35)

(38)

(39)

Alleyne, A. G. Stamp Collapse in Soft Lithography. Langmuir 2005, 21, 8058– 8068. Hua, F.; Sun, Y.; Gaur, A.; Meitl, M. A.; Bilhaut, L.; Rotkina, L.; Wang, J.; Geil, P.; Shim, M.; Rogers, J. A.; et al. Polymer Imprint Lithography with Molecular-Scale Resolution. Nano Lett. 2004, 4, 2467–2471. Jayasena, B.; Melkote, S. N. An Investigation of PDMS Stamp Assisted Mechanical Exfoliation of Large Area Graphene. Procedia Manuf. 2015, 1, 840– 853.

(40)

Vijh, A. K. The Influence of Metal-Metal Bond Energies on the Adhesion,

(41)

Hardness, Friction and Wear of Metals. J. Mater. Sci. 1975, 10, 998–1004. Wang, J.; Zhu, L.; Chen, J.; Li, B.; Thong, J. T. L. Suppressing Thermal Conductivity of Suspended Tri-Layer Graphene by Gold Deposition. Adv. Mater.

(42)

(43)

2013, 25, 6884–6888. Liao, A.; Alizadegan, R.; Ong, Z. Y.; Dutta, S.; Xiong, F.; Hsia, K. J.; Pop, E. Thermal Dissipation and Variability in Electrical Breakdown of Carbon Nanotube Devices. Phys. Rev. B - Condens. Matter Mater. Phys. 2010, 82, 1–9. Liao, A. D.; Wu, J. Z.; Wang, X.; Tahy, K.; Jena, D.; Dai, H.; Pop, E. Thermally 15

ACS Paragon Plus Environment

ACS Applied Nano Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Limited Current Carrying Ability of Graphene Nanoribbons. Phys. Rev. Lett. (44)

2011, 106, 2–5. Chauhan, J.; Guo, J. High-Field Transport and Velocity Saturation in Graphene.

(45)

Appl. Phys. Lett. 2009, 95, 1–4. Borysenko, K. M.; Mullen, J. T.; Barry, E. A.; Paul, S.; Semenov, Y. G.; Zavada, J. M.; Nardelli, M. B.; Kim, K. W. First-Principles Analysis of Electron-Phonon

(46)

Interactions in Graphene. Phys. Rev. B - Condens. Matter Mater. Phys. 2010, 81, 3–6. Hirai, H.; Tsuchiya, H.; Kamakura, Y.; Mori, N.; Ogawa, M. Electron Mobility

(47)

Calculation for Graphene on Substrates. J. Appl. Phys. 2014, 116. Ni, Z.; Wang, Y.; Yu, T.; Shen, Z. Raman Spectroscopy and Imaging of

(48)

Graphene. Nano Res. 2008, 1, 273–291. López-Corral, I.; Germán, E.; Juan, A.; Volpe, M. A.; Brizuela, G. P. DFT Study of Hydrogen Adsorption on Palladium Decorated Graphene. J. Phys. Chem. C

(49)

2011, 115, 4315–4323. Nagashio, K.; Nishimura, T.; Kita, K.; Toriumi, A. Mobility Variations in Mono-

(50)

and Multi-Layer Graphene Films. Appl. Phys. Express 2009. Kumar, R.; Malik, S.; Mehta, B. R. Interface Induced Hydrogen Sensing in Pd

(51)

Nanoparticle/Graphene Composite Layers. Sensors Actuators, B Chem. 2015, 209, 919–926. Xia, F.; Perebeinos, V.; Lin, Y. M.; Wu, Y.; Avouris, P. The Origins and Limits

(52)

of Metal-Graphene Junction Resistance. Nat. Nanotechnol. 2011, 6, 179–184. Dorgan, V. E.; Bae, M. H.; Pop, E. Mobility and Saturation Velocity in Graphene

(53)

(54)

on SiO2. Appl. Phys. Lett. 2010, 97, 2010–2012. Mao, R.; Kong, B. D.; Gong, C.; Xu, S.; Jayasekera, T.; Cho, K.; Kim, K. W. First-Principles Calculation of Thermal Transport in Metal/Graphene Systems. Phys. Rev. B 2013, 87, 165410. Schmidt, A. J.; Collins, K. C.; Minnich, A. J.; Chen, G. Thermal Conductance and Phonon Transmissivity of Metal–Graphite Interfaces. J. Appl. Phys. 2010, 107, 1–5.

(55)

(56)

Goler, S.; Piazza, V.; Roddaro, S.; Pellegrini, V.; Beltram, F.; Pingue, P. Self-Assembly and Electron-Beam-Induced Direct Etching of Suspended Graphene Nanostructures. J. Appl. Phys. 2011, 110, 1–17. Jang, B.; Kim, W.; Song, M. J.; Lee, W. Reliability and Selectivity of H2 Sensors Composed of Pd Film Nanogaps on an Elastomeric Substrate. Sensors Actuators, B Chem. 2017, 240, 186–192. 16

ACS Paragon Plus Environment

Page 16 of 25

Page 17 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Nano Materials

(57)

(58)

Jung, S.; Baik, K. H.; Ren, F.; Pearton, S. J.; Jang, S. Pt-AlGaN/GaN Hydrogen Sensor with Water-Blocking PMMA Layer. IEEE Electron Device Lett. 2017, 38, 657–660. Krško, O.; Plecenik, T.; Roch, T.; Grančič, B.; Satrapinskyy, L.; Truchlý, M.; Ďurina, P.; Gregor, M.; Kúš, P.; Plecenik, A. Flexible Highly Sensitive Hydrogen Gas Sensor Based on a TiO2thin Film on Polyimide Foil. Sensors Actuators, B

(59)

Chem. 2017, 240, 1058–1065. Mubeen, S.; Zhang, T.; Yoo, B.; Deshusses, M. A.; Myung, N. V. Palladium Nanoparticles Decorated Single-Walled Carbon Nanotube Hydrogen Sensor. J. Phys. Chem. C 2007, 111, 6321–6327.

17

ACS Paragon Plus Environment

ACS Applied Nano Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 25

Table 1. Hydrogen sensor properties in humid ambient. The sensor of this study is compared with previous reports22,56–59

H2

Material

Concentration

56

Pd

57

Pt/AlGaN/GaN TiO2

58

CNT/Pd59 Silicon Nanowire22 Graphene/Pd (this study)

RH

Reference

Response

Recovery

Gas

Time

Time

400 ppm

90% (const.)

N2

< 30 s

1s

500 ppm

100% (const.)

N2

4s

< 30 s

300 ppm

32% (const.)

Air

< 30 s

< 20s

300 ppm

80% (const.)

Air

18 min

20 min

5000 ppm

85% (const.)

Air

6s

11 s

Air

15 s

16 s

10 ppm

5080% (variation)

18

ACS Paragon Plus Environment

Page 19 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Nano Materials

Figure 1. (a) Schematic diagram of the graphene hydrogen sensor. (b) Measurement setup of gas molecular sensing. (c) Typical time dependence of the resistance change. The sensor response was defined as the resistance change ratio. The response time was defined as the time necessary for the sensor response to reach half of its maximum value during the sensing duration. The sensing duration was 3 min. 170x109mm (300 x 300 DPI)

ACS Paragon Plus Environment

ACS Applied Nano Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 2. (a) Optical and SEM images of the suspended graphene sensor. (b) Raman spectrum of graphene on Au (black line) and graphene (red line). The Raman spectrum of graphene was obtained by subtracting the Raman spectrum of the Au from that of the graphene on the Au. (c) Cross-sectional TEM image of the Pd NPs on the graphene and Au. (d) AFM image of Au electrode without Pd deposition. (e) AFM image of Au electrode with transferred graphene and Pd deposition. In the lower areas, graphene was transferred onto Au electrode before the Pd deposition. 170x109mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 20 of 25

Page 21 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Nano Materials

Figure 3. (a) ID versus VD characteristics of unfunctionalized and Pd-functionalized graphene. (b) Temperature dependence of the sensor response. The device was heated by an external heater (hot chuck). Initial resistances of R0 are 486, 463, 488, 508,538 Ω in the temperature ascending order. Sensor response as a function of time at VD (c) lower than and (d) greater than 0.8 V. Initial resistances of R0 are 481, 522, 525, 526, 526, 533, 548, 564 Ω in the VD ascending order. (e) Relationship between sensor response and H2 concentration. The inset shows the time dependence of sensor response for various H2 concentrations, which were used to extract the relationship. 170x119mm (300 x 300 DPI)

ACS Paragon Plus Environment

ACS Applied Nano Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 4. (a) Normalized sensor responses of graphene sensors heated by self-heating (black) at VD of 0.9 V and the hot chuck (red) at 135°C at VD of 0.01 V. During the self-heating measurements, the hot chuck temperature was kept at 300 K. Sensor responses normalized by the maximum values matched with each other. The close agreement suggests that the temperature increase by self-heating was successfully realized. (b) Dependences of the response time on the hot chuck temperature (blue crosses) and power used by self-heating (red squares). The close agreement indicates that the self-heated graphene temperature can be estimated from the response time. (c) Temperature as a function of self-heating power for the experimental data extracted from Figure 3b and simulation data. The simulation data were obtained using the finite element method simulator COMSOL Multiphysics. (d) Drain current versus drain voltage characteristics of suspended graphene functionalized with Pd NPs. Breakdown voltage was determined to be 1.5 V. (e) Simulated temperature distribution in suspended graphene shown in Figure 4d. The simulation was performed with COMSOL Multiphysics. The graphene temperature was increased to approximately 900 K when the interfacial thermal resistance was set to 2500 µm2·K/mW. (f) TEM images of Pd/graphene on an Au electrode. Numerous gaps were formed between the graphene and Au; the areas of contact between the graphene and Au were small. We expect the small contact areas to result in large thermal contact resistance between the graphene and Au. (g) Simulated position dependence of graphene channel temperature using the self-heating technique (red) and hot chuck (black). 170x170mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 22 of 25

Page 23 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Nano Materials

ACS Paragon Plus Environment

ACS Applied Nano Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 5. (a) Temperature dependences of the sensor response using the self-heating technique (red squares) and hot chuck (black squares). The sensor response in the case of self-heating is smaller than that resulting from using the hot chuck. (b) Lateral electric field dependence of the sensor response calculated by Monte Carlo simulation. (c) Schematic diagrams showing the differences between the effects of Coulomb scattering in a low lateral electric field (left) and a high lateral electric field (right). 170x129mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 24 of 25

Page 25 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Nano Materials

Figure 6. (a) RT time dependence of the sensor response to a RH change from 50% to 80% at a VD of 0.01 V. Initial resistance of R0 is 456 Ω. The humidity decreases the sensor resistance. (b) Sensor response to 10-ppm H2 in humid air with constant RH of 0% (black line) and 50% (red line) as a function of time. Initial resistances of R0 are 519 Ω and 520 Ω, respectively. (c) RT time dependence of the sensor response to 10ppm H2 with a RH increase from 50% to 80% at a VD of 0.1 V (upper) and 0.9 V (lower). Initial resistances of R0 when applied VD of 0.1 and 0.9 V are 471 Ω and 520 Ω, respectively. Because of the RH increase, the sensor resistance is decreased at the lower VD of 0.1 V, which agrees with the data shown in Figure 5a. Meanwhile, at the higher VD of 0.9 V, thanks to the temperature increase induced by self-heating, the sensor resistance was increased by the H2. (d) Channel temperature change as a function of time following the VD change. Using the self-heating technique, the temperature both increases and decreases in 1 µs. 170x180mm (300 x 300 DPI)

ACS Paragon Plus Environment

ACS Applied Nano Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

TOC Graphics 83x35mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 26 of 25