Micropatternable Double-Faced ZnO Nanoflowers for Flexible Gas

Sep 8, 2017 - Micropatternable double-faced (DF) zinc oxide (ZnO) nanoflowers (NFs) for flexible gas sensors have been successfully fabricated on a po...
1 downloads 10 Views 2MB Size
Subscriber access provided by GRIFFITH UNIVERSITY

Article

Micro-patternable double-faced ZnO nanoflowers flexible gas sensor Jong Woo Kim, Yoann Porte, Kyung Yong Ko, Hyungjun Kim, and Jae-Min Myoung ACS Appl. Mater. Interfaces, Just Accepted Manuscript • DOI: 10.1021/acsami.7b09251 • Publication Date (Web): 08 Sep 2017 Downloaded from http://pubs.acs.org on September 10, 2017

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 free 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 accessible to all readers and 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.

ACS Applied Materials & Interfaces 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 36

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 Materials & Interfaces

Micro-patternable double-faced ZnO nanoflowers flexible gas sensor Jong-Woo Kim,† Yoann Porte,† Kyung Yong Ko,‡ Hyungjun Kim‡ and Jae-Min Myoung*,†



Department of Materials Science and Engineering, Yonsei University, 50 Yonsei-ro, Soedaemun-gu, Seoul 03722, Republic of Korea



School of Electrical and Electronic Engineering, Yonsei University, 50 Yonsei-ro, Soedaemungu, Seoul 03722, Republic of Korea

*E-mail: [email protected]

ACS Paragon Plus Environment

1

ACS Applied Materials & Interfaces

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 2 of 36

ABSTRACT

Micro-patternable double-faced (DF) zinc oxide (ZnO) nanoflowers (NFs) flexible gas sensors have been successfully fabricated on a polyimide (PI) substrate with single wall carbon nanotubes (SWCNTs) as electrode. The fabricated sensors are constituted of ZnO nanoshells layed out on a PI substrate at regular intervals, on which ZnO nanorods (NRs) were grown inand out-side the shells to maximize the surface area and form a connected network. This threedimensional (3D) network structure possesses multiple gas diffusion channels and the micropatterned island structure allows to enhance the stability of flexible devices by dispersing the strain into the empty spaces of the substrate. Moreover, the micro-patterning technique on a flexible substrate enables fabricating highly integrated nanodevices. The SWCNTs were chosen as electrode for their flexibility and the Schottky barrier they form with ZnO, improving the sensing performances. The devices exhibited high selectivity towards NO2 as well as outstanding sensing characteristics with a stable response of 218.1, fast rising and decay times of 25.0 and 14.1 sec, respectively, and percent recovery greater than 98% upon NO2 exposure. The superior sensing properties arose from a combination of high surface area, numerous active junction points, donor point defects in ZnO NRs, and the use of SWCNTs electrode. Furthermore, the DF-ZnO NFs gas sensors showed sustainable mechanical stability. Despite the physical degradation observed, the devices still demonstrated outstanding sensing characteristics after 10000 bending cycles at a curvature radius of 5 mm.

KEYWORDS: NO2 gas sensor, ZnO nanoflower, SWCNT, micro-patterned array, networkstructured flexible device

ACS Paragon Plus Environment

2

Page 3 of 36

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 Materials & Interfaces

INTRODUCTION With the development of technologies such as automation, virtual reality, and augmented reality, the demand for accurate and reliable sensor technology is expanding. Among the various types of sensors, gas sensors have attracted considerable attention due to their role in the monitoring of atmospheric environment and human health for medical diagnosis, as well as in the detection of pollutants and toxic gases such as nitrogen dioxide (NO2), ammonia (NH3), carbon monoxide (CO), and sulfur dioxide.1-3 Especially, NO2 is the most common air pollutant produced by the fossil-fuel combustion and vehicle emissions. Recently, it has been identified as a major component of fine particulate matter which is a global issue, and efforts are underway to reduce the amount of emissions.4 In addition to environmental problems, NO2 causes skin damage even on very small amount of exposures, and critical respiratory disorder when exposed to about 50 parts per million (ppm).5 Therefore, it is necessary to develop an improved gas sensor using various techniques and materials for detecting trace of NO2. Metal-oxide semiconductor-based gas sensors have been widely investigated owing to their remarkable sensing property as well as simple sensing mechanism relying on changes of the resistivity.6 Moreover, in the past few years, with the recent advances in nanotechnology, researches have been focused on the use of nanomaterials with high surface-to-volume ratio, which highly affected the characteristics of gas sensors.7 Among metal-oxides, zinc oxide (ZnO) can be easily synthesized as nanomaterials such as nanobelts, nanorods (NRs), and nanowires using a low-temperature solution process.8,9 These ZnO nanomaterials have already been applied to various gas sensors including hydrogen, NO2, NH3, CO, and ethanol due to their excellent properties; i.e. high electrochemical and environmental stability, high crystallinity, low cost, no toxicity, easy synthesis, and large specific surface area.10-12 Despite these advantages, the ZnO-

ACS Paragon Plus Environment

3

ACS Applied Materials & Interfaces

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 4 of 36

based gas sensors generally operate at high temperatures, and their sensing performances still need to be improved in terms of response, saturation behavior, characteristic times, recovery, and selectivity.13 Since most of the conventional ZnO-based gas sensors generally have crosssensitivity to various gases, it is necessary to develop highly selective sensors with improved gas response.14 Recently, various studies have been carried out to improve the characteristics of ZnO-based gas sensors such as metal-assisted functionalization, application of porous morphology materials, and structural modulation with multiple junctions by tetrapod structure and nanobrunchs.15-17 These techniques allowed to improve the performances of the sensors through the effects of large effective area which induces multiple gas diffusion channels and tunes the concentration of charge carriers.18 However, the high temperature fabrication process and the post-treatments remain critical for their use in industry as they increase the process costs. Moreover, the structures using porous nanomaterials or multiple junctions have limited application in the field of flexible devices due to fragile geometry and random arrangement of nanomaterials.15-17 Moreover, despite advances in nanotechnology, there still exists a problem of evaluating the properties of nanomaterials and fabricating highly integrated nanodevices.19 With the demand for more compact devices, as well as the application to large surface areas, accurate position control technology of nanomaterials such as vertical stacking or self-assembly techniques is essential.20 In a recent study, a rubbing technique for precise alignment of fine particles was investigated. This rubbing technique facilitates accurate positioning and assembly of microparticles (MPs) in a short time without additional solvent, post- and surface-treatments.21 The precise alignment of various functional particles can be controlled by using the rubbing technique, thereby realizing a highly integrated device.

ACS Paragon Plus Environment

4

Page 5 of 36

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 Materials & Interfaces

In this study, we have developed highly integrated micro-patternable ZnO gas sensors with improved sensing properties and stable characteristics under flexibility tests. To maximize the surface area of nanomaterials, ZnO NRs were grown in- and out-side the ZnO shells using lowtemperature solution process. The shells were arranged in island shape using the rubbing technique to relax the strain of the device and ensure stable mechanical characteristics. The NRs grown on both sides of the shells form a myriad of junctions and develop the double-faced (DF) ZnO nanoflowers (NFs) network structure. Furthermore, single wall carbon nanotubes (SWCNTs) were applied as electrode to improve the flexibility and form a Schottky barrier with ZnO in order to enhance the gas sensor performances. This highly integrated DF-ZnO NFs gas sensor exhibits outstanding sensing characteristics, especially in decay time, recovery, and selectivity, which are a problem in ZnO-based gas sensors. Moreover, the devices were still operating without significant degradation of the performances in terms of response, characteristic times, and recovery, even after dynamic bending tests of 10000 cycles at a curvature radius of 5 mm.

RESULTS AND DISCUSSION Figure 1 schematically describes the fabrication steps of micro-patternable DF-ZnO NFs gas sensors. Firstly, as a template for DF-ZnO NFs, the polystyrene MPs with an average diameter of 5 μm were transferred by polydimethylsiloxane (PDMS)-transfer technique on a circular wellpatterned glass substrate with a diameter of 5 μm and a height of 3 μm prepared by photolithography. Then, the MPs were aligned by being pushed into the circular well patterns by unidirectional PDMS rubbing technique. This technique is a simple method to align particles in a short time without any solvent or treatment. It facilitates the precise alignment of MPs over spinor spray-coating.21 Next, the 100 nm-thick ZnO films were deposited on the MPs array by using

ACS Paragon Plus Environment

5

ACS Applied Materials & Interfaces

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 6 of 36

radio-frequency magnetron sputtering. Then, the ZnO-coated MPs were transferred on polyvinylphenol (PVP) adhesive-coated polyimide (PI) substrate. This was followed by calcination at 250 °C for 3 h in air on a hot plate to improve the crystallinity of the ZnO films and remove the polystyrene to obtain high surface area to volume ratio. After removal of the polystyrene MPs, the remaining bowl-shaped ZnO shells possess a large surface area, which can be enhanced further by growing ZnO NRs on their surfaces. Following the growth of ZnO NRs on the in- and out-side surfaces of the shells by using the low-temperature aqueous solution process, micro-patterned DF-ZnO NFs network structures were formed, as shown in Figure S1. Finally, in order to obtain a high performance gas sensor with gas selectivity and high flexibility, SWCNTs were spray coated as finger-shaped electrode patterns. Due to the maximized surface area, improved crystallinity, and micro-patterned network structure, DF-ZnO NFs gas sensor can exhibit optimized performances. Furthermore, the micro-patterned network structures are effective for flexible devices by reducing the strain.22

ACS Paragon Plus Environment

6

Page 7 of 36

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 Materials & Interfaces

Figure 1. Schematic illustration of the fabrication process for DF-ZnO NFs network structure gas sensor.

The improved crystallinity after calcination and growth of NRs was confirmed by X-ray diffraction (XRD) and photoluminescence (PL) data analyses. Figure S2a and b show respectively the XRD and PL data before and after calcination, and after growth of NRs. The XRD data confirms the growth of ZnO with a wurzite structure as the diffraction peaks observed correspond to the Miller indices from the joint committee on powder diffraction standards #361541 data sheet. Before calcination, the ZnO film exhibits a (0002) preferred orientation with a secondary (21� 19) orientation. The crystallinity of the ZnO shells was improved after calcination. Finally, after the growth of NRs on the shells, a strong (0002) diffraction peak was measured

ACS Paragon Plus Environment

7

ACS Applied Materials & Interfaces

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 36

along with a minor peak for the (21� 10) plane, as shown in Figure S2a. This indicates the ZnO

NRs were successfully grown along the c-axis, and were highly crystalline. The improvement in crystallinity was also confirmed by PL analyses. Figure S2b shows the evolution of the PL spectra of ZnO before calcination, after calcination, and after growth of NRs. As the process progresses, the intensity of the near-band edge (NBE) emission at 378.61 nm as well as the broad deep-level emission (DLE) increase considerably due to improved crystallinity of ZnO. The broad DLE is mainly caused by the various defects induced by solution process and has significant impact on the sensor response. A more detailed PL analyses are discussed later.

Figure 2. (a) Top-view SEM image of MPs array in circular well patterns. Inset: Cross-sectional SEM image showing MPs array in height-optimized well patterns, (b) Tilted-view SEM image of aligned ZnO shells after transfer and calcination at 250 °C for 3 h in air on a hot plate. Inset: Top-view SEM image showing polystyrene-removed ZnO shells after calcination, (c) SEM image of DF-ZnO NFs network structure after growth of ZnO NRs on in- and out-side the shells, and (d) Magnified SEM image of the junctions between NRs.

ACS Paragon Plus Environment

8

Page 9 of 36

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 Materials & Interfaces

Scanning electron microscopy (SEM) analyses of the different step of the fabrication process have been carried and are presented in Figure 2. Figure 2a shows top-view SEM image of micropatterned MPs in circular well patterns using PDMS rubbing technique. The optimum height of the well pattern was optimized to stop the MPs from getting out of the wells, and has been identified to be about 0.6D (D: particle diameter, 5 μm), as shown in the inset of Figure 2a. Optical microscopy (OM) images before and after placing the MPs in the wells are shown in Figure S3a. After transfer and calcination, the polystyrene was completely removed and leaving only the aligned ZnO shells with large surface area on which the ZnO NRs could grow, as shown in the inset of Figure 2b. Figure 2b shows the aligned ZnO shells partially embedded in the PVPadhesive layer to form a point contact that minimizes the contact area with the substrate. The island configuration on the flexible substrate relaxes strain by dispersing it into the empty spaces of the substrate from the edge of the pattern. This strain relaxation effect is more effective as the size of the pattern is reduced.23,24 Therefore, the island configuration at regular intervals of the ZnO shells embedded in PVP is favorable for the mechanical flexibility of the device. Figure 2c shows the DF-ZnO NFs network structure after growth of ZnO NRs both in- and out-side the shells to maximize the surface area. The alignment of the shells on the PVP-adhesive was maintained during the growth of the NRs. The growth time and concentration were optimized to form a network structure through physical contact of the NRs between the shells. The NRs grown in- and out-side the shells differ slightly in shape owing to the irregular source supply inside the shells. Therefore, the NRs grown outside the shell are needle-shaped and dense (Figure 2d), while those grown inside the shells are much shorter with a lower density, as shown in Figure S3b. Due to the growth of NRs on both sides of the shells and the island configuration,

ACS Paragon Plus Environment

9

ACS Applied Materials & Interfaces

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 10 of 36

the DF-ZnO NFs network structure possesses a large surface area favorable to gas sensor and is highly resistant to strain, which is advantageous for application in flexible devices.

Figure 3. Gas sensing characteristics of DF-ZnO NFs gas sensor upon NO2 exposure at 270 °C and relative humidity of 34%. Characteristics under different gas concentrations (50, 100, 200, 300, 400, and 500 ppm): (a) Gas response as a function of time, (b) Response, rising and decay times, and (c) Percent recovery. Repetitive operating characteristics at 500 ppm: (d) Gas response as a function of time, (e) Response, rising and decay times, and (f) Percent recovery.

The gas sensing performances of DF-ZnO NFs gas sensors were systematically investigated with a constant voltage bias of 1 V applied between each pair of SWCNTs electrodes and under

ACS Paragon Plus Environment

10

Page 11 of 36

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 Materials & Interfaces

an average relative humidity of 34%. The response of the DF-ZnO NFs gas sensor was calculated to conduct a quantitative analysis and is defined as follows:25 response =

𝑅𝑅𝑔𝑔 𝑅𝑅0

(1)

Here, Rg and R0 are the resistance of the gas sensor after and before exposure to the target gas, respectively. The response represents the reactivity of the sensors and is a fundamental index for evaluating the selectivity and response rate through comparative analysis. First, the operating temperature for gas sensing was optimized as shown in Figure S4. Due to the thermal stability of the PI substrate and the degradation of the SWCNTs, the temperature did not exceed 300 °C. The devices exhibited the highest response, as well as the lowest deviation at an operating temperature of 270 °C. Therefore, all subsequent characteristic analyses of gas sensors were conducted at this temperature. The comparative analyses of NO2, NH3, and CO gas sensing performances were then carried out to select the optimal target gas and evaluate the operating characteristics of the DF-ZnO NFs gas sensors. Figure 3 shows the sensing properties of the DFZnO NFs gas sensors upon exposure to NO2. The exposure under the target gas and air was maintained for 500 sec, while the exposure concentration of NO2 was varied from 50 to 500 ppm. The sensor showed an increase in the resistance upon NO2 exposure, and the response of the sensors increased proportionally with increasing concentration of NO2, as shown in Figure 3a. The gas sensor showed perfect rising and decay saturation behavior even at a low NO2 concentration of 50 ppm, and the response of the gas sensor increased linearly from 29 to 218.1 as the NO2 gas concentration increased, as shown in Figure 3b. As the NO2 concentration increased from 50 to 500 ppm, the rising time was increased slightly from 25.0 to 30.8 sec but the decay time was decreased by half, from 28.1 to 14.1 sec, improving the sensing performances. The decay time corresponds to the time it takes for the resistance to returns to its original state. It

ACS Paragon Plus Environment

11

ACS Applied Materials & Interfaces

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 12 of 36

represents an important index in the evaluation of the sensor stability during repetitive operation. Therefore, this short decay time indicates that the DF-ZnO NFs gas sensor is capable of stable repetitive measurements. The recovery is an essential factor for gas sensing as it significantly affects the reliability and sustainability of the device. As metal oxide-based gas sensors often fail to recover, the degree of recovery is an important indicator of the quality of the gas sensor. Therefore, the recovery characteristics of the DF-ZnO NFs gas sensors were investigated by calculating the percent recovery, defined as follows:26 recovery (%) =

𝑅𝑅𝑔𝑔 − 𝑅𝑅𝑟𝑟 × 100 𝑅𝑅𝑔𝑔 − 𝑅𝑅0

(2)

Here, Rg is the resistance value after 500 sec of target gas exposure, Rr is the recovered

resistance value after 500 sec of air exposure, and R0 is the resistance value before exposure to the target gas. As shown in Figure 3c, the DF-ZnO NFs gas sensor showed an outstanding percent recovery value greater than 98% regardless of NO2 concentration. This demonstrates the reliability of the devices towards NO2 sensing. To evaluate the sustainability of the sensors, a repeatability test was conducted at fixed NO2 concentration of 500 ppm, at which the sensing characteristics such as response, rising and decay times are optimal. The sensor characteristics showed a stable saturation behavior without degradation even after several cycles, as shown in Figure 3d. The response and decay time did not show significant variations, with values around 210 and 17 sec, respectively, as shown in Figure 3e. As for the rising time, it gradually stabilized, reaching 27 sec. Figure 3f shows a stable recovery above 98%. Additionally, to evaluate the repeatability of the sensors under different concentrations of NO2, the characteristics of the sensor were measured while the exposure concentration of NO2 was decreased from 500 to 50 ppm. As shown in Figure S5, the response characteristics of the DF-ZnO NFs sensor remained excellent without showing signs of degradation even after the NO2 concentration was decreased

ACS Paragon Plus Environment

12

Page 13 of 36

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 Materials & Interfaces

to 50 ppm from a high concentration of 500 ppm. The response of the sensors was decreased proportionally with decreasing concentration of NO2. These results confirm the ability of the DFZnO NFs gas sensors to detect the NO2 gas while demonstrating outstanding and stable characteristics.

Figure 4. Gas sensing characteristics of DF-ZnO NFs gas sensor upon NH3 and CO exposure at 270 °C and relative humidity of 34%. Upon NH3 exposure under different concentrations (50, 100, 200, 300, 400, and 500 ppm): (a) Gas response as a function of time, (b) Response, rising and decay times, and (c) Percent recovery. Upon CO exposure under different concentrations (50, 100, 200, 300, 400, and 500 ppm): (d) Gas response as a function of time, (e) Response, rising and decay times, and (f) Percent recovery.

ACS Paragon Plus Environment

13

ACS Applied Materials & Interfaces

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 14 of 36

The performances of DF-ZnO NFs gas sensors upon NH3 and CO exposure are both shown in Figure 4. In contrast to the NO2 gas response behavior, the DF-ZnO NFs gas sensors showed a decrease in the resistance under NH3 and CO exposure. Such difference occurs due to NO2 acting as an electron acceptor when the gas molecule is adsorbed on the ZnO surface, while NH3 and CO act as donors.27 Unlike the NO2 gas response, the gas sensor did not show the saturation behavior of rising and decay upon the NH3 exposure regardless of the gas concentration, as shown in Figure 4a. Moreover, the response of the sensor was very low, below 1.4. The response increased from 1.24 to 1.37 in proportion to the NH3 concentration, but was abnormally high in the first cycle. Even though the rising and decay times varied from 87.4 to 130.5 sec and from 177 to 204 sec, respectively, no significant changes were observed with increasing concentration of NH3, as shown in Figure 4b. The devices also exhibited a lower recovery value, below 80%, as shown in Figure 4c. Figure 4d-f shows the sensor characteristics upon CO gas exposure with increasing concentration. The sensor showed a saturation behavior for the CO at high concentrations, but little response at low concentrations, as shown in Figure 4d. The sensor characteristics upon CO exposure seem to be similar to NH3 exposure. The response of the sensor was low and increased from 1.19 to 1.46 proportionally to the concentrations, but was abnormally high at 1.33 in the first cycle as in the case of NH3. The rising time decreased from 121 to 46 sec with increasing CO concentration. The decay time was not measureable at concentrations below 200 ppm, and was almost unchanged in at about 23 sec for concentration above 300 ppm, as shown in Figure 4e. Although the recovery gradually increased with the CO concentration, it remained weak with a peak value of only 56% at a maximum concentration of 500 ppm, as shown in Figure 4f. Additionally, when comparing Figure 4a and 4d, a shift in the

ACS Paragon Plus Environment

14

Page 15 of 36

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 Materials & Interfaces

baseline of the resistance was observed. This baseline shift depends on the degree of recovery of the device. As shown in Figure 4c and 4f, the percent recoveries of the devices upon NH3 and CO exposure are low. The characteristics of the device were evaluated by continuously exposing NO2, NH3, and CO under the same conditions and the baseline shift is observed when a gas exchange without a fully recovery. Thus, the baseline shift occurs due to a repetitive exposure to a low recovery gases. It is clear from these results that the DF-ZnO NFs gas sensors exhibited low sensitivity towards NH3 and CO gases. The responses measured were almost 2-orders of magnitude lower than with the NO2 gas. Furthermore, the recovery for the sensors under NH3 and CO exposure was below 80%, while under NO2 exposure the sensors showed an outstanding recovery of 98%. Therefore, the DF-ZnO NFs gas sensors have a high selectivity towards NO2 gas. The DF-ZnO NFs gas sensors owe their high selectivity and superior sensing performances for NO2 to several factors, such as the morphology, native point defects, and SWCNTs electrodes. The first factor is the maximized surface area obtained through the growth of ZnO NRs in- and out-side of the ZnO shells. It facilitates the diffusion of gas molecules and the surface reaction on ZnO.28 In addition, the three-dimensional (3D) network structures of ZnO shells and NRs improve the gas response.29,30 Generally, the sensing mechanisms of resistive change-type metaloxide semiconductor sensors were described using the grain-boundary model and the surfacedepletion model.31,32 The grain-boundary model is applicable for polycrystalline grain boundaries in films. The response measured is usually low since only a small fraction near the grain is involved in the reaction. In contrast, the surface-depletion model is applicable for nanomaterials such as nanoparticles, NRs, and nanoneedles. The response is higher due to the large surface area of nanomaterials which contributes to the formation of the surface depletion

ACS Paragon Plus Environment

15

ACS Applied Materials & Interfaces

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 16 of 36

region. This surface depletion region is caused by the adsorption of oxygen species acting as electron acceptors at the surface of oxide semiconductors. The adsorbed gases can interact with the electron donor on the surface to form two ionic states below the edge of the conduction band. The electron donor states on the oxide semiconductor surface are usually oxygen vacancies (VO), which possess a large secondary ionization energy.33 Therefore, the donor level becomes an ionic state of +1 valence and the surface depletion region is formed.34,35 When the sensor is exposed to NO2, NO2 is adsorbed on the surface of ZnO forming a surface depletion layer and reducing the concentration of electrons by acting as an acceptor. The reaction formulas can be described as follows:17,36 − 𝑁𝑁𝑁𝑁2(𝑔𝑔) + 𝑒𝑒 − → 𝑁𝑁𝑁𝑁2(𝑎𝑎𝑎𝑎)

− + 𝑂𝑂 − (𝑎𝑎𝑎𝑎) + 2𝑒𝑒 − → 𝑁𝑁𝑁𝑁(𝑔𝑔) + 2𝑂𝑂2− (𝑎𝑎𝑎𝑎) 𝑁𝑁𝑁𝑁2(𝑎𝑎𝑎𝑎)

(3)

(4)

Here, g and ad correspond to gas and adsorbate, respectively. After adsorption of the NO2, the surface depletion layer is formed over the entire surface of DF-ZnO NFs. In this work, the ZnO shells with a film-like structure correspond to the first model, while the NRs were applied to the

latter model. The DF-ZnO NFs gas sensor combines both models and allows complementary adsorption of the NO2 molecules on the shell grain boundaries and on the large surface area provided by the NRs.

ACS Paragon Plus Environment

16

Page 17 of 36

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 Materials & Interfaces

Figure 5. Schematic energy-band diagrams of DF-ZnO NFs gas sensor (a) before and (b) after exposure of NO2, and (c) Room-temperature PL spectra with deconvoluted Gaussian sub-peaks of the gas sensor.

An important parameter to consider for performances improvement is the presence of multiple active points in the DF-ZnO NFs network structure, corresponding to the junction points between the NRs. These junction points can be assimilated to grain boundaries and induce the formation of potential barriers, as illustrated in Figure 5a. These junctions correspond to active points where electrons from the conduction band of ZnO NRs can be transferred to the adsorbed NO2 acting as an electron acceptor. This leads to the formation of a depletion layer near the junction, and to an increase of the potential barrier, Figure 5b.37 The electron transport is affected by the height of the potential barriers and can be described by the following equation:17,32 𝑅𝑅 = 𝑅𝑅0 𝑒𝑒𝑒𝑒𝑒𝑒 �

−𝑒𝑒 ∆𝑉𝑉𝑏𝑏 � 𝑘𝑘𝐵𝐵 𝑇𝑇

(5)

ACS Paragon Plus Environment

17

ACS Applied Materials & Interfaces

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 36

Here, ∆𝑉𝑉𝑏𝑏 is the change in potential barrier defined as the potential in air minus that in target

gas, 𝑅𝑅0 is a factor including the intergranular resistance and geometrical effect, 𝑒𝑒 is the charge of an electron, 𝑘𝑘𝐵𝐵 is Boltzmann’s constant, and T is the absolute temperature. The electron transport

decreases when the potential barrier increases, therefore the adsorption of NO2 at junctions of NRs induce a decrease of the measured current. As a result, the presence of numerous junctions in DF-ZnO NFs network structures is crucial to demonstrate the superior properties of the gas sensors. Another factor affecting the characteristics of the ZnO-based gas sensors is the presence of native point defects as well as surface depletion layers. In particular, the NO2 gas sensing performances of ZnO are related to donor defects, which give rise to free electrons, such as zinc interstitial (Zni) and VO.38 The high selectivity for NO2 in the ZnO-based gas sensor is influenced by the donor defect state VO. Through the density function theory (DFT), it was found that the adsorption energy of NO2 on the VO site was larger than that of NH3 and CO.27 This indicates that NO2 adsorption on the ZnO surface would be improved in the presence of VO. As shown in Figure 5c, the PL spectrum analysis of the DF-ZnO NFs was carried out to identify point defects contributing to the properties of the gas sensor. The room-temperature PL spectrum consists of a sharp peak centered at 378.6 nm and a broad peak centered at about 560 nm corresponding to NBE emission and DLE, respectively. While the NBE is related to the bandgap of the material, the DLE is caused by the presence of multiple point defects in the material and whispering gallery mode (WGM) resonances. The WGM resonance is a phenomenon in which light waves circulate around at the resonator boundaries due to multiple internal reflections. The hexagonal needle-shaped 1D ZnO can act as a resonator, and the WGM phenomenon can be observed in the PL spectrum.39 In the case of ZnO, the DLE observed is generally attributed to native point

ACS Paragon Plus Environment

18

Page 19 of 36

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 Materials & Interfaces

defects such as Zni, VO, oxygen interstitial (Oi), and zinc vacancy (VZn) in ZnO.40 Their presence is caused by the nature of the aqueous solution process which induces the formation of multiple defects during the growth of the NRs.41 The broad DLE peak was deconvoluted into four Gaussian sub-peaks centered at 552.5, 606.7, 702.8, and 749.3 nm, corresponding to the photon energy of 2.24, 2.04, 1.76, and 1.65 eV, respectively. These sub-peaks are attributed to the electronic transitions from Zni to VO, Zni to Oi, VO to valence-band maximum (VBM), and conduction-band minimum (CBM) to VO, respectively.42 Among the deconvoluted sub-peaks, those corresponding to Zni and VO donor defects transitions are predominant, which confirms they are the main defects in the ZnO structure. The PL analyses allowed to demonstrate the presence of VO and Zni in the ZnO NFs, which significantly affect the gas sensing performances by improving the reactivity of NO2 on the ZnO surface. In addition to the presence of the native point defects, the use of SWCNTs also plays a positive role in increasing the selectivity to NO2 in the DF-ZnO NFs gas sensors. DFT studies on the interaction of various gas molecules with SWCNTs have been carried out, showing each molecule possesses different adsorption energy and degree of charge transfer to SWCNTs.43,44 While the electronic properties of SWCNTs are sensitive to NO2 adsorption, the SWCNTs were not only inactive with NH3 and CO gases but also demonstrated repulsive forces.45 Additionally, it is also known that the junction barrier height between the electrode and the semiconductor sensing materials plays an important role in the resistive change-type sensors.46 In the case of a Schottky barrier contact, the barrier height is directly dependent on the electron affinity (𝜒𝜒) of a semiconductor and the work function of a metal (∅𝑚𝑚 ). The Schottky barrier height can be described as follows:47,48

∅𝐵𝐵 = ∅𝑚𝑚 − 𝜒𝜒

(6)

ACS Paragon Plus Environment

19

ACS Applied Materials & Interfaces

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 20 of 36

Here, ∅𝐵𝐵 is the Schottky barrier under zero bias. The work function (∅𝑚𝑚 ) of the SWCNTs was

measured to be 5.02 eV using surface-analysis photoelectron spectrometer, as shown in Figure

S6. The electron affinity (𝜒𝜒) of ZnO is considered to be 4.35 eV from the reported literature.49 Considering ZnO as the sensing material, the Schottky barrier height with the SWCNT electrode is 0.67 eV, as schematically illustrated in Figure 5a. Therefore, it is believed that SWCNTs are a suitable electrode material for ZnO NRs to form a Schottky barrier, and thus improving the characteristics of the DF-ZnO NFs gas sensor.

Figure 6. Gas sensing properties of the DF-ZnO NFs flexible gas sensor during the dynamic bending test up to 10000 cycles with a curvature radius of 5 mm upon NO2 exposure at 500 ppm, 270 °C and relative humidity of 34%: (a) Camera image of the device on the bending machine, (b) Gas response as a function of time, (c) Response, rising and decay times, and (d) Percent recovery.

Lastly, in order to evaluate the performances of the DF-ZnO NFs gas sensor as a flexible device, a mechanical bending test was conducted under NO2 exposure. Figure 6a shows a camera

ACS Paragon Plus Environment

20

Page 21 of 36

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 Materials & Interfaces

image of the flexible DF-ZnO NFs gas sensor on a bending machine. The bending test was performed at a curvature radius of 5 mm. As shown in the response-time characteristic curve from Figure 6b, even though the electrical fluctuation of the device gradually increases with increasing number of bending cycles, the response characteristic appears to improve. In quantitative analyses, the response of the sensors was increased from 202.2 to 297.6 after 10000 cycles, as shown in Figure 6c. Moreover, the characteristic times were also improved. The rising and decay times were decreased from 36.0 to 19.2 and 20.3 to 11.5 sec, respectively. Furthermore, even the recovery increased from 89.4 to 94.8%, as shown in Figure 6d. Overall, the characteristics of the gas sensor appear to improve as the bending cycle increases from 0 to 10000. However, despite an improvement of the numerical values during the evaluation of the sensor characteristics, the resistance gradually increased while under the same conditions, along with significant fluctuations. This behavior implies a deterioration of the sensors. As shown in Figure S7, the degradation aspects of the sensor with increasing bending cycle were confirmed by SEM analyses. Overall, the aligned network structure embedded in the PVP-adhesive layer was maintained but partial fractures occurred. After flexibility tests of 10000 cycles, the NRs appeared as bent, indicating that a constant strain was applied to the NRs during the repetitive flexibility test. As a result of repetitive strain, the breakaway of NRs from shells occurred and is the first sign of device failure. On the other hand, the shells appeared to maintain their structure, and cracked shells were very rarely observed. Despite the deterioration of the sensor after bending cycles observed under SEM, the DF-ZnO NFs gas sensors showed saturation behavior and were still operating as NO2 detector, even after 10000 cycles of bending test. These results demonstrate the stability and high performances of the DF-ZnO NFs gas sensors.

ACS Paragon Plus Environment

21

ACS Applied Materials & Interfaces

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 22 of 36

Table 1. Comparison of the NO2 gas sensing properties of the ZnO-based gas sensor.

Synthesis method

Operating Concentration Temp. (ppm) (°C)

Gas response (R/R0)

Rising time (s)

Decay time (s)

Radius Number of of curvature bending (mm) cycles

Ref.

Low temp. Solution process

270

50 – 500

29 – 218.1

25 – 30.8

28.1 – 14.1

5

10000

This work

Low temp. Solution process

250

0.01

824

1200

3000

-

-

50

Low temp. Solution process

200

200

64

41

125

-

-

17

Low temp. Solution process

200

100

37.2

6.72

52.62

-

-

51

Low temp. Solution process

290

40

264

-

-

-

-

52

Low temp. Solution process

25

1

2.61

180

300

-

-

53

In order to relate the performances of the DF-ZnO NFs gas sensors presented in this work, their characteristics are compared to reported ZnO nanomaterial-based NO2 gas sensors from the literature in Table 1. Other comparisons with NO2 sensors using ZnO film or other metal-oxide nanostructures are shown in Table S1. Numerous reported sensor devices showed sensitive

ACS Paragon Plus Environment

22

Page 23 of 36

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 Materials & Interfaces

properties at very low concentrations, but had long characteristic times and did not exhibit saturation behavior. Some reports showed considerable failure in recovery. Moreover, although sensors fabricated on flexible substrates have been reported, the stability of the sensors after repetitive deformation has not been evaluated. The comparative assessment reveals that the DFZnO NFs gas sensors possess superior sensing characteristics towards NO2 and sustainable mechanical stability by maintaining sensing properties even after 10000 cycles of bending test with a curvature radius of 5 mm. Therefore, the combination of NFs network structure with maximizing surface area, donor point defects, numerous active points, and the use of SWCNTs as electrode in the DF-ZnO NFs sensors provide excellent sensing characteristics, stable mechanical properties, and enables high high integration of nanodevices.

CONCLUSIONS Micro-patterned DF-ZnO NFs flexible gas sensors with maximized surface area were successfully fabricated on PI substrates by growth of ZnO NRs on ZnO shells. As a sacrificial layer to form the ZnO shells, polystyrene MPs were arranged at regular intervals by PDMS rubbing technique and the arrangement was maintained after transfer and calcination. This island configuration at regular intervals allows to improve the mechanical flexibility by dispersing the strain into the empty spaces of the substrate. The maximized surface area of the DF-ZnO NFs network structure was obtained by growth of NRs on ZnO shells. The operating temperature of the sensor was optimized at 270 °C considering the stability of the SWCNTs electrode and the PI substrate. The DF-ZnO NFs gas sensors under NO2 exposure showed perfect saturation behavior and outstanding sensing performances. The sensors showed a response of 218.1, rising time of 25.0 sec, decay time of 14.1 sec, and percent recovery greater than 98% upon NO2 exposure.

ACS Paragon Plus Environment

23

ACS Applied Materials & Interfaces

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 24 of 36

Interestingly, the sensors showed higher selectivity towards NO2 than NH3 and CO. The comparative analyses revealed that the high response and selectivity of the sensor were due to combination of the high surface area and numerous active points from the NRs, as well as the ZnO point defects of donor state, and the SWCNTs electrode. Furthermore, the DF-ZnO NFs gas sensors demonstrated sustainable mechanical stability. Although some physical degradation of the device was observed under SEM after the repeated flexible test, the sensor functioned even after 10000 cycles with a curvature radius of 5 mm. These results are important for the development of nanomaterial-based sensor devices as we successfully improved the performances in NO2 sensing by using a 3D micro-patterned network structure. This structure with maximized surface area not only greatly improves the response and characteristics times, but also enhances the selectivity and recovery in a metal oxide-based gas sensor. Moreover, the micro-patterning technique on a flexible substrate enables fabricating highly integrated nanodevices and application to flexible devices. Finally, it provides stable mechanical sustainability in flexible electronic devices based on nanomaterial semiconductors.

EXPERIMENTAL SECTION Micro-pattern arrays of polystyrene MPs. First, the glass substrates were cleaned by sonication in acetone, methanol, and deionized (DI) water consecutively for 15 min each. Then, circular patterns with a diameter of 5 μm and a height of 3 μm were defined on the substrates by photolithography. Next, the MPs with a diameter of 5 μm purchased by Sigma-Aldrich were placed on the patterned substrates by using PDMS-transfer. After transfer, the MPs were aligned by being pushed into circular patterns by PDMS rubbing technique. The PDMS rubbing technique was performed unidirectionally.

ACS Paragon Plus Environment

24

Page 25 of 36

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 Materials & Interfaces

Construction of micro-patterned ZnO shells on flexible substrate. The ZnO film with a thickness of 100 nm was deposited on the micro-patterned MPs by using radio-frequency magnetron sputtering with a plasma power of 150 W and a pressure of 15 mtorr in Ar environment. As a flexible receiver substrate, a PVP-coated PI substrate was prepared by spin coating of a 10% PVP solution with a cross-linking agent, poly(melamine-co-formaldehyde) in propylene glycol monomethyl ether acetate, and was softly baked at 80 °C for 1 min. Then, the ZnO-coated MPs were transferred on the PVP-coated PI substrate by applying a pressure of 7.5 g/cm2, and then heating at 120 °C for 1 min. After that, in order to finalize the micro-patterned ZnO shells, the transferred MPs were calcined at 250 °C for 3 h on a hot plate. Fabrication of micro-patterned ZnO NFs flexible gas sensor. The ZnO NRs were grown on the micro-patterned ZnO shells by using an aqueous solution process. The ZnO shell arrays on the flexible substrate were immersed in a chemical bath containing an equimolar solution of 10 mM

of

zinc

acetate

dihydrate

(Zn(CH3COO)2·2H2O,

ACS

reagent

96459)

and

hexamethylenetetramine (C6H12N4, ACS reagent 398160). The reaction was carried out at 95 °C for 3 h. After the reaction was complete, SWCNTs were deposited using spray coater and maintained at 120 °C in air to remove the solvent. In order to get rid of residual surfactant and dispersant, SWCNTs electrodes were rinsed with DI water for 5 min. Characterization and Measurement. The surface morphologies of all materials used in this work were characterized by using SEM (Hitachi S-5000). The structural properties of the materials were investigated by using XRD (Rigaku SmartLab). The work function of SWCNTs electrode was analyzed by a surface-analysis photoelectron spectrometer (AC-2, Riken Keiki Co. Ltd). To investigate the optical properties of the ZnO, room-temperature PL spectroscopy measurements were performed using IK3252R-E He-Cd laser (325 nm) source coupled with

ACS Paragon Plus Environment

25

ACS Applied Materials & Interfaces

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 26 of 36

MonoRa 320i monochromator (Dongwoo Optron) and Andor SOLIS simulation package. The gas sensing performances were monitored in a sealed sensing chamber coupled with an electrical feed-through and gas inlet and outlet. The target gas was diluted with nitrogen gas (concentration: 50-500 ppm) and dry air was used as the purging gas. A more detailed description of the gas sensor measurement can be found in Figure S8.

ACS Paragon Plus Environment

26

Page 27 of 36

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 Materials & Interfaces

ASSOCIATED CONTENT Supporting Information The Supporting Information is available free of charge on the ACS Publications website (http://pubs.acs.org) at DOI: SEM images of DF-ZnO NFs formed by the growth of ZnO NRs on both sides of the ZnO shells, XRD and PL peaks of ZnO film (before calcination), shells (after calcination), and NRs, OM image after placing the MPs in circular well patterns, Response at different operating temperatures, Work function measurement result of SWCNTs, SEM images after dynamic bending test of 10000 cycles with a curvature radius of 5 mm, Schematic illustration of gas sensing system, Table by comparing the ZnO film to other metal-oxide nanostructures sensor for detecting NO2 gases.

AUTHOR INFORMATION Corresponding Author *E-mail: [email protected] Notes The authors declare no competing financial interest.

ACKNOWLEDGMENT

ACS Paragon Plus Environment

27

ACS Applied Materials & Interfaces

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 28 of 36

This work was supported by Samsung Research Funding Center of Samsung Electronics under Project Number SRFC-MA1301-07.

ACS Paragon Plus Environment

28

Page 29 of 36

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 Materials & Interfaces

REFERENCES (1)

Ryabtsev, S.; Shaposhnick, A.; Lukin, A.; Domashevskaya, E. Application of Semiconductor Gas Sensors for Medical Diagnostics. Sens. Actuators B Chem. 1999, 59, 26–29.

(2)

Li, H.-Y.; Huang, L.; Wang, X.-X.; Lee, C.-S.; Yoon, J.-W.; Zhou, J.; Guo, X.; Lee, J.-H. Molybdenum Trioxide Nanopaper as a Dual Gas Sensor for Detecting Trimethylamine and Hydrogen Sulfide. RSC Adv. 2017, 7 (7), 3680–3685.

(3)

Barbosa, M. S.; Suman, P. H.; Kim, J. J.; Tuller, H. L.; Varela, J. A.; Orlandi, M. O. Gas Sensor Properties of Ag- and Pd-Decorated SnO Micro-Disks to NO2, H2 and CO: Catalyst Enhanced Sensor Response and Selectivity. Sens. Actuators B Chem. 2017, 239, 253–261.

(4)

Anenberg, S. C.; Miller, J.; Minjares, R.; Du, L.; Henze, D. K.; Lacey, F.; Malley, C. S.; Emberson, L.; Franco, V.; Klimont, Z.; Heyes, C. Impacts and Mitigation of Excess Diesel-related NOx Emissions in 11 Major Vehicle Markets. Nature 2017, 545, 467–471.

(5)

Kleinerman, J.; Rynbrandt, D. Lung Proteolytic Activity and Serum Protease Inhibition after NO2 Exposure. Arch. Environ. Health 2013, 31, 37–41.

(6)

Wang, C.; Yin, L.; Zhang, L.; Xiang, D.; Gao, R. Metal Oxide Gas Sensors: Sensitivity and Influencing Factors. Sensors 2010, 10 (3), 2088–2106.

(7)

Lu, Y.; Ma, Y.; Ma, S.; Yan, S. Hierarchical Heterostructure of Porous NiO Nanosheets on Flower-like ZnO Assembled by Hexagonal Nanorods for High-Performance Gas Sensor. Ceram. Int. 2017, 43 (10), 7508–7515.

ACS Paragon Plus Environment

29

ACS Applied Materials & Interfaces

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

(8)

Page 30 of 36

Baek, S. D.; Biswas, P.; Kim, J. W.; Kim, Y. C.; Lee, T. Il; Myoung, J. M. LowTemperature Facile Synthesis of Sb-Doped P-Type ZnO Nanodisks and Its Application in Homojunction Light-Emitting Diode. ACS Appl. Mater. Interfaces 2016, 8 (20), 13018– 13026.

(9)

Zhang, J.; Yang, Y.; Xu, B.; Jiang, F.; Li, J. Shape-Controlled Synthesis of ZnO Nanoand Micro-Structures. J. Cryst. Growth 2005, 280 (3), 509–515.

(10)

Das, S. N.; Kar, J. P.; Choi, J.-H.; Lee, T. Il; Moon, K.-J.; Myoung, J.-M. Fabrication and Characterization of ZnO Single Nanowire-Based Hydrogen Sensor. J. Phys. Chem. C 2010, 114 (3), 1689–1693.

(11)

Hjiri, M.; El Mir, L.; Leonardi, S.; Donato, N.; Neri, G. CO and NO2 Selective Monitoring by ZnO-Based Sensors. Nanomaterials 2013, 3 (3), 357–369.

(12)

Wang, J. X.; Sun, X. W.; Yang, Y.; Huang, H.; Lee, Y. C.; Tan, O. K.; Vayssieres, L. Hydrothermally Grown Oriented ZnO Nanorod Arrays for Gas Sensing Applications. Nanotechnology 2006, 17 (19), 4995–4998.

(13)

Hsiao, C. C.; Luo, L. S. A Rapid Process for Fabricating Gas Sensors. Sensors 2014, 14 (7), 12219–12232.

(14)

Postica, V.; Gröttrup, J.; Adelung, R.; Lupan, O.; Mishra, A. K.; Leeuw, N. H. De; Ababii, N.; Carreira, J. F. C.; Rodrigues, J.; Sedrine, N. Ben; Correia, M. R.; Monteiro, T.; Sontea, V.; Mishra, Y. K. Multifunctional Materials : A Case Study of the Effects of Metal Doping on ZnO Tetrapods with Bismuth and Tin Oxides. Adv. Funct. Mater. 2017, 27, 1604676.

ACS Paragon Plus Environment

30

Page 31 of 36

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 Materials & Interfaces

(15)

Zhang, J.; Liu, X.; Wu, S.; Cao, B.; Zheng, S. One-Pot Synthesis of Au-Supported ZnO Nanoplates with Enhanced Gas Sensor Performance. Sens. Actuators B Chem. 2012, 169, 61–66.

(16)

Mishra, Y. K.; Modi, G.; Cretu, V.; Postica, V.; Lupan, O.; Reimer, T.; Paulowicz, I.; Hrkac, V.; Benecke, W.; Kienle, L.; Adelung, R. Direct Growth of Freestanding ZnO Tetrapod Networks for Multifunctional Applications in Photocatalysis, UV Photodetection, and Gas Sensing. ACS Appl. Mater. Interfaces 2015, 7 (26), 14303–14316.

(17)

Pawar, R. C.; Lee, J. W.; Patil, V. B.; Lee, C. S. Synthesis of Multi-Dimensional ZnO Nanostructures in Aqueous Medium for the Application of Gas Sensor. Sens. Actuators B Chem. 2013, 187, 323–330.

(18)

Paulowicz, I.; Hrkac, V.; Kaps, S.; Cretu, V.; Lupan, O.; Braniste, T.; Duppel, V.; Tiginyanu, I.; Kienle, L.; Adelung, R. Three-Dimensional SnO2 Nanowire Networks for Multifunctional Applications : From High-Temperature Stretchable Ceramics to Ultraresponsive Sensors. Adv. Electron. Mater. 2015, 1, 1500081.

(19)

Kaps, S.; Bhowmick, S.; Gro, J.; Hrkac, V.; Stau, D.; Guo, H.; Warren, O. L.; Adam, J.; Kienle, L.; Minor, A. M.; Adelung, R.; Mishra, Y. K. Piezoresistive Response of QuasiOne-Dimensional ZnO Nanowires Using an in Situ Electromechanical Device. ACS Omega 2017, 2 (6), 2985–2993.

(20)

Park, H.; Afzali, A.; Han, S.-J.; Tulevski, G. S.; Franklin, A. D.; Tersoff, J.; Hannon, J. B.; Haensch, W. High-Density Integration of Carbon Nanotubes via Chemical Self-Assembly. Nat. Nanotechnol. 2012, 7 (12), 787–791.

ACS Paragon Plus Environment

31

ACS Applied Materials & Interfaces

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

(21)

Page 32 of 36

Koh, K.; Hwang, H.; Park, C.; Lee, J. Y.; Jeon, T. Y.; Kim, S. H.; Kim, J. K.; Jeong, U. Large-Area Accurate Position Registry of Microparticles on Flexible, Stretchable Substrates Using Elastomer Templates. ACS Appl. Mater. Interfaces 2016, 8 (41), 28149– 28158.

(22)

Kim, D.-H.; Song, J.; Choi, W. M.; Kim, H.-S.; Kim, R.-H.; Liu, Z.; Huang, Y. Y.; Hwang, K.-C.; Zhang, Y.; Rogers, J. A. Materials and Noncoplanar Mesh Designs for Integrated Circuits with Linear Elastic Responses to Extreme Mechanical Deformations. Proc. Natl. Acad. Sci. U. S. A. 2008, 105 (48), 18675–18680.

(23)

Hsu, P. I.; Huang, M.; Xi, Z.; Wagner, S.; Suo, Z.; Sturm, J. C. Spherical Deformation of Compliant Substrates with Semiconductor Device Islands. J. Appl. Phys. 2004, 95 (2), 705–712.

(24)

Gu, D.; Baumgart, H.; Naumann, F.; Petzold, M. Finite Element Modeling and Raman Study of Strain Distribution in Patterned Device Islands on Strained Silicon-on-Insulator (sSOI) Substrates. ECS Trans. 2010, 33 (4), 529–535.

(25)

Xu, H.; Liu, X.; Cui, D.; Li, M.; Jiang, M. A Novel Method for Improving the Performance of ZnO Gas Sensors. Sens. Actuators B Chem. 2006, 114 (1), 301–307.

(26)

Chatterjee, A. P.; Mitra, P.; Mukhopadhyay, A. K. Chemically Deposited Zinc Oxide Thin Film Gas Sensor. J. Mater. Sci. 1999, 34 (17), 4225–4231.

(27)

An, W.; Wu, X.; Zeng, X. C. Adsorption of O2, H2, CO, NH3, and NO2 on ZnO Nanotube: A Density Functional Theory Study. J. Phys. Chem. C 2008, 112 (15), 5747–5755.

ACS Paragon Plus Environment

32

Page 33 of 36

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 Materials & Interfaces

(28)

Huang, J.; Wu, Y.; Gu, C.; Zhai, M.; Yu, K.; Yang, M.; Liu, J. Large-Scale Synthesis of Flowerlike ZnO Nanostructure by a Simple Chemical Solution Route and Its Gas-Sensing Property. Sens. Actuators B Chem. 2010, 146 (1), 206–212.

(29)

Ponzoni, A.; Comini, E.; Sberveglieri, G.; Zhou, J.; Deng, S. Z.; Xu, N. S.; Ding, Y.; Wang, Z. L. Ultrasensitive and Highly Selective Gas Sensors Using Three-Dimensional Tungsten Oxide Nanowire Networks. Appl. Phys. Lett. 2006, 88 (20), 203101.

(30)

Yi, J.; Lee, J. M.; Park, W. Il. Vertically Aligned ZnO Nanorods and Graphene Hybrid Architectures for High-Sensitive Flexible Gas Sensors. Sens. Actuators B Chem. 2011, 155 (1), 264–269.

(31)

Zhou, Z. G.; Tang, Z. L.; Zhang, Z. T. Studies on Grain-Boundary Chemistry of Perovskite Ceramics as CO Gas Sensors. Sens. Actuators B Chem. 2003, 93 (1), 356–361.

(32)

Feng, P.; Wan, Q.; Wang, T. H. Contact-Controlled Sensing Properties of Flowerlike ZnO Nanostructures. Appl. Phys. Lett. 2005, 87 (21), 213111.

(33)

Janotti, A.; Walle, C. G. V. D. Fundamentals of Zinc Oxide as a Semiconductor. Rep. Prog. Phys. 2009, 72, 126501.

(34)

Lenaerts, S.; Division, P.; Group, M.; Division, C. S. FT-IR Characterization of Tin Dioxide Gas Sensor Materials under Working Conditions. Spectroc. Acta Pt. A-Molec. Biomolec. Spectr. 1995, 51 (5), 883–894.

(35)

Noboru, Y.; Jun, F.; Masato, K.; Tetsuro, S. Interactions of Tin Oxide Surface with O2, H2O and H2. Surf. Sci. 1979, 86, 335–344.

ACS Paragon Plus Environment

33

ACS Applied Materials & Interfaces

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

(36)

Page 34 of 36

Kumar, R.; Al-Dossary, O.; Kumar, G.; Umar, A. Zinc Oxide Nanostructures for NO2 Gas sensor Applications: A Review. Nano-Micro Lett. 2015, 7 (2), 97–120.

(37)

Lupan, O.; Chow, L.; Pauporté, T.; Ono, L. K.; Roldan Cuenya, B.; Chai, G. Highly Sensitive and Selective Hydrogen Single-Nanowire Nanosensor. Sens. Actuators B Chem. 2012, 173, 772–780.

(38)

Chen, M.; Wang, Z.; Han, D.; Gu, F.; Guo, G. Porous ZnO Polygonal Nanoflakes: Synthesis, Use in High-Sensitivity NO2 Gas Sensor, and Proposed Mechanism of Gas Sensing. J. Phys. Chem. C 2011, 115 (26), 12763–12773.

(39)

Reimer, T.; Paulowicz, I.; Der, R. R.; Kaps, S.; Lupan, O.; Benecke, W.; Ronning, C.; Adelung, R.; Mishra, Y. K. Single Step Integration of ZnO Nano- and Microneedles in Si Trenches by Novel Flame Transport Approach : Whispering Gallery Modes and Photocatalytic Properties. ACS Appl. Mater. Interfaces 2014, 6 (10), 7806-7815.

(40)

Vempati, S.; Mitra, J.; Dawson, P. One-Step Synthesis of ZnO Nanosheets: A Blue-White Fluorophore. Nanoscale Res. Lett. 2012, 7 (1), 470.

(41)

Tang, Q.; Zhou, W.; Shen, J.; Zhang, W.; Kong, L.; Qian, Y. A Template-Free Aqueous Route to ZnO Nanorod Arrays with High Optical Property. Chem. Commun. 2004, 1 (6), 712–713.

(42)

Alvi, N. H.; Ul Hasan, K.; Nur, O.; Willander, M. The Origin of the Red Emission in NZnO Nanotubes/p-GaN White Light Emitting Diodes. Nanoscale Res. Lett. 2011, 6 (1), 130.

(43)

Peng, S.; Cho, K. Chemical Control of Nanotube. Nanotechnology 2000, 11, 57–60.

ACS Paragon Plus Environment

34

Page 35 of 36

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 Materials & Interfaces

(44)

Zhao, J.; Buldum, A.; Han, J.; Lu, J. P. Gas Molecule Adsorption in Carbon Nanotubes and Nanotube Bundles. Nanotechnology 2002, 13 (2), 195–200.

(45)

Chang, H.; Lee, J. Do; Lee, S. M.; Lee, Y. H. Adsorption of NH3 and NO2 Molecules on Carbon Nanotubes. Appl. Phys. Lett. 2001, 79 (23), 3863–3865.

(46)

Skucha, K.; Fan, Z.; Jeon, K.; Javey, A.; Boser, B. Palladium/silicon Nanowire Schottky Barrier-Based Hydrogen Sensors. Sensors Actuators, B Chem. 2010, 145 (1), 232–238.

(47) Freeouf, J.; Woodall, J. Schottky Barriers: An Effective Work Function Model. Appl. Phys. Lett. 1981, 10598, 727–729. (48)

Kwon, D.-K.; Lee, S. J.; Myoung, J.-M. High-Performance Flexible ZnO Nanorod UV Photodetectors with a Network-Structured Cu Nanowire Electrode. Nanoscale 2016, 8 (37), 16677–16683.

(49) Alivov, Y. I.; Kalinina, E. V.; Cherenkov, A. E.; Look, D. C.; Ataev, B. M.; Omaev, A. K.; Chukichev, M. V.; Bagnall, D. M. Fabrication and Characterization of N-ZnO/p-AlGaN Heterojunction Light-Emitting Diodes on 6H-SiC Substrates. Appl. Phys. Lett. 2003, 83 (23), 4719–4721. (50) Oh, E.; Choi, H.; Jung, S.; Cho, S.; Chang, J.; Lee, K.; Kang, S.; Kim, J.; Yun, J.; Jeong, S. Sensors and Actuators B : Chemical High-Performance NO2 Gas Sensor Based on ZnO Nanorod Grown by Ultrasonic Irradiation. Sens. Actuators B Chem. 2009, 141 (1), 239– 243. (51)

Chougule, M. A.; Sen, S.; Patil, V. B. Fabrication of Nanostructured ZnO Thin Film Sensor for NO2 Monitoring. Ceram. Int. 2012, 38 (4), 2685–2692.

ACS Paragon Plus Environment

35

ACS Applied Materials & Interfaces

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

(52)

Page 36 of 36

Chem, J. M.; Bai, S.; Hu, J.; Li, D.; Luo, R.; Chiun, C. Quantum-Sized ZnO Nanoparticles : Synthesis , Characterization and Sensing. J. Mater. Chem. 2011, 21, 12288–12294.

(53)

Marc, D.; Xin, G.; Chao, Z.; Yifan, L. Flexible NO 2 Gas Sensors Based on Sheet-like Hierarchical ZnO 1 − X Coatings Deposited on Polypropylene Papers by Suspension Flame Spraying. J. Taiwan Inst. Chem. Eng. 2017, 75 (2), 280–286.

Table of Contents

ACS Paragon Plus Environment

36