This article appeared in a journal published by Elsevier. The attached copy is furnished to the author for internal non-commercial research and education use, including for instruction at the authors institution and sharing with colleagues. Other uses, including reproduction and distribution, or selling or licensing copies, or posting to personal, institutional or third party websites are prohibited. In most cases authors are permitted to post their version of the article (e.g. in Word or Tex form) to their personal website or institutional repository. Authors requiring further information regarding Elsevier’s archiving and manuscript policies are encouraged to visit: http://www.elsevier.com/authorsrights Author's personal copy Journal of Hazardous Materials 265 (2014) 124–132 Contents lists available at ScienceDirect Journal of Hazardous Materials journal homepage: www.elsevier.com/locate/jhazmat Effective decoration of Pd nanoparticles on the surface of SnO2 nanowires for enhancement of CO gas-sensing performance Do Dang Trung a , Nguyen Duc Hoa a,∗ , Pham Van Tong a , Nguyen Van Duy a , T.D. Dao b , H.V. Chung b , T. Nagao b , Nguyen Van Hieu a,∗ a b International Training Institute for Materials Science, Hanoi University of Science and Technology, No. 1 Dai Co Viet Road, Hanoi, Viet Nam National Institute for Materials Science, 1-1 Namiki, Tsukuba 305-0044, Japan h i g h l i g h t s g r a p h i c a l a b s t r a c t • Pd nanoparticles decoration on nanowire’s surface was done via a simple route. • The size and distribution of Pd nanoparticles was controlled by in situ reduction of the Pd complex. • The coating Pd enhanced capable detection of hazardous CO gas of onchip growth SnO2 nanowires. • Pd-decorated SnO2 nanowires sensors have good stability to hazardous CO gas. a r t i c l e i n f o Article history: Received 20 August 2013 Received in revised form 18 October 2013 Accepted 24 November 2013 Available online 2 December 2013 Keywords: Pd NPs decoration SnO2 nanowire CO gas sensor a b s t r a c t Decoration of noble metal nanoparticles (NPs) on the surface of semiconducting metal oxide nanowires (NWs) to enhance material characteristics, functionalization, and sensing abilities has attracted increasing interests from researchers worldwide. In this study, we introduce an effective method for the decoration of Pd NPs on the surface of SnO2 NWs to enhance CO gas-sensing performance. Singlecrystal SnO2 NWs were fabricated by chemical vapor deposition, whereas Pd NPs were decorated on the surface of SnO2 NWs by in situ reduction of the Pd complex at room temperature without using any linker or reduction agent excepting the copolymer P123. The materials were characterized by advanced techniques, such as high-resolution transmission electron microscopy, scanning transmission electron microscopy, and energy-dispersive X-ray spectroscopy. The Pd NPs were effectively decorated on the surface of SnO2 NWs. As an example, the CO sensing characteristics of SnO2 NWs decorated with Pd NPs were investigated at different temperatures. Results revealed that the gas sensor exhibited excellent sensing performance to CO at low concentration (1–25 ppm) with ultrafast response-recovery time (in seconds), high responsivity, good stability, and reproducibility. © 2013 Elsevier B.V. All rights reserved. 1. Introduction Carbon monoxide (CO) is a colorless, odorless, and tasteless but highly toxic gas. The toxicity of CO gas is mainly caused by the inhalation. Breathing CO gas can result in health problem because ∗ Corresponding authors. Tel.: +84 4 38680787; fax: +84 4 38692963. E-mail addresses: ndhoa@itims.edu.vn (N.D. Hoa), hieu@itims.edu.vn (N.V. Hieu). 0304-3894/$ – see front matter © 2013 Elsevier B.V. All rights reserved. http://dx.doi.org/10.1016/j.jhazmat.2013.11.054 it displaces oxygen in the blood and deprives the heart, brain, and other vital organs of oxygen. In blood, CO is believed to combine with hemoglobin to produce carboxyhemoglobin, which usurps the space in hemoglobin leading to ineffective delivering oxygen to bodily tissues [1]. Carbon monoxide is a byproduct of an incomplete combustion process of organic substances, such as petroleum and fuel, in internal combustion engines caused by the insufficient supply of oxygen. Exposure to a low concentration of approximately 35 ppm can cause headache and dizziness, whereas exposure to Author's personal copy D.D. Trung et al. / Journal of Hazardous Materials 265 (2014) 124–132 a higher concentration (∼0.64%) can result in convulsion, respiratory arrest, and even death. Thus, Occupational Safety and Health Administration limits CO concentration for long-term exposure to less than 50 ppm. Early monitoring or alarm of CO at low concentration is very important not only in the environment protection but also for human health safety. The current trend for CO sensor development is to increase sensitivity and reduce response time, as well as lowering the detection limit. Nanostructured metal oxide semiconductors (MOS) are fundamental in studies and practical applications of photocatalysts [2], batteries [3], biosensors [4], and gas sensors because of their unique optical, electrical, and functional properties [5,6]. Many attempts have been devoted for the design, synthesis, processing, characterization, and device applications of MOS at the nanoscale. Extensive studies have been concentrated on the development of new materials with novel nanostructures and excellent properties to improve the performance of gas sensors [7–11]. Different types of gas sensors have been developed including of optical sensors, electrochemical sensors, calorimetric sensors, and resistive gas sensors. The resistive gas sensors which operated based on the variation in their electrical resistance upon chemical reaction and/or adsorption between the analytical gas molecules and the surface of sensing layers, appear to be advantages not only because of their low cost, simple completion, and good reliability for realtime control systems, but also the diverse practical applications in environmental monitoring, transportation, security, defense, space missions, energy, agriculture, medicine, etc. [7]. Among the materials used in resistive gas sensors, MOS such as SnO2 , ZnO, WO3 , CuO, Co3 O4 , and NiO have attracted great attention because of their reasonable sensitivities to various gases, such as NO2 , NH3 , CO, H2 , and C2 H5 OH [8–10]. Comparing with the thin film or the bulk forms of metal oxide, the nanowire structure was reported to exhibit significant advantages such as the small diameter compatible with Debye length and the anisotropic quantum confinement effect, those promise excellent sensitivity as well as cost effective for devices fabrication [7,8]. Generally, resistive type of MOS-based gas sensors cannot sufficiently identify an unknown gas or mixture of gases because of their poor selectivity, but they are designed for monitoring some reducing gases and usually respond to many compounds [10,11]. High detection limit also reduces the number of potential applications of conventional gas sensors. One of the key techniques to improve the functionalities, sensitivity, and selectivity of MOS is decoration or doping of noble metal nanoparticles (NPs) on their surface [12–14]. Different metal NPs such as Pt, Au, Ag, and Pd have been used in the decoration or functionalization on the surface of MOS to enhance sensor performance [15]. Each metal exhibits significantly enhanced response to a specific gas or to some gases [16,17]. For instance, decoration of Pt NPs on the surface of bead-like SnO2 nanowires (NWs) significantly enhances the H2 sensing performance of the NWs but reduces the response to NO2 [18]. Pt and Pd NPs were decorated on the surface of In2 O3 flowerlike nanobundles to enhance their sensitivities and responses to CO gas at room temperature [19]. The decoration of Pd nanodots on the surface of networked ZnO nanowires was also reported to exhibit significant high response to CO gas at room temperature [20]. The room temperature CO sensing was also demonstrated on the Ptloaded SnO2 porous nanosolid [21]. Pd NPs and NiO were decorated on the surface of SnO2 NWs to significantly enhance their NO2 and H2 S gas-sensing responses [22,23]. Highly responsive H2 gas sensor based on the metal–insulator transition mechanism was also observed in V2 O5 NWs decorated with Pd NPs [24]. Decoration or functionalization of noble metal NPs on the surface of nanomaterials can be performed by a vacuum deposition [25], drop-casting [26], supercritical fluid deposition [27], thermal pyrolysis [28], or self-assembly [29]. Each decoration method has some advantages and disadvantages. The self-assembly method 125 is complicated because noble metal NPs are first prepared and stabilized in polymer before they can be decorated on the surface of nanostructured metal oxides [30]. Kolmakov et al. [31] demonstrated an in situ vacuum deposition of Pd on the surface of an individual SnO2 NW and investigated its gas-sensing characteristics. Moon et al. [32] reported a new route to enhance the NO2 sensitivity of nanocrystalline TiO2 fiber-based gas sensors by doping with Pd, in which the Pd-doped TiO2 fibers were fabricated by electrospinning and subsequent annealing. Yuasa et al. [33] prepared Pd-loaded SnO2 nanocrystals by photochemical deposition of Pd2+ onto SnO2 in an aqueous solution to achieve a highly sensitive hydrogen gas sensor. Kim et al. [34] prepared In2 O3 -Pd core–shell NWs by direct sputter deposition of Pd layer on bare In2 O3 NWs to enhance their NO2 sensing characteristics. They demonstrated that the improvement of sensing properties is caused by the enhanced adsorption or dissociation of NO2 and the associated spillover effects. One-step hydrothermal synthesis was also applied to fabricate Pd-ZnO nanoflowers, wherein Pd NPs and ZnO nanoflowers were simultaneously grown in a sealed Teflon-lined stainless steel autoclave [35]. However, this method has difficulties in decoration of noble metal NPs on the surface of different MOS. Zhang et al. reported the self-assemblies of Pd NPs on the surfaces of single crystal ZnO nanowires for enhanced H2 S gas performances [36], whereas Li et al. demonstrated an enhancement of H2 S gas-sensing properties by assembly of Pd NPs on the surface of SnO2 NWs [37]. Chang et al. [38] decorated ZnO nanorod arrays with Pd NPs by solution routes of aqueous chemical growth and successive photochemical reduction to improve CO gas-sensing characteristics. In their study, ZnO nanorods were prepared first by a hexamethylenetetramine-assisted hydrothermal method, and the decoration of Pd NPs was performed by ultraviolet irradiation reduction of PdCl2 in ethanol solution. Development of a general, low-temperature facile and effective method that does not require a vacuum, linker, or reduction agent for the decoration of Pd NPs on a wide range of MOS in gas-sensing applications is still a challenge and an interesting task. In this study, we introduce a vacuum-less method for the decoration of Pd NPs on the surface of SnO2 NWs at room temperature to enhance their CO gas-sensing performance. Decoration of Pd NPs on the surface of MOS was archived by direct reduction of Pd complex in the presence of pluronic surfactant. This method is a facile and effective route that enables the decoration of Pd NPs on the surface of different MOSs, such as SnO2 , ZnO, Co3 O4 , and NiO, for gassensing applications. The CO-sensing properties of the fabricated SnO2 NWs decorated with Pd NPs were investigated and found to exhibit significantly enhanced performance. The fabricated materials also have potential applications in catalytic and/or bio-sensing applications. 2. Experimental 2.1. On-chip growth of SnO2 NWs Single-crystal SnO2 NWs were prepared directly on the sensor chips by thermal evaporation. The sensor chips have two Au electrodes and heater that were screen-printed on the front and back sides of an Al2 O3 substrate (Fig. 1A). The sensor chips used for the in situ growth of SnO2 NWs were pre-deposited with a ∼5 nm thick gold layer on the front side. The SnO2 NW process was conducted as previously described [23,39]. In a typical synthesis, 0.3 g Sn powder was loaded on an alumina boat, which was then loaded in a quartz tube of a thermal evaporation system [23]. The sensor chips were placed on the same quartz tube ∼2 cm away from the boat. The temperature in the furnace was rapidly ramped up to 750 ◦ C for 15 min. During the process, the vacuum of the quartz Author's personal copy 126 D.D. Trung et al. / Journal of Hazardous Materials 265 (2014) 124–132 Fig. 1. (A) Al2 O3 substrate with screen-printing Au electrodes, (B) electrodes after growth of SnO2 NWs; (C) low and (D) high magnifications of FE-SEM images of the SnO2 NW grown on the substrate. tubes was maintained at ∼2 × 10−2 torr by a rotary vacuum pump. When the furnace reached the set-point temperature, 0.5 sccm oxygen gas was introduced through the quartz tube for 15 min. The furnace was cooled down to room temperature to collect the samples. spectroscopy (EDS) using an X-ray analyzer integrated with a TEM instrument. Crystal structures of materials were studied through a wide-angle powder X-ray diffraction (XRD) using Cu K␣ X-radiation with a wavelength of 0.154178 nm. 2.3. Gas-sensing characterizations 2.2. Decoration of Pd NPs on the surface of SnO2 NWs Decoration of the Pd NPs on the surface of nanostructured MOS was performed by in situ reduction of the Pd complex in the presence of pluronic (P123) surfactant at room temperature without using any further linker and reducing agent [40] and/or assistance of photochemical reduction [41]. Copolymer P123 acts as a surfactant and a reducing agent to reduce Pd complex and stabilize the NPs, thereby preventing aggregation and enhancing the active surface of materials. The Pd complex was first prepared by dissolving PdCl2 in aqueous NaCl solution to obtain a complex solution of 20 mg/mL (Na2 [PdCl4 ]). At least three SnO2 NW sensors were added to the complex solution with further stirring. After a few minutes, a solution of 2 g pluronic dissolved in 40 mL H2 O was added to the solution to reduce Pd2+ into Pd NPs. The reduction process was performed for 4 h in ambient condition. After the products were collected and washed with deionized water and ethanol, the samples were dried in an oven at 45 ◦ C overnight. Characterization of the synthesized materials was performed using several techniques, such as high-resolution transmission electron microscopy (HRTEM), scanning transmission electron microscopy (STEM), and selective area electron diffraction (SAED). Elemental analyses were performed by energy-dispersive X-ray The sensors were dried and then heat-treated at 500 ◦ C for at least 1 h to stabilize sensor resistance. The gas-sensing properties were then measured by a flow-through technique with a standard flow rate of 400 sccm for both dry air balance and analytic gases using a home-made system. This sensing system enabled rapid switching (in a second) from dried air to analytic gas to investigate the response-recovery time of the sensors. Prior to measurements, the dry air was flown through the sensing chamber until the stability of sensor resistance was reached. During sensing measurement, resistance of the sensors was continuously measured using a Keithley (2700) instrument interfaced with a computer while the dried air and analytic gases were switched on/off in each cycle. In our experiment, we used the standard gas concentration of 1000 ppm CO balanced in nitrogen and mixed with dry air as carrier using a series of mass flow controllers to obtain a lower concentration. Gas concentration was calculated as follows: C(ppm) = Cstd (ppm) × f/(f + F), where f and F are the flow rates of analytic gas and dry air, respectively, and Cstd (ppm) is the concentration of the standard gas used in the experiment. Sensor response was defined as S = Rair /Rgas , where Rair and Rgas are sensor resistances in dry air and analytic CO gas, respectively [10]. Author's personal copy D.D. Trung et al. / Journal of Hazardous Materials 265 (2014) 124–132 3. Results and discussion 3.1. Material characterizations Optical images of sensor chips before and after the growth of SnO2 NWs are shown in Fig. 1A and B, respectively. The gap between two electrodes was about 400 m, and they were stable after the growth of SnO2 NWs. White wool-like products were found on the sensor chip substrates as a result of the catalytic growth of the nanowires. The SnO2 nanowires grew on the electrodes and on the substrate (gap between two electrodes) to form a mat-like nanowires. These entangled mat-like nanowires will act as conducting paths for electrons to flow in the sensing characterization 127 (Fig. S1, ESI). The morphology of SnO2 NWs grown on the chips was characterized by FE-SEM (Fig. 1C and D). As shown in the figures, a thick layer of SnO2 NWs was successfully deposited on the chips. The as-grown SnO2 NWs have an average diameter of about 50–150 nm, and a length between 5 and 50 m. The NW surface was very smooth because of the VLS growth mechanism for the crystalline SnO2 NWs with gold as the catalyst in NW growth [39]. Morphological characterization of Pd NPs decorated on the surface of SnO2 NWs was investigated using HRTEM, STEM, and EDS. Fig. 2 shows the HRTEM images of the undecorated (A and B) and Pd NP-decorated (C and D) SnO2 NWs. The undecorated SnO2 NWs had a smooth surface with a uniform diameter along their axis despite the variation in size of each NW from ∼40 nm to 80 nm (Fig. 2A). Fig. 2. HRTEM images of pristine SnO2 NWs (A and B) and Pd NCs decorated on the surface of SnO2 NWs (C and D); Pd NC (E and F). Author's personal copy 128 D.D. Trung et al. / Journal of Hazardous Materials 265 (2014) 124–132 The HRTEM image and respective SAED pattern demonstrated that the SnO2 NW was a single crystal with tetragonal rutile structure (Fig. 2B) [42]. After decoration with Pd NPs, the shape of the SnO2 NW was not damaged, indicating that the SnO2 material is stable in the synthesis environment. Small Pd NPs were homogenously decorated on the surface of the SnO2 NW (Fig. 2C). In the photoreduction method for the decoration of Pd NPs on the surface of MOS nanorods, the Pd NPs agglomerated on the surface of the NWs but were not homogenously decorated on the surface of materials. This method is ineffective for gas sensor application because the agglomeration of Pd NPs might decrease the height of Schottky barrier formed at the interface between Pd NPs and NWs, thereby reducing sensor performance [33,38]. In our study, neither the surface functionalization nor the linker reagent was applied for the decoration of MOS [43] with Pd NPs excepting the P123 surfactant, but the Pd NPs were homogenously and strongly attracted on the surface of SnO2 NWs. Strong binding between Pd NPs and the surface of SnO2 NWs generated the barrier height at the interface, thus forming the conduction depletion region that contributed to the sensitivity of materials. In the vacuum deposition of Pd on the surface of SnO2 NWs, controlling the deposition rate and time is important to prevent the formation of a continuous layer of noble metal, just the core–shell structure, which lowers the sensing performance of the device [18,31]. As shown in the HRTEM image (Fig. 2C), the density of Pd NPs dispersed on the surface of SnO2 NW was approximately 0.22 NPs/m2 . The size, morphology and dispersion of Pd NPs on the surface of SnO2 nanowires could be controlled by varying the concentration of palladium complex and the synthesis time, and such characteristics may influent on the gas-sensing performances, but they are out of the context of this study. Pd NPs have a nearly spherical shape with an average diameter of approximately 6 nm. A high-magnification HRTEM image (Fig. 2E) demonstrated that Pd NP is highly crystalline and that the gap between the lattice fringes was approximately 0.22 nm, which belongs to the space of the (1 1 1) planes (d(1 1 1) = 0.2246 nm). SAED analysis (Fig. 2F) indicated that Pd NP is a single crystal, in which the diffraction pattern was indexed to the (1 1 1) reflection of a cubic structure of Pd (JCPDS card No. 05-0681) [44]. Some stacking faults were also observed in the HRTEM images of SnO2 NW and Pd NP. Further characterization of Pd NPs decorated on the surface of SnO2 NWs by STEM image and EDS mapping is shown in Fig. 3. STEM image also confirmed that Pd NPs were homogenously decorated on the surface of the SnO2 NWs (Fig. 3A). Elemental analysis by EDS mapping revealed the existence of Sn, O, and Pd, in which Sn and O were originally from the SnO2 NWs, whereas Pd was originally from the Pd NPs. O is expected to be higher than Sn in the EDS mapping, but the results indicated otherwise (Fig. 3B and C). This instance has not been clarified. Thus, the use of EDS mapping to calculate the composition of material was omitted in this study. The distribution of Pd NPs on the surface of SnO2 NWs was homogenously, and no aggregation of Pd NPs was observed. Crystal structures of the SnO2 NWs decorated with Pd NPs and the bare Pd NPs investigated by XRD are shown in Fig. 4. The XRD pattern of the Pd-SnO2 NWs exhibited strong diffraction peaks that were perfectly indexed to the tetragonal rutile structure of SnO2 , which is consistent with the reported data from the JCPDS card (77-0450) [45]. The standard profiles of cubic Pd (JCPDS, 05-0681) and PdO (JCPDS, 41-1107) are also potted in Fig. 4. However, none diffraction peak of the standard cubic Pd is observed marching with the XRD pattern of Pd-SnO2 NWs, whereas the strongest peak (1 0 1) of PdO is overlapped with the (2 0 0) of SnO2 thus it is difficult to index or confirm the existence of Pd related phases in this XRD pattern. This result can be attributed to the small size and low content of Pd NPs in the sample, resulting in a much lower intensity at the diffraction peaks compared with those of single crystal SnO2 NWs. To confirm the existence and successful synthesis of Pd NPs, an XRD pattern of the pure Pd NPs synthesized under the same condition as the Pd decoration process but without the addition of SnO2 NWs is shown in Fig. S2 (ESI). The XRD pattern indicated the face-centered cubic crystal structure of Pd NPs, with a space group of Fm3m (space Fig. 3. STEM image (A) and EDS mapping (C and D) of Pd-decorated SnO2 NWs. Author's personal copy 129 Pd-SnO 2 NWs JCPDS, No. 77-0450 40 30 (310) (112) (301) 60 70 2θ (deg.) 40 50 60 (220) (103) (200) (112) (200) Pd (JCPDS 05-0681) PdO (JCPDS 41-1107) (110) (002) Intensity 20 50 (111) 30 (101) 20 (002) (220) (110) (111) (211) (200) Intensity (a.u.) (101) D.D. Trung et al. / Journal of Hazardous Materials 265 (2014) 124–132 70 Fig. 4. XRD patterns of (A) SnO2 NWs decorated with Pd NPs, and standard profiles of Pd and PdO. group number: 225) and cell parameters a = b = c ≈ 0.388 nm. These data are consistent with the standard JCPDS card No. 05-0681. Moreover, the intensity of the (1 1 1) diffraction peak was higher than that of the (2 0 0) diffraction peak. This result suggests that the Pd NPs are highly crystalline and mainly bound by {1 1 1} facets [46]. These features are advantageous for the CO-sensing application because the Pd NPs with preferred {1 1 1} crystal facets exhibit significantly better catalytic activity for CO catalytic oxidation than the {1 0 0} facets [47]. 3.2. CO gas-sensing performance In our study, the CO gas-sensing characteristics of the synthesized Pd NP-decorated SnO2 NWs were investigated and compared with those of pristine SnO2 NWs (Fig. 5(A)). The transient resistance response to CO gas of the pristine SnO2 NWs and the Pd-decorated SnO2 NWs demonstrated that both sensors can detect CO gas. Both sensors exhibited decreased resistance upon exposure to CO gas indicating that the Pd-SnO2 NWs possess typical n-type gas-sensing properties similar to pristine SnO2 NWs [48]. However, decoration of Pd NPs on the surface of SnO2 NWs strongly enhanced the CO gassensing performance of SnO2 NW sensors. The response S (Rair /Rgas ) to 400 ppm CO at 400 ◦ C was 7 for the Pd NP-decorated SnO2 NWs and 3.9 for the pristine SnO2 NWs. Pd NP decoration also accelerates the adsorption and desorption rates of analytic gas molecules, thereby resulting in ultrafast response-recovery time of the sensors. The 90% response and recovery time of the pristine SnO2 NW sensors at 400 ◦ C were 10 and 60 s, respectively, whereas those of the Pd NP-decorated SnO2 NW sensor were approximately 5 and 40 s, respectively. The improvement of response-recovery speed of the Pd-SnO2 NW sensor was possibly due to the acceleration of interaction rate between CO molecules and pre-adsorbed oxygen (O2 − , O− , or O2− ) under the catalytic activity of Pd NPs. The enhanced responsivity of the Pd-SnO2 NWs compared with that of the pristine SnO2 NWs can be explained based on the spillover and the formation of Schottky barrier between Pd NPs and SnO2 Fig. 5. Comparison of the CO-sensing characteristics of the pristine SnO2 NW and Pd-decorated SnO2 NW sensors. NWs, which is similar to the mechanism of the Pd-ZnO arrays sensor [40]. A cartoon illustrates the depletion region and the interaction between analytic gas molecules and pre-adsorbed oxygen on the surface of Pd-SnO2 NW is shown in Fig. 5(B). The catalytic activity of Pd NPs accelerates the dissociation of oxygen molecules and causes a spillover of the ion absorbed oxygen on the surface of SnO2 . More ion adsorbed oxygen on the surface of SnO2 provides more sensing sites and enhances the sensitivity. In addition, SnO2 is an ntype semiconductor with a work function of 4.5 eV, which is higher than that of Pd (∼5.5 eV). Therefore, when Pd NPs make contact with SnO2 , electrons flow from the SnO2 NWs to the Pd NPs and an electron-depletion region or the Shottky barrier forms between SnO2 and Pd. The interaction of CO molecules with pre-adsorbed oxygen releases electrons back to Pd-SnO2 NWs and significantly reduces the height of Schottky barrier, resulting in a decrease in sensor resistance [31]. For practical applications, we considered the response-recovery time, responsivity, stability, and reproducibility of the sensors. Thus, three sensors were fabricated under optimum synthesis conditions, and their CO gas-sensing characteristics were investigated and compared. The transient resistance response to CO (25–400 ppm) measured at different temperatures (350–450 ◦ C) are shown in Fig. 6A–C and in Fig. S3 (ESI). The three sensors showed similar trends of sensing behavior at all temperatures, in which the resistance of the sensors decreased rapidly upon exposure to CO. After refreshing the sensing chamber with dry air, the sensor resistance recovered to the initial values. This result indicates that the adsorption of CO molecules on the surface of Pd-decorated SnO2 NWs is reversible. The response and recovery times of the sensors calculated for the 90% of the variation measured at different temperatures ranged from 4 s to 6 s and from 20 s to 40 s, respectively. The response time of about 210 s was reported in Pd-ZnO nanowires for the CO gas at room temperature [20]. Higher working temperature required shorter recovery time in the measured ranges. The temperature dependence of sensor response to 100 ppm CO shown Author's personal copy D.D. Trung et al. / Journal of Hazardous Materials 265 (2014) 124–132 400ppm 400 (A) 140 120 Sensor1 @ CO 200ppm 200 100ppm 50ppm 100 25ppm 75 Resistance(kΩ) 50 o 25 @350 C 150 100 50 o @400 C (A) 100 80 60 10 ppm 25 ppm 40 200 150 o 20 o @450 C 0 100 200 300 400 500 600 700 800 0 5 400 425 450 (B) @ 100 ppm 4 Response, S (Rair/Rgas) 375 Sensor1 Sensor2 Sensor3 2 10 (C) @400oC 6 Sensor1 Sensor2 Sensor3 4 2 0 50 100 150 5 10 15 20 CO conc. (ppm) 25 o 120 3 8 1.6 SnO2-Pd, 25 ppm CO @ 400 C (B) 140 Resistance (kΩ) 350 2.0 Time (s) 160 o 2.4 100 200 300 400 500 600 700 800 Time (sec.) Operating Temp. ( C) 1 ppm 2.8 SnO2-Pd, CO @ 400 C 100 50 2.5 ppm 5 ppm Response, S(Rair /Rgas) 600 Resistance (k Ω) Conc. (ppm) 130 200 250 300 350 400 CO conc. (ppm) Fig. 6. The CO response of the Pd-SnO2 NWs sensors: (A) the change in resistance of sensors upon exposure to different concentrations of CO measured at 350, 400, and 450 ◦ C; (B) temperature dependence and (C) concentration dependence of sensors response, respectively. in Fig. 6B demonstrated that the optimal working temperature is 400 ◦ C. Higher or lower than this working temperature, the sensor showed a decrease in responsivity. Based on the gas-sensing characteristics in Fig. 6A and in Fig. S3, the responsivities of three sensors were plotted as functions of CO concentration (Fig. 6C). All the sensors showed a nearly linear dependence of responsivity to CO concentration, and no significant difference in responsivity of the three sensors was found, indicating a good reproducibility of sensors. Development of gas sensors that can detect CO gas at low concentration is also important. In this study, the low-concentration CO-sensing characteristics of the Pd NP-decorated SnO2 NW sensor from 1 ppm to 25 ppm were also investigated. The transient resistance responses to CO gas measured at an optimal temperature of 400 ◦ C are shown in Fig. 7A. The sensor exhibited very good sensing characteristics even at a low concentration of CO (e.g., 1 ppm). The sensor also showed very fast response-recovery time from 10 s to 40 s. Summary of CO sensing performances of different sensors is presented in Table 1 (see ESI). Our developed PdSnO2 nanowire sensor exhibited a better CO sensing performance 100 80 60 40 7 cycles 20 0 200 400 600 800 1000 Time (s) Fig. 7. Low concentration CO sensing characteristics of the Pd-SnO2 NWs at 400 ◦ C: (A) the change in sensor resistance upon exposure to different concentrations of CO and (B) stability of the sensor after seven cycles of switching on and off, from dry air to CO gas and back to dry air. than ZnO nanorods, SnO2 single nanowire, Pd/SnO2 nanoparticles and NiO/SnO2 single nanowire, but comparable with In/Pd-doped SnO2 nanoparticles and Pd/ZnO nanowire junctions. For instance, according to literature, CO sensing based on SnO2 NW sensor has a detection limit of approximately 5–10 ppm. A single SnO2 NW sensor has a detection limit of 5 ppm at approximately 300 ◦ C with a response time of 100 s (AC measurement) [49]. Kuang et al. [50] reported that the detection limit of the pristine and NiOfunctionalized SnO2 NW-based sensors (FET type) to CO is 100 ppm at 250 ◦ C. Qian et al. [51] demonstrated that the CO detection limit of a sensor with a single SnO2 nanobelt decorated with Au NPs can reach approximately 5 ppm. Our results demonstrated that Pd NPs decorated on the surface of SnO2 NWs are not only promising in screening CO at high concentration but also effective in detecting at low ppm level (1–25 ppm), thereby satisfying the environmental monitoring requirement. In addition to high responsivity, stability and response-recovery time are two of the most important parameters of gas sensors for real-life applications. The stability of the Pd-SnO2 NW sensor was investigated for seven switching on/off cycles, from dry air to CO gas and back to dry air, as shown in Fig. 7B. Author's personal copy D.D. Trung et al. / Journal of Hazardous Materials 265 (2014) 124–132 131 Table 1 Summary of CO sensing performances of different sensors. Materials Methods SnO2 single nanowire SnO2 quantum dots ZnO nanorods NiO/SnO2 FET single nanowire Pt/SnO2 porous solid Pd/ZnO nanowire junctions Pd/SnO2 NPs In/Pd-doped SnO2 NPs Pd/SnO2 nanowires Gas (ppm) Spray pyrolysis Sono-chemical Hydrothermal Sputtering deposition of NiO Solvothermal hot press ␥-Ray radiolysis of Pd Infiltration process Sol–gel In situ reduction of Pd NPs CO, 4 ppm CO, 1000 ppm CO, 30 ppm CO, 500 ppm CO, 100 ppm CO, 0.1 ppm CO, 18 ppm CO, 1 ppm CO, 1 ppm The sensor exhibited stable signals after seven cycles, and no significant distortion in responsivity was observed. This result might be attributed to the high crystalline feature and high thermal stability of Pd-SnO2 NWs [31]. Selectivity is important in gas sensors. Thus, we present the response of Pd-decorated SnO2 NW sensors to various types of gases, including CO (1–25 ppm), H2 (10–50 ppm), NH3 (10–50 ppm), and CO2 (1000–15,000 ppm) at 400 ◦ C. The response as a function of gas concentration is presented in Fig. 8. Apparently, the Pd-decorated SnO2 NW sensor exhibited a remarkably improved selectivity to the investigated interference gases. The responsivities to CO gas (1–25 ppm) ranged from 1.8 to 2.8, whereas these values to H2 , NH3 , and CO2 with higher concentrations only ranged from 1.1 to 1.6. Our recent work has reported that the pristine SnO2 NW sensor has very poor selectivity to the measured gases [52,53]. If we define the selectivity factor = S/C where S is sensor response, and C is gas concentration, then the selectivity factor to CO (0.24) is higher than that of H2 (0.17), NH3 (0.106) and CO2 (0.001). This result confirms that Pd decoration enhances the response and selectivity to CO gas of SnO2 NW sensors. The better response toward CO than that of H2 and other gases of the synthesized Pd-SnO2 NWs was not clear yet and this needs further investigation for clarifying. However, Liewhiran et al. reported the higher response to H2 than to CO of the Pd-loaded SnO2 nanoparticles prepared by flame-spray method [54], whereas Zhang et al. reported a much higher response to CO of SnO2 doped with Pd and In than other gases such as C2 H5 OH, CH4 , H2 , CH3 COCH3 , NO2 , C2 H2 , NH3 [55]. Temp. ◦ 250 C 225 ◦ C 400 ◦ C 250 ◦ C Room temp. Room temp. 60 ◦ C 140 ◦ C 400 ◦ C Response (S = Rair /Rgas ) Refs. 1.019 (Rgas /Rair ) 147 1.1 15.9 64.5 1.02 ∼1.9 ∼3 1.8 Sens. Actuators B 138 (2009) 160–167 Sens. Actuators B 145 (2010) 7–12 Sens. Actuators B 181 (2013) 529–536 J. Phys. Chem. C 112 (2008) 11,539–11,544 Sens. Actuators B, 184 (2013) 33–39 Sen. Actuators B, 168 (2012) 8 Sens. Actuators B 177 (2013) 770–775 Sens. Actuators B 139 (2009) 287–291 This work 4. Conclusions We have introduced an effective method of decorating Pd NPs on the surface of SnO2 NWs to enhance CO-sensing performance. SnO2 NWs were directly grown on the sensor chips, whereas the decoration of Pd NPs was performed by in situ reduction of the Pd complex in the presence of pluronic P123 surfactant. This method is very facile and effective, and can be applied to decorate Pd NPs on the surface of various MOSs, such as ZnO and NiO. Gas-sensing characterization demonstrated enhanced CO-sensing performance of SnO2 NWs decorated with Pd NPs. Pd NP-decorated SnO2 NW sensors can detect a very low concentration of CO, down to below ppm level, with good response-recovery time and high stability. Decoration of Pd NPs on the surface of MOSs is effective for detection of CO and potential for the next generation of gas sensors. It can also be used in catalytic applications. Acknowledgments This work was finished through the collaboration between the International Training Institute for Materials Science (iSensors group, ITIMS) and the National Institute for Materials Science (MANA) under the Memorandum of Understanding signed in 2012. This research was funded by the Vietnam National Foundation for Science and Technology Development (Code: 103.02-2011.40) and VLIR-UOS under the Research Initiative Project (ZEIN2012RIP20). Appendix A. Supplementary data 3.2 CO H2 Response, S(Rair/Rgas) 2.8 NH3 CO2 2.4 2.0 1.6 1.2 0 20 40 5000 10000 15000 CO conc. (ppm) Fig. 8. The CO, H2 , NH3 , and CO2 gases response of the Pd-SnO2 NWs sensor as a function of gases concentration at temperature of 400 ◦ C. Supplementary data associated with this article can be found, in the online version, at http://dx.doi.org/10.1016/j.jhazmat. 2013.11.054. References [1] N.V. Hieu, N.D. Khoang, D.D. Trung, L.D. Toan, N.V. Duy, N.D. Hoa, J. Hazard. Mater. 209 (2013) 244–245. [2] Y. Liang, Y. Li, H. Wang, J. Zhou, J. Wang, T. Regier, H. Dai, Nat. Mater. 10 (2011) 780. [3] H.B. Wu, J.S. Chen, H.H. Hng, X.W. Lou, Nanoscale 4 (2012) 2526. [4] P.R. Solanki, A. Kaushik, V.V. Agrawal, B.D. Malhotra, NPG Asia Mater. 3 (2011) 17. [5] X. Chen, P. Li, H. Tong, T. Kako, J. Ye, Sci. Technol. Adv. Mater. 12 (2011) 044604. [6] B.H. Kim, S.Y. Oh, H.Y. Yu, W.G. Hong, Y.J. Yun, Y.Y. Kim, H.J. Kim, Sci. Technol. Adv. Mater. 11 (2010) 065003. [7] N.D. Hoa, S.A. El-Safty, Nanotechnology 22 (2011) 485503. [8] H.V. Han, N.D. Hoa, P.V. Tong, H. Nguyen, N.V. Hieu, Mater. Lett. 94 (2013) 41. [9] A. Wisitsoraat, A. Tuantranont, V. Patthanasettakul, T. Lomas, P. Chindaudom, Sci. Technol. Adv. Mater. 6 (2005) 261. [10] (a) N.D. Hoa, V.V. Quang, N.V. Hieu, D.J. Kim, J. Alloys Compd. 549 (2013) 260; (b) N.D. Khoang, H.S. Hong, D.D. Trung, N.V. Duy, N.D. Hoa, D.D. Thinh, N.V. Hieu, Sens. Actuators B 181 (2013) 529. [11] K.Y. Choi, J.S. Park, K.B. Park, H.J. Kim, H.D. Park, S.D. Kim, Sens. Actuators B 150 (2010) 65. [12] Y. Yamada, C.K. Tsung, W. Huang, Z. Huo, S.E. Habas, T. Soejima, C.E. Aliaga, G.A. Somorjai, P. Yang, Nat. Chem. 3 (2011) 372. Author's personal copy 132 D.D. Trung et al. / Journal of Hazardous Materials 265 (2014) 124–132 [13] B. Liu, S. Yu, Q. Wang, W. Hu, P. Jing, Y. Liu, W. Jia, Y. Liu, L. Liu, J. Zhang, Chem. Commun. 49 (2013) 3757. [14] C.J. Arbiol, J.R. Morante, U. Weimar, N. Bârsan, W. Göpel, Sens. Actuators B 70 (2000) 87. [15] M. Stankova, X. Vilanova, J. Calderer, E. Llobet, J. Brezmes, I. Gracia, C. Can, X. Correig, Sens. Actuators B 113 (2006) 241. [16] S. Ren, G. Fan, S. Qu, Q. Wang, J. Appl. Phys. 110 (2011) 084312. [17] Y. Xiao, L. Lu, A. Zhang, Y. Zhang, L. Sun, L. Huo, F. Li, ACS Appl. Mater. Interfaces 4 (2012) 3797. [18] N.V. Duy, N.D. Hoa, N.V. Hieu, Sens. Actuators B 173 (2012) 211. [19] H.Y. Lai, C.H. Chen, J. Mater. Chem. 22 (2012) 13204. [20] S.W. Choi, S.S. Kim, Sens. Actuators B 168 (2012) 8. [21] K. Wang, T. Zhao, G. Lian, Q. Yu, C. Luan, Q. Wang, D. Cui, Sens. Actuators B 184 (2013) 33. [22] S.W. Choi, S.H. Jung, S.S. Kim, Nanotechnology 22 (2011) 225501. [23] N.V. Hieu, P.T.H. Van, L.T. Nhan, N.V. Duy, N.D. Hoa, Appl. Phys. Lett. 101 (2012) 253106. [24] J.M. Baik, M.H. Kim, C. Larson, C.T. Yavuz, G.D. Stucky, A.M. Wodtke, M. Moskovits, Nano Lett. 9 (12) (2009) 3980. [25] H.B.R. Lee, S.H. Baeck, T.F. Jaramillo, S.F. Bent, Nano Lett. 13 (2) (2013) 457. [26] F. Chávez, G.F. Pérez-Sánchez, O. Goiz, P. Zaca-Morán, R. Peña-Sierra, A. Morales-Acevedo, C. Felipe, M. Soledad-Priego, Appl. Surf. Sci. 275 (2013) 28. [27] C.Y. Chen, J.K. Chang, W.T. Tsai, C.H. Hung, J. Mater. Chem. 21 (2011) 19063. [28] O. Mekasuwandumrong, J. Panpranot, Ind. Eng. Chem. Res. 48 (6) (2009) 2819. [29] Y. Fang, S. Guo, C. Zhu, S. Dong, E. Wang, Chem. Asian J. 5 (8) (2010) 1838. [30] Y. Zhang, J. Xu, P. Xu, Y. Zhu, X. Chen, W. Yu, Nanotechnology 21 (2010) 285501. [31] A. Kolmakov, D.O. Klenov, Y. Lilach, S. Stemmer, M. Moskovits, Nano Lett. 5 (4) (2005) 667. [32] J. Moon, J.A. Park, S.J. Lee, T. Zyung, I.D. Kim, Sens. Actuators B 149 (2010) 301. [33] M. Yuasa, T. Kida, K. Shimanoe, ACS Appl. Mater. Interfaces 4 (8) (2012) 4231. [34] S.S. Kim, J.Y. Park, S.W. Choi, H.G. Na, J.C. Yang, H.W. Kim, J. Alloys Compd. 509 (2011) 9171. [35] L.L. Xing, C.H. Ma, Z.H. Chen, Y.J. Chen, X.Y. Xue, Nanotechnology 22 (2011) 215501. [36] Y. Zhang, Q. Xiang, J. Xu, J. Xu, P. Pan, Q.F. Li, J. Mater. Chem. 19 (27) (2009) 4701. [37] H. Li, J. Xu, Y. Zhu, X. Chen, Q. Xiang, Talanta 82 (2010) 458. [38] C.M. Chang, M.H. Hon, I.C. Leu, RSC Adv. 2 (2012) 2469. [39] N.V. Hieu, Sens. Actuators B 144 (2010) 425. [40] B. Lim, M. Jiang, P.H.C. Camargo1, E.C. Cho, J. Tao, X. Lu, Y. Zhu, Y. Xia, Science 324 (2009) 1302. [41] H. Chen, G. Wei, A. Ispas, S.G. Hickey, A. Eychmüller, J. Phys. Chem. C 114 (50) (2010) 21976. [42] S. Phadungdhitidhada, S. Thanasanvorakun, P. Mangkorntong, S. Choopun, N. Mangkorntong, D. Wongratanaphisan, Curr. Appl. Phys. 11 (2011) 1368. [43] C.M. Chang, M.H. Hon, I.C. Leu, ACS Appl. Mater. Interfaces 5 (1) (2013) 135. [44] C. Zhu, J. Zeng, P. Lu, J. Liu, Z. Gu, Y. Xia, Chem. Eur. J. 19 (2013) 5127. [45] D.D. Trung, N.V. Toan, P.V. Tong, N.V. Duy, N.D. Hoa, N.V. Hieu, Ceram. Int. 38 (2012) 6557. [46] H. Chen, G. Wei, A. Ispas, S.G. Hickey, A. Eychmüller, J. Phys. Chem. C 114 (2010) 21976. [47] R. Wang, H. He, L.C. Liu, H.X. Dai, Z. Zhao, Catal. Sci. Technol. 2 (2012) 575. [48] A. Kolmakov, Y. Zhang, G. Cheng, M. Moskovits, Adv. Mater. 15 (2003) 997. [49] F. Hernández-Ramírez, A. Tarancón, O. Casals, J. Arbiol, A. Romano-Rodríguez, J.R. Morante, Sens. Actuators B 121 (2007) 3. [50] Q. Kuang, C.S. Lao, Z. Li, Y.Z. Liu, Z.X. Xie, L.S. Zheng, Z.L. Wang, J. Phys. Chem. C 112 (2008) 11539. [51] L.H. Qian, K. Wang, Y. Li, H.T. Fang, Q.H. Lu, X.L. Ma, Mater. Chem. Phys. 100 (2006) 82. [52] N.D. Khoang, D.D. Trung, N.V. Duy, N.D. Hoa, N.V. Hieu, Sens. Actuators B 174 (2012) 594. [53] D.D. Trung, L.D. Toan, H.S. Hong, D.T. Lam, T. Trung, N.V. Hieu, Talanta 88 (2012) 152. [54] C. Liewhiran, N. Tamaekong, A. Wisitsoraat, A. Tuantranont, S. Phanichphant, Sens. Actuators B 176 (2013) 893. [55] T. Zhang, L. Liu, Q. Qi, S. Li, G. Lu, Sens. Actuators B 139 (2009) 287.