Taehun
Im
ac,
Juyun
Lee
abc,
Sung-Chul
Kim
d,
Joharimanitra
Randrianandraina
e,
Joo-Won
Lee
a,
Myoung Won
Chung
f,
Taesung
Park
c,
Kam-Hung
Low
g,
Seungkyu
Lee
g,
Soong Ju
Oh
c,
Yun Chan
Kang
c,
Seunghyun
Weon
f,
Jung-Hoon
Lee
e,
Seon Joon
Kim
*abh and
Sohee
Jeong
*a
aMaterials Architecturing Research Center, Korea Institute of Science and Technology, Seoul, 02792, South Korea. E-mail: seonjkim@kist.re.kr; soheejeong@kist.re.kr
bConvergence Research Center for Solutions to Electromagnetic Interference in Future-mobility, Korea Institute of Science and Technology, Seoul, 02792, Republic of Korea
cDepartment of Materials Science and Engineering, Korea University, Seoul, 02841, Republic of Korea
dAdvanced Analysis and Data Center, Korea Institute of Science and Technology, Seoul, 02792, Republic of Korea
eComputational Science Research Center, Korea Institute of Science and Technology, Seoul, 02792, Republic of Korea
fSchool of Health and Environmental Science & Department of Health and Safety Convergence Science, Korea University, Seoul, 02841, Korea
gDepartment of Chemistry, The University of Hong Kong, Pokfulam Road, Hong Kong SAR, China
hDivision of Nanoscience and Technology, KIST School, University of Science and Technology, Seoul, 02792, Republic of Korea
First published on 22nd July 2024
With the increasing demand for ammonia applications, there is a significant focus on improving NH3 detection performance at room temperature. In this study, we introduce a groundbreaking NH3 gas sensor based on Cu(I)-based coordination polymers, featuring semiconducting, single stranded 1D-helical nanowires constructed from Cu–Cl and N-methylthiourea (MTCP). The MTCP demonstrates an exceptional response to NH3 gas (>900% at 100 ppm) and superior selectivity at room temperature compared to current materials. The interaction mechanism between NH3 and the MTCP sensor is elucidated through a combination of empirical results and computational calculations, leveraging a crystal-determined structure. This reveals the formation of NH3–Cu and NH3–H3C complexes, indicative of a thermodynamically favorable reaction. Additionally, Ag-doped MTCP exhibits higher selectivity and a response over two times greater than the original MTCP, establishing it as a prominent NH3 detection system at room temperature.
New conceptsIn this manuscript, we introduce a concept to synthesize and utilize a novel Cu(I)-based coordination polymer (CP) that exhibits exceptional gas sensing sensitivity and selectivity toward NH3 at room temperature. On many efforts to develop novel nanomaterials that enable high-performance chemical sensing, polymers, 2D materials, and their hybrids have been studied as an alternative to conventional metal oxide semiconductors, which are limited by low selectivity and high operating temperature. In this manner, CPs have gained attention as they offer tunable electrical properties that can be adjusted through the modulation of metal and ligand components. Here, we highlight the development of a novel sensing material, featuring a unique structure consisting of bundles of semiconducting, single stranded 1D-helical nanowires each with Cu(I) centers bonded to S atoms in N-methylthiourea ligand and Cl atoms (referred to as MTCP). The very narrow electrical conduction pathway in each CP single strand provides an outstanding platform to realize ultrasensitive molecule detection. Remarkably, sensors based on MTCP exhibited up to a 2000% gas response toward 100 ppm NH3, respectively, which is one of the highest responses ever reported at room temperature among all sensing materials. |
Solid-state-type gas sensors based on semiconductors have been favored and commercialized due to their small size and compatibility with existing electronic devices. To date, metal oxide semiconductor (MOS) based gas sensors have been extensively employed for ammonia gas detection among various materials due to their high sensitivity and material accessibility. However, MOS gas sensors typically require elevated operating temperatures to enhance their sensitivity, which results in increased power consumption, thermal stress on the sensor, and the need for heating elements.2,3 Such limitations can be critical in the fabrication of integrative small devices or portable devices which need to operate on low power and in a safer environment. Additionally, MOS sensors exhibit low selectivity toward specific target gases due to the difficulty in engineering molecule-nanoparticle interactions, as gas molecules indirectly interact with the MOS nanoparticles through an oxygen-mediated electron transfer mechanism. As an alternative, gas sensors based on polymers, 2D materials, and composites offer the advantage of operating at room temperature. For instance, polyaniline (PANI) which has some drawbacks like low sensitivity, has showed enhanced NH3 sensing performance when combined with other functional materials (e.g. carbon materials, MXene, metal oxides).4–8 However, it is important that their sensitivity is notably lower compared to MOS sensors.3,9
Recently, coordination polymers (CPs) have emerged as a potential solution for room temperature gas sensing.10,11 CPs represent a hybrid material, combining advantageous properties from both inorganic and organic materials, not only in terms of chemical composition but also overall characteristics. Notably, CPs offer tunable electric properties that can be adjusted by changing the metal centers or ligands, thereby controlling their band gap and conductivity. However, CPs face certain limitations as gas-sensing materials, such as the complexity of synthesis, requiring meticulous control over reactant ratios, reaction conditions, and purification steps, leading to higher production costs. Additionally, most CPs do not exhibit proper electrical conductivity like semiconductors, which hampers efficient signal transduction in response to the target gases.12 Although a few coordination polymers exhibit a certain level of conductivity, their gas selectivity and sensitivity significantly lack when compared to other materials such as MOS. This deficiency can be attributed to the ineffective overlap between the d orbitals in metals and p orbitals in oxygen- or nitrogen-based redox-interactive ligands bonds, which is commonly found in coordination polymers.13,14 Furthermore, CPs predominantly form a 3D network, leading to limited charge transport in a single direction. As a result, the formation of electronic conduction network become challenging. To address this limitation, previous studies have suggested using metal–sulfur bonds and 1D network as potential solutions.15–17 It is believed that these bonds can enhance electronic bonding and charge transport, making coordination polymers more akin to semiconducting or metallic materials like SnO2 and ZnO. Unlike 3D and 2D pathways, the presence of narrow molecular conduction pathways (e.g. 1D network) allows for efficient signal transduction by maximizing changes in electrical resistance upon exposure to the target gas. Among various CPs, coordination polymers incorporating copper metals are a promising for gas sensor material because copper has low toxicity, low cost, and abundance compared to precious and rare-earth metals, and well-established interactions to ammonia molecules. For example, copper sulfide and halide materials were used for ammonia sensing, showing improved selectivity and sensitivity due to the stable interaction between Cu ions and NH3 molecules through electron transfer from the lone pair electron of N.18–22 Therefore, copper halide CPs, combined with diverse ligands, as well as copper-organic network CPs utilizing organic-containing ligands, have been employed in gas sensor applications.23–25
In this study, we developed an innovative NH3 gas sensor based on Cu(I) CPs with semiconducting 1D-helical nanowires (referred to as MTCP {[Cu(Cl)(MT)], where MT = N-methylthiourea}) (Scheme 1). This MTCP offers several novel features, including Cu–S bonding for semiconductor-like electrical conductivity, a 1D network for efficient signal transduction, and a selective Cu–NH3 interaction. This sensor exhibits an exceptionally high response to NH3 gas and superior selectivity at room temperature compared to other existing materials to date. Notably, this coordination polymer can be easily synthesized in ethanol solvents within 10 minutes at room temperature. We have successfully determined the structure of a new coordination polymer consists of a 1D-helical chain constructed with Cu centers bonded to S atoms in MT ligands and Cl atoms, as confirmed by X-ray crystallographic and X-ray spectroscopic analysis. The molecular-level study of the NH3 reaction mechanism in this MTCP reveals that the effective NH3 sensing arises from the formation of NH3–Cu and NH3–H3C complexes during exposure of NH3 gas that is computationally predicted to be highly exothermic. Additionally, Ag-doped MTCP improves both selectivity and reactivity, demonstrating one of the highest reported gas responses toward NH3.
To elucidate the structural characteristics of MTCP, we conducted a comprehensive analysis of its composition and bonding states using Fourier transform infrared spectroscopy (FTIR), X-ray photoelectron spectroscopy (XPS), and X-ray absorption spectroscopy (XAS). In Fig. S4 (ESI†), upon comparing the FTIR peaks of MTCP with those of the MT precursor, significant shifts were observed in the peaks corresponding to N–H stretching vibrations from 3254 cm−1 to 3290 cm−1. This shift can be attributed to the highly polar structure resulting from the formation of Cu–S bonds.29,30 Furthermore, the intensity of strong peaks within the range of 1100 to 900 cm−1, corresponding to CS vibrations, decreases due to the generation of Cu–S bonds.29–31 The generation of these bonds also causes shift in certain peaks related to N–H vibrational properties from 1296 cm−1, 1404 cm−1, and 1556 cm−1 to 1302 cm−1, 1420 cm−1, and 1585 cm−1.
In order to better understand of the coordination of Cu and ligand in MTCP, we also performed Cu K-edge and L-edge XAS analyses (Fig. 2b and c). For 3d metals, K-edge XAS primarily probes 1s → 4p transition, while L3,2-edge XAS probes 2p → 3d transition. The white-line absorption of Cu K-edge of MTCP, primary indicator of the Cu oxidation, resembles that of CuCl rather than CuCl2. This alignment suggests that the oxidation state of Cu in MTCP is +1, which is also supported by the result obtained by XPS analyses. Furthermore, L-edge spectra, including both surface-sensitive total electron yield (TEY) and bulk-sensitive total fluorescence yield (TFY) spectra, was investigated to observe the details of ligand coordination (Fig. 2c). For copper chlorides, the peak at 930.8 eV was detected in both TEY and TFY spectra, indicating the oxidation of the Cu species and showing a local structure consistent with CuO.36,37 On the other hand, the TEY and TFY spectra of MTCP show a very weak signal in this region in the both the surface and bulk of MTCP, suggesting that the Cu(I) in MTCP maintains a robust coordination with the ligands with good oxidation resistance. It is important to note that Cu(I) is known to be well oxidized and can easily be formed CuO.38,39
Moreover, the Cu L-edge X-ray absorption near edge structure (XANES) spectra shows distinctive hybrid state in MTCP compared to copper chlorides. In the TEY and TFY spectra of MTCP, a new peak was observed at 936 eV which is higher energy than the 933.5 eV typically associated with Cu(I) compound.37 Previous reports have established that the interaction between a metal with filled d-orbitals and ligands, facilitated by π back-bonding contribution, leads to the appearance of a high-energy peak in L-edge spectra.40,41 Interestingly, this spectral feature is also observed in MTCP, which suggests the strong interaction between the central Cu(I) atom and the thione ligand, as confirmed by IR and X-ray spectroscopic analyses as well.42 Additionally, the peak at 936 eV exhibits a comparable intensity to the lower energy feature at 933.5 eV, which suggests electron transitions into an antibonding orbital with a significant Cu-d character, associated with thione π back-bonding interaction.43,44 In overall, these spectroscopic distinctions of MTCP from other copper–chlorine compounds, such as CuCl or CuCl2, signifies the formation of a new bond between Cu–Cl and MT ligand.
To further comprehend the electrical properties of MTCP, electrical conductivity measurements were conducted using a 2-point probe current measurement technique (Fig. S6, ESI†). The conductivity and resistivity of MTCP were calculated by employing Ohm's law, wherein the dimensions of sample and the measured current were considered as variables. As a results of these calculations, it was determined that MTCP exhibits a temperature coefficient of resistivity (TCR) of −0.020 K−1. Notably, the negative TCR aligns with the characteristic behavior commonly associated with semiconductors, suggesting that MTCP possess semiconducting properties. To further elucidate the band diagram of the semiconducting MTCP, ultraviolet-visible spectroscopy (UV-vis) (Fig. S7a, ESI†) and ultraviolet photoelectron spectroscopy (UPS) (Fig. S8, ESI†) were carried out. The Tauc plot in Fig. S7b (ESI†) revealed that the bandgap of MTCP was 3.76 eV, which is similar to those of semiconducting metal oxides commonly used in gas sensors, such as SnO2 (3.6 eV).45 UPS measurements revealed that the work function (WF) of MTCP was 4.7 eV, while the valence band maximum (VBM) was 0.87 eV below from the Fermi level. Combining these results, a band diagram for MTCP could be drawn as shown in Fig. 2d. The band diagram clearly shows that MTCP would be a p-type semiconductor as the Fermi level is closer to the VBM than the conduction band minimum (CBM).
Using DFT calculations, we elucidated energetically favorable NH3 binding sites in MTCP to further understand the adsorption mechanism at the molecular level. As spectroscopy analyses revealed the chemical binding of NH3 with Cu, this case was initially investigated. DFT calculations revealed that there are two favorable binding sites for this case: (i) NH3 binding to Cu within the Cu–S channel (panel (i) of Fig. 4b), leading to a N (in NH3)–Cu distance of 2.08 Å with a binding energy of −0.71 eV, and (ii) NH3 binding to Cu from outside of the Cu–S channel (panel (ii) of Fig. 4b) where NH3 replaces with Cl atoms with a stronger binding energy of −1.50 eV at a N (in NH3)–Cu distance of 2.01 Å. Such binding behaviors are consistent with a previous study on Cu(I)–halide compounds for NH3-selective gas sensors, showing that Cu ions and NH3 are likely to form ammine complexes.20
Moreover, another stable NH3 binding site was revealed using ab initio molecular dynamics (AIMD) calculations (see panel (iii) of Fig. 4b and Fig. S12, ESI†). As shown in Fig. S12 (ESI†), NH3 initially interacts with amine units in MTCP at a calculated binding energy of −0.25 eV, which then moves in between two methyl groups of the ligand side. The calculated binding energy of this latter configuration was −0.38 eV. Both binding energies were obtained by optimizing a snapshot of the configuration using DFT calculations. Then eventually, NH3 further diffuses into a position in between four methyl groups each originating from different MTCP backbones as shown in panel (iii) of Fig. 4b and Fig. S12 (ESI†). The NH3 binding energy of this geometry is −0.37 eV, where NH3 entrapment is favored by the environment provided by the surrounding methyl groups. The calculated NH3 binding energies of three main binding sites are summarized in Fig. 4c. Our DFT calculations showed that ammonium ion-like components (NH3⋯H) and NH3–Cu(I) can form in the presence of NH3, which agrees well with our in situ DRIFT measurements. In overall, these calculations show that MTCPs have sufficient binding sites that can facilitate the efficient and strong adsorption of NH3 molecules.
The dynamic gas sensing response of Ag-MTCP toward 0.25–100 ppm NH3 is shown in Fig. 5c and d where the graphs based on actual resistance values is in Fig. S15 (ESI†). Indeed, Ag-MTCP showed a gas response of 1925% for 100 ppm NH3, which is twice higher than that of MTCP. Here, the Ag-MTCP sensor exhibited a response time of 378 s and a recovery time of 215 s toward 100 ppm NH3 at room temperature. Fig. 5e compares the maximal gas response of Ag-MTCP toward 0.25–100 ppm NH3, clearly showing the linear correlation similar to what was observed for MTCP. Furthermore, the gas response of Ag-MTCP under humid environments was also investigated, as shown in Fig. S16 (ESI†). While the gas response of Ag-MTCP toward 20 ppm NH3 in a dry environment was 300%, the gas response was decreased to 160% in a 40% RH environment. A similar phenomenon was observed in MTCP where the gas response in a dry environment was 190%, decreasing to 120% in a 40% RH environment. This is expected to be due to partial screening of adsorption sites by water molecules.
We estimate that two factors contribute to the enhanced sensitivity after Ag doping. First, in a previous study, a polaron-hopping based conduction mechanism was proposed for this type of molecule where the transformation between Cu+ and Cu2+ plays a major role.53 While the redox behavior between Cu+ and Cu2+ is known to be relatively reversible, Ag ions strongly prefer to exist as Ag+ compared to Ag2+, based on reduction potentials. Such a behavior can obstruct the hopping-based conduction mechanism, in which a large decrease in electrical conductivity was observed after Ag doping, as measured with the assistance of focused ion beam (FIB) analysis. (Fig. S17 and S18, ESI†). Second, Ag ions are also well known to bind with NH3 to form ammine complexes, which can provide a synergistic effect.
In addition, MTCP and Ag-MTCP was tested against gases with various types of chemical moieties to investigate their gas sensing selectivity where the average gas response of multiple samples toward acetone, ethanol, hexane, H2, and NO2 was measured (Fig. 5f). Remarkably, this sensor exhibited a significant gas response, primarily to NH3, with the exception of NO2, although the respond is marginal. For example, the gas response of MTCP toward 100 ppm acetone and ethanol was −6.4% and −4.0%, respectively, which is a negligible value compared to the gas response toward NH3 gas. These results show that MTCP-based sensors have a superior selectivity toward NH3. Furthermore, the positive gas response toward NH3 and negative gas response toward NO2 suggests a p-type semiconductor-like behavior, which is in correspondence with the revealed electronic structure in Fig. 2d. Fig. 5g and Table S2 (ESI†) shows the state-of-the-art comparison of the gas sensing performance of various nanomaterials toward NH3 under room temperature including metal oxides, polymers, and 2D materials. Detailed comparisons are shown in Table S2 (ESI†). It can be clearly seen that MTCP and Ag-MTCP is among the most sensitive NH3 gas sensors to the best of our knowledge with a maximal gas response of 915% and 1925%, respectively, where previous reports show gas responses lower than 300% for 100 ppm NH3. These results highlight the exceptional gas response of MTCPs at room temperature.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4mh00651h |
This journal is © The Royal Society of Chemistry 2024 |