Coordinated water molecule-induced solid-state superprotonic conduction by a highly scalable and pH-stable coordination polymer (CP)

Rupam Sahoo a, Shaozhen Luo b, Naresh Kumar Pendyala b, Santanu Chand a, Zhi-Hua Fu b and Madhab C. Das *a
aDepartment of Chemistry, Indian Institute of Technology Kharagpur, Kharagpur 721302, WB, India. E-mail: mcdas@chem.iitkgp.ac.in; Web: https://www.chemiitkgp-mcdaslab.com/
bState Key Laboratory of Structural Chemistry, Fujian Institute of Research on the Structure of Matter, Chinese Academy of Sciences (CAS), 155 Yangqiao Road West, Fuzhou, Fujian 350002, China

Received 3rd January 2023 , Accepted 5th May 2023

First published on 9th May 2023


Abstract

The effective design of new efficient crystalline solid-state proton conductors (SSPCs) is relevant to the advancement of clean energy applications. Although it is well known that metal ions enhance the acidity of their coordinated water molecules, proton-conducting coordination polymers (CPs) and metal–organic frameworks (MOFs) in which coordinated water molecules act as sole intrinsic proton sources are largely unexplored. Considering such a design principle, we present a highly scalable and highly robust (stable over a wide pH range of 2–10, in open air for six months, water and boiling water) CP, IITKGP-101, in which Cu43-OH)4 SBUs are coordinated with ample water molecules and display superprotonic conductivity in both the single crystal and pellet forms. The 1D chains with hydrophilic interlayer spaces within the framework housing abundant coordinated water as the sole proton source, which make an extended H-bonding pathway for efficient proton migration and thus exhibit a conductivity value of 5.2 × 10−4 S cm−1 in single crystal form and 2.45 × 10−4 S cm−1 in pellet form at 80 °C and 98% RH. The easy scalability in gram scale in an aqueous medium and highly robust nature make this framework a promising candidate as an SSPC.


Introduction

The depletion of non-renewable fossil fuels necessitates alternative clean and sustainable energy resources for sustaining human life.1,2 The proton exchange membrane fuel cell (PEMFC) has received extensive attention as the most favorable candidate for portable energy-related devices and vehicle applications due to its ultra-low emission features and high power density.3,4 Although the previously developed and commonly used state-of-the-art proton-conducting materials, Nafion and Flemion, displayed superior proton conductivity, the high cost associated with per-fluorination limits their widespread usage.3,5–7 In addition, the amorphous nature of these traditional polymers is unable to provide accurate structural information for in-depth investigation on proton-conduction mechanisms and pathways. Therefore, the development of crystalline, structurally greatly stable and highly conductive solid-state proton conductors (SSPCs) is in demand.5 In this regard, coordination polymers (CPs) and metal–organic frameworks (MOFs) assembled by metal ions and organic linkers having variability of the structural architecture8–39 and superior tunability40–61 have displayed excellent proton conductivity, even in ultra-high superprotonic regions (>10−2 S cm−1 or higher).6,62 Most importantly, the stronger coordination bonds and highly crystalline nature enable accurate structural characterization, which is beneficial for understanding the plausible proton-conduction mechanism. It should be noted that CPs are comparatively less explored than their MOF counterparts as SSPCs.6,8,11,62–73 Again, by taking advantage of their structural superiority, ∼300 proton-conducting CPs have been developed so far; however, CPs displaying conductivity at the superprotonic region (≥10−4 S cm−1) are even rarer.6,62

Surprisingly, in most cases, proton conductivity has been analyzed in the pellet form rather than in single crystals.74 Due to random orientation during conduction measurement of the pelletized sample, the characterization of anisotropic effects is precluded. To solve this issue, proton conduction measurements on single crystals with ordered and well-defined guest molecules are indispensable. The insufficient crystal length, irregular shape, inadequate stability, and complex sample processing are the major barriers to proton conduction measurements on single crystals.74 The proton conduction measurements of single crystals could provide a basis for understanding the essence of complex conductive mechanisms. To the best of our knowledge, there are still very few reports where proton conductivity has been analyzed in the single crystal form or both single crystal and pellet forms,22,74–84 and thus this demands more attention. On the other hand, CPs are infamous for their instability under water/moist conditions, which limits them in practical applications, especially toward water containing media such as in PEMFCs where a high level of humidification is essential for cyclic operation in FCs. Also, easy bulk-scale synthesis is indeed a prerequisite criterion for successful industrial deployment. Thus, the effective design and development of chemically robust SSPCs with obvious scalability are of significant interest.

In terms of design principles, depending upon the proton sources, CPs are generally categorized into two types: (i) intrinsic and (ii) extrinsic proton conductors.9,62,63 Although both types of proton sources are efficient to achieve proton-conductivity in the ultra-high superprotonic regions,6,62 the developed materials having extrinsic proton sources may suffer from a leaching effect, which thus can limit their practical utility.5 In these circumstances, the design and development of highly stable CPs having intrinsic proton sources (such as uncoordinated –SO3H, –PO3H2, and –COOH groups flanked toward the pore channels or protic H3O+, Me2NH2+, NH4+, H2PO4, etc. counter-ions present in the pore channels) is of significant interest. Recently, the crucial roles of coordinated water molecules in bringing the protonic conductivity on CPs and MOFs have been highlighted, where these coordinated water molecules are shown to act as the sole intrinsic proton sources by virtue of their enhanced acidity, while bound to the metal centers.9,63 To construct CPs and MOFs with a higher number of coordinated water molecules and to make the frameworks superprotonic using coordinated water molecules as the only protonic source, exclusively aqueous medium or aqueous DMF solution (usage of minimal DMF to resolve the precursors’ solubility issue) for crystallization has been proposed.63 However, such types of proton-conducting CPs and MOFs are still limited in the literature, among which CPs are even rarer and demand further development.9,10,16,63 Besides imparting protonic conductivity, such an approach could bring inherent water stability because of the usage of an aqueous crystallization medium, which is again beneficial for being used as an SSPC in real applications.

On the other hand, employing aromatic acid ligands to construct robust CPs is undoubtedly a good choice as they possess strong coordination ability with diverse coordination modes.65 It should be noted that the aromatic rings of the constructed CPs have a significant contribution to the framework's overall stability such as (i) π–π stacking interactions between the aromatic rings, which provide extra stabilization energy and (ii) hydrophobicity of the aromatic rings that enhances the water stability. In addition, uncoordinated acid groups (deprotonated carboxylates, sulfonates, and phosphonates) of the ligands do act as proton hopping sites and participate in extensive H-bonding construction, which is further beneficial for achieving proton conductivity of such CPs.

Considering the aforementioned aspects, we have designed and developed a highly scalable and wide range pH-stable (2–10) Cu(II)-bpy chain-based coordination polymer, IITKGP-101, with molecular formula {[Cu43-OH)4(4,4′-bpy)2(2,6-NDS)(H2O)7]·2,6-NDS·6H2O}n (where bpy = 4,4′-bipyridine, 2,6-NDS = disodium 2,6-naphthalenedisulfonate), which displayed superprotonic conductivity in both the single crystal and pellet forms. The extensive H-bonded chains inside the 1D interlayer spaces are constructed among the well-ordered ample coordinated water molecules, lattice water molecules, free lattice 2,6-NDS units, and uncoordinated O atoms of the metal-bound 2,6-NDS units. The maximum proton conductivity was obtained as 5.2 × 10−4 S cm−1 in the single crystal form at 80 °C and 98% RH, whereas in the pellet form a conductivity of 2.45 × 10−4 S cm−1 could be achieved under the same conditions. IITKGP-101 was synthesized in an aqueous medium using a room temperature slow evaporation crystallization technique, possessing a high density of coordinated water molecules, and exhibited protonic conductivity driven by ample coordinated water molecules as the sole intrinsic source. Multiple π⋯π stacking interactions between the bpy units along with aromatic 2,6-NDS acid ligands and bpy units not only enhanced the chemical stability (water, boiling water and pH) of the developed framework but also exhibited framework robustness even after treatment under the extreme conditions of conductivity measurement (80 °C and 95–98% RH) for a longer time. Most importantly, the ease of G-scale synthesis, open-air simple handling (stable as long as for six months in an open lab atmosphere), superior stability, and superprotonic conductivity with extended durability made IITKGP-101 promising toward this projected application.

Results and discussion

Room-temperature crystallization of Cu(NO3)2·3H2O, 4,4′-bpy, and disodium salt of 2,6-NDS under an aqueous medium yielded block-shaped blue crystals of IITKGP-101 (Fig. S1, ESI). It crystallizes in the triclinic crystal system with the P[1 with combining macron] space group (Table S1, ESI) in which a high density of coordinated water molecules containing Cu43-OH)4 SBUs are connected via 4,4′-bpy bridging ligands and construct a 1D polymeric infinite chain (Fig. 1a). The asymmetric unit consists of two halves of the Cu43-OH)4 SBU coordinated with seven water molecules, two bridging 4,4′-bpy units, and one coordinated and one lattice 2,6-NDS unit along with six lattice water molecules (Fig. S2, ESI). Among the four Cu(II) centers, three (Cu2, Cu3, and Cu4) are coordinated with two water molecules each whereas the remaining one (Cu1) is coordinated with one water molecule and one deprotonated 2,6-NDS unit (Fig. 1b). Each of the Cu(II) centers possesses a distorted octahedral coordination environment: coordinated with two crystallographically independent water molecules (except one Cu1-center that is coordinated with one water and one 2,6-NDS unit) along with one 4,4′-bpy unit and three oxygen atoms from the three different μ3-OH bridges. In Cu43-OH)4 SBUs, the Cu–O bond distances are significantly shorter and lie in the range of 1.943(8)–1.994(1) Å, whereas the Cu–O bond distances in the mono-coordinated 2,6-NDS unit are relatively higher (2.682(1) Å). The Cu–N bond distances for bidentate bridging bpy ligands lie in the range of 2.010(1)–2.044(1) Å, whereas for coordinated water molecules the Cu–O bond distances lie in the range of 1.973(1)–2.395(1) Å (Table S2, ESI). As mentioned earlier, aromatic acid ligands play a significant role in the framework's stability through π⋯π stacking interactions; in IITKGP-101, the coordinated 2,6-NDS units also display strong π⋯π stacking interactions with the bridging bpy units having planar separation distances of 3.747 Å (Fig. S3, ESI). In addition, the bidentate bridging bpy units in two consecutive layers also exhibit π⋯π stacking interactions with a centroid-to-centroid distance of 3.653 Å, as displayed in Fig. S3 (ESI). Moreover, each 1D chain is further interconnected by H-bonding interactions among the coordinated water molecules, coordinated and lattice 2,6-NDS units, along with lattice water molecules, thus leading to the formation of an extensive H-bonded supramolecular architecture. The overall H-bonding pattern is displayed in Fig. S4 (ESI) and all such H-bonding interactions are listed in Table S3 (ESI). Besides extensive H-bonding, several non-bonding interactions are also present throughout the framework, as displayed in Table S4 (ESI). It should be noted that the hydrophilic channels between the 1D layers are extensively occupied by lattice water molecules and lattice 2,6-NDS units, which enabled the construction of continuous H-bonding throughout the framework. The lattice 2,6-NDS units could connect two different 1D chains via H-bonding interactions, as displayed in Fig. S5 (ESI). The two oxygen atoms (O18 and O21) of lattice 2,6-NDS units are hydrogen bonded with the hydrogens of the coordinated water molecules (H5A, H23B), whereas another oxygen atom (O17) along with O21 constructed H-bonds with the hydrogens (HW6A and HW1B) of the lattice water molecules.
image file: d3qm00007a-f1.tif
Fig. 1 (a) 1D polymeric infinite chain of IITKGP-101 (color codes: Cu, cyan; O, red; S, yellow; N, blue; C, gray; and H, white); (b) representation of Cu43-OH)4 SBUs displaying a high-density of coordinated water; (c) comparison the PXRD plots of IITKGP-101 for the as-synthesized, water-treated, pH-treated and bulk-scale synthesized samples.

As displayed in Fig. 1c, the PXRD pattern of the as-synthesized IITKGP-101 is in good agreement with the simulated pattern obtained from the single-crystal structure and confirms the phase purity of the as-synthesized sample. To verify the preliminary stability of IITKGP-101 in open air, the as-synthesized sample was left open in the laboratory atmosphere for a prolonged time and the crystallinity was monitored by PXRD analysis. No noteworthy change even after six months of open-air exposure signifies an open atmosphere easy accessibility of the developed material (Fig. S6, ESI). Furthermore, to confirm the water stability of the developed framework, IITKGP-101 was immersed not only in distilled water but also treated with boiling water for 24 h. The excellent retention of the PXRD peak positions in both cases confirms the structural integrity of IITKGP-101 (Fig. 1c). Besides hydro-stability, the chemical stability of the developed framework was also evaluated. No significant change in PXRD patterns even after immersion in pH 2 and pH 10 aqueous solutions for 24 h proves the wide range of chemical stability of IITKGP-101 (Fig. 1c). Such high framework robustness could be attributed to the presence of ample stronger metal–N coordination (soft–soft acid–base combination) with a large number of π⋯π stacking and extensive H-bonding interactions throughout the framework. It may be noted that, although, quite a few MOFs have been reported to show a wide range of pH stability by now and the reports are gradually increasing, while considering their potential industrial usage such numbers are considerably limited for CPs. The bulk material could be easily synthesized in gram scale through a simple room temperature crystallization method from a green aqueous medium. Most importantly, the crystallinity and phase purity of the G-scale as-synthesized sample could also be confirmed by PXRD analysis, as depicted in Fig. 1c. The characteristically strong and broad peak at ∼3500–3000 cm−1 in the FT-IR spectrum demonstrates the presence of water molecules (Fig. S7, ESI). Thermogravimetric analysis (TGA) under an N2 atmosphere has been performed to verify the thermal stability of the developed material. IITKGP-101 was thermally stable up to 250 °C with an initial wt loss of 7.3% (calcd 7.5%) at 100 °C followed by 7.9% wt loss (calcd 8.7%) at 240 °C, which could be attributed to the expelling of both crystallized and coordinated water molecules, respectively (Fig. S8, ESI). Differential scanning calorimetry (DSC) measurements ruled out the possibility of phase transitions under the proton conduction measurement conditions (Fig. S9, ESI). To have more information on the framework robustness, a VTPXRD (variable-temperature powder X-ray diffraction) study was performed, which displays that IITKGP-101 maintains its crystallinity until 175 °C (Fig. S10, ESI).

During the synthesis of IITKGP-101, the disodium salt of 2,6-NDS was used; thus, it is essential to confirm that the developed material does not contain any Na+ ions that can lead to ionic conductivity. The EDS analysis revealed the complete absence of sodium ions in the synthesized sample and confirmed that the obtained conductivity solely arises because of the protons (Fig. S11a, ESI). On the other hand, the field emission scanning electron microscopy (FESEM) image displays the dispersed phase crystalline nature of the IITKGP-101 sample (Fig. S11b, ESI).

Before proton conductivity evaluation, the moisture stability of IITKGP-101 was analyzed by exposing the as-synthesized sample under 97% RH for 3 days, which displayed complete retention of the crystallinity. Furthermore, the as-synthesized sample was also exposed to 80 °C and 98% RH for 16 h. No significant change in PXRD patterns under extreme measurement conditions confirmed the excellent moisture stability with extended durability of the developed material (Fig. S12, ESI). These are the preliminary criteria to be fulfilled for a crystalline material for further evaluation of its proton conduction properties. Such high stability coupled with the presence of several proton-conducting species (abundant coordinated and lattice water molecules, cationic [Cu43-OH)4(4,4′-bpy)2(2,6-NDS)(H2O)7]2+ units and anionic charge balancing 2,6-NDS units) bearing copious proton hopping sites interconnected via a well-defined H-bonding network as previously described prompted us to analyze the proton conductivity of IITKGP-101. As mentioned earlier, the metal center enhances the acidity of the coordinated water molecules, which can act as efficient proton sources and thus may lead to higher proton conductivity. A proton conduction study was performed via alternating-current (AC) impedance analysis in both the single crystal and pellet forms (Fig. S13, ESI). Proton conduction measurements were performed by varying both the temperature (30–80 °C) and humidity (40–98%). At ambient temperature (30 °C) in the single crystal form, IITKGP-101 displayed a proton conductivity of 4.25 × 10−5 S cm−1 under 98% RH (Table S5, ESI), which gradually increased with the increase of temperature and reached the superprotonic region with the value of 5.22 × 10−4 S cm−1 at 80 °C (Fig. 2). The higher proton conductivity at higher temperatures could be attributed to the generation of more active proton-conducting species due to the bond cleavage and the mobility increment of the protonic species. In the Nyquist plots, two different components were observed in which semicircles attribute the bulk resistance of the material, whereas the tails of the curves represent the accumulation of protons at the electrodes (Fig. 2). At 80 °C under lower humidity (40%), IITKGP-101 exhibited a relatively lower proton conductivity of 6.22 × 10−8 S cm−1 (Table S5, ESI); however, the conductivity increased significantly with the increase of RH and reached the superprotonic region at 98% RH with the value of 5.22 × 10−4 S cm−1 (Fig. S14 and 15, ESI). It should be noted that the proton conductivity increased by four orders of magnitude from low RH (40%) to high RH (98%). Such a significant rise in conductivity with the increment of RH demonstrates that the adsorbed water molecules act as bridges between the donor–acceptor molecules and lead to water-mediated proton diffusion efficiently into the structure. For a better understanding of the proton-conducting property, an ambient temperature water-sorption study was performed by activating the as-synthesized material at 373 K temperature under a vacuum. As displayed in Fig. S16 (ESI), the steep water uptake capacity even at lower relative humidity signifies the hydrophilic nature of the developed framework. At a lower relative pressure (P/P0 = 0.1), the H2O sorption was 34.7 cc g−1 (corresponding to uptake of two water molecules per formula unit), which increased significantly with the increase of relative pressure and reached the maximum of 100.1 cc g−1 at P/P0 = 0.83 (corresponding to ∼5.4 water molecules per formula unit). The absorbed water molecules further act as efficient proton carriers in the channel and displayed a superprotonic conductivity at higher RH. For a better understanding of this phenomenon, we have correlated the proton conductivity with the number of absorbed water molecules present in the system (Fig. 3). The lower proton conductivity of 6.22 × 10−8 S cm−1 at low humidity (40% RH) could be attributed to the lower number of absorbed water molecules (4.42) per formula unit, which increased with the increase of humidity and reached the maximum of ∼5.4 water molecules per formula unit at 98% RH to lead the superprotonic conductivity. The maximum proton conductivity of 5.22 × 10−4 S cm−1 at 80 °C and 98% RH is comparable to those of benchmark MOFs and CPs in which proton conductivity has been tested in the single crystal form (Table S7, ESI). Although the proton conductivity value of IITKGP-101 is slightly lower than those of Co-MOF-74 (6.1 × 10−4 S cm−1 at 90 °C, 95% RH),841·3H2O (1.43 × 10−3 S cm−1 at 80 °C, 95% RH),82PCC-72 (1.2 × 10−2 S cm−1 at 95 °C, 98% RH),81CFA-17 (2.1 × 10−3 S cm−1 at 22 °C, 95% RH),74In(5-Hsip)2 (1.25 × 10−3 S cm−1 at 25 °C, 40% RH),83CPM-103a (5.8 × 10−2 S cm−1 at 22.5 °C, 98.5% RH),80Pt2(MPC)4Cl2Co(DMA)(HDMA)·guest (2.2 × 10−2 S cm−1 at 60 °C, 95% RH),61CoCa·nH2O (1 × 10−3 S cm−1 at 25 °C, 95% RH),79 and In(imdcH)(ox) (1.11 × 10−2 S cm−1 at 22.5 °C, 98.5% RH),22 the conductivity is higher than those of CoLa-II (3.05 × 10−4 S cm−1 at 25 °C, 95% RH),76Zn(H2PO4)2(TzH)2 (1.1 × 10−4 S cm−1 at 130 °C),75 and Eu2(CO3)(ox)2 (9.02 × 10−5 S cm−1 at 80 °C).77


image file: d3qm00007a-f2.tif
Fig. 2 (a) Temperature-dependent (30–80 °C) Nyquist plots for the single crystal form of IITKGP-101 at 98% RH; (b) proton conductivity vs. temperature plot for IITKGP-101 in crystal form at 98% RH.

image file: d3qm00007a-f3.tif
Fig. 3 Proton conductivity vs. water sorption isotherm correlation plots at 80 °C for IITKGP-101 with various RH values.

Besides single crystal proton conductivity measurements of IITKGP-101, the temperature and humidity-dependent proton conductivity was also investigated for the pelletized sample. As the mobility of the proton-conducting species is directly proportional to the temperature, the proton conductivity of IITKGP-101 increases with the increase of temperature. At 98% RH and ambient temperature (30 °C) the pelletized sample of IITKGP-101 displayed a proton conductivity of 3.17 × 10−7 S cm−1 (Table S6, ESI), lower than the crystal form (4.25 × 10−5 S cm−1); however with increasing temperature, the proton conductivity also increased significantly and reached the maximum of 2.46 × 10−4 S cm−1 at 80 °C and 98% RH (Fig. 4). Although this value was slightly lower than the value obtained from the proton conductivity measurement (5.22 × 10−4 S cm−1) of the crystal form under the same conditions, it still lay in the superprotonic region (≥10−4 S cm−1).3 The relatively lower proton conductivity value of the pelletized sample than the single crystal could be attributed to the random orientation of the pelletized samples during conduction measurements and grain boundary effects. Furthermore, the humidity-dependent proton conductivity was measured with the pelletized sample. From Fig. S17 (ESI), it has been observed that with the increase of R.H., the bulk resistance of the material decreases, which is reflected in the Nyquist plot. Gradually decreasing semicircles with increasing R.H. lead to the maximum conductivity of 2.46 × 10−4 S cm−1 under 98% R.H. at 80 °C (Table S6, ESI), which is comparable with those of several well-known crystalline materials (Table S8, ESI). An attempt was made to measure the proton conductivity under anhydrous conditions of activated IITKGP-101; however, the Nyquist plots showed an arc without a semicircle and the resistance was very high beyond the detection range of the instrument. Therefore the proton conductivity of IITKGP-101 is poor under anhydrous conditions (Fig. S18, ESI).


image file: d3qm00007a-f4.tif
Fig. 4 (a) Temperature-dependent (30–80 °C) Nyquist plots for the pellet form of IITKGP-101 at 98% RH (inset: enlarged portions of the high-frequency regions for clear visualization); (b) proton conductivity vs. temperature plot for IITKGP-101 in pellet form at 98% RH.

To have a better insight into the proton transportation mechanism, the activation energy (Ea) of IITKGP-101 was calculated from the perfect linear fitting of the Arrhenius plot (Fig. S19, ESI). Depending upon the activation energy (Ea), the proton conduction mechanism of the solid-state electrolytes has been classified into two types, (i) Grotthuss mechanism (Ea < 0.4 eV) in which a proton hops through the H-bonding between the hydronium ion and water. Also, the proton transfer and sequential molecular rotation occur simultaneously in the Grotthuss mechanism. (ii) Vehicular mechanism (Ea > 0.4 eV) in which the protic species (H3O+ or NH4+) diffuses by itself as a proton-attached vehicle.69 The resulting activation energy of 0.78 eV for IITKGP-101 suggests that proton transportation follows a vehicular-type mechanism in which protons transfer in the hydrated form. At higher humidity, both the coordinated and lattice water molecules interacted with the adsorbed water molecules to construct directional H-bonding networks and the dissociated protons from the coordinated water molecules passed through this network to achieve superprotonic conductivity. The adsorbed water molecules act as bridges between the donor–acceptor molecules and lead to water-mediated proton diffusion efficiently into the structure. As discussed earlier, having ample coordinated water molecules with their enhanced acidity along with high water uptake by IITKGP-101 facilitates the construction of efficient and fast proton transfer pathways, which further leads to such a high proton conductivity. After impedance measurement, the excellent retention of the PXRD patterns of the as-synthesized material confirms the framework stability throughout the proton conduction measurement (Fig. S20, ESI).

The proton conductivity value reported herein is on the order of only 10−4 S cm−1, which is considerably low compared to the recently reported MOFs and CPs; however, such a basic concept of applying the enhanced acidity of coordinated water molecules to induce proton conductivity, its wide range pH stability, hydro and boiling water stability, open air stability for as long as six months, ease of scalability through simple crystallization from green solvents like water, and conductivity measurements in both single crystal and pellet forms made this framework truly interesting and of high potential as a solid-state proton conductor.

Conclusions

In summary, taking advantage of coordinated water molecules’ enhanced acidity, we have constructed a pH-stable coordination polymer, IITKGP-101, in which coordinated water molecules act solely as proton sources, thus imparting the proton conductivity. A logical approach of design strategy and synthesis conditions could bring the robustness of the developed framework. The presence of abundant coordinated water molecules along with lattice water molecules, lattice 2,6-NDS units, and uncoordinated O atoms of the metal-bound 2,6-NDS unit in the hydrophilic interlayer spaces of the developed framework constructs an effective proton-conducting pathway responsible for exhibiting superprotonic conductivity. Both the temperature and humidity-dependent conductivity have been measured in single crystal and pellet form as well. The moderate conductivity, framework's ultrahigh robustness, and easy aqueous medium scale-up synthesis process made IITKGP-101 a promising proton conductor. We believe that the design principle presented in this work will mobilize the researchers to develop more proton-conducting MOFs and CPs having ample coordinated water molecules as intrinsic protonic sources and will be beneficial for further development of this burgeoning field.

Conflicts of interest

The authors declare no conflicts of interest.

Acknowledgements

R. S. thanks CSIR, New Delhi, India, for his Senior Research Fellowship. We acknowledge Prof. Gang Xu (Fujian Institute of Research on the Structure of Matter, CAS) for conductivity measurements. M. C. D. acknowledges the SERB, New Delhi for financial support via the Core Research Grant (CRG/2019/001034).

References

  1. W.-H. Huang, X.-M. Li, X.-F. Yang, X.-X. Zhang, H.-H. Wang and H. Wang, The recent progress and perspectives on metal- and covalent-organic framework based solid-state electrolytes for lithium-ion batteries, Mater. Chem. Front., 2021, 5, 3593–3613 RSC .
  2. A. Kirubakaran, S. Jain and R. K. Nema, A review on fuel cell technologies and power electronic interface, Renewable Sustainable Energy Rev., 2009, 13, 2430–2440 CrossRef CAS .
  3. P. Colomban, Proton Conductors: Solids, Membranes and Gels – Materials and Devices, Cambridge University Press, Cambridge, 1992 Search PubMed .
  4. J. P. Lemmon, Energy: Reimagine fuel cells, Nature, 2015, 525, 447–449 CrossRef CAS PubMed .
  5. R. Sahoo, S. Mondal, S. C. Pal, D. Mukherjee and M. C. Das, Covalent-Organic Frameworks (COFs) as Proton Conductors, Adv. Energy Mater., 2021, 11, 2102300 CrossRef CAS .
  6. S. C. Pal and M. C. Das, Superprotonic Conductivity of MOFs and Other Crystalline Platforms Beyond 10−1 S cm−1, Adv. Funct. Mater., 2021, 31, 2101584 CrossRef CAS .
  7. K. A. Mauritz and R. B. Moore, State of Understanding of Nafion, Chem. Rev., 2004, 104, 4535–4586 CrossRef CAS PubMed .
  8. D.-W. Lim and H. Kitagawa, Proton Transport in Metal–Organic Frameworks, Chem. Rev., 2020, 120, 8416–8467 CrossRef CAS PubMed .
  9. S. Chand, S. C. Pal, D.-W. Lim, K. Otsubo, A. Pal, H. Kitagawa and M. C. Das, A 2D Mg(II)-MOF with High Density of Coordinated Waters as Sole Intrinsic Proton Sources for Ultrahigh Superprotonic Conduction, ACS Mater. Lett., 2020, 2, 1343–1350 CrossRef CAS .
  10. S. C. Pal, S. Chand, A. G. Kumar, P. G. M. Mileo, I. Silverwood, G. Maurin, S. Banerjee, S. M. Elahi and M. C. Das, A Co(ii)-coordination polymer for ultrahigh superprotonic conduction: an atomistic insight through molecular simulations and QENS experiments, J. Mater. Chem. A, 2020, 8, 7847–7853 RSC .
  11. T. Yamada, K. Otsubo, R. Makiura and H. Kitagawa, Designer coordination polymers: dimensional crossover architectures and proton conduction, Chem. Soc. Rev., 2013, 42, 6655–6669 RSC .
  12. Y. He, F. Chen, B. Li, G. Qian, W. Zhou and B. Chen, Porous metal–organic frameworks for fuel storage, Coord. Chem. Rev., 2018, 373, 167–198 CrossRef CAS .
  13. M. Bazaga-García, R. M. P. Colodrero, M. Papadaki, P. Garczarek, J. Zoń, P. Olivera-Pastor, E. R. Losilla, L. León-Reina, M. A. G. Aranda, D. Choquesillo-Lazarte, K. D. Demadis and A. Cabeza, Guest Molecule-Responsive Functional Calcium Phosphonate Frameworks for Tuned Proton Conductivity, J. Am. Chem. Soc., 2014, 136, 5731–5739 CrossRef PubMed .
  14. H. Gao, Y.-B. He, J.-J. Hou and X.-M. Zhang, In Situ Aliovalent Nickle Substitution and Acidic Modification of Nanowalls Promoted Proton Conductivity in InOF with 1D Helical Channel, ACS Appl. Mater. Interfaces, 2021, 13, 38289–38295 CrossRef CAS PubMed .
  15. J.-P. Zhang, Y.-B. Zhang, J.-B. Lin and X.-M. Chen, Metal Azolate Frameworks: From Crystal Engineering to Functional Materials, Chem. Rev., 2012, 112, 1001–1033 CrossRef CAS PubMed .
  16. S. M. Elahi, S. Chand, W.-H. Deng, A. Pal and M. C. Das, Polycarboxylate-Templated Coordination Polymers: Role of Templates for Superprotonic Conductivities of up to 10−1 S cm−1, Angew. Chem., Int. Ed., 2018, 57, 6662–6666 CrossRef CAS PubMed .
  17. A. Pal, S. C. Pal, K. Otsubo, D.-W. Lim, S. Chand, H. Kitagawa and M. C. Das, A Phosphate-Based Silver–Bipyridine 1D Coordination Polymer with Crystallized Phosphoric Acid as Superprotonic Conductor, Chem. – Eur. J., 2020, 26, 4607–4612 CrossRef CAS PubMed .
  18. M. Y. B. Zulkifli, R. Lin, M. Chai, V. Chen and J. Hou, Transport tuning strategies in MOF film synthesis – a perspective, J. Mater. Chem. A, 2022, 10, 14641–14654 RSC .
  19. J. Guo, X. Xue, H. Yu, Y. Duan, F. Li, Y. Lian, Y. Liu and M. Zhao, Metal–organic frameworks based on infinite secondary building units: recent progress and future outlooks, J. Mater. Chem. A, 2022, 10, 19320–19347 RSC .
  20. L. Feng, K.-Y. Wang, J. Willman and H.-C. Zhou, Hierarchy in Metal–Organic Frameworks, ACS Cent. Sci., 2020, 6, 359–367 CrossRef CAS PubMed .
  21. S. Wang, M. Wahiduzzaman, L. Davis, A. Tissot, W. Shepard, J. Marrot, C. Martineau-Corcos, D. Hamdane, G. Maurin, S. Devautour-Vinot and C. Serre, A robust zirconium amino acid metal-organic framework for proton conduction, Nat. Commun., 2018, 9, 4937 CrossRef PubMed .
  22. X. Zhao, C. Mao, X. Bu and P. Feng, Direct Observation of Two Types of Proton Conduction Tunnels Coexisting in a New Porous Indium–Organic Framework, Chem. Mater., 2014, 26, 2492–2495 CrossRef CAS .
  23. H. Wu, F. Yang, X.-L. Lv, B. Wang, Y.-Z. Zhang, M.-J. Zhao and J.-R. Li, A stable porphyrinic metal–organic framework pore-functionalized by high-density carboxylic groups for proton conduction, J. Mater. Chem. A, 2017, 5, 14525–14529 RSC .
  24. Z. Li, Z. Zhang, Y. Ye, K. Cai, F. Du, H. Zeng, J. Tao, Q. Lin, Y. Zheng and S. Xiang, Rationally tuning host–guest interactions to free hydroxide ions within intertrimerically cuprophilic metal–organic frameworks for high OH− conductivity, J. Mater. Chem. A, 2017, 5, 7816–7824 RSC .
  25. S. Natarajan and P. Mahata, Metal–organic framework structures – how closely are they related to classical inorganic structures?, Chem. Soc. Rev., 2009, 38, 2304–2318 RSC .
  26. R. Sahoo and M. C. Das, C2s/C1 hydrocarbon separation: The major step towards natural gas purification by metal-organic frameworks (MOFs), Coord. Chem. Rev., 2021, 442, 213998 CrossRef CAS .
  27. Y. Ye, L. Gong, S. Xiang, Z. Zhang and B. Chen, Metal–Organic Frameworks as a Versatile Platform for Proton Conductors, Adv. Mater., 2020, 32, 1907090 CrossRef CAS PubMed .
  28. R. Sahoo, S. Mondal, D. Mukherjee and M. C. Das, Metal–Organic Frameworks for CO2 Separation from Flue and Biogas Mixtures, Adv. Funct. Mater., 2022, 32, 2207197 CrossRef CAS .
  29. X. Han, S. Yang and M. Schröder, Porous metal–organic frameworks as emerging sorbents for clean air, Nat. Rev. Chem., 2019, 3, 108–118 CrossRef CAS .
  30. N. Stock and S. Biswas, Synthesis of Metal-Organic Frameworks (MOFs): Routes to Various MOF Topologies, Morphologies, and Composites, Chem. Rev., 2012, 112, 933–969 CrossRef CAS PubMed .
  31. Ü. Kökçam-Demir, A. Goldman, L. Esrafili, M. Gharib, A. Morsali, O. Weingart and C. Janiak, Coordinatively unsaturated metal sites (open metal sites) in metal–organic frameworks: design and applications, Chem. Soc. Rev., 2020, 49, 2751–2798 RSC .
  32. S. S. Nagarkar, S. M. Unni, A. Sharma, S. Kurungot and S. K. Ghosh, Two-in-One: Inherent Anhydrous and Water-Assisted High Proton Conduction in a 3D Metal–Organic Framework, Angew. Chem., Int. Ed., 2014, 53, 2638–2642 CrossRef CAS PubMed .
  33. O. Basu, S. Mukhopadhyay, S. Laha and S. K. Das, Defect Engineering in a Metal–Organic Framework System to Achieve Super-Protonic Conductivity, Chem. Mater., 2022, 34, 6734–6743 CrossRef CAS .
  34. E. Pardo, C. Train, G. Gontard, K. Boubekeur, O. Fabelo, H. Liu, B. Dkhil, F. Lloret, K. Nakagawa, H. Tokoro, S.-I. Ohkoshi and M. Verdaguer, High Proton Conduction in a Chiral Ferromagnetic Metal–Organic Quartz-like Framework, J. Am. Chem. Soc., 2011, 133, 15328–15331 CrossRef CAS PubMed .
  35. T. N. Tu, N. Q. Phan, T. T. Vu, H. L. Nguyen, K. E. Cordova and H. Furukawa, High proton conductivity at low relative humidity in an anionic Fe-based metal–organic framework, J. Mater. Chem. A, 2016, 4, 3638–3641 RSC .
  36. D. Gui, X. Dai, Z. Tao, T. Zheng, X. Wang, M. A. Silver, J. Shu, L. Chen, Y. Wang, T. Zhang, J. Xie, L. Zou, Y. Xia, J. Zhang, J. Zhang, L. Zhao, J. Diwu, R. Zhou, Z. Chai and S. Wang, Unique Proton Transportation Pathway in a Robust Inorganic Coordination Polymer Leading to Intrinsically High and Sustainable Anhydrous Proton Conductivity, J. Am. Chem. Soc., 2018, 140, 6146–6155 CrossRef CAS PubMed .
  37. V. G. Ponomareva, K. A. Kovalenko, A. P. Chupakhin, D. N. Dybtsev, E. S. Shutova and V. P. Fedin, Imparting High Proton Conductivity to a Metal–Organic Framework Material by Controlled Acid Impregnation, J. Am. Chem. Soc., 2012, 134, 15640–15643 CrossRef CAS PubMed .
  38. Q.-x Luo, B.-w An, M. Ji and J. Zhang, Hybridization of metal–organic frameworks and task-specific ionic liquids: fundamentals and challenges, Mater. Chem. Front., 2018, 2, 219–234 RSC .
  39. J. H. Choe, D. W. Kang, M. Kang, H. Kim, J. R. Park, D. W. Kim and C. S. Hong, Revealing an unusual temperature-dependent CO2 adsorption trend and selective CO2 uptake over water vapors in a polyamine-appended metal–organic framework, Mater. Chem. Front., 2019, 3, 2759–2767 RSC .
  40. W. Lu, Z. Wei, Z.-Y. Gu, T.-F. Liu, J. Park, J. Tian, M. Zhang, Q. Zhang, T. Gentle Iii, M. Bosch and H.-C. Zhou, Tuning the structure and function of metal–organic frameworks via linker design, Chem. Soc. Rev., 2014, 43, 5561–5593 RSC .
  41. X. Zhang, Z. Chen, X. Liu, S. L. Hanna, X. Wang, R. Taheri-Ledari, A. Maleki, P. Li and O. K. Farha, A historical overview of the activation and porosity of metal–organic frameworks, Chem. Soc. Rev., 2020, 49, 7406–7427 RSC .
  42. N. Li, R. Feng, J. Zhu, Z. Chang and X.-H. Bu, Conformation versatility of ligands in coordination polymers: From structural diversity to properties and applications, Coord. Chem. Rev., 2018, 375, 558–586 CrossRef CAS .
  43. Z. Chen, H. Jiang, M. Li, M. O’Keeffe and M. Eddaoudi, Reticular Chemistry 3.2: Typical Minimal Edge-Transitive Derived and Related Nets for the Design and Synthesis of Metal–Organic Frameworks, Chem. Rev., 2020, 120, 8039–8065 CrossRef CAS PubMed .
  44. S. B. Peh, Y. Wang and D. Zhao, Scalable and Sustainable Synthesis of Advanced Porous Materials, ACS Sustainable Chem. Eng., 2019, 7, 3647–3670 CrossRef CAS .
  45. A. Karmakar, E. Velasco and J. Li, Metal-organic frameworks as effective sensors and scavengers for toxic environmental pollutants, Natl. Sci. Rev., 2022, 9 Search PubMed .
  46. B. Li, M. Chrzanowski, Y. Zhang and S. Ma, Applications of metal-organic frameworks featuring multi-functional sites, Coord. Chem. Rev., 2016, 307, 106–129 CrossRef CAS .
  47. E.-S. M. El-Sayed and D. Yuan, Waste to MOFs: sustainable linker, metal, and solvent sources for value-added MOF synthesis and applications, Green Chem., 2020, 22, 4082–4104 RSC .
  48. R. Sahoo, S. Ghosh, S. Chand, S. Chand Pal, T. Kuila and M. C. Das, Highly scalable and pH stable 2D Ni-MOF-based composites for high performance supercapacitor, Composites, Part B, 2022, 245, 110174 CrossRef CAS .
  49. R. Sahoo, S. Chand, M. Mondal, A. Pal, S. C. Pal, M. K. Rana and M. C. Das, A “Thermodynamically Stable” 2D Nickel Metal–Organic Framework over a Wide pH Range with Scalable Preparation for Efficient C2s over C1 Hydrocarbon Separations, Chem. Eur. J., 2020, 26, 12624–12631 CrossRef CAS PubMed .
  50. H.-G. Zhou, R.-Q. Xia, J. Zheng, D. Yuan, G.-H. Ning and D. Li, Acid-triggered interlayer sliding of two-dimensional copper(i)–organic frameworks: more metal sites for catalysis, Chem. Sci., 2021, 12, 6280–6286 RSC .
  51. M. Gupta and J. J. Vittal, Control of interpenetration and structural transformations in the interpenetrated MOFs, Coord. Chem. Rev., 2021, 435, 213789 CrossRef CAS .
  52. S. S. Park, A. J. Rieth, C. H. Hendon and M. Dincă, Selective Vapor Pressure Dependent Proton Transport in a Metal–Organic Framework with Two Distinct Hydrophilic Pores, J. Am. Chem. Soc., 2018, 140, 2016–2019 CrossRef CAS PubMed .
  53. M. Liu, L. Chen, S. Lewis, S. Y. Chong, M. A. Little, T. Hasell, I. M. Aldous, C. M. Brown, M. W. Smith, C. A. Morrison, L. J. Hardwick and A. I. Cooper, Three-dimensional protonic conductivity in porous organic cage solids, Nat. Commun., 2016, 7, 12750 CrossRef CAS PubMed .
  54. W.-L. Xue, W.-H. Deng, H. Chen, R.-H. Liu, J. M. Taylor, Y.-K. Li, L. Wang, Y.-H. Deng, W.-H. Li, Y.-Y. Wen, G.-E. Wang, C.-Q. Wan and G. Xu, MOF-Directed Synthesis of Crystalline Ionic Liquids with Enhanced Proton Conduction, Angew. Chem., Int. Ed., 2021, 60, 1290–1297 CrossRef CAS PubMed .
  55. H. Furukawa, K. E. Cordova, M. O’Keeffe and O. M. Yaghi, The Chemistry and Applications of Metal-Organic Frameworks, Science, 2013, 341, 1230444 CrossRef PubMed .
  56. S. M. G. E. Cohen and N. L. G. E. Rosi, Postsynthetic Modification of Metal–Organic Frameworks, Inorg. Chem., 2021, 60, 11703–11705 CrossRef PubMed .
  57. F. Afshariazar, A. Morsali, S. Sorbara, N. M. Padial, E. Roldan-Molina, J. E. Oltra, V. Colombo and J. A. R. Navarro, Impact of Pore Size and Defects on the Selective Adsorption of Acetylene in Alkyne-Functionalized Nickel(II)-Pyrazolate-Based MOFs, Chem. Eur. J., 2021, 27, 11837–11844 CrossRef CAS PubMed .
  58. T. T. M. Nguyen, H. M. Le, Y. Kawazoe and H. L. Nguyen, Reticular control of interpenetration in a complex metal–organic framework, Mater. Chem. Front., 2018, 2, 2063–2069 RSC .
  59. Y.-Y. Wang, X.-Y. Ji, M. Yu and J. Tao, Confining lead-free perovskite quantum dots in metal–organic frameworks for visible light-driven proton reduction, Mater. Chem. Front., 2021, 5, 7796–7807 RSC .
  60. Z.-H. Li, H. Zeng, G. Zeng, C. Ru, G. Li, W. Yan, Z. Shi and S. Feng, Multivariate Synergistic Flexible Metal-Organic Frameworks with Superproton Conductivity for Direct Methanol Fuel Cells, Angew. Chem., Int. Ed., 2021, 60, 26577–26581 CrossRef CAS PubMed .
  61. K. Otsubo, S. Nagayama, S. Kawaguchi, K. Sugimoto and H. Kitagawa, A Preinstalled Protic Cation as a Switch for Superprotonic Conduction in a Metal–Organic Framework, JACS Au, 2022, 2, 109–115 CrossRef CAS PubMed .
  62. S. Chand, S. M. Elahi, A. Pal and M. C. Das, Metal–Organic Frameworks and Other Crystalline Materials for Ultrahigh Superprotonic Conductivities of 10−2 S cm−1 or Higher, Chem. Eur. J., 2019, 25, 6259–6269 CrossRef CAS PubMed .
  63. R. Sahoo, S. C. Pal and M. C. Das, Solid-State Proton Conduction Driven by Coordinated Water Molecules in Metal–Organic Frameworks and Coordination Polymers, ACS Energy Lett., 2022, 7, 4490–4500 CrossRef CAS .
  64. S.-S. Bao, G. K. H. Shimizu and L.-M. Zheng, Proton conductive metal phosphonate frameworks, Coord. Chem. Rev., 2019, 378, 577–594 CrossRef CAS .
  65. X.-X. Xie, Y.-C. Yang, B.-H. Dou, Z.-F. Li and G. Li, Proton conductive carboxylate-based metal–organic frameworks, Coord. Chem. Rev., 2020, 403, 213100 CrossRef CAS .
  66. D.-W. Lim and H. Kitagawa, Rational strategies for proton-conductive metal–organic frameworks, Chem. Soc. Rev., 2021, 50, 6349–6368 RSC .
  67. W.-H. Li, W.-H. Deng, G.-E. Wang and G. Xu, Conductive MOFs, EnergyChem, 2020, 2, 100029 CrossRef .
  68. G. K. H. Shimizu, J. M. Taylor and S. Kim, Proton Conduction with Metal-Organic Frameworks, Science, 2013, 341, 354–355 CrossRef CAS PubMed .
  69. D.-W. Lim, M. Sadakiyo and H. Kitagawa, Proton transfer in hydrogen-bonded degenerate systems of water and ammonia in metal–organic frameworks, Chem. Sci., 2019, 10, 16–33 RSC .
  70. F. Xiang, S. Chen, Z. Yuan, L. Li, Z. Fan, Z. Yao, C. Liu, S. Xiang and Z. Zhang, Switched Proton Conduction in Metal–Organic Frameworks, JACS Au, 2022, 2, 1043–1053 CrossRef CAS PubMed .
  71. X. Meng, H.-N. Wang, S.-Y. Song and H.-J. Zhang, Proton-conducting crystalline porous materials, Chem. Soc. Rev., 2017, 46, 464–480 RSC .
  72. P. Ramaswamy, N. E. Wong and G. K. H. Shimizu, MOFs as proton conductors – challenges and opportunities, Chem. Soc. Rev., 2014, 43, 5913–5932 RSC .
  73. D. W. Kang, M. Kang and C. S. Hong, Post-synthetic modification of porous materials: superprotonic conductivities and membrane applications in fuel cells, J. Mater. Chem. A, 2020, 8, 7474–7494 RSC .
  74. H. Bunzen, A. Javed, D. Klawinski, A. Lamp, M. Grzywa, A. Kalytta-Mewes, M. Tiemann, H.-A. K. von Nidda, T. Wagner and D. Volkmer, Anisotropic Water-Mediated Proton Conductivity in Large Iron(II) Metal–Organic Framework Single Crystals for Proton-Exchange Membrane Fuel Cells, ACS Appl. Nano Mater., 2019, 2, 291–298 CrossRef CAS .
  75. D. Umeyama, S. Horike, M. Inukai, T. Itakura and S. Kitagawa, Inherent Proton Conduction in a 2D Coordination Framework, J. Am. Chem. Soc., 2012, 134, 12780–12785 CrossRef CAS PubMed .
  76. S.-S. Bao, K. Otsubo, J. M. Taylor, Z. Jiang, L.-M. Zheng and H. Kitagawa, Enhancing Proton Conduction in 2D Co–La Coordination Frameworks by Solid-State Phase Transition, J. Am. Chem. Soc., 2014, 136, 9292–9295 CrossRef CAS PubMed .
  77. Q. Tang, Y. Liu, S. Liu, D. He, J. Miao, X. Wang, G. Yang, Z. Shi and Z. Zheng, High Proton Conduction at above 100 °C Mediated by Hydrogen Bonding in a Lanthanide Metal–Organic Framework, J. Am. Chem. Soc., 2014, 136, 12444–12449 CrossRef CAS PubMed .
  78. S. Tominaka and A. K. Cheetham, Intrinsic and extrinsic proton conductivity in metal-organic frameworks, RSC Adv., 2014, 4, 54382–54387 RSC .
  79. S.-S. Bao, N.-Z. Li, J. M. Taylor, Y. Shen, H. Kitagawa and L.-M. Zheng, Co–Ca Phosphonate Showing Humidity-Sensitive Single Crystal to Single Crystal Structural Transformation and Tunable Proton Conduction Properties, Chem. Mater., 2015, 27, 8116–8125 CrossRef CAS .
  80. Q.-G. Zhai, C. Mao, X. Zhao, Q. Lin, F. Bu, X. Chen, X. Bu and P. Feng, Cooperative Crystallization of Heterometallic Indium–Chromium Metal–Organic Polyhedra and Their Fast Proton Conductivity, Angew. Chem., Int. Ed., 2015, 54, 7886 CrossRef CAS PubMed .
  81. L. Qin, Y.-Z. Yu, P.-Q. Liao, W. Xue, Z. Zheng, X.-M. Chen and Y.-Z. Zheng, A “Molecular Water Pipe”: A Giant Tubular Cluster {Dy72} Exhibits Fast Proton Transport and Slow Magnetic Relaxation, Adv. Mater., 2016, 28, 10772–10779 CrossRef CAS PubMed .
  82. R. Li, S.-H. Wang, X.-X. Chen, J. Lu, Z.-H. Fu, Y. Li, G. Xu, F.-K. Zheng and G.-C. Guo, Highly Anisotropic and Water Molecule-Dependent Proton Conductivity in a 2D Homochiral Copper(II) Metal–Organic Framework, Chem. Mater., 2017, 29, 2321–2331 CrossRef CAS .
  83. B. Joarder, J.-B. Lin, Z. Romero and G. K. H. Shimizu, Single Crystal Proton Conduction Study of a Metal Organic Framework of Modest Water Stability, J. Am. Chem. Soc., 2017, 139, 7176–7179 CrossRef CAS PubMed .
  84. S. Hwang, E. J. Lee, D. Song and N. C. Jeong, High Proton Mobility with High Directionality in Isolated Channels of MOF-74, ACS Appl. Mater. Interfaces, 2018, 10, 35354–35360 CrossRef CAS PubMed .

Footnotes

Electronic supplementary information (ESI) available: The Electronic supplementary information contains instrument details, synthesis details, table of crystallographic data, PXRD patterns, TGA profiles, DSC patterns, calculation details, proton conduction results, and proton conduction comparison tables. CCDC 2220922. For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d3qm00007a
These authors contributed equally to this work.

This journal is © the Partner Organisations 2023