Xuan
Zhang
ab,
Kai
Wan
b,
Palaniappan
Subramanian
b,
Maowen
Xu
a,
Jiangshui
Luo
*cd and
Jan
Fransaer
*b
aKey Laboratory of Luminescent and Real-Time Analytical Chemistry (Southwest University), Ministry of Education, School of Materials and Energy, Southwest University, Chongqing 400715, PR China. E-mail: xumaowen@swu.edu.cn
bDepartment of Materials Engineering, KU Leuven, Leuven 3001, Belgium. E-mail: jan.fransaer@kuleuven.be
cCollege of Materials Science and Engineering, Sichuan University, 610065, Chengdu, China
dLaboratory for Soft Matter and Biophysics, Department of Physics and Astronomy, KU Leuven, Leuven, 3001, Belgium. E-mail: jiangshui.luo@kuleuven.be
First published on 30th March 2020
Metal–organic framework (MOF) thin films have received increasing attention for many applications, such as chemical sensors and membranes. Several techniques have been developed for the deposition of MOF films. In particular, the processing of electrochemical deposition of MOFs has only recently been initiated but stands out from other methods due to the mild preparation conditions, monitoring of continuous processes, precisely controllable synthesis parameters and potential for large-scale production. Recently, interest in the electrochemical deposition of MOF films has started to expand from their preparation to applications. In this review, we summarize and critically assess the state of the mechanism, the influence of parameters on electrochemical deposition of MOFs and their corresponding applications in different areas. Moreover, the strengths and shortcomings of different electrochemical deposition methods and their suitable scopes are discussed. Finally, the urgent challenges and future opportunities of electrochemical deposition of MOFs are highlighted.
Compared with the above methods, electrochemical methods are considered as some of the most promising methods to prepare MOF films inspired by the first patent proposed by BASF.20 On the one hand, electrochemical synthesis allows the large-scale production of MOFs in powder form, while on the other, it is also an effective method to synthesize thin films and coatings. Some of the salient features of the electrochemical deposition method are (i) there is no pre-treatment required except for simple cleaning of the substrate, (ii) the possibility to operate under milder conditions with short synthesis time, (iii) the possibility to monitor the process in real time and continuously which is useful for industrial scale operations, and (iv) the self-closing ability of electrochemical deposition ensures high throwing power and fewer cracks in the film. More importantly, the parameters during electrochemical deposition can be easily and precisely controlled. Combined with the designable characteristics of MOFs, the highly controllable electrochemical deposition procedure is a promising strategy for tailor-making MOF films for desired applications. The salient features and advantages outlined above have evoked interest in electrodeposited MOF films for different applications in recent years as shown in Fig. 1.
Although some reviews related to the electrochemical synthesis of MOFs have been published, they are mostly focused on the synthesis and only a few targeted applications and related discussions are reviewed.21,22 In the last few years, progress was made not only in the electrochemical deposition of MOFs, but also in various device fabrications using electrodeposited MOF films. The research on electrochemical deposition of MOFs has started to move from synthesis towards applications (see Table 1). For different applications, the requirements and challenges for MOF films are not the same. Therefore, this review focuses on three aspects: (1) the latest progress in the mechanism and the influence of parameters on MOF films associated with three types of electrochemical deposition of MOFs, (2) key developments in the applications of electrochemically deposited MOF films, and (3) the strengths and shortcomings of these three types of electrochemical deposition methods for different applications.
Sample | Method | Application |
---|---|---|
Cu-BTC | AED | Synthesis32,34,36,38,40,44,47–50 |
Cu-BTC | AED | Sensor31,33,35 |
Cu-BTC | AED | Catalysis57 |
Cu-BTC | AED | Separation37,102 |
Cu-BTC | AED | Template39,117 |
Cu-BTC | CED | Sensor84,87–89 |
Cu-BTC | EPD | Synthesis67,72,91 |
Cu-BTC | EPD | Sensor92 |
Cu-BDC, Cu-BTEC | AED | Synthesis91 |
Cu-BTEC | CED | Supercapacitor106 |
Cu(CHDA) | AED | Synthesis44 |
Cu(INA)2 | AED | Synthesis44,50 |
Cu-TDPAT/CCQDs | CED | Sensor90 |
Cu-TCA | AED | Sensor78 |
Co/Ni-BTC-DMF | AED | Template116 |
Co-BDC-DABCO | AED | Template120 |
Co-CA | CED | Template62 |
Ni-BTC, Fe-BTC, Fe/Ni-BTC | CED | Catalysis96,97 |
Ni-BDC-NH3/CCQDs | CED | Sensor89 |
Zn-BTC, Eu-HBPTC | AED | Sensor77 |
Zn-BTC | EPD | Catalysis68 |
Mn/Mn-BTC | AED | Template118 |
Al-MIL-53, NU-1000 | EPD | Synthesis67 |
Eu-TDC, Tb-BDC | CED | Sensor64 |
Eu-HBPTC, Tb-HBPTC, Gb-HBPTC | CED | Sensor65 |
Tb-BTC, Eu-BTC, Tb/Eu-BTC | EPD | Sensor81 |
UiO-66 | AED | Synthesis43 |
UiO-66 | EPD | Synthesis67 |
UiO-66/Ln, UiO-66-(COOH)2, UiO-66-hybrid | EPD | Sensor82 |
UiO-66/C-QDs | EPD | Sensor83 |
Ir–Zne | AED | Sensor76 |
MIL-100 | AED | Separation53 |
MIL-100 | AED | Supercapacitor105 |
MOF-5 | CED | Synthesis55,56,58,59 |
MOF-525 | EPD | Catalysis95 |
NENU-3 | AED | Sensor35 |
ZIF-4, ZIF-7, ZIF-8, ZIF-14, ZIF-67, ZIF-8 | AED | Supercapacitor107 |
ZIF-67 | AED | Template119 |
ZIF-8, ZIF-67, ZIF-71 | CED | Synthesis61 |
ZIF-8 | CED | Separation103 |
ZIF-7, ZIF-8 | EPD | Separation104 |
The crystal size and thickness of the MOF films could be varied with the applied voltage (related to the concentration of metal ions near the surface). Also, a higher concentration of water in the synthesis mixture led to larger MOF crystals. In subsequent work, the AED conditions were studied for this archetypal MOF (Cu-BTC).32–40 The charge required to oxidize the anode can be supplied in both amperometric (constant current) and potentiometric (constant voltage) modes. According to nucleation theory, higher current densities lead to smaller particle size. As the reaction proceeds, the average crystal size and thickness increase with increasing deposition time. In the end, the films are one to a few crystals thick. During prolonged growth, however, MOF films detach from the substrate, due to the anodic dissolution of the underlying substrate. The organic electrolyte often has low conductivity with a low energetic efficiency, which can be improved by increasing the temperature or through the addition of conductive salts. MTBS is the most common conductive salt for the electrochemical deposition of MOFs. Tetrabutylammonium tetrafluoroborate (TBATFB),27,29 LiClO4 (ref. 41) and NaNO3 (ref. 42) were also used as conductive salt in some cases. It is worth mentioning that these conductive salts can get trapped in the pores of MOFs, leading to a decrease of the accessible surface area. Interestingly, it was found that TBATFB is not trapped inside the pores during anodic electrodeposition in spite of its large size.29 In some studies, other kinds of additives (for example, modulators to control the crystal size and morphology) were added to the electrolyte, which affected the formation of intermediates or led to competition between the additives and the ligand towards the metal ions.41,43 In the case of AED, the temperature exerts a significant influence on the solubility of linkers, conductivity of the electrolyte, kinetics of MOF formation and morphology of the resulting MOFs. Some of these factors (pH, deprotonation of linkers, hydration of metal ions, the solubility of linkers and conductivity of electrolyte…) can be tuned by varying the solvent composition.31,34,36,38,44
Bearing all this in mind, a comprehensive study of the mechanism of MOF crystal nucleation and growth during the anodic electrodeposition process was undertaken by researchers at KU Leuven.45 As shown in Fig. 2A, the AED of MOF films consists of four phases:
Fig. 2 (A) Proposed mechanism of anodic electrodeposition of MOFs and SEM pictures, taken at a 75° angle with the normal, of the four phases: (I) initial nucleation (a), (II) growth of islands (b), (III) intergrowth (c), and (IV) detachment (d). Copper-coated wafer substrate, 2 V vs. counter electrode, after 10 s, 10 min, 60 min and 125 min. Reproduced with permission from ref. 45. Copyright 2016, the Royal Society of Chemistry. (B) Mechanism of cathodic electrodeposition of crystalline MOFs. Cathodic electrodeposition of MOFs relies on the accumulation of OH− ions. Reproduced with permission from ref. 55. Copyright 2011, American Chemical Society. (C) Schematic illustration of EPD of MOFs, showing the attraction of charged MOF particles toward an oppositely charged electrode using an applied electric field and illustration of the patterning of the MOF EPD film procedure: (I) spin coating a layer of photoresist on a bare FTO substrate, (II) using photolithography to create patterned squares of photoresist, (III) deposition of MOF particles on the conductive exposed FTO areas using EPD, and (IV) removal of the remaining photoresist squares to obtain the desired pattern. Reproduced with permission from ref. 67. Copyright 2014, Wiley-VCH. |
(i) Initial nucleation: a critical concentration of metal ions forms upon anodic dissolution of a metal substrate. This stage can be clearly observed by detecting the mass change using EQCM. Once MOF crystals nucleate on the substrate, no new nucleation is needed for the deposition to proceed.
(ii) Growth of islands: the small MOF nuclei progressively grow into big crystals within minutes. In time, the size distribution and the average size of MOF crystals become broader and larger, respectively.
(iii) Intergrowth: there are two steps in this phase. The self-closing behavior of the electrochemical synthesis of MOF film can be understood in part by the different contributions of the applied potential (E):
E = E0 + ηactivation + ηohmic + ηconc |
(iv) Detachment: as the linkers cannot migrate through the MOF layer but copper ions can, the MOF layer grows at the MOF–solution interface. This implies that after a certain time, the copper underneath the MOF at the metal–MOF interface has dissolved so much that the MOF crystals detach from the substrate. Very recently, this four-step process was strongly supported by Dryfe et al. via in situ electrochemical atomic force microscopy.46
For the application of MOF films, the electrode structure and the mechanical properties of the MOF films are also critical. The most commonly reported MOFs show porosities restricted to the microporous regime (<2 nm), which makes the applications of MOF films suffer from diffusion limitations.47 An important approach to reduce this diffusion limitation is the introduction of mesopores (2–50 nm) and macropores (>50 nm) into the MOF films that can shorten the diffusion paths and enlarge the effective surface area by enhanced mass and ion transport. Researchers at KU Leuven prepared a 3D hierarchical structure MOF coated electrode by combining a dealloying procedure with anodic electrodeposition.48 A Ag–Cu alloy was used as an anode for preparing Cu-BTC. After application of a suitable voltage, the “noble” adatoms (i.e. Ag) are left behind and form a porous matrix structure during the dissolution of Cu, which forms MOFs inside the porous matrix. This 3D hierarchical structure MOF coated matrix electrode showed much higher surface area (7000 m2 m−2) than the dealloyed structure without a linker (470 m2 m−2) or pure copper plate derived MOF layers (2433 m2 m−2). This 3D hierarchical structure not only increased the surface area of the electrode but also reduced the risk of detachment of the MOF film. In order to achieve a similar goal, Kuhn et al. used colloidal crystals of silica spheres as a template to prepare hierarchical macro-/microporous Cu-BTC composite electrodes.49 Firstly, silica spheres were coated on a Au plate by a Langmuir–Blodgett (LB) technique. Thereafter, Cu was deposited on the surface of the silica sphere covered electrode by electrodeposition. After removal of silica spheres, an ordered macroporous Cu scaffold with open pores was obtained. Finally, dissolution of this porous Cu scaffold by electrochemical anodic oxidation in a linker-containing electrolyte created the desired MOF structure. In an attempt to further downsize MOF structures, Wouter et al. established a method to synthesize nanowires and three-dimensionally interconnected nanowire networks of Cu-MOFs, including both Cu(isonicotinate)2[C12H8O4N2Cu, Cu(INA)2] and Cu-BTC, by a combination of ion-track technology and electrochemistry.50 They first prepared porous polycarbonate membranes via ion-track technology. Then metallic Cu was electrodeposited inside these porous polycarbonate membranes. In a last step, the metallic Cu nano-wires were anodically dissolved in a linker-containing solution, forming the desired MOFs with the structure of nanowires and nanowire networks. Similarly, using the two-step method ((I) deposition of metal and (II) anodic dissolution), MOFs can be anodically deposited on different kinds of substrates without the limitation of the corresponding metal substrates.51,52
The mechanical properties of MOF films can strongly affect their stability in device structures. Through nanoindentation and nanoscratch experiments, Tan and co-workers found that the particle size and the type of linker exert a large influence on the mechanical properties of Cu-based MOFs.44 Among the three kinds of Cu-based MOFs {Cu(trans-cyclohexane-1,4-dicarboxylate)[C8H10O4Cu, Cu(CHDA)], Cu(INA)2 and Cu-BTC}, Cu(CHDA) exhibits the best mechanical performance with the highest stiffness (10.9 GPa) and hardness (0.46 GPa). In their more recent work, they modified the mechanical properties of Cu-BTC films by tuning the parameters of the anodic electrodeposition process. Not only does the reaction time (affecting the etching of the substrate) have an influence on the mechanical properties of the MOF film, the roughness of the substrate also affects the adhesion strength (the change of lift-off force).32 Based on this, the group at KU Leuven compared the adhesion strength of different Zr6O4(OH)4(1,4-benzenedicarboxylate)6 (C48H28O32Zr6, UiO-66) films, prepared by anodic or cathodic electrodeposition.43 Significantly enhanced adhesion strength of UiO-66 film could be achieved by anodic electrodeposition, caused by a metal oxide layer bridging the MOF layer and substrate.
In addition, anodic electrochemical synthesis offers additional degrees of freedom in the synthesis of MOFs, which allows it to be combined with other methods. The synergistic effect of two methods further facilitates the industrial production of MOFs and broadens the possibility of MOF film synthesis. A high-temperature high-pressure (HT-HP) electrochemical cell was used for combining the solvothermal method with anodic electrodeposition to prepare crystalline Fe3O(H2O)2OH(benzene-1,3,5-tricarboxylate)2 (C18H8O16Fe3, MIL-100) films,53 which is hard to synthesize at low temperatures. The accessible synthesis temperature range can be enlarged by this method. Therefore, they used this cell to explore the temperature effect on the morphology of the well-studied Cu-BTC in their work as well. Cu-BTC crystals with an unusual cubic morphology deposited on the copper plate at 200 °C. Another benefit of this method is that the conductive salt, nitrates or acids can be avoided in the electrolyte due to the increase of conductivity and higher solubility of linkers. In addition, a new synergic strategy combining sonochemistry and electrodeposition for the synthesis of MOF powders was reported by Júnior et al.54
NO3− + H2O + 2e− → NO2− + 2OH− | (1) |
H2BDC + OH− → HBDC+ + H2O | (2) |
In CED, the applied potential plays a critical role in the formation of MOF films. For example, metallic Zn deposits together with the MOF film, because the applied potential is sufficiently negative in this process. In their follow-up studies,56 in order to avoid the electrodeposition of Zn, the reduction of triethylammonium (Et3NH+, see eqn (3)) which has a more positive reduction potential than the reduction of NO3− was chosen as an alternative to increase the local pH at the electrode surface. The product is trimethylamine and H2. H2 is a relatively inert molecule that has less influence on the formation of MOFs.
2R3NH+ (R = alkyl) + 2e− → 2R3N + H2 | (3) |
R3N + H2O ⇌ R3NH+ + OH− | (4) |
Meanwhile, although the substrate has no relation to the metal source during CED, it still has a large influence on the reduction potential and the affinity between the MOF particles and the substrate, which is related to the formation and adhesion of MOF films.56,57 Therefore, to minimize the overpotential for the reduction of Et3NH+, Pt was chosen as a working electrode. In a solution of Et3NHCl and H2BDC in DMF, the onset of proton reduction was at around −0.5 V [vs. Ag/Ag(cryptand)+] with a peak current of −1.0 V [vs. Ag/Ag(cryptand)+]. At this potential, no electrodeposition of Zn occurs. More interestingly, the MOF phase can be controlled by varying the concentration of Et3NH+ and the applied potential. At a concentration of 300 mM Et3NH+, the only phase that is formed is (Et3NH)2Zn3(BDC)4. When the concentration of Et3NH+ was reduced to 100 mM, exclusively (Et3NH)2Zn3(BDC)4 deposited at −1.10 V and a mixture of (Et3NH)2Zn3(BDC)4 and MOF-5 was obtained at −1.50 V. Based on this, either mixed films or sandwich-type bilayer surface structures could be deposited. Following these studies, Li and Dincă focused on studying the different CED parameters of MOF-5.58 They found that Cl− inhibits the formation of MOF-5. Also, water is not required initially, as it is generated during the process. In depositions conducted without nitrate-containing Zn2+ precursors, Zn2(BDC)(OH)2 would be formed. Based on these results, they proposed a possible transformation scheme shown in Fig. 3. Similar to the solvothermal method,59 the MOF-5 formation is accelerated by nitrate, as it leads to the formation of an important intermediate of layered Zn5(OH)8(H2O)2·(NO3)2 phase. Therefore, NO3− not only facilitates the formation of a base for the deprotonation of the linker but also structurally forms the essential intermediate of layered Zn5(OH)8(H2O)2·(NO3)2 phase during the CED process. Furthermore, they explored the influence of different oxygen sources, such as ClO4−, water and oxygen, on the formation of μ4-O centers in MOF-5.60 The relative crystallinity of these MOF films with other oxygen sources was lower than the crystallinity observed with nitrate, which is due to the role played by the layered zinc hydroxyl-nitrates as intermediates. Very recently, Guang et al. reported a comprehensive study of oxygen-assisted cathodic deposition of a series of ZIFs [including Zn(2-methylimidazole)2(C8H10N4Zn, ZIF-8), Co(2-methylimidazole)2(C8H10N4Co, ZIF-67) and Zn(4,5-dichloroimidazolate)2(C6H2Cl4N4Zn, ZIF-71] to avoid the plating of reduced metals.61 The reaction arises from the reduction of O2 to O2− on the electrode surface. Through the formation of the superoxide ion O2− or superoxo-ZnII complex, the deprotonation of the linker will happen as shown in eqn (5):
2(Zn2+)(O2−) + 2[Zn(HmiM)mLn] → 2[Zn(miM)(HmiM)m−1Ln] + 2Zn2+ + H2O2 + O2 | (5) |
Fig. 3 Proposed transformation scheme to account for the different crystalline phases observed during the cathodic electrodeposition of MOF-5. The inset shows the experimental PXRD pattern of a sample deposited at −1.50 V on FTO in water-containing DMF (1.31 M H2O) for 15 min with reagent concentrations of [ZnCl2] = 150 mM, [NaNO3] = 300 mM, and [H2BDC] = 50 mM. The PXRD patterns of Zn5(OH)8(H2O)2·(NO3)2 and Zn5(OH)8(Cl)2·(H2O)2 were simulated. Reproduced with permission from ref. 58. Copyright 2015, American Chemical Society. |
The deprotonated complex will subsequently evolve into a ZIF as proven by previous research on the hydrothermal synthesis.62 Through this procedure, they deposited large-area uniform ZIF films of various thicknesses on different substrates without metallic zinc impurities. Additionally, the applied potential is not the only parameter influencing the CED of MOF films; deposition time, supporting electrolyte, substrate and additives also affect the MOF film formation. Yang et al. studied the influence of these parameters during the CED of terbium-succinate (TB-SA) from DMF containing ammonium nitrate onto transparent FTO electrodes.63 Deposition time controls the thickness and mass loading of the MOF films. For the same deposition time, the thickness of the film can also be controlled by current density. But, when the current density was too large (>0.6 mA cm−2 in this case), the adhesion of the MOF films to the FTO was reduced. A similar situation was also observed in their later research on the CED of Eu(thiophene-2,5-dicarboxylate)(C6H2O4SEu, Eu-TDC).64 Moreover, the higher current density will lead to more negative potentials during cathodic electrodeposition, which can lead to the formation of other impurity compounds on the substrate, for example reduction of metal ions. The conductivity and viscosity of an electrolyte depend in part on the concentration and/or nature of the supporting electrolyte and additives, which are key factors determining the current efficiency.63,65 In addition, similar to the above discussion presented for the MOF-5 case, different supporting electrolytes and additives facilitate the formation of specific intermediates affecting the kinetics, and the nucleation and growth of MOFs.
Fig. 4 (A) Electrochemically grown [Cu3(BTC)2] coatings in QCMB measurement of water adsorption. Schematic representation of the setup and signal upon adsorption of water from nitrogen streams at different RH values illustrating the reversibility and reproducibility. Reproduced with permission from ref. 31. Copyright 2009, Wiley-VCH. (B) Fabrication of composite films via the EPD method and emission spectra of the Eu@UiO-66-hybrid film recorded in the temperature range of 273–403 K with excitation at 330 nm. Reproduced with permission from ref. 82. Copyright 2018, American Chemical Society. (C) Schematic sensor setup used for VOC detection and fabricated capacitive nanosensor with Cu-BTC film and capacitance variation of sensor vs. methanol concentration (0–1000 ppm) upon applying 1 MHz, and LCR meter frequencies with the corresponding cycling behavior of the Cu-BTC-based capacitive sensor after introducing analytes with 500 ppm of methanol. Reproduced with permission from ref. 33. Copyright 2016, Elsevier. |
Compared with other sensing mechanisms, changes of luminescence (including enhancement, shifting, or quenching) in MOFs have been the most widely used mechanism in MOF based sensors until now due to their wide range of applications (such as temperature, oxygen, pH, biomolecules and so on).74,75 The first electrochemically synthesized, luminescent MOF thin film sensor was prepared by Mei-Lin Ho et al.76 The Ir–Zne coordination compound was synthesized using Ir(ppy)2 (H2dcbpy)PF6(L-H2, ppy = 2-phenylpyridine, H2dcbpy = 4,4′-dicarboxy-2,2′-bipyridine) as a ligand and Zn as metal nodes. The Zn ions were produced by the anodic dissolution of Zn metal, predeposited on a stainless steel mesh. The influence of electrodeposition conditions such as voltage, time, and temperature on the MOF structures, particle size and photoluminescence properties, oxygen-sensing performance and other factors is explored in detail. Finally, MOFs synthesized at 5 V and 30 °C for 1 h exhibited the best performance with a deduced KSV (associated Stern–Volmer quenching constant) value of 3.55, a detection limit of 0.050%, a recovery time of 21 s and a response time of 23 s toward oxygen. After 11 cycles of testing, >70% intensity could be maintained. Subsequently, a combination of glucose oxidase with this electrode could be used as a glucose sensor electrode exploiting the simultaneous depletion of oxygen with a concomitant increase of phosphorescence. The linear range for the determination of glucose was 0.1–6.0 mM with a limit of detection (LOD) of 0.05 mM and a response time of less than 120 s. Moreover, the biosensor showed similar results to isotope-dilution gas chromatography mass-spectrometry for detection of glucose in human serum. The fluorescent properties of Zn-based MOFs are also useful for the detection of nitro-explosives. Fluorescent Zn-MOF films were explored by Rong Cao and his co-workers.77 The Zn3(1,3,5-benzenetricarboxylate)2(C18H6O12Zn3, Zn-BTC) film was prepared via AED with two zinc plates as the anode and cathode in an aqueous solution (containing H3BTC and NH4F). 4-Nitrotoluene (4-NT), 2,3-dimethyl-2,3-dinitrobutane (DMNB), 2,4,6-trinitrotoluene (TNT), nitrobenzene (NB) and 1,3-dinitrobenzene (1,3-DNB) were chosen as target chemicals. TNT and 1,3-DNB (200 ppm) exhibited a significant quenching effect on the fluorescence of these MOF films with quenching percentages of 66% and 62%, respectively. The detection limits of these nitro explosives are as low as 0.5 ppm in solution. Compared with transition metal based MOFs, lanthanide MOFs attract more attention as luminescent sensors due to large Stokes shifts, intense sharp emissions, high color purity and a long lifetime.74 These appealing luminescent properties of lanthanide MOFs originate from f–f transitions with organic linkers acting as antennas. Interestingly, optical color changes of specific MOF films, when a change of applied potential happens, can be observed even by the naked eye. Lu et al. developed MOF-based electrochromic films with redox-active ligands [Cu3(tricarboxytriphenyl amine)2 (C42H24O12N2Cu3, CuTCA)] on FTO (with pre-deposition of Cu particles on FTO) by anodic electrodeposition.78 The fabricated film can achieve rapid switching speed (both coloration and bleaching time < 5 s), a high optical contrast of 65% @ 700 nm and long cycling ability (1000 cycles with <5% contrast attenuation). However, because of the high cost of these rare earth metals, AED is not suitable for preparing lanthanide MOFs. Yangyi Yang and his co-workers engineered a series of lanthanide MOF based sensors by cathodic electrodeposition. Their first work is related to carbonate detection.65 A thin film of Eu(benzophenone-3,3′,4,4′-tetracarboxylate)(H2O)2 (C17H11O11Eu, Eu-HBPTC) was coated on FTO by cathodic electrodeposition. Based on the same method, Tb-HBPTC and Gb-HBPTC were also synthesized successfully. The luminescent properties of Eu-HBPTC film for sensing carbonate in aqueous solution were then studied. When this film was immersed in 10−3 M CO32− for two hours, the emission at 617 nm almost disappeared. The quenching effect can still be observed even in the presence of 10−6 M CO32−. Common interferents such as Cl−, ClO4−, BrO3−, IO3−, SO42−, HPO42− and PO43− anions were tested under the same conditions. No significant change was observed for the fluorescence intensity, indicating the high selectivity of this carbonate sensor. In their follow-up work, Eu based MOF films with different linkers [thiophene-2,5-dicarboxylate (TDC) and naphthalene-2,6-dicarboxylate (NDC)] were studied for the sensing of nitroaromatic explosives64 and picric acid,79 respectively. Another lanthanide element used for luminescent sensors is Tb. Yangyi Yang et al. used different Tb based MOF films for the detection of Cu2+.63 A flower-shaped thin sheet comprising clusters of Tb-succinate MOFs was prepared by CED on FTO. Under optimal electrodeposition conditions (current density, time, supporting electrolyte, the addition of NaOH), a dense and smooth layer was obtained. The stability of this thin film in water and DMF was explored by immersing it into a solution for 3 days. There was no obvious decline in DMF, and more than 75% emission intensity (5D4 → 7F5 emission at 545 nm) could be retained in water. After immersion of this film in 1 × 10−3 M Cu2+, the emission intensity decreased more than 85% in 10 min. A linear relationship between the emission intensity and the concentration of Cu2+ could be observed in the range from 1 × 10−5 M to 1 × 10−3 M with a KSV of 6298 M−1. Compared to powder samples, thin film samples showed a faster response. For the sensing of Cu2+ by Tb-SA films, interfering cations (K+, Ba2+, Mg2+, Cd2+, Ca2+, Co2+, Ni2+, Zn2+ and Fe3+) showed no obvious influence on the detection of Cu2+. Based on the luminescence lifetime results with and without the addition of Cu2+, the behavior of the quenching process is found to be static quenching. In addition, the color change under 254 nm UV-light and the peak changes in XRD indicate that this quenching effect could be caused by ion-exchange of Cu2+ into the TB-succinate structure. Following the same strategy, smooth and adherent thin films of Tb(benzene-1,3,5-tricarboxylate)(C9H3O6Tb, Tb-BTC) were coated on FTO as well, which showed an even higher KSV (1.7 × 104 M−1) than that of TB-succinate.80 These studies indicate that linkers and metal ions in MOF structures can be easily tuned by CED under suitable electrodeposition conditions.
EPD was also used to prepare thin films of lanthanide MOFs. Recently, Rong Cao and his co-workers developed a sensor made by EPD. Continuous and dense Ln-BTC MOF films [including Tb-BTC MOF, Eu(benzene-1,3,5-tricarboxylate)(C9H3O6Eu, Eu-BTC) MOF and Tb0.55Eu0.45-BTC MOF] were rapidly (within 5 min) and easily coated on different substrates such as Zn, ITO and FTO.81 Firstly, MOFs were prepared by a solvothermal method using different sources of metal ions. Then, these MOFs were dispersed into a CH2Cl2 solution for EPD using different substrates, voltages and deposition times. A layer of lanthanide MOFs was deposited on the positive electrode due to the free carboxylate group of these MOFs (negative charge). The as-prepared Tb-BTC films were used for the detection of different analytes: (1) Cr3+ in aqueous solution, (2) NB in an ethanol solution and (3) NB and TNT in the gas phase. All analytes showed clear luminescence quenching for the Tb-BTC films. Because of the separation of the synthesis and coating steps in EPD, predesigned and functionalized MOFs prepared by other methods can be deposited. The flexibility in choosing the desired MOF offered by the EPD method allows for further improvement in the sensing performance of MOF films. For example, Cao et al. reported a procedure to prepare lanthanide loaded UiO-66-(COOH)2 film with hybrid linkers by EPD for temperature sensing.82 They used post-synthetic exchange (PSE) and post-synthetic modification (PSM) to introduce a luminescent ligand (1,4-naphthalene dicarboxylic acid) and luminescent metal ions (Eu or Tb) into the UiO-66 structure, respectively. Moreover, the particle size and Ln content of Ln@UiO-66-hybrid MOFs can be easily controlled during the solvothermal process for the pristine UiO-66-(COOH)2 MOF and PSM step. The resultant Ln@UiO-66-hybrid MOFs exhibited a blue emission from H2NDC and a green emission from Tb3+ or red emission from Eu3+ (Fig. 4B). Under optimal conditions (type of metal precursor, Ln content and particle size), Eu@UiO-66-hybrid MOF film showed the best temperature-sensing ability with a relative sensitivity of 4.26% K−1, between 303 K and 403 K. Carbon quantum dots (C-QDs) can also be doped into the UiO-66-(COOH)2 structure to prepare C-QDs@UiO-66-(COOH)2 film based on the same EPD procedure.83 This film can be used to detect temperature changes in a lower temperature range (97–297 K) in comparison to the Eu@UiO-66-hybrid MOF with a relative sensitivity of 1.3% K−1. In order to explore the possibility of AED of lanthanide MOF films without using costly lanthanide metal plates, a two-step method was proposed recently.51 In the first step, a ZnO–Tb mixture was deposited on the surface of the substrate (Zn or Al). Then a layer of Tb-BTC was deposited on the substrate via a typical anodic electrodeposition of MOFs using a constant current in a water–absolute ethanol electrolyte containing H3BTC and MTBS. The resulting Tb-BTC film showed the same morphology and luminescence properties as Tb-BTC films derived from a Tb plate. This Tb-BTC film has been successfully tested for the detection of 2,4-dinitrotoluene (DNT) by the quenching of luminescence effect. Surface-enhanced Raman scattering (SERS) of noble metal nanoparticles can be used to detect the “fingerprint” signature of an analyte. In order to prevent the noble metal nanoparticles from aggregating, MOFs can be used as good carriers for loading these particles due to their high surface area and porosity. However, the balance between maintaining the structure of the MOF and the increasing number of Raman “hot spots” at the junctions of these nanoparticles is the bottleneck. In order to solve this problem, Haitao and his co-workers constructed core–shell Cu-BTC@Ag nanoparticles on a screen-printed carbon electrode with a controllable structure by a two-step CED,84i.e. CED of Cu-BTC on a screen-printed carbon electrode with subsequent CED of Ag nanoparticles. After CED, the structure of the Cu-BTC can be maintained well. The resulting samples showed high SERS activity for detecting 4-aminothiophenol (LOD of 0.5 nM) and a series of polycyclic aromatic hydrocarbons (LOD from 0.15 nM to 20 nM for different molecules) with high stability and reproducibility.
Electrochemical sensors are the largest and the most versatile group of chemical sensors and are widely used in daily life. These sensors use conductometric, capacitive, and amperometric detection methods and impedance analysis.85 M. H. Sheikhi et al. fabricated a MOF based capacitive nanosensor by anodic electrodeposition of Cu-BTC.33 A thin film (5 μm) of Cu-BTC was coated on a patterned Cu plate as exhibited in Fig. 4C. This sensor was used to detect ethanol and methanol vapors based on changes in capacitance caused by the absorption of volatile organic compounds (which changes the dielectric constant). The linear range of this sensor for ethanol and methanol is 0–1000 ppm with a LOD of 39.1 ppm and 130.0 ppm for methanol and ethanol, respectively. Recently, an amperometric-based sensor was prepared by Qu and his co-workers.35 In their work, anodic electrodeposition was employed to prepare an electro-active metal–organic framework film (NENU-3), which is composed of Cu-BTC encapsulating phosphotungstic acid, on a copper plate. This electrode was used to detect bromate, which is correlated with certain health and environmental issues,86 by the electrocatalytic reduction of bromate. Some kinetic parameters of the electrocatalytic reduction of bromate, such as the catalytic rate constant (ks) and the electron transfer coefficient (α), were derived from CV and LSV data. At a potential of −0.3 V, a wide linear range (0.05–72.74 mM) with an LOD of 12 μM was achieved by this chronoamperometric method. Moreover, inorganic anti-interferents such as Na+, K+, Cl−, NO3−, NO2−, SO42−, ClO3−, NH4+ and CO32− do not have an obvious influence on the signal of the electrocatalytic reduction of bromate. Except for bromate, Nianjun et al. studied the sensing performance of Cu-BTC films on glassy carbon electrodes (deposited by CED) for different kinds of organic molecules, including nicotinamide adenine dinucleotide, diethylstilbestrol, estradiol, hypoxanthine, tartrazine, sunset yellow, and xanthine.87 They found that the applied cathodic potential has a clear influence on the morphology, thickness and sensing performance of the Cu-BTC films. The same method was used by Ji and his co-workers for electrochemical determination of bisphenol A as well.88 The highest peak currents of Cu-BTC films for determination of bisphenol A were achieved at a potential of 0.496 V (vs. SCE). At this potential, the Cu-BTC films showed a large linear range (5.0 to 2000 nM) with a low LOD (0.72 nM). The results show that Cu-BTC films can reach a comparable analytical performance (for determination of bisphenol A) with high-performance liquid chromatography. Electrochemical sensors based on MOFs can be used for chiral recognition and resolution of organic molecules as well. Xuan Kuang et al. successfully co-deposited chiral carbon quantum dots (CCQDs) and MOFs [Ni(1,4-benzenedicarboxylate)-NH3, C8H4O4Ni-NH3, and Ni-BDC-NH3] on Cu foil by CED.89 This electrode exhibited good selectivity and high sensitivity to penicillamine enantiomers with a good linear relationship between current density and L-PA% in the racemic mixture. In their following work,90 the preparation of a CCQDs/[Cu3(2,4,6-tris(dimethylamino)-1,3,5-triazine)(H2O)3]·10H2O·5DMA (C27H18O15N6Cu3, Cu-TDPAT) composite electrode was performed following the same procedure. The composite electrode not only shows quantitative analysis towards tyrosine (Tyr) enantiomers but also presents the ability to determine L-Tyr% in racemic mixtures. The same detection method for H2O2 sensing was explored by Nianjun Yang et al.91 In their paper, Cu-based MOFs with different organic ligands were deposited on a glassy carbon electrode by the CED method. Based on their theoretical and experimental results, not only the morphology but also the electrochemical behavior and active sites of the Cu-MOFs can be tuned by different organic ligands, including 1,3,5-benzenetricarboxylic acid, 1,4-benzenedicarboxylic acid, and 1,2,4,5-benzenetetracarboxylic acid. The Cu-MOF with 1,3,5-benzenetricarboxylic acid linkers exhibited the highest current response for H2O2 at a potential of −0.7 V (vs. SCE). This work also contains stability results after the electrochemical test, which is important for practical application of such sensors but is rarely reported. FT-IR, XPS, and XRD data and SEM images of these Cu-MOFs before and after measurements were recorded. After 10 min testing time, the structure of the Cu-MOFs was not changed, but decomposition and collapse of these MOFs were observed after long duration stability tests (e.g. 24 h). In order to improve the conductivity of MOF based electrodes for sensors, Jianshan et al. prepared Cu-BTC/multi-walled carbon nanotube (MWNT) multilayer films by a two-step EPD from two different solutions, including Cu-BTC and MWNTs, respectively.92 The electrode with eight layers (four layers of Cu-BTC and four layers of MWNT films) showed the best performance for detecting glucose with a sensitivity of 3878 μA cm−2 mM−1.
Fig. 5 (A) (a) Optical photograph and (b) SEM image of the MOF-199/Ni film prepared by electrodeposition with (c) the schematic of the eosin-Y sensitized photocatalytic process on the MOF-199/Ni film. Reproduced with permission from ref. 57. Copyright 2016, the Royal Society of Chemistry. (B) (a) ECD process of Fe/Ni-BTC MOF thin films. (b) Design and synthesis of Fe/Ni-BTC MOFs. (c) Optical image, (d) scanning electron microscopy (SEM) image, and (e) higher-magnification image of Fe/Ni-BTC@NF. (f) LSV curves of ECD Ni-BTC@NF, ECD Fe-BTC@NF, ECD Fe/Ni@NF, ECD Fe/Ni-BTC@NF, and NF measured at 2 mV s−1. (g) Tafel plots of ECD Ni-BTC@NF, ECD Fe-BTC@NF, ECD Fe/Ni@NF, and ECD Fe/Ni-BTC recorded at 1 mV s−1. (h) Comparison of ECD Ni-BTC, ECD Fe-BTC, ECD Fe/Ni, ECD Fe/Ni-BTC, bulk Ni–Fe oxides, and porous Ni–Fe oxides. Reproduced with permission from ref. 96. Copyright 2016, American Chemical Society. |
In situ synthesis of catalyst layers is highly recommended for electrochemical or photoelectrochemical reactions not only because of the enhancement of reactivity and ease of recycling but also because it help in avoiding the use of binders. For electrochemical or photoelectrochemical reactions, the active materials must be coated on a conductive substrate using additional binders, such as Nafion and PTFE. These non-conductive additives increase the resistance, cause partial blocking of pores and active sites, hamper the transport of ions and increase the processing time and cost. More importantly, these issues increase with increasing aerial mass loading of the catalyst. To address these challenges, binder-free electrodes are highly desired. Electrochemical deposition is one of the most promising ways to solve/circumvent the above problems. Hod and co-workers proposed a strategy to immobilize an Fe-porphyrin catalyst into Zr6O4(OH)4(4,4′,4′′,4′′′-(porphyrin-5,10,15,20-tetrayl)tetrabenzoate-H2)3(C144H82O32Zr6, MOF-525) thin films for the electrochemical reduction of CO2.95 First, MOF-525 was coated on an FTO substrate by EPD. After a solvothermal post-metalation with iron chloride, the target electrode configuration could be accomplished with high areal concentrations equivalent to ∼900 monolayers of surface-adsorbed Fe-TPP. At an overpotential of 650 mV for the reduction of CO2 to CO in DMF electrolyte containing 1 M TBAPF6, the main product is a mixture of CO and H2 with a 100% faradaic efficiency (54 ± 2 and 45 ± 1% for CO and H2 formation, respectively). The current was limited by the rate of change of diffusion through the MOF structure. For future work, improvement of the rate of diffusion could be a key to realizing better catalysts. In the following year, a pure MOF catalyst for electrochemical reduction of CO2 was reported by Kang et al.68 A solvothermal method was used to synthesize Zn-BTC with different morphologies by changing the mass fraction of ZnCl2 in the synthesis solution. These Zn-BTC MOFs with different morphologies were deposited onto carbon paper by EPD from a methanol/DFG solution, using a constant potential (10–50 V) for one hour. They found that not all the morphologies of Zn-BTC form smooth layers. When the mass fractions of ZnCl2 were 0.44 and 0.5, the irregular surface is dominant. This is most likely caused by the particle size of the suspension and the uneven current distribution at the edges. Then these electrodes were used as a cathode in the electrochemical reduction of CO2 with different ionic liquid electrolytes. Under the same conditions, the sheet-like Zn-BTC covered electrode showed the highest current density with the highest electrochemical surface area. Gas chromatography (GC) and 1H-NMR identified CH4 as the main reaction product with a selectivity of 80%. Such a high selectivity is rare. This study also explored in more detail the effect of different electrolytes and metal electrodes. A higher current density and higher selectivity of the sheet-like Zn-BTC covered electrode than other metal electrodes for the reduction of CO2 were observed. Due to the strong interaction between fluorine and CO2, BmimBF4 was found to be the most effective electrolyte. Another example of a binder-free thin film MOF electrode for electrochemical reactions is related to the oxygen evolution reaction (OER), which is a critical half-reaction for electrochemical water splitting and in rechargeable metal–air batteries due to the sluggish kinetics of the OER. In order to avoid binder issues, CED was used by Bo Wang et al. for preparing Ni3(benzene-1,3,5-tricarboxylate)2(C18H6O12Ni3, Ni-BTC) MOF layers.96 To further improve the catalytic ability of Ni-BTC, 10% (based on the molar ratio of two metal sources) of iron ions were introduced into the electrolyte. At a potential of −1.5 V vs. Ag/AgCl with a deposition time of 30 s, Ni-BTC, MIL-100 and Fe/Ni-BTC MOF thin films were coated on NF as shown in Fig. 5B. The NF-Fe/Ni-BTC MOF shows the best OER performance with an overpotential of only 270 mV at 10 mA cm−2 and with the lowest Tafel slope (47 mV dec−1). At 1.5 V vs. RHE, the current density of NF-Fe/Ni-BTC thin films was around 100 times higher than that of Fe/Ni-BTC powder (prepared by a solvothermal method) coated on glassy carbon electrodes. The NF-Fe/Ni-BTC electrode also showed no signs of degradation at 1.53 V vs. RHE during 15 h. As mentioned in Section 2.2, the applied potential has a large influence on the formation of MOF films during CED. Chen and his co-workers reported that a metallic Fe nanostructure was formed when CED of Fe-BTC was performed at −1.5 V (vs. SCE).97 They prepared different Fe-BTC films at different negative potentials (−1.1 V, −1.5 V and −1.9 V). The Fe/Fe-BTC film deposited at −1.5 V exhibited the loosest morphology and best electrochemical catalytic performance for H2O2 reduction, because of more electrochemically active sites originating from the loosest morphology of the Fe/Fe-BTC film deposited at −1.5 V.
Fig. 6 (A) Overview of the fabrication steps for a MOF coated microseparator and the corresponding binary breakthrough profiles of a MOF coated microseparator with n-hexane (10.21% in feed stream) and methanol (2.11% in feed stream) vapor. Reproduced with permission from ref. 37. Copyright 2013, Elsevier. (B) Overview of the setup for a MOF coated microseparator and breakthrough curves of a MIL-100(Fe) microfluidic device fed by a 40/60 methanol–n-hexane blend in nitrogen carrier gas. Reproduced with permission from ref. 53. Copyright 2013, the Royal Society of Chemistry. |
Fig. 7 (A) (a) Perspective views of the crystal structure showing the coordinative environment around the metal center in M(HBTC)(DMF)2 (all H atoms are omitted for clarity); (b) packing of three 2D layers of M(HBTC)(DMF)2 to generate a supramolecular framework by short contact stacking from the c axis (different layers are denoted in different colors); (c) view of the structure from the b axis; (d) schematic representation of anodic electrodeposition of MOFs; (e) schematic illustration of the process to prepare 2D-CMO material on NF. Reproduced with permission from ref. 116. Copyright 2017, Elsevier. (B) Schematic illustration of the synthesis of a Cu@porous carbon matrix on a copper foam substrate. Reproduced with permission from ref. 117. Copyright 2018, Elsevier. (C) Preparation of MOF-derived carbon composite electrodes by either thermolysis of electrochemically deposited MOF films in one step (red) or thermolysis of MOF powder and conventional electrode preparation in several steps (grey). Reproduced with permission from ref. 118. Copyright 2017, the Royal Society of Chemistry. |
Applicability | Purity | Continuity and compactness | Adhesion strengths | Limitation of the substrate | |
---|---|---|---|---|---|
AED | Medium | High | High | High | Limited |
CED | Low | Medium | High | Medium | No limit (conductive) |
EPD | High | High | Low | Low | No limit (conductive) |
(1) AED: because of the high current density attainable, high adaptability and high purity of the resulting MOF films based on anodic electrodeposition, AED exhibited promising potential for use in modern industrial processes. Anodically electrosynthesized MOF films are more compact than films derived from the other two electrochemical methods. The films also display on average higher adhesion strengths, which is perhaps caused by the formation of an oxide layer between the substrate and the MOF films serving as a bridging layer. This bridging layer can prevent detachment during subsequent processing steps or tests (e.g. pyrolysis). Therefore, anodic electrodeposition could be a better approach than the other two electrochemical methods for in situ growth of MOF-derived materials on the substrate. In addition, if the MOF film based devices are operated under harsh conditions (e.g. strong agitation, gas evolution, …), strong adhesion strength is necessary. It should be noticed that the enhanced adhesion strengths have a prerequisite relationship with the detachment of MOF crystals caused by anodic corrosion of the substrate. This is a balance between anodic corrosion of the substrate and formation of MOFs. Currently, there are no better methods except for optimizing the experimental conditions to find the balance point. Using a metallic electrode as the source of metal cations (instead of metal salts), the formation of corrosive ions or by-products can be avoided. However, the correspondence between the metallic electrode and the resulting MOF node is the main limitation of AED.
(2) CED: similarly, a dense and continuous MOF film can be prepared by CED as well. Unlike AED, any conductive substrate can be used as an electrode for CED of MOF films. Furthermore, the incorporation of a second (or more) metal node into the framework can be easily achieved by CED if the formation of these multi-metallic MOFs should be thermodynamically and kinetically feasible. In the case of CED, the potential plays a critical role in the resulting MOF films. Compared with AED, there are more competition reactions on the cathode. In most cases, the impurity of the metallic structure (originating from cathodic deposition of metals) is hard to avoid especially for the metal ions with more positive reduction potential. Moreover, the structures of the resulting MOF films are sensitive to reduction potential as well.
(3) EPD: theoretically speaking, any kinds of MOFs can be deposited on a substrate by EPD. If a specific or complex MOF structure is useful for the desired applications, EPD could be a better choice. In contrast to AED and CED, EPD is a coating technique rather than a synthesis method. An additional synthesis step is necessary for EPD of MOFs. Therefore, the advantages inherent to the electro-synthesis of MOFs such as milder synthesis conditions and shorter synthesis time are not present in the EPD of MOF films. On the other hand, the separation of synthesis and deposition steps means that the MOFs can be functionalized in advance. Voids and cracks are hard to avoid in EPD, which will be the Achilles' heel for some applications (e.g. separations). When the continuous and dense film is not necessary (e.g. catalysis), EPD could be used for coating MOF film easily and efficiently. However, the mechanical fragility issues of MOF films prepared by EPD should be carefully considered.
From a broader point of view, at the early stage of research, most studies focus on some archetypal MOFs, such as Cu-BTC and MOF-5, which are notoriously simple to prepare. Recently, based on the better understanding of the principles of electrodeposited MOFs, such as critical concentration during AED and the formation of an essential intermediate during CED, more and more different kinds of MOFs are deposited by electrodeposition as shown in Table 1. However, due to the large number of members in the MOF family (more than 70000), it is necessary and urgent to know the commonalities and differences behind the electrodeposition processes of these MOFs, for instance, how the deposition parameters (such as linkers, nodes, additives, pH…) affect the formation of MOF layers (the nucleation and crystal growth). Meanwhile, even though controllable thickness and crystal size have been achieved by electrodeposition, optimization of microstructures, such as research on defects, grain boundaries and channel orientations, is rare. In addition, with the development of MOF research, new types of MOFs have exhibited promising performance for desired applications, such as conductive MOFs, multi-metallic MOFs (consisting of multiple metal nodes within a single crystal) and multivariate metal–organic frameworks (consisting of multiple linkers of different functionalities within a single crystal). Electrodeposition could be a promising method to prepare these new types of MOF films. Last but not least, the stability of MOF films (both mechanical and chemical) is always a challenge before taking technology from laboratory to applications. Therefore, we recommend an analysis of stability for all reported studies.
Nowadays, studies of MOF films are shifting from synthesis towards their applications. The versatility of electrochemical deposition is a powerful asset in the preparation of MOF films for different applications. We sincerely hope that this review will facilitate the discovery and understanding of electrochemical deposition of MOF films and inspire more researchers to join this area.
This journal is © The Royal Society of Chemistry 2020 |