Next Article in Journal
Enhanced Antimicrobial Activity through Synergistic Effects of Cold Atmospheric Plasma and Plant Secondary Metabolites: Opportunities and Challenges
Previous Article in Journal
Poria cocos (Schw.) Wolf, a Traditional Chinese Edible Medicinal Herb, Promotes Neuronal Differentiation, and the Morphological Maturation of Newborn Neurons in Neural Stem/Progenitor Cells
Previous Article in Special Issue
Influence of Co3O4 Nanostructure Morphology on the Catalytic Degradation of p-Nitrophenol
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Azo-Linkage Redox Metal–Organic Framework Incorporating Carbon Nanotubes for High-Performance Aqueous Energy Storage

1
College of Materials, Xiamen University, Xiamen 361005, China
2
School of Chemistry and Chemical Engineering, Nanjing University of Science and Technology, Nanjing 210094, China
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(22), 7479; https://doi.org/10.3390/molecules28227479
Submission received: 20 September 2023 / Revised: 24 October 2023 / Accepted: 31 October 2023 / Published: 8 November 2023
(This article belongs to the Special Issue Exclusive Feature Papers in Physical Chemistry)

Abstract

:
The design of well-defined hierarchical free-standing electrodes for robust high-performance energy storage is challenging. We report herein that azo-linkage redox metal–organic frameworks (MOFs) incorporate single-walled carbon nanotubes (CNTs) as flexible electrodes. The in situ-guided growth, crystallinity and morphology of UiO-66-NO2 MOFs were finely controlled in the presence of CNTs. The MOFs’ covalent anchoring to CNTs and solvothermal grafting anthraquinone (AQ) pendants endow the hybrid (denoted as CNT@UiO-66-AQ) with greatly improved conductivity, charge storage pathways and electrochemical dynamics. The flexible CNT@UiO-66-AQ displays a highest areal specific capacitance of 302.3 mF cm−2 (at 1 mA cm−2) in −0.4~0.9 V potential window, together with 100% capacitance retention over 5000 cycles at 5 mA cm−2. Its assembled symmetrical supercapacitor (SSC) achieves a maximum energy density of 0.037 mWh cm−2 and a maximum power density of 10.4 mW cm−2, outperforming many MOFs-hybrids-based SSCs in the literature. Our work may open a new avenue for preparing azo-coupled redox MOFs hybrids with carbaneous substrates for high-performance robust aqueous energy storage.

1. Introduction

Metal-organic frameworks (MOFs) have flourished as a fascinating class of electrochemical energy storage electrode materials [1,2,3,4,5,6,7,8]. Featuring porous crystalline structures, they are composed of metal cluster nodes and organic linkers/ligands. Their diversified functionalities, unique morphologies and tunable porosities endow them with superior charge storage capability. However, the low conductivity of pristine MOFs is acknowledged as the dominant limitation for their practical applications. One facile strategy to address this issue is to incorporate conductive components (e.g., carbon-based materials and conducting polymers) into MOFs skeletons, affording improved conductivity and possible extra charge storage capacity [9,10,11,12,13,14]. Furthermore, the functional surface of carbon materials can guide the uniform growth of MOFs along them and stabilize the as-formed hierarchical structures [15,16,17]. For example, ultrathin nickel MOF-based two-dimensional (2D) nanosheets were prepared with finely tuned thickness and specific surface with the aid of in situ interpenetration of carboxylated carbon nanotubes (CNTs) [16]. The 2D integrated hybrid electrode delivered a superior specific capacity of 680 C g−1 at 1 A g−1 with 68% capacity retention at 10 A g−1.
Besides considering conductivity, it is crucial to improve MOFs’ capacity by introducing new conjugated ligands or redox functionalities. The incorporation of redox ligands or pendants in/onto MOFs skeletons efficiently improves electrochemical charge storage through reversible redox reactions in aqueous electrolytes [18,19,20]. In accordance with this, Chen’s group recently reported a two-dimensional copper–benzoquinoid MOF to serve as lithium-ion battery cathode [21]. The hybrid electrode delivered a high reversible capacity (387 mAh g−1) and good cycling stability, where a three-electron redox reaction for each coordination unit was further unveiled using comprehensive spectroscopic techniques. However, for redox ligand-based MOFs, the change in the oxidation state of ligands may impact the coordination geometry of metal nodes. This creates great challenges in selecting suitable building blocks for stabilizing the structure under electrochemical conditions.
On the other hand, the grafting of functional pendants via post-synthetic modification has matured into an efficient and facile strategy to broaden the scope of MOFs [22,23]. The diverse functional groups (e.g., -NH2, -OH, -SH, et al.) on ligands afford great possibilities for chemical modification to generate multi-purpose structures. Meanwhile, the covalent bonding can further prevent the electrodes from dissolving into electrolytes. Post-synthetic coupling of redox pendants can endow MOFs with redox activity to improve their charge storage capacity and extra ion transfer to promote efficient electrochemical dynamics. For instance, quinones and their derivatives, showcasing reversible redox reactions and high theoretical capacity, have been successfully explored for batteries and capacitors [24,25,26,27]. However, the post-synthetic grafting of MOFs using redox quinones has rarely been reported, which is mainly attributed to a lack of reliable synthetic approaches for linkage building. Zirconium-based MOFs are most commonly used due to the availability of a large variety of benzene dicarboxylic acids for ligands, and their high chemical stability can endure further chemical modifications.
In this work, we report the covalent bonding of redox pendants onto MOFs via a facile post-synthetic approach and their incorporation with CNTs as hybrid self-standing electrodes for high-performance and robust energy storage in aqueous electrolytes (Figure 1). The stable azo-coupled MOF was readily achieved via a one-pot solvothermal reaction between 1-nitroanthraquinone and nitro-substituted UiO-66 (denoted as UiO-66-NO2) in a mixture solution of dimethyl formamide and water. More importantly, we further incorporate carboxylated carbon nanotubes CNTs for the in-situ growth of MOFs to fine tune their crystallinity and morphology as well as their conductivity. The as-obtained redox MOFs hybrid (denoted as CNTs@UiO-66-AQ) with CNTs features azo-linked anthraquinone (AQ) pendants and well-aligned CNTs as the charge transfer channels, where capacitance and redox reaction contribute to significantly improved charge storage. Consequently, the self-standing CNT@UiO-66-AQ electrode delivers a maximum areal specific capacitance of 302.3 mF cm−2 at 1 mA cm−2 in 1 M H2SO4 electrolyte. Excellent rate performance is revealed by the undecayed capacitance after 5000 cycles at 5 mA cm−2. Importantly, two film electrodes-assembled symmetrical supercapacitors (SSCs) exhibit a maximum energy density of 0.037 mWh cm−2 and a highest power density of 10.4 mW cm−2, outperforming most MOFs-based SSCs in the literature.

2. Results and Discussion

As depicted in Figure 1a, UiO-66-NO2 MOF is steadily anchored onto the CNT surface via coordinative interactions. The post-synthetic solvothermal reaction converts two nitro groups into robust azo-linkage between AQ pendants and UiO-66 skeletons. The modification of redox quinones extended the potential window and enhanced the charge storage. The CNT incorporated redox MOF hybrids (CNT@UiO-66-AQ) can be easily prepared using two-step solvothermal reactions (Figure 1a). The direct azo-coupling redox AQ pendants onto MOF skeletons afforded a darker-colored product, implying successful chemical modifications. X-ray diffraction (XRD) spectroscopy was used to reveal the evolution of crystal structures of the MOFs accompanying the post-synthetic functionalization. As depicted in Figure 2a, UiO-66-NO2 exhibits peak patterns consistent with the simulated results for UiO-66, indicating that the nitro substituents have no effect on the crystalline structure of the MOF skeleton. After the azo-coupling AQ was added to the Zr-MOF skeleton, UiO-66-AQ still exhibit the typical crystalline peaks despite the decreased intensities compared with UiO-66-NO2, confirming the unchanged framework structures [28]. Importantly, new strong peaks appear in the region of 28°~36°, 50° and 60°, which are attributed to the packing and stacking of AQ pendants within MOF holes. It should be noted that the incorporation of CNTs results in overlapping and broadening of MOFs’ peaks, indicating a certain decrease in the crystallinity of MOFs. However, with the grafting of redox pendants onto the electron-donating MOF skeleton, the resulting intermolecular charge transfer and packing pendants induced charge hopping that will significantly improve the charge transport in CNT@UiO-66-AQ hybrid [27]. The four-point probe measurement shows CNT@UiO-66-AQ hybrid exhibits nearly 62% improved conductivity (~164.7 S cm−1) than CNT@UiO-66-NO2 (100.1 S cm−1) (Figure 2b). Fourier transform infrared spectroscopy (FT-IR) was applied to confirm the successful azo-coupling of AQ pendants. As demonstrated in Figure 2c and Figure S1, the characteristic vibration of -NO2 can be observed at peak 1530 and 1380 cm−1 for 1-nitroanthraquinone and UiO-66-NO2, respectively [29]. Such peak assigned to -NO2 vibration disappears; instead, UiO-66-AQ exhibits a new peak at 1620 cm−1, representing the azo group [30]. This suggests the successful conversion of nitro into the azo group in our post-synthetic modification. However, the nitrogen-containing functionality change is difficult to detect within the FT-IR spectra of CNT@UiO-66-NO2 and CNT@UiO-66-AQ (Figure S1, Supplementary Materials), which can be attributed to the blending of CNTs.
The formation of azo linkage is clearly confirmed with X-ray photoelectron spectroscopy (XPS). The N 1s spectrum of CNT@UiO-66-NO2 shows two obvious peaks at 405.6 and 400.0 eV, corresponding to -NO2 and -NH2, respectively (Figure 2d). The -NH2 peaks originate from 3-aminobenzoic acid for UiO-66 [31]. After nitro–azo conversion, the characteristic peak of -NO2 from UiO-66-NO2 vanishes, and a new peak at 399.1 eV assigned to the azo group appears for CNT@UiO-66-AQ [32]. At the same time, CNT@UiO-66-AQ displays the peaks for C=O at 529.7 and 287.1 eV in the high-resolution O 1s and C 1s spectra, respectively (Figure 2e and Figure S2a, Supplementary Materials). This further confirms the covalent bonding of AQ pendants onto the UiO-66-based MOF skeleton [33,34]. As expected, both CNT@UiO-66-NO2 and CNT@UiO-66-AQ exhibit the same peaks at 185.2 and 182.6 eV for Zr 3d spectra, indicating the unchanged charge state of Zr during nitro–azo conversion (Figure S2b, Supplementary Materials) [35]. Similar phenomena are observed for the evolution of four elements (ca. C 1s, N 1s, O 1s and Zr 3d) valences accompanying UiO-66-NO2 transform into UiO-66-AQ (Figure S3, Supplementary Materials).
The morphology of MOFs and their CNTs hybrids is assessed using scanning electron microscopy (SEM) and a transmission electron microscope (TEM). UiO-66-NO2 exhibits heavily aggregated particles with irregular shapes, which are ~1 μm in diameter (Figure 3a). This coincides with the TEM image (Figure 3c), wherein low transmittance is observed for large stacked particles. After azo-coupling AQ, UiO-66-AQ takes the form of soft and fluffy cottons (Figure 3b), and nanosized particles can be observed in the TEM image (Figure 3d). This can explain the decreased peak intensity in XRD patterns for UiO-66-AQ due to its loosely packed particle structure. It should be noted that UiO-66-NO2 and UiO-66-AQ present almost the same homogeneous and fine distribution of four elements (including C, Zr, N and O), as shown in Figure S4 (Supplementary Materials). The guided growth of MOFs along 2D CNTs is demonstrated in Figure 3e,f. UiO-66-NO2 nanoparticles (upto ~100 nm in diameter) homogeneously attach to CNTs to construct a porous hierarchical hybrid. Even smaller-sized UiO-66-AQ nanoparticles fill the cavities of well-aligned CNTs.
The porous structures of UiO-66 derived MOFs and their growth-guided CNTs hybrids were further studied using nitrogen adsorption–desorption analysis. As depicted in Figure 4, UiO-66-NO2 exhibits a typical type-I isotherm for nitrogen adsorption, revealing a specific surface area of 462.6 m2 g−1. The AQ-bonded UiO-66-AQ behaves similarly to a type-Ⅳ isotherm to afford a reduced surface area of 129.2 m2 g−1. However, UiO-66-AQ shows significantly increased pore sizes (from 1.7 to 9.1 nm) and pore volume (from 0.2 to 0.3 cm3g−1) in comparison to UiO-66-NO2. These results indicate the azo-coupled AQ modification can greatly relieve the stacking of UiO-66 based MOFs, which agrees well with the morphology study in Figure 3.
In terms of porosity investigation, both CNT@UiO-66-NO2 and CNT@UiO-66-AQ display typical type-Ⅳ isotherms (Figure 4c). CNT@UiO-66-NO2 showcases a specific surface area (274.5 m2 g−1) larger than CNT@UiO-66-AQ (127.1 m2 g−1) but lower than UiO-66-NO2. Correspondingly, the pore size is increased from 8.1 nm for CNT@UiO-66-NO2 to 14.3 nm for CNT@UiO-66-AQ (Figure 4d). Such microporous and mesoporous structures of our as-developed redox MOFs and CNT hybrid films can offer sufficient low-resistance channels for electrolyte penetration when used as electrodes [36]. Besides increased pore size and volume, the azo-coupling post-synthetic modification also leads to improved thermal stability for UiO-66-AQ and CNT@UiO-66-AQ compared to their nitro counterparts, as indicated by the thermogravimetric analysis (TGA) in Figure S6 (Supplementary Materials).
The coordinative effect of the carboxyl group toward Zr ions renders guided and uniform distribution of MOF crystals along CNT nanofibers for CNT@UiO-66-NO2 (Figure 3e), which can be more clearly observed in the TEM image (Figure 5). The heteronuclear and crystal growth of UiO-66-NO2 MOFs on CNTs is thus concluded [35]. As a comparison, CNT@UiO-66-AQ shows a smoother surface with several neighboring CNTs aligning together into a nano-belt and MOFs filled within the cavities (Figure 3f). Such morphology can be clearly observed via high-resolution TEM (HRTEM) (Figure 5a,b). CNT@UiO-66-AQ nanoparticles upto ~10 nm in diameter tightly attach to CNTs’ surfaces after grafting AQ pendants. Interestingly, the HRTEM images demonstrate clear crystal lattice fringes for CNT@UiO-66-AQ with a d-spacing of 0.36 nm (Figure 5b,c). However, such a phenomenon cannot be observed for CNT@UiO-66-NO2 (Figure S5, Supplementary Materials), indicating the restacking of organic pendants. Nonetheless, typical high-angle annular dark field (HAADF) images reveal the uniform distribution of C, N, O, and Zr for both samples (Figure 5e–i and Figure S5c, Supplementary Materials).
The electrochemical performance of a redox MOFs hybrid with CNTs was first evaluated in a three-electrode cell with 1 M H2SO4 electrolyte. Figure S7a (Supplementary Materials) shows the typical CV curves of UiO-66-NO2 and UiO-66-AQ electrodes at a scan rate of 30 mV s−1. In comparison to UiO-66-NO2, UiO-66-AQ shows a larger area under CV curve and a pair of redox peaks in the region of −0.3~−0.1 V, corresponding to the proton-coupled faradaic reactions of carbonyl groups. UiO-66-AQ also displays a wider potential window (−0.4~1 V vs. 0~0.8 V) and longer charge/discharge time (50 s vs. 220 s), as depicted by the galvanostatic charge–discharge (GCD) curves in Figure S7b (Supplementary Materials). Based on the discharge curve, UiO-66-AQ presents a 4-fold specific capacitance (40 F g−1) compared to UiO-66-NO2 (10 F g−1) at a current density of 0.5 A g−1.
For flexible energy storage, carboxylated CNTs-incorporated MOF hybrids are designed in our case as free-standing film electrodes, wherein the carboxyl groups serve as coordination sites for the metal nodes and anchor to improve the bonding strength between components in hybrids. The CV and GCD curves of CNT@UiO-66-NO2 and CNT@UiO-66-AQ (10 mV s−1) are shown in Figure 6a and Figure S8 (Supplementary Materials), respectively. As expected, CNT@UiO-66-AQ boosts significantly improved charge storage by taking advantage of both capacitance and redox reactions, when compared with CNT@UiO-66-NO2 or UiO-66-AQ. Three pairs of redox peaks can be observed in the CV curve of CNT@UiO-66-AQ. We thus explore the reaction kinetics through the equation i = aνb, where i and ν are the oxidation peak currents and scan rates. The b equal to 1 represents the capacitive process, while b of 0.5 refers to the redox process. In our cases, we obtain straight lines with the slope as b when plotting logarithm i against logarithm ν (Figure 6b). The b value is 0.74 for major peak 1 for CNT@UiO-66-AQ, suggesting a synergetic charge storage of capacitive and redox of AQ. The b values for peak 2 and 3 for CNT@UiO-66-AQ are quite close to 1, which is mainly attributed to the capacitance contribution of CNTs. Interestingly, the redox peak 4 accounting for ligands is marginal for UiO-66-NO2. CNT@UiO-66-AQ presents an excellent areal specific capacitance upto 302.3 mF cm−2 at 1 mA cm−2, which is nearly three-fold higher than that of CNT@UiO-66-NO2 (80.1 mF cm−2) (Figure S9, Supplementary Materials). By increasing the current density from 1 to 20 mA cm−2 (Figure 6c), CNT@UiO-66-AQ was found to retain 71.6% of its initial areal specific capacitance of (216.5 mF cm−2), outperforming CNT@UiO-66-NO2 with a capacitance retention rate of 49.9% under the same conditions. All the results reveal that the incorporation of redox pendants and CNTs substrate is beneficial, improving the conductivity, charge storage, and rate performance of hybrid electrodes simultaneously.
Electrochemical impedance spectroscopy (EIS) was used to reveal the charge and mass transport behavior of the electrodes. Generally, the intercept of the semicircle line at the x-axis represents the resistance (Rs) of the electrolyte contacting with the current collector and electrode materials [37]. The charge transfer resistance (Rct) is calculated based on the diameter of the semicircle in the high-frequency region, while the straight line at low frequencies demonstrates the ion diffusion within electrode [38]. The steeper shape of the sloped line represents an ideal capacitive behavior with the faster diffusion of ions in electrolyte [39]. As expected, CNT@UiO-66-NO2 and CNT@UiO-66-AQ electrodes exhibit a straight line in the low-frequency region and a semicircle in the high-frequency region in the EIS spectra (Figure 7). CNT@UiO-66-AQ exhibits a dramatically lower Rct (4.4 Ω) than CNT@UiO-66-NO2 (23.4 Ω), which is consistent with their conductivity results. Importantly, both EIS spectra show almost vertical slopes in the low-frequency regions for two electrodes, indicating efficient ion diffusion.
In order to quantitatively evaluate the diffusion impedance, we further plot the real part of the collected impedance (Z′) versus the reciprocal square root of the sampling frequency (Figure 6d). From the slope (0.82 vs. 0.35) of as-obtained lines, we can conclude that CNT@UiO-66-AQ has much lower ion diffusion resistance than CNT@UiO-66-NO2 to promote more efficient mass transfer. This agrees well with ordered morphology and redox-assisted electrochemical kinetics for CNT@UiO-66-AQ electrodes (Figure 5e,f). Besides the superior capacitance and rate performance, CNT@UiO-66-AQ also shows good long-term electrochemical stability. As shown in Figure 6e, there was almost no drop in its initial specific capacitance after 5000 cycles, while only 93% retention of initial capacitance was obtained for CNT@UiO-66-NO2 after 1600 cycles. More importantly, during the continuous cycling of CNT@UiO-66-AQ, no AQ pendants cleavage or organic intermediate precipitation was observed, as traced by the color and composition of the electrolyte. This implies the excellent durability of azo linkage and its tolerance to strong acidic aqueous conditions.
In order to evaluate the applicability of our designed CNT@UiO-66-AQ as a self-standing film electrode for flexible energy storage, we fabricated symmetrical supercapacitors (SSCs) by assembling two identical electrodes on both sides of a filter paper separator with 1 M H2SO4 as the electrolyte. Figure 8a shows the CV curves of an SSC at a scan rate that increased from 1 to 200 mV s−1. A pair of redox peaks can clearly be observed at operating voltage regions of 0~0.3 V, corresponding to the redox reactions of AQ moiety. According to the discharge curve, our SSC showcases an excellent areal capacitance of 155.4 mF cm−2 (1 mA cm−2). The capacitance can be retained by 48% when increasing the current density to 16 mA cm−2 (Figure 8b). The cyclic stability of the as-fabricated SSC device was revealed over 10,000 GCD cycles at 1 mA cm−2 (Figure S10, Supplementary Materials). During the cycling process, the specific capacitance declines by 20% after the first 4000 cycles before it slowly drops to 71.9% of the initial value at the end of 10,000 cycles. Such a high capacitance retention over such a long cycle life demonstrates the excellent cycling stability of our CNT@UiO-66-AQ-based SSCs. Figure 8c depicts the Nyquist plots of the initial cycle and after 10,000 cycles for the SSC. The longer Warburg line after long cycling implies sufficient electrolyte permeation and diffusion within the porous electrodes [40]. Meanwhile, the lower charge transport resistance (indicated by a smaller diameter of the semicircle) reveals the activation of the electrodes during cycling [41].
Power density (P) and energy density (E) are two key parameters used to evaluate the electrochemical performance for practical SC devices. Our SSC device can achieve a maximum P of 10.4 mW cm−2 and a highest E of 0.037 mWh cm−2 (Figure 8d), outperforming most MOFs-based SSCs in the literature. This can be attributed to the extended voltage window and improved specific capacitance of the CNT@UiO-66-AQ electrode. Representative device data (P, E) for the high-performance SSCs are listed as follows for comparison: NiCo-MOF/MWCNT (0.499 mW cm−2, 0.027 mWh cm−2) [39], CFs@UiO-66/PPy (0.26275 mW cm−2, 0.0128 mWh cm−2) [42], PET/MOF-1/rGO/PPy (0.03 mW cm−2, 0.0029 mWh cm−2) [43], ZIF-PPy (0.12 mW cm−2, 0.0113 mWh cm−2) [44], CNTs@Mn-MOF (0.12260 mW cm−2, 0.00699 mWh cm−2) [45], PANI-ZIF-67-CC (0.82 mW cm−2, 0.3396 mWh cm−2) [46], PANI/UiO-66 (0.05 mW cm−2, 0.0197 mWh cm−2) [47], Cu-CAT-NWAs/PPy (0.4 mW cm−2, 0.0224 mWh cm−2) [48]. In a word, the MOFs used in our electrode have played at least three roles: (i) providing the anchoring sites to bond with CNTs and AQ to improve both conductivity and energy storage capability, (ii) affording a highly porous framework to promote electrochemical dynamics, and (iii) separating CNTs and redox functionalities to allow different energy storage mechanisms to work for optimal performance.

3. Experimental Section

3.1. Synthesis of UiO-66-NO2

UiO-66-NO2 was synthesized according to the reported procedure with slight modification [15]. Briefly, ZrCl4 (233 mg), 2-nitroterepthalic acid (233 mg) and 3-aminobenzoic acid (244 mg) were added into DMF (20 mL) and dissolved by the addition of 0.16 mL of concentrated hydrochloric acid. The solution was then transferred into a Teflon-lined stainless-steel autoclave (40 mL) and heated at 120 °C for 2 days. After cooling to room temperature, the reaction mixture was filtered to wash the precipitate with DMF and methanol in sequence three times. The crude product was then stirred in methanol for three days. Further filtration and drying under vacuum at 120 °C for 12 h afforded the title product as a pale-yellow powder (190 mg, 65% yield) (inset photo in Figure 1b).

3.2. Synthesis of UiO-66-AQ

The grafting of redox pendants onto nitro-functionalized MOF via azo linkage was realized using solvothermal reaction in the mixture solution of DMF and water. In our case, UiO-66-NO2 (100 mg) was first dispersed in DMF (10 mL), then 1-nitroanthraquinone (100 mg) and deionized water (30 mL) were added in sequence. The mixture was transferred into a Teflon-lined stainless-steel autoclave and heated at 180 °C for 12 h. The collected dark purple solid was washed several times using DMF until the filtrate became colorless. The crude product was stirred in methanol for three days, and the product (inset photo in Figure 1c) was obtained via filtration and dried under a vacuum at 120 °C for 12 h.

3.3. Preparation of Self-Standing CNT@UiO-66-NO2

Carboxylated single-walled CNTs (80 mg) were dispersed in DMF (16 mL) and sonicated for 30 min. Then, the MOF precursor solution [ZrCl4 (85 mg), 2-nitroterepthalic acid (85 mg), and 3-aminobenzoic acid (89 mg) mixed in DMF (8 mL) with concentrated hydrochloric acid (60 μL)] was added. The mixture was transferred into a Teflon-lined stainless-steel autoclave and heated at 120 °C for 2 days. The suspension was centrifugated three times using DMF. The obtained black powders were dispersed in ethanol and filtered into homogenous films (Figure 1d). A further freeze-thawing readily afforded a flexible and stretchable thin film (ca. 70 μm), which was cut into 1 cm × 1 cm size to use as self-standing electrodes.

3.4. Preparation of CNT@UiO-66-AQ Self-Standing Film Electrode

CNT@UiO-66-NO2 black powder was dispersed into DMF (10 mL), then 1-nitroanthraquinone (100 mg) and deionized water (30 mL) was added in sequence. After sonication for 30 min, the mixture was transferred into a Teflon-lined stainless-steel autoclave and heated to 180 °C for 12 h. The black powders were obtained after centrifugations by using DMF, dichloromethane, and methanol, respectively. The obtained black powders were dispersed in ethanol and filtered into homogenous films (ca. 40 μm) (Figure 1e). Similarly, the flexible CNT@UiO-66-AQ self-standing film electrodes were prepared via freeze-thawing and cut procedures.

3.5. Preparation of Reference Electrode

Electroactive powder materials (UiO-66-NO2 and UiO-66-AQ) were mixed with acetylene black and poly(tetrafluoroethylene) with a weight ratio of 8:1:1. The as-obtained paste was pressed onto carbon cloth. The as-prepared composite membrane was cut into square pieces with 1 × 1 cm size and used directly.

3.6. Assembly of Symmetric Supercapacitors (SSCs)

The coin-cell shape SSCs were fabricated by sandwiching two electrodes with one filter paper as the separator, where aqueous 1 M H2SO4 was used as the electrolytes.

4. Conclusions

In summary, we have successfully designed redox MOFs incorporating CNTs hybrid for high-performance and cycling-stable electrodes for flexible energy storage. The azo-coupling AQ onto UiO-66 MOFs was realized via facile solvothermal reaction. The growth and morphology of crystalline MOFs was finely controlled in the presence of CNTs. The hybrid showcases a three-dimensional interpenetrating network, affording improved porosity and electric conductivity. The as-designed CNT@UiO-66-AQ electrode offers an extended potential window (−0.4~1 V), excellent energy storage (to 302.3 mF cm−2 at 1 mA cm−2), and negligible loss of capacitance over 5000 cycles in aqueous acidic electrolyte. Electrochemical kinetics reveal the interplay of capacitance and redox reaction, accompanying efficient charge and ion transfer. The symmetrical supercapacitor assembled by two CNT@UiO-66-AQ self-standing film electrodes deliver a maximum energy density of 0.037 mW h cm−2 and a highest power density of 10.4 mW cm−2, together with 71.9% capacity retention after 10,000 GCD cycles. Such device performance, to the best of our knowledge, is superior to most analogue MOFs electrode-based SSCs. Our post-synthetic azo-coupling redox pendants approach can be extended to different MOFs skeletons and redox nitroaromatics for the design of high-performance flexible energy storage.

Supplementary Materials

The following Supplementary Materials can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28227479/s1.

Author Contributions

H.Z. conducted most research work, X.W. helped in some characterization, J.Z. helped in manuscript preparation, W.T. supervised the whole project. All authors have read and agreed to the published version of the manuscript.

Funding

This research is funded by the National Natural Science Foundation of China (Grant No. 51861145401) and the Priority Academic Program Development (PAPD) of Jiangsu Higher Education Institutions.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Baumann, A.E.; Burns, D.A.; Liu, B.; Thoi, V.S. Metal-organic framework functionalization and design strategies for advanced electrochemical energy storage devices. Commun. Chem. 2019, 2, 86. [Google Scholar] [CrossRef]
  2. Zheng, S.; Li, Q.; Xue, H.; Pang, H.; Xu, Q. A highly alkaline-stable metal oxide@metal–organic framework composite for high-performance electrochemical energy storage. Natl. Sci. Rev. 2020, 7, 305–314. [Google Scholar] [CrossRef] [PubMed]
  3. Xiao, P.; Bu, F.; Zhao, R.; Aboud, M.F.A.; Shakir, I.; Xu, Y. Sub-5 nm ultrasmall metal–organic framework nanocrystals for highly efficient electrochemical energy storage. ACS Nano 2018, 12, 3947–3953. [Google Scholar] [CrossRef] [PubMed]
  4. Wang, D.-G.; Liang, Z.; Gao, S.; Qu, C.; Zou, R. Metal-organic framework-based materials for hybrid supercapacitor application. Co-ord. Chem. Rev. 2020, 404, 213093. [Google Scholar] [CrossRef]
  5. Qiu, T.; Liang, Z.; Guo, W.; Tabassum, H.; Gao, S.; Zou, R. Metal–organic framework-based materials for energy conversion and storage. ACS Energy Lett. 2020, 5, 520–532. [Google Scholar] [CrossRef]
  6. Mehtab, T.; Yasin, G.; Arif, M.; Shakeel, M.; Korai, R.M.; Nadeem, M.; Muhammad, N.; Lu, X. Metal-organic frameworks for energy storage devices: Batteries and supercapacitors. J. Energy Storage 2019, 21, 632–646. [Google Scholar] [CrossRef]
  7. Hua, Y.; Li, X.; Chen, C.; Pang, H. Cobalt based metal-organic frameworks and their derivatives for electrochemical energy conversion and storage. Chem. Eng. J. 2019, 370, 37–59. [Google Scholar] [CrossRef]
  8. Zhao, Y.; Liu, J.; Horn, M.; Motta, N.; Hu, M.; Li, Y. Recent advancements in metal organic framework based electrodes for su-percapacitors. Sci. China Mater. 2018, 61, 159–184. [Google Scholar] [CrossRef]
  9. Rajak, R.; Saraf, M.; Mobin, S.M. Robust heterostructures of a bimetallic sodium–zinc metal–organic framework and reduced graphene oxide for high-performance supercapacitors. J. Mater. Chem. A 2019, 7, 1725–1736. [Google Scholar] [CrossRef]
  10. Yu, D.; Ge, L.; Wei, X.; Wu, B.; Ran, J.; Wang, H.; Xu, T. A general route to the synthesis of layer-by-layer structured metal organic framework/graphene oxide hybrid films for high-performance supercapacitor electrodes. J. Mater. Chem. A 2017, 5, 16865–16872. [Google Scholar] [CrossRef]
  11. Wei, T.; Zhang, M.; Wu, P.; Tang, Y.-J.; Li, S.-L.; Shen, F.-C.; Wang, X.-L.; Zhou, X.-P.; Lan, Y.-Q. POM-based metal-organic framework/reduced graphene oxide nanocomposites with hybrid behavior of battery-supercapacitor for superior lithium storage. Nano Energy 2017, 34, 205–214. [Google Scholar] [CrossRef]
  12. Xia, W.; Qu, C.; Liang, Z.; Zhao, B.; Dai, S.; Qiu, B.; Jiao, Y.; Zhang, Q.; Huang, X.; Guo, W.; et al. High-performance energy storage and conversion materials derived from a single metal–organic framework/graphene aerogel composite. Nano Lett. 2017, 17, 2788–2795. [Google Scholar] [CrossRef] [PubMed]
  13. Yuan, Y.; Zhou, J.; Rafiq, M.I.; Dai, S.; Tang, J.; Tang, W. Growth of Ni-Mn layered double hydroxide and polypyrrole on bacterial cellulose nanofibers for efficient supercapacitors. Electrochim Acta 2019, 295, 82–91. [Google Scholar]
  14. Zhou, Y.; Wang, X.; Acauan, L.; Kalfon-Cohen, E.; Ni, X.; Stein, Y.; Gleason, K.K.; Wardle, B.L. Ultrahigh-areal-capacitance flexible supercapacitor electrodes enabled by conformal P3MT on horizontally aligned carbon-nanotube arrays. Adv. Mater. 2019, 31, e1901916. [Google Scholar] [CrossRef] [PubMed]
  15. Dai, S.; Han, F.; Tang, J.; Tang, W. MOF-derived Co3O4 nanosheets rich in oxygen vacancies for efficient all-solid-state symmetric supercapacitors. Electrochim. Acta 2019, 328, 135103. [Google Scholar] [CrossRef]
  16. Ran, F.; Xu, X.; Pan, D.; Liu, Y.; Bai, Y.; Shao, L. Ultrathin 2D Metal–Organic Framework Nanosheets In Situ Interpenetrated by Functional CNTs for Hybrid Energy Storage Device. Nano-Micro Lett. 2020, 12, 46. [Google Scholar] [CrossRef]
  17. Zhou, J.; Yuan, Y.; Tang, J.; Tang, W. Metal-organic frameworks governed well-aligned conducting polymer/bacterial cellulose membranes with high areal capacitance. Energy Storage Mater. 2019, 23, 594–601. [Google Scholar] [CrossRef]
  18. Ding, B.; Solomon, M.B.; Leong, C.F.; D’Alessandro, D.M. Redox-active ligands: Recent advances towards their incorporation into coordination polymers and metal-organic frameworks. Co-ord. Chem. Rev. 2021, 439, 213891. [Google Scholar] [CrossRef]
  19. Rambabu, D.; Lakraychi, A.E.; Wang, J.; Sieuw, L.; Gupta, D.; Apostol, P.; Chanteux, G.; Goossens, T.; Robeyns, K.; Vlad, A. An electrically conducting li-ion metal–organic framework. J. Am. Chem. Soc. 2021, 143, 11641–11650. [Google Scholar] [CrossRef]
  20. Gui, B.; Meng, X.; Chen, Y.; Tian, J.; Liu, G.; Shen, C.; Zeller, M.; Yuan, D.; Wang, C. Reversible tuning hydroquinone/quinone reaction in metal–organic framework: Immobilized molecular switches in solid state. Chem. Mater. 2015, 27, 6426–6431. [Google Scholar] [CrossRef]
  21. Jiang, Q.; Xiong, P.; Liu, J.; Xie, Z.; Wang, Q.; Yang, X.; Hu, E.; Cao, Y.; Sun, J.; Xu, Y.; et al. A Redox-Active 2D metal–organic framework for efficient lithium storage with extraordinary high capacity. Angew. Chem. Int. Ed. 2020, 59, 5273–5277. [Google Scholar] [CrossRef] [PubMed]
  22. Chen, S.; Song, Z.; Lyu, J.; Guo, Y.; Lucier, B.E.G.; Luo, W.; Workentin, M.S.; Sun, X.; Huang, Y. Anhydride post-synthetic modification in a hierarchical metal-organic framework. J. Am. Chem. Soc. 2020, 142, 4419–4428. [Google Scholar] [CrossRef] [PubMed]
  23. Yin, Z.; Wan, S.; Yang, J.; Kurmoo, M.; Zeng, M.-H. Recent advances in post-synthetic modification of metal–organic frameworks: New types and tandem reactions. Coord. Chem. Rev. 2019, 378, 500–512. [Google Scholar] [CrossRef]
  24. Wang, X.; Zhou, J.; Tang, W. Poly(dithieno[3,2-b:2’,3’-d]pyrrole) twisting redox pendants enabling high current durability in all-organic proton battery. Energy Storage Mater. 2021, 36, 1–9. [Google Scholar] [CrossRef]
  25. Sheng, L.; Fang, D.; Wang, X.; Tang, J.; Han, Q.; Zhou, J.; Tang, W. Boosting PEDOT energy storage with redox dopant and electrolyte additive. Chem. Eng. J. 2020, 401, 126123. [Google Scholar] [CrossRef]
  26. Wang, X.; Zhou, J.; Tang, W. Emerging polymer electrodes for aqueous energy storage. Mater. Horizons 2021, 8, 2373–2386. [Google Scholar] [CrossRef]
  27. Fang, D.; Zhou, J.; Sheng, L.; Tang, W.; Tang, J. Juglone bonded carbon nanotubes interweaving cellulose nanofibers as self-standing membrane electrodes for flexible high energy supercapacitors. Chem. Eng. J. 2020, 396, 125325. [Google Scholar] [CrossRef]
  28. Fan, W.; Peh, S.B.; Zhang, Z.; Yuan, H.; Yang, Z.; Wang, Y.; Chai, K.; Sun, D.; Zhao, D. Tetrazole-functionalized zirconium metal-organic cages for efficient C2H2/C2H4 and C2H2/CO2 separations. Angew. Chem. Int. Ed. 2021, 60, 17338–17343. [Google Scholar] [CrossRef]
  29. Xie, Y.; Li, D.; Jin, B. In situ FT-IR spectroelectrochemical study of the reduction of 1,4-dinitrobenzene in room-temperature ionic liquids. J. Electroanal. Chem. 2016, 774, 1–6. [Google Scholar] [CrossRef]
  30. Chandra, S.; Kundu, T.; Kandambeth, S.; BabaRao, R.; Marathe, Y.; Kunjir, S.M.; Banerjee, R. Phosphoric acid loaded azo (−N=N−) based covalent organic framework for proton conduction. J. Am. Chem. Soc. 2014, 136, 6570–6573. [Google Scholar] [CrossRef]
  31. Jayaramulu, K.; Horn, M.; Schneemann, A.; Saini, H.; Bakandritsos, A.; Ranc, V.; Petr, M.; Stavila, V.; Narayana, C.; Scheibe, B.; et al. Covalent graphene-MOF hybrids for high-performance asymmetric supercapacitors. Adv. Mater. 2021, 33, 2004560. [Google Scholar] [CrossRef] [PubMed]
  32. Zhao, G.; Zhang, Y.; Gao, Z.; Li, H.; Liu, S.; Cai, S.; Yang, X.; Guo, H.; Sun, X. Dual active site of the azo and carbonyl-modified covalent organic framework for high-performance li storage. ACS Energy Lett. 2020, 5, 1022–1031. [Google Scholar] [CrossRef]
  33. Girardeaux, C.; Druet, E.; Demoncy, P.; Delamar, M. The polyimide (PMDA-ODA) titanium interface. Part 2. XPS study of polyimide treatments and ageing. J. Electron Spectrosc. Relat. Phenom. 1995, 74, 57–66. [Google Scholar] [CrossRef]
  34. Li, Q.; Dai, Z.; Wu, J.; Liu, W.; Di, T.; Jiang, R.; Zheng, X.; Wang, W.; Ji, X.; Li, P.; et al. Fabrication of Ordered Macro-Microporous Single-Crystalline MOF and Its Derivative Carbon Material for Supercapacitor. Adv. Energy Mater. 2020, 10, 1903750. [Google Scholar] [CrossRef]
  35. Yang, S.J.; Choi, J.Y.; Chae, H.K.; Cho, J.H.; Nahm, K.S.; Park, C.R. Preparation and Enhanced Hydrostability and Hydrogen Storage Capacity of CNT@MOF-5 Hybrid Composite. Chem. Mater. 2009, 21, 1893–1897. [Google Scholar] [CrossRef]
  36. Li, Y.; Yang, W.; Yang, W.; Wang, Z.; Rong, J.; Wang, G.; Xu, C.; Kang, F.; Dong, L. Towards High-Energy and Anti-Self-Discharge Zn-Ion Hybrid Supercapacitors with New Understanding of the Electrochemistry. Nano-Micro Lett. 2021, 13, 95. [Google Scholar] [CrossRef] [PubMed]
  37. Peng, H.; Raya, J.; Richard, F.; Baaziz, W.; Ersen, O.; Ciesielski, A.; Samorì, P. Synthesis of Robust MOFs@COFs Porous Hybrid Materials via an Aza-Diels–Alder Reaction: Towards High-Performance Supercapacitor Materials. Angew. Chem. Int. Ed. 2020, 59, 19602–19609. [Google Scholar] [CrossRef]
  38. Cong, H.-P.; Ren, X.-C.; Wang, P.; Yu, S.-H. Flexible graphene–polyaniline composite paper for high-performance supercapacitor. Energy Environ. Sci. 2013, 6, 1185–1191. [Google Scholar] [CrossRef]
  39. Wang, X.; Yang, N.; Li, Q.; He, F.; Yang, Y.; Wu, B.; Chu, J.; Zhou, A.; Xiong, S. Solvothermal synthesis of flower-string-like NiCo-MOF/MWCNT composites as a high-performance supercapacitor electrode material. J. Solid State Chem. 2019, 277, 575–586. [Google Scholar] [CrossRef]
  40. Wen, P.; Gong, P.; Sun, J.; Wang, J.; Yang, S. Design and synthesis of Ni-MOF/CNT composites and RGO/carbon nitride com-posites for an asymmetric supercapacitor with high energy and power density. J. Mater. Chem. A 2015, 3, 13874–13883. [Google Scholar] [CrossRef]
  41. Zhao, J.; Li, Z.; Yuan, X.; Yang, Z.; Zhang, M.; Meng, A.; Li, Q. A high-energy density asymmetric supercapacitor based on Fe2O3 nanoneedle arrays and NiCo2O4/Ni(OH)2 hybrid nanosheet arrays grown on SiC nanowire networks as free-standing advanced electrodes. Adv. Energy Mater. 2018, 8, 1702787. [Google Scholar] [CrossRef]
  42. Qi, K.; Hou, R.; Zaman, S.; Qiu, Y.; Xia, B.Y.; Duan, H. Construction of metal–organic framework/conductive polymer hybrid for all-solid-state fabric supercapacitor. ACS Appl. Mater. Interfaces 2018, 10, 18021–18028. [Google Scholar] [CrossRef] [PubMed]
  43. Wang, L.; Feng, X.; Ren, L.; Piao, Q.; Zhong, J.; Wang, Y.; Li, H.; Chen, Y.; Wang, B. Flexible solid-state supercapacitor based on a metal–organic framework interwoven by electrochemically-deposited PANI. J. Am. Chem. Soc. 2015, 137, 4920–4923. [Google Scholar] [CrossRef]
  44. Hou, R.; Miao, M.; Wang, Q.; Yue, T.; Liu, H.; Park, H.S.; Qi, K.; Xia, B.Y. Integrated Conductive Hybrid Architecture of Metal–Organic Framework Nanowire Array on Polypyrrole Membrane for All-Solid-State Flexible Supercapacitors. Adv. Energy Mater. 2019, 10, 1901892. [Google Scholar] [CrossRef]
  45. Barakzehi, M.; Montazer, M.; Sharif, F.; Norby, T.; Chatzitakis, A. MOF-modified polyester fabric coated with reduced graphene oxide/polypyrrole as electrode for flexible supercapacitors. Electrochim. Acta 2020, 336, 135743. [Google Scholar] [CrossRef]
  46. Xu, X.; Tang, J.; Qian, H.; Hou, S.; Bando, Y.; Hossain, M.S.A.; Pan, L.; Yamauchi, Y. Three-dimensional networked metal–organic frameworks with conductive polypyrrole tubes for flexible supercapacitors. ACS Appl. Mater. Interfaces 2017, 9, 38737–38744. [Google Scholar] [CrossRef]
  47. Zhang, Y.; Lin, B.; Sun, Y.; Zhang, X.; Yang, H.; Wang, J. Carbon nanotubes@metal–organic frameworks as Mn-based symmetrical supercapacitor electrodes for enhanced charge storage. RSC Adv. 2015, 5, 58100–58106. [Google Scholar] [CrossRef]
  48. Shao, L.; Wang, Q.; Ma, Z.; Ji, Z.; Wang, X.; Song, D.; Liu, Y.; Wang, N. A high-capacitance flexible solid-state supercapacitor based on polyaniline and Metal-Organic Framework (UiO-66) composites. J. Power Sources 2018, 379, 350–361. [Google Scholar] [CrossRef]
Figure 1. (a) Schematic route of the post-synthetic grafting AQ pendants onto MOFs with in situ growth guided by carboxylated single-walled CNTs. TEM images of (b) UiO-66-NO2, (c) UiO-66-AQ, (d) CNT@UiO-66-NO2, and (e) CNT@UiO-66-AQ. The insets of (be) show the photos of the corresponding filter cakes.
Figure 1. (a) Schematic route of the post-synthetic grafting AQ pendants onto MOFs with in situ growth guided by carboxylated single-walled CNTs. TEM images of (b) UiO-66-NO2, (c) UiO-66-AQ, (d) CNT@UiO-66-NO2, and (e) CNT@UiO-66-AQ. The insets of (be) show the photos of the corresponding filter cakes.
Molecules 28 07479 g001
Figure 2. (a) XRD patterns of as-synthesized MOFs and their hybrids with CNTs. (b) Conductivity of MOF hybrids with CNTs. (c) FT-IR spectra of 1-nitroanthraquinone (bottom curve), UiO-66-NO2 (middle curve) and UiO-66-AQ (top curve). High-resolution XPS spectra for (d) N 1s and (e) O 1s regions as well as their peak deconvolution (color lines) in CNT@UiO-66-AQ and CNT@UiO-66-NO2.
Figure 2. (a) XRD patterns of as-synthesized MOFs and their hybrids with CNTs. (b) Conductivity of MOF hybrids with CNTs. (c) FT-IR spectra of 1-nitroanthraquinone (bottom curve), UiO-66-NO2 (middle curve) and UiO-66-AQ (top curve). High-resolution XPS spectra for (d) N 1s and (e) O 1s regions as well as their peak deconvolution (color lines) in CNT@UiO-66-AQ and CNT@UiO-66-NO2.
Molecules 28 07479 g002
Figure 3. SEM images of (a) UiO-66-NO2, (b) UiO-66-AQ, (e) CNT@UiO-66-NO2 and (f) CNT@UiO-66-AQ. TEM images of (c) UiO-66-NO2 and (d) UiO-66-AQ.
Figure 3. SEM images of (a) UiO-66-NO2, (b) UiO-66-AQ, (e) CNT@UiO-66-NO2 and (f) CNT@UiO-66-AQ. TEM images of (c) UiO-66-NO2 and (d) UiO-66-AQ.
Molecules 28 07479 g003
Figure 4. N2 adsorption and desorption isotherms for (a) UiO-66-NO2 and UiO-66-AQ and (c) CNT@UiO-66-NO2 and CNT@UiO-66-AQ, and pore size distributions for (b) UiO-66-NO2 and UiO-66-AQ and (d) CNT@UiO-66-NO2 and CNT@UiO-66-AQ.
Figure 4. N2 adsorption and desorption isotherms for (a) UiO-66-NO2 and UiO-66-AQ and (c) CNT@UiO-66-NO2 and CNT@UiO-66-AQ, and pore size distributions for (b) UiO-66-NO2 and UiO-66-AQ and (d) CNT@UiO-66-NO2 and CNT@UiO-66-AQ.
Molecules 28 07479 g004
Figure 5. (a) TEM and (b,c) HRTEM images CNT@UiO-66-AQ. (d) TEM image of CNT@UiO-66-NO2. (e) HAADF image of CNT@UiO-66-AQ and (fi) corresponding element mapping of C, Zr, N and O, respectively.
Figure 5. (a) TEM and (b,c) HRTEM images CNT@UiO-66-AQ. (d) TEM image of CNT@UiO-66-NO2. (e) HAADF image of CNT@UiO-66-AQ and (fi) corresponding element mapping of C, Zr, N and O, respectively.
Molecules 28 07479 g005
Figure 6. Electrochemical performance of CNT@UiO-66-NO2 and CNT@UiO-66-AQ in three-electrode electrochemical cells with 1 M H2SO4 aqueous electrolyte. (a) CV curves at a scan rate of 30 mV s−1. (b) The relationship between logarithm oxidation peak current and logarithm scan rate. (c) Variation of specific capacitance against the current densities. (d) Variations and fittings between Zre and the reciprocal square root of the angular frequency in the low-frequency region of EIS plots. (e) Cycling performance of CNT@UiO-66-AQ at 5 mA cm−2.
Figure 6. Electrochemical performance of CNT@UiO-66-NO2 and CNT@UiO-66-AQ in three-electrode electrochemical cells with 1 M H2SO4 aqueous electrolyte. (a) CV curves at a scan rate of 30 mV s−1. (b) The relationship between logarithm oxidation peak current and logarithm scan rate. (c) Variation of specific capacitance against the current densities. (d) Variations and fittings between Zre and the reciprocal square root of the angular frequency in the low-frequency region of EIS plots. (e) Cycling performance of CNT@UiO-66-AQ at 5 mA cm−2.
Molecules 28 07479 g006
Figure 7. EIS plots of CNT@UiO-66-NO2 and CNT@UiO-66-AQ.
Figure 7. EIS plots of CNT@UiO-66-NO2 and CNT@UiO-66-AQ.
Molecules 28 07479 g007
Figure 8. Electrochemical performance of CNT@UiO-66-AQ SSCs. (a) CV curves recorded at different scan rates. (b) GCD curves measured at different current densities. (c) EIS plots. (d) Ragone plots in comparison to representative high-performance MOF-based SSCs in the literature.
Figure 8. Electrochemical performance of CNT@UiO-66-AQ SSCs. (a) CV curves recorded at different scan rates. (b) GCD curves measured at different current densities. (c) EIS plots. (d) Ragone plots in comparison to representative high-performance MOF-based SSCs in the literature.
Molecules 28 07479 g008
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhang, H.; Wang, X.; Zhou, J.; Tang, W. Azo-Linkage Redox Metal–Organic Framework Incorporating Carbon Nanotubes for High-Performance Aqueous Energy Storage. Molecules 2023, 28, 7479. https://doi.org/10.3390/molecules28227479

AMA Style

Zhang H, Wang X, Zhou J, Tang W. Azo-Linkage Redox Metal–Organic Framework Incorporating Carbon Nanotubes for High-Performance Aqueous Energy Storage. Molecules. 2023; 28(22):7479. https://doi.org/10.3390/molecules28227479

Chicago/Turabian Style

Zhang, Hualei, Xinlei Wang, Jie Zhou, and Weihua Tang. 2023. "Azo-Linkage Redox Metal–Organic Framework Incorporating Carbon Nanotubes for High-Performance Aqueous Energy Storage" Molecules 28, no. 22: 7479. https://doi.org/10.3390/molecules28227479

Article Metrics

Back to TopTop