Tapan
Dey
a,
Nitish
Kumar
be,
Rahul
Patil
a,
Prakash Kumar
Pathak
b,
Sudip
Bhattacharjee
c,
Praveen
Yadav
d,
Asim
Bhaumik
c,
Rahul R.
Salunkhe
*b and
Saikat
Dutta
*a
aElectrochemical Energy & Sensor Research Laboratory, Amity Institute of Click Chemistry Research and Studies, Amity University Noida, Uttar Pradesh 201313, India. E-mail: sdutta2@amity.edu
bMaterials Research Laboratory, Department of Physics, Indian Institute of Technology Jammu, Jammu and Kashmir, India. E-mail: rahul.salunkhe@iitjammu.ac.in
cSchool of Materials Science, Indian Association for Cultivation of Science, Jadavpur, Kolkata, India
dRaja Ramanna Centre for Advanced Technology (RRCAT), Dept. of Atomic Energy, Indore 452013, MP, India
eDepartment of Industrial and Materials Science, Chalmers University of Technology, Gothenburg, SE-41296, Sweden
First published on 8th January 2025
Electrochemical supercapacitors and the electrochemical oxidation of biomass-derived oxygenates have great significance for long-term high-performance devices. However, appropriate sites with redox features remain a bottleneck for electrochemical oxidation and capacitance retention. Herein, N-doped carbon sheets with Mn-phosphate-doping and Co-metal nanoparticles were synthesized via a facile one-pot activation and calcination of the layered potassium phthalimide salt without inclusion of any additional activators or template. The unique 2D-structure of the obtained microporous carbon flakes with a layered structure provides a sturdy N-C matrix for prolonged charging/discharging with abundant active adsorption sites and an effective route for rapid electrolyte ion transport with a shorter diffusion distance for the adsorption/desorption of ions. Through these merits, K-Ph-NC offers high capacitance and outstanding rate performance with an incredible energy density in capacitor devices, and the specific capacitance of the as-prepared K-Ph-NC is proportional to the number of micropores. K-Ph-NC was further transformed to a K-Ph-Oxide, a graphene oxide version of K-phthalimide, by using an improved Hummer's method by using Mn-salt and phosphoric acid, which resulted in a phthalimene oxide doped with Mn-phosphate. In addition, a composite of K-Ph-NC with ZIF-67 was thermally calcined at 700 °C under an Ar atmosphere, which resulted in e-ZIF-67/K-Ph-NC with an etched surface. A comparative electronic and structural analysis followed by a capacitance retention and electrochemical oxygen evolution reaction study revealed the role of Co-nanoparticles as compared to the Mn-phosphate doping in the resulting materials. A symmetric supercapacitor device exhibited a maximum SE value of 22.7 W h kg−1 with a maximum SP of 10
416.7 W kg−1, which is mainly due to the favorable microporous pore architecture in e-ZIF-67/K-Ph-NC as compared to K-Ph-NC and K-Ph-Oxide. This highlights the role of cobalt nanoparticles in e-ZIF-67/K-Ph-NC with an etched outer surface. A promising overpotential of 450 mV at 10 mA cm−2 in the OER by e-ZIF-67/K-Ph-NC can be correlated to the charge transfer resistance across the electrode–electrolyte interface.
Herein, we demonstrate a series of N-doped carbon sheets with Mn-phosphate and Co-metal nanoparticle (CoNP) doping via a facile one-pot activation and calcination of the layered potassium phthalimide salt without inclusion of any additional activators or templates (Fig. 1). The optimal K-Ph-NC sample produces a large 2D nanosheet with a uniform sub-nanometer microporous distribution of roughly 1.06 nm and an immense specific surface area of 1060 m2 g−1. A robust N-C matrix for extended charging/discharging with a large number of active adsorption sites and a successful pathway for quick electrolyte ion transport with a reduced diffusion distance for ion adsorption/desorption are provided by the special microporous carbon flakes with layered architectures. Through these merits, K-Ph-NC offers high capacitance and outstanding rate performance with incredible energy density in capacitor devices with the specific capacitance proportional to the number of micropores at 0.80 nm. This established a relation between the capacitive performance and critical pore size below 1 nm. The high surface area K-Ph-NC was further transformed to a K-Ph-Oxide, a graphene version of K-phthalimide, by using an improved Hummer's method with Mn-salt and phosphoric acid, which resulted in a phthalimene oxide (K-Ph-Oxide) doped with Mn-phosphate. To explore the ability of K-Ph-NC to be functionalized with transition metal nanoparticles, a composite with ZIF-67 was made by calcination at 700 °C under an Ar atmosphere to give an etched e-ZIF-67/K-Ph-NC (Fig. 1). The corresponding symmetric supercapacitor device exhibited a maximum SE value of 22.7 W h kg−1 with a maximum SP of 10
416.7 W kg−1, which is mainly due to the favorable microporous pore architecture in e-ZIF-67/K-Ph-NC as compared to K-Ph-NC and K-Ph-Oxide. This highlights the role of cobalt nanoparticles in e-ZIF-67/K-Ph-NC with an etched outer surface. A promising overpotential of 450 mV at 10 mA cm−2 in the OER by e-ZIF-67/K-Ph-NC can be correlated to the charge transfer resistance across the electrode–electrolyte interface.
The X-ray diffraction (XRD) pattern of K-Ph-NC showed graphitic planes at 2θ 26.7° and 44.1° corresponding to the (001) and (100) planes, respectively. The XRD pattern reveals that the K-Ph-Oxide consists of manganese phosphate planes of (002), (221), (302), (003), (402), and (512) at 2θ of 18.14°, 19.01°, 25.31°, 27.16°, 29.95°, and 35.61°, respectively (JCPDS No. 029-0893) (Fig. 2b). The presence of Co metallic particles is confirmed from the (200), (220), (400), (420), (422), (440), (600) and (620) planes at 2θ of 17.35°, 24.58°, 35.04°, 39.34°, 44.54°, 50.51°, 53.59°, 56.74°, respectively (JCPDS No. 024-0327) (Fig. 2c). The graphitic carbon planes for K-Ph NC are confirmed at 23.44° and 44.14° (Fig. 2c). However, the nature of the XRD plots for K-Ph NC and e-ZIF-67/K-Ph-NC are similar. The peaks of the Co nanoparticles in e-ZIF-67/K-Ph-NC can be distinguished from those of K-Ph-NC. N2 adsorption–desorption analysis of K-Ph-NC revealed a high BET surface area of 1060 m2 g−1 resulting from the K-phthalimide pyrolysis, which forms an extended 2D-surface of N-doped C (Fig. 2d) with a significant microporous network as revealed from the pore width versus pore volume profile (Fig. 2e). The pore analysis suggests the presence of significant micropores in K-Ph-NC with an average micropore size of 1.06 nm. Therefore, metal center functionalization on both of the above materials would give a microporous network that facilitates many electrochemical processes. Moreover, ex situ SEM and Raman spectral investigation after 250 cycles of galvanometric charge–discharge reveals the extent of the intensity drop of the e-ZIF-67/K-Ph-NC. But there is no significant change in the d-band and g-band shifts. Even though the material starts degrading after 250 cycles, there are no morphological changes as revealed from the scanning electron microscopy (SEM) analysis (Fig. S4a–f†) of e-ZIF-67/K-Ph-NC before cycling and after 250 cycles of CV, and the ex situ SEM images of e-ZIF-67/K-Ph-NC after 250 cycles (Fig. S4g–k†).
Mn3+ centers are decorated in K-Ph-Oxide as confirmed by soft X-ray absorption spectroscopy measurement. The peak at the Mn-L3 edge confirmed the presence of the Mn3+ oxidation state in K-Ph-Oxide, which represents Mn-phosphate (Fig. 2f).25,26 The incorporation of Mn-phosphate in the carbon framework is supported by the XRD pattern (Fig. 2b). Survey spectra of K-Ph-Oxide are shown in Fig. 3a. The C 1s XPS deconvoluted spectrum reveals the presence of three kinds of C centers, two of which are oxygen-containing C on the surface of K-Ph-Oxide (Fig. 3b). The N 1s XPS spectral pattern suggests that both pyrrolic and pyridinic Ns are available in K-Ph-NC in a 60
:
40 ratio approximately (Fig. 3c). The comparison of the elemental atom% composition suggests a difference in the C% of K-Ph-Oxide compared to K-Ph-NC. Moreover, the 532 eV peak in the deconvoluted O K-edge XPS spectrum confirms the presence of a graphene-like oxide structure (Fig. 3d).26
XPS was employed to further investigate the elemental composition and electronic structure of the e-ZIF-67/K-Ph-NC. The survey spectrum (Fig. 3e) of e-ZIF-67/K-Ph-NC revealed the coexistence of C, Co, N, and O elements. The deconvoluted C 1s spectra (Fig. 3f) support the presence of sp2 and sp3 carbon atoms with the detection of Co2+ 2p, proving the successful formation of Co NPs (Fig. 3g). N 1s spectrum deconvoluted curve fitting shows the pyrrolic and pyridinic N in the framework of e-ZIF-67/K-Ph-NC (Fig. 3h). The O 1s spectrum (Fig. 3i) showed overlaps of C–O and C
O groups.
The high-resolution transmission electron microscopy (HR-TEM) image of K-Ph-Oxide shows a CNT-type multiwalled architecture that originates from K+-ion based etching at 700 °C, producing a 2-D network of oxide functionality resembling graphene oxide. The process of K-Ph-Oxide formation from K-Ph (Fig. 1) by the improved Hummer's method resembles graphene oxide formation from graphite.24 It is envisaged that the intra-molecular chemical interaction between the oxygen sp2 lone pair (lp) and carbon π orbitals is the structural origin for the stabilization of K-Ph-Oxide.27 The high-resolution TEM image (Fig. 4a) for K-Ph-Oxide shows multiwall-type channels as in multiwalled carbon nanotubes, which suggests a tubular multiwalled structure. The unique 2D-structure of K-Ph-Oxide is clear from the TEM images (Fig. 4b–d) at various resolutions. A certain variation of the density of dark spots corresponding to the carbon nanotube-like architecture was found at all resolutions, confirming a non-uniform distribution of this architecture throughout the entire matrix of K-Ph-Oxide. The TEM images of e-ZIF-67/K-Ph-NC (Fig. 4e–h) show a box-type shape of ZIF-67-NC particles originating from ZIF-67, wherein 2D-sheet-like K-Ph-NC is clear. At slightly higher resolution at the 100 nm scale, partially occupied K-Ph-NC was seen with e-ZIF-67 particles (Fig. 4g). The dark-field TEM image (Fig. 4h) confirms the presence of e-ZIF-67-NC in the composite, wherein the ZIF-67-NC surface is clear as compared to the bright field images (Fig. 4e–g).
The comparative CV profiles for a set of samples at a scan rate of 100 mV s−1 are shown in Fig. 5a. As evident from the comparative CV shapes, the e-ZIF-67/K-Ph-NC exhibited a more enclosed area under the CV curve, explaining its improved electrochemical activity.28 The CV curves of e-ZIF-67/K-Ph-NC at varying scan rates from 5 to 200 mV s−1 are shown in Fig. 5b, while for samples K-Ph-NC and K-Ph-Oxide they are shown in Fig. S4a and c.† e-ZIF-67/K-Ph-NC exhibited a very high capacitance of 385.4 F g−1 at a scan rate of 5 mV s−1, while K-Ph-NC and K-Ph-Oxide exhibited a capacitance of 242 and 185.9 F g−1, respectively. The capacitance values for these three samples at different scan rates are displayed in Fig. 5c. The e-ZIF-67/K-Ph-NC and K-Ph-NC samples exhibited good retention at higher scan rates as well. In contrast, the performance off the K-Ph-Oxide sample degraded drastically at higher scan rates, owing to its low surface area. Interestingly, the e-ZIF-67/K-Ph-NC sample showed high performance compared to different carbon-based materials, as tabulated in Table S1.† Moreover, this improved performance for e-ZIF-67/K-Ph-NC can also be explained by its very low charge transfer resistance, as studied by impedance spectroscopy (Fig. S5†). The e-ZIF-67/K-Ph-NC sample exhibited a small bulk resistance of 0.59 Ω and a negligible charge transfer resistance of 0.07 Ω compared to 0.6 Ω (bulk) and 0.24 Ω (charge transfer) for K-Ph-NC. The highly reduced charge transfer resistance further supports the improved supercapacitive performance of the e-ZIF-67/K-Ph-NC sample. Further insights into the charge storage mechanism were acquired by utilizing the power law and Dunn's method for the e-ZIF-67/K-Ph-NC sample, as displayed in Fig. S6.† The “b” value calculated using the power law (i = avb) exhibited a value of 0.82, which indicates that both electric double layer capacitance (EDLC) and pseudocapacitance (Fig. S6a†) processes are involved. Moreover, the contribution of these two processes at different scan rates is plotted in Fig. S6b.† The EDLC capacitance increases on increasing the scan rate, which is consistent with earlier reports.29,30
Following the assessment of the three-electrode performance, the e-ZIF-67/K-Ph-NC substance was extended to examine the packed two-electrode performance. The active material was coated over carbon cloth, and then packed in a symmetric configuration using filter paper as the separator and 2 M tetraethylammonium tetrafluoroborate in propylene carbonate (PC) as the electrolyte.31 The voltage window for the device was finalized by performing CV curves at 100 mV s−1 at different voltages varying from 1 V to 3.4 V, as displayed in Fig. 6a. The voltage window was fixed at 3 V, and further CV and GCD performances were recorded. The CV curves at different scan rates for the assembled symmetric supercapacitor device (SSD) are shown in Fig. 6b. As anticipated, the device exhibited a nearly rectangular CV shape, indicating that most of the charge is being stored by the electrostatic adsorption–desorption of electrolyte ions.
Further, the GCD curves at different current densities, namely 0.033, 0.05, 0.1, 0.2, 0.3, 0.5, 1, 3, and 5 A g−1, are displayed in Fig. 6c. The material exhibited triangular charge–discharge curves without any visible IR drop. Based on the capacitance values calculated from the GCD curves, the specific energy (SE) and specific power (SP) values were calculated and plotted in the Ragone plot. A comparison of different two-dimensional carbon materials for SSDs with our device is shown in the Ragone plot in Fig. 6d. These reported materials include carbon sheets,32 two-dimensional graphitic carbon,33 carbon nanosheets,34 and sponge-like graphene sheets.35 Interestingly, our device exhibited a maximum specific energy (SE) value of 22.7 W h kg−1 with a maximum specific power (SP) of 10
416.7 W kg−1 (Fig. 6d). A possible reason for the excellent performance of e-ZIF-67/K-Ph-NC can be attributed to the 2D morphology and etched outer surface, along with high surface area. Moreover, the microporous pore architecture in e-ZIF-67/K-Ph-NC plays a major role in enhancing efficiency as compared to other ZIF-67 and K-phthalimide-derived composites. The symmetric supercapacitor device was connected in series, and could provide power to a 1.5 V red color light emitting diode (LED) with high brightness. This supports the experimental data of energy storage. The images of the LED glowing when connecting two SSDs in series (Fig. S8†) suggest its real field application towards supercapacitor devices.
A strong redox feature with an E1/2 value of 1.28 V (vs. RHE) was observed in the CV cycles for e-ZIF-67/K-Ph-NC, which can be ascribed to the oxidation of Co embedded on the e-ZIF-67/K-Ph-NC framework. There was an increment of 0.2 mA current in the highest current density with the progress of the CV cycles from the 1st to 30th cycle, accompanied by a constant peak position of the redox couple after cycle 1. The cathodic shift of the redox feature after the 1st cycle indicates the cooperative activation of the catalyst. Linear sweep voltammetry (LSV) was recorded within a potential range of 1.0–2.2 V (vs. RHE) with e-ZIF-67/K-Ph-NC after the activation of the electrode (by 10 CV cycles at 5 mV s−1). The redox peak within the 1.2 to 1.4 V range in the LSV curves indicates the in situ formation of Co(III)-species, followed by a steady increase in the catalytic current after 1.5 V (vs. RHE), which reached up to 80 mA cm−2 at 2.2 V (vs. RHE) indicating a catalytic OER (Fig. 7a). The best activity was obtained with 1 mg of mass loading, which implied monolayer formation on the electrode surface to achieve enhanced OER performance. The electrode shows a promising and better overpotential of 450 mV compared to K-Ph-NC at 10 mA cm−2 towards an effective electrocatalyst. The improved OER activity of e-ZIF-67/K-Ph-NC can be correlated to the charge transfer resistance across the electrode–electrolyte interface. The lower Tafel value in e-ZIF-67/K-Ph-NC reflects its better charge transfer kinetics during reaction and efficient electrocatalyst than K-Ph-NC (Fig. 7b). An electrochemical impedance spectroscopic (EIS) study was performed at a constant potential of 1.54 V (vs. RHE) and the semicircular Nyquist plot and corresponding equivalent circuit fitting provided the charge-transfer resistance (Rct) of the electrode and indicated that the electrode–electrolyte junction follows a double layer formation (Fig. 7c). The Rct value obtained for e-ZIF-67/K-Ph-NC was 41.48 Ω, indicating the rapid charge transfer kinetics across the electrode–electrolyte junction, which greatly influences the electrocatalytic activity. The better catalytic performance of e-ZIF-67/K-Ph-NC for the OER and supercapacitor activity is mainly attributed to the large specific surface area and good electrical conductivity. Besides faster kinetics, another crucial factor that greatly influences the catalyst performance is the number of active sites participating in the reaction, which together constitute the electrochemically active surface area (ECSA). The ECSA is directly related to the double-layer capacitance (Cdl) value of the material. The Cdl can be calculated from the non-faradaic current response of the electrode (Fig. S8†). The ECSA can be determined from the Cdl value and the specific capacitance (Cs) of the electrode being investigated. The obtained Cdl value for e-ZIF-67/K-Ph-NC is 0.414 mF, which corresponds to an ECSA value of 20.7 cm2 (Fig. S8a†). The ECSA normalized activity was also in good agreement with the intrinsic OER activity of the catalyst.
416.7 W kg−1, which is mainly due to the favorable microporous pore architecture in e-ZIF-67/K-Ph-NC as compared to K-Ph-NC and K-Ph-Oxide. This highlights the role of cobalt nanoparticles in e-ZIF-67/K-Ph-NC with an etched outer surface. A promising overpotential of 450 mV at 10 mA cm−2 in the OER by e-ZIF-67/K-Ph-NC can be correlated to the charge transfer resistance across the electrode–electrolyte interface.
Footnote |
| † Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4se00979g |
| This journal is © The Royal Society of Chemistry 2025 |