Defect-Driven Adsorption of NiPc-based Catalyst on CNTs
We chose three CNTs for our investigations, GCNT with a high graphitized degree (GCNT-H), GCNT with a low graphitized degree (GCNT-L), and nGCNT. TEM and scanning electron microscope (SEM) images (Supplementary Fig. 2) revealed that all the CNTs exhibit a similar physical nature, with negligible differences in the CNT morphology, diameter, and size. Besides, the morphology of nGCNT and GCNTs remained after the immobilization of NiTAPc (Supplementary Fig. 3). The successful immobilization of NiTAPc was confirmed by elemental energy dispersion spectroscopy (EDS, Supplementary Fig. 4). Notably, high-angle annular dark-field (HAADF) scanning transmission electron microscope (STEM) from aberration-corrected (AC)-HRTEM images revealed a notable finding regarding the dispersion of NiTAPc on different CNTs. As shown in Fig. 1a, GCNT-H exhibits the best-graphitized surface among the three CNTs, which is expected to be the most optimized surface for π-π interactions. However, we observed that NiTAPc preferentially absorbed at the less perfect graphitized domains of GCNT-H (i.e., intersections), contradicting the commonly held perception that strong π-π interaction promotes the dispersion of NiPc catalysts. A similar phenomenon was also observed for NiTCPc and CoTAPc on GCNT-H (Supplementary Fig. 5). Instead, much more uniform dispersions of NiTAPc were observed on GCNT-L and nGCNT (Fig. 1b-c). Note, we observed comparable dispersion of metal Pc on nGCNT with relevant literature work24,32,33. Besides, the loadings of Ni on GCNT-H, GCNT-L, and nGCNT were determined to be 0.22%, 0.22%, and 0.21% in weight percentage, respectively, using inductively coupled plasma optical emission spectrometer (ICP-OES, Supplementary Table 1). These comparable Ni loadings exclude the possibility of that the observed difference in absorption behaviors is due to the differences in catalyst loading. Thus, we propose that the defects on the CNT surface are primarily responsible for the absorption of molecular catalysts such as NiPc.
Raman spectra were recorded to investigate the intrinsic properties of the different CNTs. As shown in Fig. 1d, the GCNT-H exhibits the smallest ID/IG ratio, validating that the process of graphitization has significantly reduced the defect level of CNTs35,36. After the immobilization of NiTAPc, the representative D and G bands in Raman spectra remained unchanged (Supplementary Fig. 6), indicating no chemical changes occurred to the CNTs during the loading of molecular catalysts. Besides, the results of X-ray photoelectron spectroscopy (XPS) (Supplementary Fig. 7) indicate the same valence state for the Ni atoms in these three NiTAPc/CNTs and free NiTAPc. Overall, we can conclude that NiPc-based catalysts predominate interact with the CNT via defects rather than π-π stacking, and these interactions are likely moderate based on our simulations.
Effects of CNT Defects on Electrochemical CO2R
We first examined the CO2R performance of these NiTAPc/CNTs catalysts using a conventional H-cell. Since CO and H2 were the only two detectable products in our systems, we introduced the CO/H2 molar ratio as an indicator for the selectivity of CO2R. We believe this is a more effective method to reveal the intrinsic CO2R selectivity of a given catalyst, as compared to the conventionally reported Faradaic efficiency (FE). For instance, a 99% FE of CO does not warrant a low HER FE due to the experimental error-induced uncertainties that could cause total FE to exceed 100%. Notably, NiTAPc/GCNT-H outperformed the other two catalysts significantly (Supplementary Fig. 8), both in terms of selectivity and activity towards CO. NiTAPc/GCNT-H reached a superior CO/H2 molar ratio of nearly 1900:1, and exhibited substantially higher CO partial current densities at identical cathodic potentials compared to the NiTAPc/nGCNT electrode. Next, we assessed the CO2R performance of these catalysts in a gas diffusion electrode (GDE) based flow cell (Supplementary Fig. 9) under industrially relevant current densities. Note, we collected gaseous products from both the gas and cathode chambers to avoid underestimating the H2 evolution reaction (HER), especially in cases of minor flooding37,38. Nevertheless, we have also reported all the measured FEs in our study (Supplementary Table 5). Additionally, we conducted meticulous calibration of the gas chromatograph (GC) to ensure the accuracy of the measurements towards a trace amount of H2 (Supplementary Fig. 10–11). As shown in Fig. 1e, NiTAPc/GCNT-H exhibits a remarkable 6000:1 CO/H2 molar ratio and an overpotential nearly 400 mV lower than that of NiTAPc/nGCNT (with a CO/H2 molar ratio of only ~ 580) at 300 mA cm− 2, under typical CO2R conditions. NiTAPc/GCNT-H retained its exceptionally high selectivity towards CO even at higher current densities of 400 and 500 mA cm− 2 (Supplementary Fig. 12). In contrast, significant HER took place on NiTAPc/nGCNT at current densities of 400 mA cm− 2 or higher, consistent with the above H-cell result and previous literature reports24,39,40. Meanwhile, NiTAPc/GCNT-L shows moderate CO2R selectivity and overpotential, further confirming this strong correlation between CO formation selectivity/activity and the defect level of the CNTs supports. We further assessed our catalyst compositions in a membrane-electrode-assembly (MEA) reactor at ampere-level current. As shown in Fig .1e, NiTAPc/GCNT-H again exhibits superior CO selectivity and reduced cell voltage. In contrast, the NiTAPc/nGCNT counter sample exhibits much lower CO selectivity and larger cell voltage, especially at elevated current density of 400 mA cm− 2 (Supplementary Fig. 12d).
Metal residues and amorphous carbon species are generally used as catalysts in preparing nGCNT. To exclude their influence on our electrochemical testing, the pristine nGCNT was purified before use (see Method section). We believe that most metal residues were removed by acid washing, and any remaining metal impurities are encapsulated in CNTs and thus inaccessible during electrocatalysis (Supplementary Table 3). Supplementary Fig. 13a-b show that the HER kinetics of purified nGCNT resemble those of GCNTs but are in sharp contrast to that of pristine nGCNT, indicating that no obvious electrochemically exposed metal exists in purified nGCNT. The introduction of heteroatoms via acid washing may also influence the catalyst/substrate interactions33. Supplementary Fig. 13c-e and the corresponding note indicate that our treatment has a negligible effect on pristine CNT nature, further excluding the possibility that different behavior of NiTAPc on nGCNT and GCNT-H corresponds to oxygen-group interaction.
Moreover, we tested the CO2R performance of CoTAPc/GCNT-H and CoTAPc/nGCNT and compared them with NiTAPc counterparts (Supplementary Fig. 14). The much lower CO/H2 molar ratio of CoTAPc compared to the NiTAPc, supporting our hypothesis that HER primarily arises from the metal center in the molecular catalysts rather than the CNTs substrates, at least under the CO2R conditions. Finally, we conducted Ar-feed electrolysis on the NiTAPc/CNTs. Supplementary Fig. 15 shows all three catalysts exhibit higher overpotential for HER than for CO2R, with the potential trend being opposite to that of CO2R, and NiTAPc/GCNT-H shows the largest overpotential increase. The well-matched overpotential trend and CO2R selectivity reveal the high intrinsic activity of NiTAPc for CO2R over HER, resulting in a much lower overpotential for CO2R than that of HER at the same current densities. The lowest HER overpotential of NiTAPc/nGCNT further indicates its highest HER activity, which is also congruent with its poorest CO selectivity. In all, through these careful control experiments, we confirm that the above remarkable CO formation selectivity/activity originates from the defect engineering of the CNT supports rather than the presence of other impurities.
Previously, it has been reported that electron-donating groups functionalized NiPc/CoPc can enhance CO2R selectivity and stability. In contrast, electron-withdrawing groups functionalized NiPc/CoPc can reduce the overpotential of CO2R, however, at the expense of the catalyst stability, especially under large current densities (> 200 mA cm− 2).24,41. On the contrary, some other studies suggested that the latter can improve both the activity and selectivity of CO2R41,42. Based on the above findings, we hypothesize that the strength of the interaction between the Ni-Pc-based catalysts and CNT substrates must also play a critical role, which has been considered previously. Hence, we also assessed NiTCPc/CNTs, which involves an electron-withdrawing group functionalized NiPc, for CO2R under identical conditions to validate our hypothesis. Notably, we observed that, compared to NiTCPc/nGCNT, both NiTCPc/GCNT-H and NiTCPc/GCNT-L are capable of achieving high CO/H2 molar ratios (> 2000) and significantly lower overpotentials even at high current densities of 400–500 mA cm− 2 (Supplementary Fig. 16 and Supplementary Table 5), further demonstrating the crucial role of the defect level on CNTs as catalyst supports. Taken together, we believe that the interactions between Ni-Pc-based catalysts and CNTs do not primarily result from the π-π interactions. Instead, the proper defect level on CNTs plays a crucial role in tuning the catalytic performance, and these defects might have stronger interaction with the molecular catalysts and influence their structures and catalytic performance during electrocatalysis.
Exploring the effects of Catalyst/CNTs Interactions during CO2R
We first employed the technique of cyclic voltammetry (CV) to investigate the differences in redox behaviors of NiPc-based catalysts on GCNTs and nGCNT. As shown in Fig. 2a, in N2 saturated electrolyte, the two reduction peaks at 0.75 V and 0.17 V vs. RHE could be assigned to the one-electron transfer process of NiTAPc− and NiTAPc2−, respectively44,45. Notably, there is a diminishing trend in the intensity of the redox peaks in the CVs, ranging from GCNT-H to nGCNT. Additionally, this trend is consistent with that of NiTCPc (Fig. 2b), when switching the substrates from GCNT-H to nGCNT. The substantially reduced redox peak intensity of NiPc molecules on nGCNT suggests that the intrinsic redox properties of the NiPc molecules are diminished upon immobilization on the defects of CNT. This difference might be associated with enhanced interactions between the molecular catalysts and the CNT substrate, as demonstrated in recent literature46.
We employed X-ray absorption spectroscopy (XAS) to investigate the coordination environment of the Ni active center in the phthalocyanine plane on CNTs, especially during the catalysis47. Take NiTAPc as an example, its ex-situ X-ray absorption near-edge structure (XANES) and R-space extended X-ray fine structure (EXAFS) spectra on different CNTs are presented in Fig. 2c and Supplementary Fig. 20. Clearly, the NiN4 active-centers in the three catalyst-composites exhibit similar coordination states. The arrow-labeled region, positioned at approximately 8351 eV and likely associated with the 1s→4px, y transition, shifts to lower energy as we move from NiTAPc/GCNT-H to NiTAPc/nGCNT. This shift confirms the deviation from perfect planar structure48–50. NiTCPc counterparts also exhibit similar trend (lower panel, Fig. 2c). To monitor the structural changes/evolutions of the NiN4 centers under real CO2R conditions, we employed the same flow cell used for assessing the CO2R performance to conduct in-situ electrochemical XAS (Supplementary Fig. 21). Encouragingly, we managed to obtain robust XAFS data under practical relevant current densities (i.e., 70 ~ 200 mA cm− 2). Note, due to the significantly low loading and the single-atom nature of the Ni-catalysts (Supplementary Table 1), we believe that the XAFS can accurately represent the authentic state of the active sites during real CO2R electrolysis.
Specifically, we focus on analyzing three moieties of the XANES of these NiTXPc/CNTs (Fig. 2d, 2g). The moieties of a (~ 8333 eV) and b (~ 8339 eV) correspond to the pre-edge associated with quadrupole-allowed 1s → 3d − 4p and dipole-allowed 1s → 4pz shakeup transition, respectively. These characteristics are the fingerprints of a typical square-planar NiN4 with D4h symmetry51–56. Moiety c (~ 8351 eV) within the white line region corresponds to the 1s→4px, y transition as mentioned above. As shown in Fig. 2e, the intensity of a, b, and position of c of the XANES spectra of the NiTAPc/GCNT-H during CO2R only exhibit subtle deviations. This indicates that the initial planar symmetry of the NiN4 is mostly retained during CO2R under a small current density (~ 145 mA cm− 2). When the current density was increased to 230 mA cm− 2 under elevated cathodic potential, the increased intensities of pre-edge a and peak c in the XANES spectra revealed a more diminished D4h symmetry of the NiN4 moiety. Considering our previous hypothesis that the interaction between molecule and defect is predominant, we can conclude that the breaking of D4h symmetry arises from the axial distortion of the Ni-N4 out-of-plane towards the substrate.
On the other hand, much more pronounced changes in the D4h symmetry of Ni-N4 were observed for NiTAPc/nGCNT under the same conditions (lower panel, Fig. 2d), as evidenced by the significantly increased intensity of pre-edge peak a, diminished intensity of peak b and deviated energy of moiety c (Fig. 2e). Additionally, the oxidation state of Ni slightly decreases, as evidenced by a red shift of its absorption edge energy. These changes suggest a more drastic distorted NiN4 plane under negative potential.
Thermal treatment has been observed to induce distortion of the M-N4 plane in M-Pc molecules on carbon substrates53,57. Therefore, we conducted a similar calcination treatment on NiTAPc/GCNT-H to verify our hypothesis regarding the assignment of symmetry breaking to the distortion of the Ni-N4 plane. As shown in Supplementary Fig. 22, after one-hour calcination at 330℃ in argon, NiTAPc/GCNT-H exhibits a XANES curve more closely resembles that of NiTAPc/nGCNT instead of the as-prepared NiTAPc/GCNT-H. This result suggests that the observed symmetry breakings in our systems likely resulted from the out-of-plane distortion of the Ni-N4 center.
Furthermore, R-space extended X-ray absorption fine structure (EXAFS) spectra of NiTAPc/GCNT-H and NiTAPc/nGCNT are analyzed. As shown in Fig. 2f, Supplementary Fig. 23, and Supplementary Table 3, both Ni − N lengths and coordination numbers (CNs) of NiTAPc/GCNT-H and NiTAPc/nGCNT exhibit a slight increase under cathodic potential, indicating that the negative potential will promote the out-of-plane distortion of Ni. In the case of NiTAPc/GCNT-L (Supplementary Fig. 24a), we observed intermediate changes in peaks a, b, and c, indicating a moderate distortion of Ni-N4 during CO2R. A similar trend is observed in the case of NiTAPc/GCNT-L (Supplementary Fig. 24b). Overall, a clear correlation between the degree of NiN4 distortion and the defects on CNTs was determined: NiTAPc/GCNT-H < NiTAPc/GCNT-L < NiTAPc/nGCNT, which corresponds well with their selectivity and activity towards CO2R.
To further consolidate the above observations, we recorded the XANES of NiTCPc on different CNTs during CO2R. In contrast to the case of NiTAPc/GCNT-H, slightly more profound changes are observed in the XANES curves of NiTCPc/GCNT-H when obtained at cathodic CO2R potentials compared to those at OCV (Fig. 2g). The pre-edge peaks a and b both deviate from the original positions (Fig. 2h), indicating a more diminished D4h symmetry. Besides, the red shift of moiety c and absorption edge position suggests the reduced oxidation state of Ni during CO2R. As anticipated, NiTCPc/nGCNT exhibits much more drastic changes in the above characteristic regions compared to that of NiTAPc/nGCNT under identical conditions (Fig. 2h). Aside from the similar red shift of the absorption edge position, the characteristic fingerprints of square planar D4h symmetry at a, b and c almost disappeared, suggesting a drastic structural change of the Ni catalyst during CO2R. As a result, we can conclude that a much stronger interaction between NiTCPc and CNT substrate defects leads to substantial structural changes in NiTCPc under cathodic bias, likely due to its electron-withdrawing functionalization. Nonetheless, we believe that, when subjected to cathodic potentials, the electric field will induce detrimental distortions to the NiN4 plane. This, in turn, is expected to result in CO2R failure, especially under larger current densities, consistent with the aforementioned instability of the NiPc catalyst substituted with electro-withdrawing groups during CO2R24.
Furthermore, we compared the XANES data of NiTAPc/CNTs with that of NiTCPc/CNTs. As shown in Fig. 2i and 2j, it is evident that both NiTAPc/GCNT-H and NiTAPc/nGCNT exhibit a great degree of preservation in their characteristic square planar structure compared to their NiTCPc counterparts under similar, if not identical, CO2R conditions. Despite the distinct functional groups, the absorption edge energy of Ni for NiTAPc/nGCNT and NiTCPc/nGCNT are very close, with an estimated average Ni oxidation state between 0 and + 2, suggesting a predominant interactions between Ni active-center and the CNT substrates58,59. Nevertheless, the more profound decrease in the valence state of the Ni active-centers during CO2R on NiTXPc on nGCNT is associated with their more distorted Ni-N4 plane. This observation unveils a stronger interaction between NiTXPc and nGCNT. Taken together, the XAFS analysis supports our hypothesis that NiTXPc molecules have higher binding affinity to defects rather than perfect graphene-like CNT surface, while Ni-Pc catalyst with electron-withdrawing groups presents stronger interaction with defects.
We also recorded XANES for the post-electrolysis sample of NiTCPc/GCNT-H. As shown in Supplementary Fig. 26, while the energy of the absorption edge almost returns to its initial state, the coordination environment of the Ni active center cannot fully recover. This result indicates that either electric field induced structural distortion or the partial reduction of ligands on the Pc ring may not be entirely reversible. However, we believe the former one has more pronounced effects on the Ni-N4 symmetry and could be the predominant reason for the observed changes. In conclusion, the in-situ XANES experiments validate our hypothesis on the correlation between interactions of the Ni-Pc catalyst/CNTs and CNTs defects. Specifically, defects on CNTs induce the distortion of the planar NiN4 coordination environment under cathodic potentials, which promotes the undesired HER and compromises the catalyst’s stability towards CO2R. Additionally, electron withdrawing group functionalized Ni-Pc catalysts exhibit a stronger affinity for defects, resulting in more pronounced structural distortions, and thereby leading to increased degradation and reduced stability when high cathodic potential is applied to afford high current densities.
We also carried out in-situ IR studies to track the reaction intermediates on Ni-Pc catalyst on different CNTs during CO2R using ATR-SEIRAS. As shown in Fig. 3a-b, both NiTAPc/GCNT-H and NiTAPc/nGCNT exhibit in-situ IR signals for the reaction intermediate of *COOH (C − O stretch signal located at a wavenumber of 1380 cm− 1, symmetric stretch of *COO− located at 1410 cm− 1)60,61, indicating the same reaction mechanism for CO2R. Note, the prominent peaks at approximately 1640 cm− 1 are H − O−H bend vibration signals from H2O60,62. To better interpret the trend of *COOH, we plot the normalized *COOH peak intensity against the cathodic potential applied. As shown in Fig. 3c, the *COOH signals for NiTAPc/GCNT-H emerge at a lower overpotential compared to that of NiTAPc/nGCNT. Additionally, they exhibit an accelerated augmentation rate, particularly at more cathodic potentials. These findings align well with previous observations, indicating that NiTAPc supported on GCNT-H displays a lower overpotential for CO2R. As anticipated, the in-situ FT-IR spectra of NiTCPc/GCNT-H and NiTCPc/nGCNT (Fig. 3d-e) exhibit a similar trend to that of the NiTAPc analogue. Notably, the much more accelerated *COOH peak intensity (Fig. 3f) provides compelling evidence that aligns with the above CO2R performance data. To exclude the interference from carbonate or bicarbonate, which also feature similar O-C-O bonds, we conducted analogous in-situ IR studies in phosphate buffer saline (PBS) with pH = 8. As shown in Supplementary Fig. 28, all four catalysts exhibit similar spectra to those recorded in KHCO3, validating that the corresponding peaks indeed originate from CO2R intermediates. Furthermore, to investigate the possible peaks covered by the strong H2O signals, we conducted deuterium isotopic in-situ IR. The spectra of NiTAPc/GCNT-H and NiTCPc/GCNT-H, presented in Supplementary Fig. 29, indicate that both catalysts show similar spectra to those observed in the protium isotopic solution in the 1370 ~ 1410 cm− 1 region. The additional appeared to signal at around 1580 cm− 1 could be indexed to the asymmetric COO− species60, which is in good agreement with protium isotopic ATR-SEIRAS.
We conducted spin-polarized DFT-based calculations to investigate the interactions between the Ni-Pc catalysts and their possible structural changes upon adsorption. Specifically, a 10×10 supercell of graphene (Gr) was constructed as the model substrate due to the large diameters of CNTs (over 15 nm) compared to Ni-Pc-based catalysts, with different amounts of carbon vacancies introduced onto the graphene supercell as defects. Figure 4a illustrates the relaxed geometry of the immobilized NiTAPc on perfect Gr (NiTAPc@Gr, Fig. 4b) with a binding energy of − 2.59 eV, including van der Waals correction. Upon introducing a single vacancy on Gr (1V-Gr), the adsorption of flat NiTAPc molecule (NiTAPc@1V-Gr-flat) is favored by approximately 0.11 eV (Fig. 4d). This result aligns with the HRTEM images in Fig. 1, confirming that NiTAPc prefers absorption on the defect-rich domains on CNTs, further ruling out the possibility of π-π stacking. Notably, we discovered a novel configuration of distorted NiTAPc on 1V-Gr (NiTAPc@1V-Gr-dis) as shown in Fig. 4c. The difference in adsorption energy between the flat and distorted structures is as small as 0.03 eV, indicating that these two configurations are energetically degenerated and can interconvert under mild conditions. Hence, we believe the above XAS characterizations record the combination of NiTAPc in two states, with the ratio between them serving as a representation of the prevailing structural composition. Similar configurations were also acquired for NiTCPc (Supplementary Fig. 30). The adsorption energy differences between flat and distorted conditions for both NiTAPc and NiTCPc were summarized in Fig. 4d. As can be seen, while both NiTAPc and NiTCPc exhibit favorable adsorptions on defected CNTs, NiTCPc@1V-Gr-dis shows slightly enhanced adsorption (in terms of adsorption energy) than its flat counterpart, in contrasts to NiTAPc, suggesting a stronger interaction between Ni center and defects on CNTs.
We then investigated the mechanism of the potential-induced catalyst distortion. As shown in Fig. 4e, the charge redistribution analysis of both NiTXPc@1V-Gr-dis shows positive charge accumulation at the Ni site, resulting in a dipole moment of 0.19 Debye and 7.97 Debye for NiTAPc and NiTCPc, respectively, pointing from the substrate to the Pc plane. Therefore, when a cathodic bias is applied, the Ni coordination center tends to move towards the CNTs substrate from the initial Pc plane, facilitating the Pc plane distortions. Additionally, compared to NiTAPc, the significantly larger dipole moment of NiTCPc indicates its stronger interaction with defects on CNTs, resulting in more drastic structural distortions under similar cathodic bias.
Based on these configurations, we simulated the corresponding XANES spectra (Supplementary Fig. 31a-b). Specifically, the intensity of the pre-edge a exhibit a slight increase for both NiTAPc@Gr and NiTAPc@1V-Gr-flat. In contrast, significant increase in the intensity of the pre-edge a was observed for the 1V-Gr-dis configuration (Note 1 and 2 in Supplementary Fig. 31). Notably, similar observations were made for the NiTCPc systems. The simulated XANES further supports that the increased intensity of pre-edge a observed experimentally (Fig. 2d, g) should be ascribed to the interaction of Ni with substrate defects. The fitting of EXAFS in R-space based on the NiTAPc@1V-Gr-dis model (Supplementary Fig. 31c) indicates that the Ni − N and Ni − C have similar coordination number (CN) and bond lengths compared with the Ni − N path fitted using NiTAPc@1V-Gr-flat as the model. The consistency in bond lengths observed in the simulated structures (Supplementary Table 4) also agrees with the above EXAFS fittings.
Moreover, our simulations also reveal that the structural distortion substantially changes the electronic structures of the Ni active center near the Fermi level (EFermi). As shown in Fig. 4f-g, the partial density of states (PDOS) of NiTAPc@1V-Gr-dis reveals that the Ni dz2 orbital has split and moved closer towards EFermi in comparison to that of NiTAPc@1V-Gr. In general, increased electronic density around EFermi benefits the corresponding electrocatalytic activity. We then calculated the CO2 molecule adsorption energy for three configurations. The results, as shown in Fig. 4h, show that CO2 binds more strongly on NiTAPc@1V-Gr-dis, with the adsorption enhanced by 0.13 eV and 0.16 eV compared to two flat configurations, respectively. Similarly, NiTCPc assembles the same trend (Fig. 4h). This enhanced CO2 adsorption would favor the subsequent CO2R, highlighting the important role of having certain defects on CNTs in improving the CO2R activity. On the other hand, our calculations revealed that in the case of HER, the Gibbs free energy of *H-NiTAPc@1V-Gr-dis also decreases when distortion occurs (Supplementary Fig. 32), potentially promoting competing HER and compromising CO2R selectivity. We believe this distortion is attributed to the interaction between the NiPc-plane and substrate defects, which is facilitated by cathodic potentials. Consequently, the defect level can be used as an effective strategy to tune the number of distorted NiPc-planes, which in turn affects the catalyst performance. The central goal of this work is to prove that both CO2R activity and selectivity can be drastically improved by controlling the substrate defect level. As for the nGCNT substrate, given its high defect level (Fig. 1c), we proposed a possible model shown in Supplementary Fig. 33. Under this circumstance, one of the adjacent NiTAPc molecules becomes substantially curved, leading to poor structural stability under more cathodic bias and promoting HER. Overall, our computational results further confirm and explain the effects of defects on CNTs in both adsorption and catalysis processes.
Enhancing CO2R Through Catalyst/CNTs Interactions
Future implementation of CO2R catalysts requires them to demonstrate great stability and reduced voltage under practical relevant operation conditions63–65. Besides, in a typical CO2R process in an alkaline electrolyte, losses of CO2 due to carbonation can lead to decreased CO2 utilization efficiency. Conducting CO2R in acidic media, on the other hand, is advantageous for reducing unnecessary losses and enhancing CO2 utilization66,67. However, a high concentration of proton in acidic electrolyte kinetically facilitates the HER, underscoring the importance of catalyst selectivity towards desired CO2R products68–70. Therefore, we performed CO2R on NiTAPc/GCNT-H(/L) in the acidic electrolyte (pH = 2). As shown in Fig. 5a, both catalyst-composites exhibit outstanding selectivity towards CO even in acidic media, close to the high selectivity we observed in alkaline electrolytes. Besides, similar trends in CNT defect-level dependent selectivity and activity were observed in acidic CO2R (Fig. 5a, Supplementary Fig. 34), demonstrating the importance of controlling the defect level of CNT substrates. Previous literature revealed the local pH increase at the cathode-electrolyte interface during electrolysis under large current densities which will lead to an increase of bulk pH67,69, therefore we used sulfuric acid as anolyte and proton exchange membrane to maintain the bulk pH of the catholyte during CO2R. Consequently, we carried out a stability assessment in acidic electrolyte for NiTAPc/GCNT-H (Supplementary Fig. 35). Notably, we observe that the catalyst maintains stability at 200 mA cm− 2 for over 20 hours with negligible changes in CO selectivity. Additionally, the use of acidic electrolytes also enabled us to precisely evaluate the single-pass conversion (SPC) of CO2 in our system71 (Supplementary Fig. 36). Encouragingly, a high SPC of approximately was achieved with constantly high CO selectivity (CO/H2 ratio > 300) at the current density of 300 mA cm− 2 using NiTAPc/GCNT-H.
Aside from the effect of defects, we also investigated the impact of functional groups on CNT during CO2R. Taking − OH and − COOH functionalized CNTs (both non-graphitized and graphitized) as examples, we immobilized NiTAPc on them via the same methods (denoted as NiTAPc/GCNT-X or NiTAPc/CNT-X, X = − OH or − COOH). Similar diameter and morphology of these CNTs were confirmed by HRTEM (Supplementary Fig. 37). Besides, the introduction of these functional groups has negligible influence on the defect level of these CNTs (Supplementary Fig. 38). As shown in Fig. 5b, both NiTAPc/GCNT-X show higher CO selectivity compared to that of NiTAPc/GCNT-H under the same conditions. Remarkably, a high CO/H2 molar ratio of 16100:1 was achieved on NiTAPc/GCNT-OH at a high current density of 500 mA cm− 2. Note, to ensure precise product quantification, we conducted careful GC calibration and optimized the product collection procedures (Supplementary Fig. 39). Hence, we could claim that this catalyst-composite technically switches off the competing HER and achieves stoichiometric CO2R. Besides, when compared to non-graphitized counterparts (Supplementary Fig. 40), NiTAPc/GCNT-X also demonstrates reduced overpotentials and significantly suppressed HER (on the orders of two magnitudes). Compared with NiTAPc/GCNT-H, we can conclude that the influence of functional groups at the terminal of GCNT on CO2R is not significant. Therefore, the different defect densities should account for the distinct CO2R selectivity and activity.
The in-situ IR and XANES spectra of the two NiTAPc/GCNT-X were collected to further explore the CNT-functional-groups effect (Supplementary Fig. 41–42). Both catalyst-composites, NiTAPc/GCNT-OH and NiTAPc/GCNT-COOH, exhibited similar *COOH intermediate IR signals at similar potentials, suggesting similar reaction kinetics with NiTAPc/GCNTs. XANES of NiTAPc/GCNT-X retains the pristine planar characteristics, which is also similar to NiTAPc/GCNT-H. However, by analyzing the water contact angle on each catalyst-composites (Supplementary Fig. 43), we found that NiTAPc/GCNT-X exhibits a higher water contact angle before and after electrolysis. Therefore, we attribute the enhanced performance of CO2R to the improved hydrophobicity of the electrode, which facilitates CO2 transport across GDE to the reaction interface, thereby reducing the concentration overpotential for CO2R.
We also calculated turnover frequency (TOF) for our catalysts firstly based on the weight percentage of Ni assuming all Ni sites are active, and compared them with the TOFs of recently reported CNT-heterogenized molecular catalysts for CO2R18,24,40,72. As shown in Fig. 5c, we plotted the TOFCO together with \(\:{\text{FE}}_{{\text{H}}_{\text{2}}}\) to demonstrate the selectivity of a given catalyst under different reaction rates. Clearly, our strategy of engineering the defect level of the CNT-substrates leads to remarkable advantages in both activity and selectivity. Additional TOF comparisons were made using surface active-sites density (Supplementary Fig. 47). In this scenario, NiTAPc/GCNT-COOH exhibits approximately four orders of magnitude inhibition in HER and the highest TOFCO of 1072 s− 1, surpassing other state-of-the-art systems, such as NiTAPc/Vulkan XC72R (Supplementary Fig. 46). Overall, we demonstrate the significance of catalyst-substrate interactions, which can significantly alter catalytic performance.
The use of zero-gap MEA electrolyzer for CO2R shows great promise towards practical applications73–76. Hence, we also evaluated the performance of NiTAPc/GCNT-H in an MEA reactor with electrode size of 100 cm2 operated at ampere-level (Fig. 5d). We also assessed the state-of-the-art Ag catalyst (~ 100 nm nanoparticles) identical conditions for comparison. Notably, an almost 100% FECO could be obtained at a total current as high as 40 A using NiTAPc/GCNT-H. Besides, the CO selectivity remains high (> 95%) at an even higher current of ~ 50 A. In contrast, CO2R selectivity of Ag-based MEA dropped dramatically at these large currents, with \(\:{\text{FE}}_{{\text{H}}_{\text{2}}}\) exceeding 20% under the same conditions. Additionally, MEA reactor based on NiTAPc/GCNT-H exhibits consistently lower cell voltage compared to the Ag-based MEA, further highlighting the promise of our catalyst composites. Moreover, we conducted a stability test in the scaled-up MEA electrolyzer and found that NiTAPc/GCNT-H could retain the nearly 100% CO selectivity for over 30 hours of continuous operation at a high current of 20 A (Fig. 5e). The minor selectivity drop at the end of the test was likely due to salt formation in the cathode chamber, a persisted challenge for anion-exchange membrane type MEA-based CO2R77–79. Nevertheless, the NiTAPc/GCNT catalyst composite clearly demonstrates significant promise for practical implementations.
Lastly, we extended our strategy to Co-Pc analog, cobalt tetraaminophthalocyanine (CoTAPc), which has demonstrated methanol production activity, to assess its efficacy for CO to methanol conversion21,80. As shown in Fig. 5f, under identical testing conditions, CoTAPc/GCNT-H exhibits a higher \(\:{\text{FE}}_{{\text{CH}}_{\text{3}}\text{OH}}\) and more importantly substantially lower overpotential at all current densities assessed, particularly at high current densities. Therefore, we believe that the catalyst-substrate interactions we have demonstrated here are applicable to other molecular catalysts and likely to other electrochemical applications as well.