2.1. Characterization
Scheme 1 illustrates the procedure to prepare Ni(Fe)DMOF and its mKB composite. Nickel nitrate, also together with iron acetate, benzene-1,4-dicarboxylic acid (H
2BDC) and 1,4 diazabicyclo[2.2.2]octane (DABCO) afford the products [Ni
2(BDC)
2DABCO] and [(Ni/Fe)
2(BDC)
2DABCO], abbreviated as NiDMOF and Ni(Fe)DMOF, respectively. In the presence of mKB the composites NiDMOF/mKBx with different mKB fractions x = 7, 14, 22 and 34 wt.%. or Ni(Fe)DMOF)/mKB14 are obtained by using a one-step solvothermal reaction at 120 °C in DMF for 48h. The synthesis of Ni(Fe)DMOF was carried out with a targeted molar Ni to Fe ratio of 32:1. A molar Ni to Fe ratio of 30:1 was achieved in the synthesized Ni(Fe)DMOF as determined from atomic absorption spectroscopy (AAS) and SEM-EDX (
Tables S2 and S3,
Figure S9, Supporting Information (SI)). The weight fractions of mKB in the MOF/mKBx composites were about 7, 14, 22, and 34 wt.%, calculated from the MOF content which was derived from nickel amount by atomic absorption spectroscopy (AAS) data (
Table S2, SI).
The experimental powder X-ray diffractogram (PXRD) of the MOF product is positively matched to the reported structure of [Ni
2(BDC)
2DABCO]·(DMF)
4·(H
2O)
1.5 (NiDMOF) (
Figure 1a and Scheme S1) [
49]. The PXRD also demonstrated that the presence of mKB influenced the crystallinity of the NiDMOF structure only at high mKB content in NiDMOF/mKB34 where peak broadening occurs (
Figure 1 and
Figure S2, SI). This is due to the formation of a larger amount of smaller MOF crystallites, also evident from scanning electron microscopy images (
Figure S7, SI). The functional groups -OH, C-O and -COOH of mKB can serve as crystallization points leading to more crystal seeds which subsequently to not grow into large crystals [
52,
53,
54]. Furthermore, the ratio of the 001/100 reflex intensities increases in the NiDMOF/mKBx composites with mKB content, indicating that in the presence of mKB the DMOF crystallites may become more oriented along the [001] direction of the NiDMOF structure (
Figure S1, SI).
The electrochemical performance is strongly correlated to the surface area, porosity and pore size distribution of electrode materials.[
55] To explore these properties, nitrogen sorption measurements were performed at 77 K and gave the expected Type I isotherms [
56] for microporous NiDMOF and Ni(Fe)DMOF (
Figure 2a and
Figure S6a). The specific BET surface area and total pore volume of NiDMOF was 2104 m
2 g
–1 and 0.82 cm
3 g
–1, respectively, in good agreement with the literature values (2050 m
2 g
–1, 0.80 cm
3 g
–1) (
Table 1) [
57]. Ni(Fe)DMOF also showed a similar high BET surface area and total pore volume with 1942 m
2 g
–1 and 0.80 cm
3 g
–1, respectively. The BET surface area and total pore volume of modified Ketjenblack (mKB) are 1234 m
2 g
–1 and 1.50 cm
3 g
–1, identical to the original Ketjenblack carbon (KB) (
Table 1 and
Figures S5a and S5b, SI). For the DMOF/mKBx composites the desorption branch additionally displays a Type H4 hysteresis [
56], as seen in the isotherm of the mesoporous mKB carbon itself (
Figure 2a,
Figures S5c and S5d). The experimental surface area and pore volume parameters of the composites NiDMOF/mKBx (x = 7, 14, 22, 34 wt.%) and Ni(Fe)DMOF/mKB14 are somewhat lower than estimated from the mass-weighted values of the neat MOF and mKB components (
Table 1). This reduction of BET value of composites was also observed for similar work and indicates that mutual pore blocking effects occur.[
48] The NiDMOF can grow into the mKB pores or on the mKB surface and thereby access to the mKB and NiDMOF pores becomes restricted. SEM images for NiDMOF/mKBx with higher amount of mKB (x ≥ 22 wt.% mKB) illustrate the surface coverage of the NiDMOF rods with mKB particles (
Figure S7, SI). At the same time, the combination of mKB with MOFs can increase the total (micro-meso)pore volume in the composites above the estimated value. The composite NiDMOF/mKB14 exhibits an experimental a total pore volume of 1.18 cm
3 g
−1, considerably higher than the estimate of 0.92 cm
3 g
−1 for a physical mixture of 86 wt.% NiDMOF and 14 wt.% mKB. This pore volume of 1.18 cm
3 g
−1 was also the maximum among the prepared DMOF/mKB composites.
The synthesized NiDMOF and Ni(Fe)DMOF samples exhibit micropores with widths of about 1.2 and 1.7 nm (
Figure 2b and
Figure S6d). In the NiDMOF/mKBx and Ni(Fe)DMOF/mKB14 composites there are additional mesopores above 2 nm from the mKB component. The increased pore volume beyond the estimate can be traced to interparticle mesopores above 5 nm (
Table 1 and
Figure S6a). Open microporous channels can provide active sites for electrochemical reaction and open mesoporous channels will improve the diffusion rate of electrolyte ions [
58].
The estimated total pore volume was calculated accordingly. NiDMOF/mKB7 stands for 93 wt.% NiDMOF and 7 wt.% mKB.
Scanning electron microscopy (SEM) of the MOFs and the composites shows rod or block microcrystals of the NiDMOF with increasing amount of fine grains of mKB from NiDMOF/mKB7 to NiDMOF/mKB34 (
Figure 3, and
Figure S6, SI). The observed NiDMOF morphology is in a good agreement with the reported one [
59]. The element mapping from SEM–energy-dispersive X-ray spectroscopy (SEM-EDX) (
Figure 3c and
Figure S8, SI) demonstrate that in Ni(Fe)DMOF the iron is homogeneously distributed in the MOF microcrystals.
The X-ray photoelectron spectroscopy (XPS) survey spectrum of Ni(Fe)DMOF/mKB14 confirmed the presence of all the elements (Ni, Fe, C, N and O) of the synthesized material (
Figure S10, SI for XPS of NiDMOF and Ni(Fe)DMOF). The high-resolution spectrum of Ni 2p (
Figure 4a) shows two characteristic peaks at 855.7 and 873.2 eV which are ascribed to Ni
2+ 2p
3/2 and Ni
2+ 2p
1/2 respectively, and two expected satellite peaks were located at 860.7 and 879 eV [
18,
60,
61,
62]. A spin-orbit coupling energy difference between the 2p
3/2 and 2p
1/2 binding energy of 17.6 eV supports the assignment of the +2 oxidation state.[
18,
63] Furthermore, in the Fe 2p spectrum, the peaks at 709.6 and 722.8 eV (
Figure 4b) confirm the +2 oxidation state of Fe [
64,
65]. The 2p
3/2 spectrum range is 710 to 720 eV including a satellite peak at 715.2 eV, while the 2p
1/2 spectrum range is 721−735 eV with a satellite peak at 734.3 eV. It should be noted that the Fe 2p spectral background has a contribution from a Ni
LMM Auger peak [
66]. More detailed information on the high-resolution XPS of C 1s, O 1s, N 1s and Fe 3p of Ni(Fe)DMOF/mKB14, Ni 2p and Fe 2p of Ni(Fe)DMOF, Ni 2p of NiDMOF are given in the SI (
Figures S11 and S12).
2.2 Electrochemical properties
The OER performance of different samples was evaluated by using a glassy carbon rotating disk electrode (GC-RDE) under alkaline conditions (1 mol L
–1 KOH). OER polarization curves were collected from linear sweep voltammetry (LSV) measurements at a sweep rate of 5 mV s
–1. As shown in the initial LSV curves (
Figure 5a), the NiDMOF/mKBx composites were more efficient for OER activity than pristine NiDMOF and mKB alone at a current density of 10 mA cm
–2. Thereby, it worth noting, that the OER activity of the NiDMOF sample is already as good as the commercial RuO
2 benchmark sample. Among all NiDMOF/mKBx composites (x = 7, 14, 22, 34 wt.% mKB), NiDMOF/mKB14 exhibits the best electrocatalytic activity with the smallest overpotential of 294 mV (vs RHE at 1.23 V) to achieve a current density of 10 mA cm
–2, which is much lower than the overpotential of mKB (375 mV), NiDMOF (315 mV), NiDMOF/mKB7 (308 mV), NiDMOF/mKB22 (302 mV) and NiDMOF/mKB34 (304 mV), and competes with the performance of the benchmark material RuO
2 (317 mV). Thereby, the overpotential of RuO
2 is in good accordance with the literature [
67]. A low overpotential means a lower demand of energy for the oxygen evolution reaction. The kinetics on the GCE toward NiDMOF and its DMOF/mKBx composites were described on the basis of the Tafel equation. The OER kinetic parameters of the samples are analysed by the Tafel equation (η = a + b log(
j)), which is used to determine the reaction mechanism and the kinetics [
68]. The Tafel slope indicates how much one has to increase the overpotential to increase the reaction rate by a factor ten, that is, it describes the influence of the overpotential on the steady-state current density and is an important parameter for the evaluation of OER kinetics. The Tafel slope has been calculated from corresponding LSV plots to get a quantitative idea about the electrocatalytic performance. Krasil´shchikov´s OER mechanism is one of the established mechanisms, which is described by reaction (3)-(6) with the corresponding Tafel slopes
b [
35,
68,
69].
The value of the Tafel slope in NiDMOF (55 mV dec
–1) is in between reaction (3) and (4), and indicates that the deprotonation of a metal hydroxide (3) and the oxidation of a metal oxide species (4) could occur together as rate-determining steps [
70,
71]. The corresponding Tafel curves derived from
Figure 5b show the Tafel slopes for NiDMOF/mKBx composites (7, 14, 22, 34 wt.% mKB ) with 53, 32, 45 and 51 mV dec
–1, respectively. Thus, integrating mKB with NiDMOF can enhance the kinetics of the catalyst. For the 14 and 22 wt.% composites there is also a change in the rate-determining step towards reaction (4) for 22 wt.% (slope of 45 mV dec
−1) and towards reaction (5) of the evolution of O
2 for 14 wt.% (slope of 32 mV dec
−1). The NiDMOF/mKB14 composite has the smallest Tafel slope (32 mV dec
–1), confirming that the interaction between NiDMOF and 14 wt.% of mKB in the composite gives the most proficient material.
It is acknowledged that due to the surface structure reconstruction of Ni-based catalysts, an activation occurs during the water oxidation process [
23]. In order to understand the electrocatalytic behavior and the activation process, 100 cyclic voltammetry (CV) scans were applied. With mKB present in the composites, the oxidation peak at 1.35-1.45 (V vs RHE) became more noticeable and shifted into positive direction, indicating the synergetic effect of mKB and NiDMOF on the oxidation of Ni
+2 to Ni
+3 (
Figure S13, SI). As seen in the
Figure S14, the CV curves of all samples after 100 CVs showed that NiDMOF/mKB14 provides the most active catalyst, while mKB alone exhibits worse OER activity due to carbon corrosion in alkaline conditions [
72]. The mKB additive leads to the electrical conductivity between the active Ni sites and the GCE and thereby enhances the electrochemical activity [
73,
74]. However, an excessive amount of mKB lowers the carbon dispersion due to agglomeration from π-π interaction between the carbon particles, which reduces the charge transfer between active Ni sites and mKB and suppresses the catalytic activity for OER [
75,
76].
Furthermore, electrochemical impedance spectroscopy (EIS) was carried out to assess the kinetics of the electrode reaction. The corresponding Nyquist plots (
Figure 5d) obtained in a frequency range from 0.01 to 100 kHz with an AC potential amplitude of 5 mV at 1.5 V (RHE) were fitted to the simplified equivalent circuit by Doyle et al. for OER (
Figure S16, SI).[
23,
77,
78,
79] From the shape of the semicircles in the Nyquist plots, it can be seen that NiDMOF and the NiDMOF/mKBx composites follow a similar trend. Smaller semicircles are displayed by the NiDMOF/mKBx composites which reflect smaller resistances at high frequency compared with both pure NiDMOF and mKB, which illustrates that the kinetic performances of the neat MOF materials were improved by integration of mKB. The low polarization resistance R
p (4Ω) and resistance adsorption components of reaction intermediates R
s (2 Ω) for NiDMOF/mKB14 indicate a superior charge transfer rate and easier formation of active species for OER, respectively, which contribute to its highest catalytic activity among the composites (
Table S4, SI). Thus, mKB improves the electrochemical conductivity which is considered to be a key factor in enhancing the electrochemical performance of MOF-based composites.
It has been intensively studied that Fe can be advantageously incorporated into Ni based catalysts. With a small amount of Fe the catalytic activity toward OER is significantly improved [
22,
23,
48]. In order to investigate the effect of Fe on the electrocatalytic NiDMOF performance a bimetallic Ni(Fe)DMOF and the Ni(Fe)DMOF/mKB14 composite were prepared and measured under the same conditions. The presence of a small amount of Fe at a Ni:Fe ratio of ~30:1 can significantly enhance the OER performance. The Ni(Fe)DMOF without mKB already exhibited an enhanced OER performance with a much lower overpotential of 301 mV at the current density of 10 mA cm
–2, which is much smaller compared to the commercial RuO
2 benchmark (317 mV) and pristine NiDMOF (315 mV). The composite of Ni(Fe)DMOF with 14 wt.% mKB further enhanced the OER activity giving the smallest overpotential of only 279 mV among the materials reported here (
Figure 6a,c). These results strongly demonstrate that Fe plays a key role in improving the OER activities of pure NiDMOF and its composites. As shown in
Figure 6b,c, the Tafel slope calculated from the corresponding LSV curve of bimetallic Ni(Fe)DMOF is 40 mV dec
–1 which is in agreement with a thin-film of Ni-Fe oxide [
80] and outperforms the monometallic NiDMOF (55 mV dec
–1). Ni(Fe)DMOF/mKB14 presents the lowest Tafel slope (25 mV dec
–1) compared to RuO
2 (56 mV dec
–1) and all DMOFs and composites investigated here (
Table S4, SI), reflecting the fastest kinetics. The results indicate a change in a mechanism where the rate−determining step becomes the evolution of O
2 (reaction 5). Furthermore, the Ni(Fe)DMOF/mKB14 composite stands out as one of the best in the recently reported most advanced Ni-based electrocatalysts with both its low overpotential (279 mV) and Tafel slope (25 mV dec
–1) (
Table S4, SI). A lower overpotential was reported for Ni(Fe)-MOF-74/KB (48 wt.% KB) (274 mV) [
48], Ni(Fe)(OH)
2/KB (47 wt.% KB) (265 mV) [
48] and Fe-doped HXP@NC800 (266 mV) [
81] (all at 10 mA cm
−2), albeit with larger Tafel slopes of 40, 55 and 49 mV dec
–1, respectively. From
Figure S15 it can be seen that even after 1000 CVs the overpotential of the catalysts have only slightly changed from 301 mV to 318 mV for the Ni(Fe)DMOF precursor and from 279 mV to 285 mV for Ni(Fe)DMOF/mKB14 (at 10 mA cm
−2), indicating their long-term activity. Furthermore, electrochemical impedance spectrometry (EIS) measurements were carried out to understand the charge-transfer kinetics during the OER process among the different samples. As shown by the Nyquist plots (
Figure 6d) and the equivalent circuit model, the resistances of charge-transfer and adsorption components of reaction intermediates of Ni(Fe)DMOF/mKB14 (R
p 4 Ω and R
s 2 Ω) was as low as NiDMOF/mKB14 (R
p 4 Ω and R
s 2 Ω), that is, lower than that of RuO
2 (R
p 23 Ω and R
s 4 Ω), NiDMOF (R
p 21 Ω and R
s 7 Ω) and Ni(Fe)DMOF (R
p 6 Ω and R
s 4 Ω) (
Table S4, SI). The catalyst from the Ni(Fe)DMOF/mKB14 precursor presents the smallest charge-transfer resistance, which indicates its faster electron transport kinetics and intrinsic excellent electrical conductivity, well in accordance with its lowest overpotential. This result is in good agreement with the identified higher catalytic activity of Ni(Fe)-MOFs compared to analogous Ni-MOFs [
25,
82,
83]. In particular, the presence of iron promotes the oxidation of nickel from +2 to +3, the latter seen as the active state of fast reaction kinetics and enhanced conductivity.
Loading a catalyst on a porous Ni foam (NF) electrode can further improve the reaction rate of OER [
84]. Because of its metallic conductivity and 3D macroporous structure, NF provides a large surface area and facilitates mass transport during OER [
84]. The catalyst ink was loaded by the drop-casting method on the surface of NF with a good distribution of the ink layer on the scaffold of the NF. The sample Ni(Fe)DMOF/mKB14, which showed the best OER activity on the glassy carbon RDE, was deposited on NF (1 cm
2) (
Figure S17, SI). The LSV curves in
Figure 7a show that the activity of Ni(Fe)DMOF/mKB14@NF is much higher than the pure NF substrate, reaching current densities of 10 and 50 mA cm
–2 with overpotentials of 247 and 291 mV, respectively. A practical current density of 400 mA cm
–2 could be delivered for Ni(Fe)DMOF/mKB14@NF at a low overpotential of 381 mV. Such performance is better than benchmark RuO
2@NF as OER catalyst with overpotentials of 278 and 340 mV at current densities of 10 and 50 mA cm
–2, respectively (
Figure 7a), the later in good accordance with the literature [
85]. The pure NF electrode shows a much lower OER activity with an overpotential of 370 mV at 10 mA cm
−2 similar as in the literature [
85].
The stability of an electrocatalyst determined by chronoamperometry is a key parameter to evaluate the practical application of a material. A long-term stability measurement was carried out by applying a potential to reach a constant current density of 50 mA cm
–2 for 30 h (
Figure 7b). The applied potential could be gradually decreased during the first 2.5 hours, indicating that the Ni-Fe catalyst was activated under the anodic potential in 1 mol L
–1 KOH electrolyte. After this time, the potential remained nearly unchanged during the rest of the 30-hour measurement. In contrast, the bare NF electrode showed an increasing potential over time for delivering a current density of 50 mA cm
–2. Pure NF also necessitates a much higher potential at the current density of 50 mA cm
–2 than Ni(Fe)DMOF/mKB14@NF during the 30 h of chronoamperometry test. The catalyst derived from Ni(Fe)DMOF/mKB14 does not only have a robust stability, but also an outstanding activity to deliver a high current density, which demonstrates that the Ni(Fe)DMOF/mKB14 precursor has the potential to serve as a good OER catalyst for practical applications.
In order to understand the transformation of the precursors Ni(Fe)DMOF and Ni(Fe)DMOF/mKB14 in the alkaline electrolyte (1 mol L
–1 KOH), we reacted macroscopic amounts of these samples in 1 mol L
–1 KOH, followed by PXRD, FTIR, SEM and N
2 sorption measurements to mimic post-mortem experiments of the minuscule electrode materials. The derived materials were collected by filtration after soaking Ni(Fe)DMOF and Ni(Fe)DMOF/mKB14 in 1 mol L
–1 KOH electrolyte for 24 h and were dried for at least 12 h at 120 °C under vacuum (< 10
−2 mbar). The PXRD pattern of the derived-Ni(Fe)DMOF/mKB14 (
Figure 8a) shows the disappearance of the crystalline MOF and suggests the formation of α/β−Ni(OH)
2, β−NiOOH and γ−NiOOH from the reflections assigned in
Figure 8a based on the patterns of nickel(II) hydroxides and nickel(III) oxide-hydroxides. This implies the transformation from Ni(Fe)DMOF/mKB14 to nickel and iron oxide-hydroxides, under the strong alkaline environment (1 mol L
–1 KOH),[
86] which has also been observed by other researchers [
87,
88,
89,
90,
91]. FTIR of the derived-materials (
Figure 8b and
Table S1) shows two broad bands at ~3400 and ~3600 cm
–1 which correspond to the stretching vibrations of adsorbed water molecules and to O−H stretching vibrations, as in Ni(OH)
2. At the same time bands due to the asymmetric vibration ν
asymCOO
− and the bending vibrations of adsorbed water molecules δO−H hydroxyl groups of the BDC phenyl ring at ~1600, and ~1570 cm
–1, along with ν
symCOO
− at ~1350 cm
–1 have decreased in intensity. New strong bands at 516 and 450 cm
–1 are attributed to Ni-OH bending vibrations in Ni(OH)
2 and the oxide-hydroxides [
92]. Interestingly, the morphology of Ni(Fe)DMOF/mKB14 as seen with SEM (
Figure 8c) does not change profoundly. The rod shape of Ni(Fe)DMOF was largely retained in the alkaline treatment. Accordingly, N
2 physisorption measurements at 77 K reveal a residual BET surface area and micro-mesoporosity in derived-Ni(Fe)DMOF and derived-Ni(Fe)DMOF/mKB14 (
Figure 8d). The derived-Ni(Fe)DMOF displays a Type IV isotherm with H2(b) hysteresis (associated with pore blocking) [
56] and a BET surface area of 222 m
2 g
–1. The adsorption isotherm of derived-Ni(Fe)DMOF/mKB14 lacks the final saturation plateau of a Type IV isotherm and appears like a Type II isotherm with an H
3 hysteresis which would be given by a largely macroporous adsorbent. The pore size distribution curves show mostly mesopores (>2 nm) for derived-Ni(Fe)DMOF and the coexistence of micropores and mesopores for derived-Ni(Fe)DMOF/mKB14 with the micropores and small mesopores due to the mKB part (the macropore size is not given by N
2 sorption). The BET surface area of derived-Ni(Fe)DMOF/mKB14 is 352 m
2 g
–1 with the increase over derived-Ni(Fe)DMOF coming from the mKB portion. Overall, a hierarchical porous nature of derived-Ni(Fe)DMOF/mKB14 was evidenced by its isotherm and hysteresis shape and the analysis of the pore size distribution. The decomposition of the Ni-MOF precursor and transformation to Ni(OH)
2/NiOOH was also observed in other work [
93,
94,
95]. The OER activity of the Ni(Fe)DMOF precursor is due to the
in situ formation of α/β-Ni(OH)
2/FeOOH followed by β/γ−NiOOH under the oxidizing anodic potential, which favors the kinetics of OER [
18,
39,
96,
97,
98]. Further, the degradation of the micro-mesoporosity of Ni(Fe)DMOF in the alkaline environment provides a large number of accessible active Ni (and promoter Fe) sites in the still porous hydroxides, which show better electrochemical activities and longer cycle numbers than crystalline and dense metal oxide catalysts [
35,
36,
99,
100,
101,
102]. In addition, the suitable amount of mKB (14 wt.%) in the composites provides electrical conductivity with its small particle size, mixes well with the Ni active sites and avoids the carbon corrosion effect of carbon materials [
73,
103].