Lettie A. Smitha,
J. Ehren Eichlera,
Kenta Kawashima
a,
Hanah Leonardb,
Franklin Tangc,
Ethan Kang Yanga,
Wuilian A. Martineza,
Yasaman Karbalaeemorad
a and
C. Buddie Mullins
*acdef
aDepartment of Chemistry, The University of Texas at Austin, Austin, TX 78712, USA
bDepartment of Biochemistry, The University of Texas at Austin, Austin, TX 78712, USA
cJohn J. McKetta Department of Chemical Engineering, The University of Texas at Austin, Austin, TX 78712, USA
dTexas Materials Institute, The University of Texas at Austin, Austin, TX 78712, USA
eAllen J. Bard Center for Electrochemistry, The University of Texas at Austin, Austin, TX 78712, USA
fH2@UT, The University of Texas at Austin, Austin, TX 78712, USA
First published on 28th July 2025
Nitrogen-doped carbons with consistent N-content and N-functionality distribution but variable porosity and surface area were prepared and studied to evaluate textural property effects on oxygen reduction reaction (ORR) electrocatalysis independent of chemical composition. By employing various characterization techniques, the similar N-content (1–2.5 at%) and N-motif distribution were confirmed between samples, as was variability in porosity and surface area. From electrochemical testing, a larger micropore volume associated with pores < 1 nm (>0.25 cm3 g−1) and surface area (>approximately 1000 m2 g−1) correlated with increased onset potential (Eonset), half-wave potential (E1/2), and an electron transfer number, n, of approximately 4, wherein the most significant effect of porous structure on electrochemical performance was observed for E1/2. With respect to Eonset and n, more subtle effects were observed regarding micropore volume (<1 nm) and surface area, as the kinetics of the ORR are predominantly tuned by the presence and type of active N-sites. An increased micropore volume and surface area realized slight increases in Eonset and n owing to improved access to these active sites. Thus, textural properties primarily affect mass transport for ORR electrocatalysis with effects on kinetic parameters only arising regarding accessibility to active sites located within micropores.
For the oxygen reduction reaction (ORR), one of the half-reactions in both direct methanol fuel cells and hydrogen fuel cells, porous and doped carbon materials have emerged as a replacement for these Pt-based materials.6–8 These carbon materials have demonstrated similar performance to platinum on carbon (Pt/C) with improved durability.6,9–12 For example, Lv and coworkers were able to obtain a half-wave potential (E1/2) of 0.85 V vs. the reversible hydrogen electrode (RHE) with their N-doped hydrogen-substituted graphdiyne prepared at 900 °C, rivaling that of platinum on carbon. Additionally, their carbon material had high selectivity with an electron transfer number, n, of approximately 4 and superior resistance to methanol crossover in comparison to Pt/C.13 The superior performance of the hydrogen-substituted graphdiyne was attributed to a large Brunauer–Emmett–Teller (BET) surface area (SA) at 1754 m2 g−1 and an abundance of pyridinic-N sites (only pyridinic N present at 2.9 ± 0.3 at%). As indicated by Lv et al.'s work, the porous structure and N motifs of N-doped carbons are considered the driving force behind their ORR electrocatalytic performance. An increased surface area has been attributed to an increased number of exposed active sites and improved mass transfer.14 The type of N groups in the carbon matrix have also been ascribed to improved E1/2 and selectivity, n, for the 4 vs. 2-electron pathway of the ORR (producing either water or hydrogen peroxide, respectively).9,10,13,15–19 Of the typical N sites detected in these carbon materials, graphitic, pyrrolic, and pyridinic N have been attributed to selectivity for the 4-electron pathway.9–12,16,20–22 These N-motifs are thought to increase performance by inducing a partial positive charge at adjacent carbon atoms, promoting the adsorption of oxygen.18,23,24
The extent to which porous structure, total N-content, and N composition affect the electrocatalytic performance of N-doped carbons independently is, however, difficult to gauge as both of these parameters tend to change simultaneously. Ferrero et al. investigated the role of pore size distribution at a constant N-content and N-distribution by using silica templates to obtain microporous and mesoporous N-doped carbon microspheres using pyrrole as a N precursor (difference in mesopore volume of 1.13 cm3 g−1 with all other adsorption parameters of nearly the same value).12 They found that the microspheres consisting mainly of mesopores exhibited better activity and selectivity for the 4-electron pathway of the ORR. However, these differences in performance were slight with E1/2-values at nearly the same potential and the onset potential (Eonset) only differing by ∼20 mV between the microporous and mesoporous microspheres.
Similarly, Dai and coworkers attributed micropores to reduced ORR performance for N-doped carbons.25 They concluded that micropores detracted from the ORR activity of their prepared carbons because such small pores were kinetically inaccessible to O2. Rather, improvements with ORR performance correlated with increases in mesopore volume, as indicated by Lv et al. above and others.13,26,27 Alternatively, Yao and coworkers,28 Eisenberg and coworkers,29 and Zhang and coworkers30 found that microporosity had a positive effect on ORR electrocatalysis, enabling a greater number of exposed active sites for electrocatalysis and improved mass transport when present in addition to mesopores in the N-doped carbon porous network.
An additional study analyzing oxidized carbons (N-free) alongside N-doped and oxidized carbons (1.4–2.1 at% N) found that N-free samples demonstrated superior performance owing to an abundance of smaller diameter micropores, as compared to the N-doped samples. Specifically, the oxidized, N-free samples predominately consisted of micropores 0.37 nm in size, close in size to that of an O2 molecule with a kinematic diameter of 0.346 nm.31 The authors proposed that these small micropores participated in and promoted the ORR because of a hierarchical porous network containing oxygen-containing functional groups in larger pores that enabled electrolyte penetration up to small micropores. They hypothesized that these micropores did not contain oxygen groups and were hydrophobic in nature, allowing extraction of O2 from the electrolyte, wherein O2 solubility is quite low.32 The highly confining pore geometry of this active site was speculated to reduce O2 to OH− via a 2-step 2-electron reduction pathway. Further evidence of this small micropore effect is provided by Gabe and coworkers, who studied porous carbon containing only C and O.33 They developed a model taking into account narrow and wide micropores, mass transport of O2 and HO2− through said pores, and the charge transfer of each reaction pathway involved in the ORR. From their modeling and experimental results, they were able to determine that micropores contribute to the ORR, that narrow micropores promote O2 reduction to HO2−, and that the rate of HO2− diffusion is suppressed such that the reduction of HO2− to OH− is favored. Thus, in the absence of N-doping, narrow micropores promote a 2-step 2-electron reduction of O2 to OH− rather than direct 4-electron reduction.
Despite the findings on micropore effects on ORR electrocatalysis for porous carbons containing oxygen functional groups, the role of micropores with respect to N-doped, porous carbons is still unclear. As discussed above and in a recent review paper,23 there exist conflicting reports on the impact of micropores on electrocatalysis in conjunction with N-sites both in terms of total N-content and the predominant N-motif. Thus, there is room for further exploration of the role of porous structure on ORR electrocatalysis for N-doped carbon electrocatalysts.
Herein, the effect of porous structure on N-doped carbon performance and selectivity as ORR electrocatalysts is investigated independent from the N-content by leveraging a synthetic method that imparts pore tunability while maintaining consistent N-content. Specifically, a N-doped carbon foam was chemically activated at 1000 °C following a dual-cation procedure enlisting the activators Na2CO3 and K2CO3.34,35 In brief, the total N-content was controlled by the activator:
N precursor and the N
:
C precursor ratios.35 Then, by varying the mole fraction of the activating mixture (Na2CO3 and K2CO3), the N-content was maintained while porosity was modulated significantly. Specifically, increasing the mol fraction of Na2CO3 to K2CO3 led to a more restricted porous network, reduced porosity (0.45 vs. ∼0.60 cm3 g−1), and smaller surface area (772.7 vs. 1202.4 m2 g−1).
Electrochemical testing of the porous, N-doped carbons revealed that improved performance and selectivity were observed when a K2CO3-rich activating mixture was employed. The N-doped carbons prepared with a greater proportion of K2CO3 vs. Na2CO3, exhibited larger E1/2 and Eonset. The improved E1/2 of the K2CO3-rich carbons was attributed to increased surface area and micropore volumes, as compared to the carbons prepared with Na2CO3-rich activation mixtures. The larger surface areas and micropore volumes of the K2CO3-rich activated carbons led to improved mass transport such that E1/2 increased. The large surface areas and micropore volumes of the K2CO3-rich activated carbons also led to increased Eonset, as a more open porous structure led to improved access to active sites for catalysis, indicating a slight effect of porosity on kinetics of the ORR. Concerning selectivity, n, a meaningful difference was only observed between the N-doped carbons prepared with a proportion of or 100% K2CO3 vs. the 100% Na2CO3 sample, as the latter resulted in a more tortuous porous network, restricting access of reactants to active sites for electrocatalysis. Thus, the volume of micropores tunes both Eonset and E1/2, with an abundance of micropores leading to improved performance. A large micropore volume and, as a result, a large surface area were obtained by employing a more K2CO3-rich activating mixture at 1000 °C. In addition, a K2CO3-rich mixture also led to increased selectivity (∼4 electrons), owing to improved access to active sites via an interconnected, open pore network and large surface area.
The presence of defects within the N-doped carbons was gauged by employing Raman spectroscopy to determine the ID to IG ratio. Specifically, a Renishaw InVia Raman system with an Ar-ion laser (514 nm, 50 mW) for excitation was focused on the sample using a microscope with a 50× objective lens. Raman spectra were collected from 500 to 3500 cm−1 with a laser exposure time of 60 s. The resulting Raman spectra were normalized and fit with six bands with Voigt peak shapes (Fig. S7). These six bands were the D4, D, D3, G, 2D, and G + D bands. The D4 band at ∼1200 cm−1 and D3 band at ∼1500 cm−1 have been attributed to amorphous character and vibrations associated with the presence of heteroatoms and other defects.37,38 The D band at ∼1350 cm−1 is also related to the presence of defects, such as heteroatom doping, vacancies, and holes. The G-band at ∼1580 cm−1 is associated with the stretching mode of sp2 bonded carbon. Then, the 2D and G + D are overtones, representing higher vibrational energy transitions that are multiples of the D and G bands.39–41 An additional peak, D2 at ∼1620 cm−1 associated with vibrations due to doping or other impurities, was not included in the fitting as it is associated with graphitic materials, and the carbons studied herein have amorphous characteristics.37,38
To observe changes in morphology, scanning electron microscopy (SEM) images were collected with an Apreo S SEM at an accelerating voltage of 10.0–20.0 kV and a beam current of 0.40 nA under vacuum. In addition, scanning transmission electron microscopy (STEM) was performed with a Hitachi S-5500 SEM/STEM to obtain sub-nanometer resolution images with accelerating voltages of 20.0–30.0 kV.
The surface composition of the prepared N-doped carbons was characterized with X-ray photoelectron spectroscopy (XPS, VersaProbe 4 X-ray Photoelectron Spectrometer) using monochromatic Al Kα radiation and by performing analysis with CasaXPS™ software. All spectra were calibrated to the sp2 carbon (CC) peak at 284.7 eV and fitted using Gaussian–Lorentzian peaks with Shirley backgrounds. High-resolution C 1s scans were deconvoluted using four peaks at 284.7 eV ± 0.0 (C
C), 285.8 ± 0.1 eV (C
N/C–C), 286.7 ± 0.1 eV (C–N), 289.2 eV ± 0.2 (N–C
O/C
O), and 291.0 ± 0.2 eV (π–π*). High-resolution N 1s spectra were deconvoluted with four peaks centered at 398.4 ± 0.1 eV (C
N–C), 399.9 ± 0.2 eV (NC3), 400.9 ± 0.1 eV (C2NH), and 402.3 ± 0.2 eV (NOx). High-resolution O 1s scans were deconvoluted with two peaks at 533.4 ± 0.2 eV (C
O) and 532.0 ± 0.3 eV (C–O).
The textural properties of the porous, N-doped carbons were determined from N2 isotherms performed at 77 K using an ASAP 2020 gas analyzer. BET SAs were determined using multi-point BET analysis via Micromeritics software. The pore size distributions were determined by the fitting of kernels derived from non-local density functional theory (NLDFT) on graphite. Prior to analysis, the carbon samples were outgassed under vacuum at 200 °C for 12 h.
The fine pore structure (0.3–1.4 nm) of the carbons was further probed by collecting CO2 isotherms at 0 °C on a Quantachrome Autosorb iQ-MP (Anton Parr). The pore size distributions were determined from these CO2 isotherms via kernels derived from non-local density functional theory (NLDFT) on carbon. To maintain the temperature at 0 °C for isotherm collection, a float sensor with a temperature-controlled glycerol/water bath (Isotemp refrigerated circulator, Fischer Scientific) was employed. Before analysis, all samples were outgassed under vacuum at 200 °C for 12 h.
![]() | (1) |
For each carbon, LSVs were collected at rotation rates ranging from 400–2500 rpm at a scan rate of 10 mV s−1. The effective number of electrons transferred was calculated using the RRDE method [eqn (2)] and the Koutecky–Leviche (KL) equation [eqn (3)]. Where J−1 is the inverse current density, Jk−1 is the inverse kinetic current density, and (Bω1/2)−1 is the inverse limiting current density, wherein ω is the rotation rate. When the current density is plotted against the inverse square root of the rotation rate, the y-intercept is Jk−1 and B−1 is the slope. By taking the reciprocal of the slope, the effective number of electrons transferred can be determined. The equation for the reciprocal of the slope, B, is shown in eqn (4). In eqn (4), n is the effective number of electrons transferred, F is Faraday's constant, CO2 is the concentration of oxygen gas, DO2 is the diffusion coefficient of O2, and v is the kinematic viscosity. The coefficient, 0.2, is used here as the rotation rate is in units of revolutions per minute (rpm).
![]() | (2) |
J−1 = Jk−1 + (Bω1/2)−1 | (3) |
B = 0.2nFDO22/3v−1/6CO2 | (4) |
While n was calculated using both the RRDE method and KL equation, results from employing the RRDE method are primarily discussed as the KL equation assumes a planar electrode surface, which has been shown to be deleterious in determining the selectivity of porous N-doped carbons with respect to the ORR.42
During activation, the porosity of the porous carbons was enabled by the decomposition of the alkali carbonates via the mechanism shown below, wherein M stands for the alkali cation [eqn (5)–(7)]. Evidence of this mechanism of pore formation is provided in the literature from extensive evolved gas analysis studies with a quadrupole mass spectrometer and X-ray diffraction studies.36,43–46 In the first step [eqn (5)], upon heating the carbon foam-carbonate mixture to several hundred degrees, the alkali carbonate decomposes into an alkali cyanate (MOCN), which further decomposes to an alkali cyanide (MCN) at ∼600 °C.36,43,46 Depending on the identity of the alkali cation, the onset temperature of M2CO3 decomposition varies.36 Specifically, when the carbon foam precursor is prepared with 80 mass% melamine, K2CO3 begins to revert to MOCN at as low as 375 °C, while such decomposition does not begin to occur for Na2CO3 until 525 °C.36 Given N is present in excess in the carbon foam, an abundant amount of these alkali cyanides form. At these elevated temperatures, the alkali cyanides melt, and the physical properties of the carbon are known to predominantly arise from interactions with the molten salt (i.e., molten salt porogenesis). Given the final synthetic temperature of 1000 °C is in excess of the melting points of KCN and NaCN by at least 400 °C, a molten salt is formed and maintained throughout the activation process. Thus, the porous structure is determined by the alkali carbonate(s) employed and occurs via molten salt templating.
M2CO3 + 2(−CN) → 2MOCN + CO | (5) |
M2CO3 + 2(−CN) → MOCN + MCN + CO2 | (6) |
MOCN + (−C) → MCN + CO | (7) |
The alkali carbonates Na2CO3 and K2CO3 were selected for activation rather than other carbonates, such as Li2CO3, because previous work36 showed that these two alkali carbonates resulted in N-doped carbons with a large proportion of pyridinic N and pore volumes composed largely of mesopores. As pyridinic N and mesopores have been attributed to improved electrocatalytic ORR performance and selectivity via the introduction of active sites and enabling improved mass transport, respectively,12,18,23,24,27,47,48 this synthetic method was uniquely posited for the preparation of N-doped carbons as ORR electrocatalysts.
As aforementioned, instead of inducing porosity with only pure Na2CO3 or K2CO3, mixtures of the two carbonates were employed because the identity of the alkali cation was found to lead to different porous structures while maintaining N composition.35 As such, porosity effects were studied independent of N-content, and by changing the mol fraction of each activating agent, differing porosities were realized. Thus, by preparing N-doped carbons using this method, the relationship between porous structure and electrochemical ORR activity and selectivity was studied irrespective of N-content, as it was held essentially constant despite modulating the activating mixture composition. Note that all N-doped carbons were activated at 1000 °C instead of lower temperatures, such as 800 °C, to overcome conductivity losses stemming from excessive N-content49,50 (i.e., a loss in covalency due to extensive disruption to the network of delocalized electrons of graphitic carbon) that would impede electrocatalytic performance. From past work,36 preparation at 800 °C led to N-contents of ∼23.5 at%, and preliminary testing (not shown) indicated that this N-content was deleterious towards ORR electrocatalysis. Activation at 1000 °C led to reduced N-content via the loss of nitrogen as gaseous species such as HCN, N2, and NCCN.46
X-ray photoelectron spectroscopy (XPS) determined the surface composition of the carbons, which is useful for understanding potential active sites for ORR electrocatalysis on the activated carbons and to confirm similar N-content between the samples. From analysis of the XPS spectra, the average N-content for 0–100Na was 1.7 ± 0.5 at% N, with an observed minimum of 1 at% N and a maximum of 2.5 at% N (Fig. 1). In addition, C-content and O-content were also very similar at 92–96 at% and 2.6–3.6 at%, respectively (Fig. 1 and S5). Fig. 1 shows the C, N, and O content for each sample. High resolution N 1s spectra revealed that there was virtually no change in the distribution of N-functionalities with increasing mol fraction of Na2CO3 until reaching 100Na (Fig. 1b). Nevertheless, the N-content between samples only varied by ∼1.5 at%, within the error of the X-ray photoelectron spectrometer (Fig. 1).51 In addition, the average N-contents fell within the range of uncertainty of one another (Fig. 1c). At 100Na, there was a clear increase in the intensity of the high-resolution N 1s XPS spectrum at ∼398.3 eV, indicating an increase in the abundance of pyridinic N sites (Fig. 1b). This result implies that the larger mol fraction of Na2CO3 compared to K2CO3 led to slightly greater selection for pyridinic N. This observation is supported by previous work, wherein K+ was speculated to preferentially form KCN adducts at pyridinic N sites.52 This selective adduct formation may be disadvantageous for ORR electrocatalysis given C-atoms adjacent to pyridinic N sites have been considered the true active sites for the ORR by many.11,18,53 However, it should be noted that such low N-contents at 1–2.5 at% make it difficult to make a steadfast conclusion from these data concerning selective tuning of the N composition by varying the mol fraction of Na2CO3 and K2CO3 because fitting of N-motifs is subject to increased uncertainty at these low concentrations. Thus, overall, the prepared N-doped carbons have nearly identical chemical composition, with activation employing only Na2CO3 possibly leading to an increased abundance of pyridinic N sites. All deconvoluted high-resolution XPS spectra can be found in the SI.
Following confirmation of similar N-content among the prepared carbon samples, X-ray diffraction (XRD) determined how varying the mol fraction of Na2CO3 in the activating salt mixture (Na2CO3, K2CO3) affected the structure of the carbons, as differences in crystallinity have been associated with changes in porous structure for carbons54 and increased stability of carbon materials under alkaline conditions.55–58 For all samples, two peaks were observed at ∼23° and ∼44° 2θ, corresponding to the 002 and 101 planes characteristic of turbostratic carbon (Fig. 2a).59,60 A “tail” was also observed at low 2θ, which is associated with the presence of micropores and a lack of long-range order.54 The 002 peak was observed because of the stacking of carbon sheets, and the 101 peak was observed because of lateral defects (pores, dopants, etc.) in the carbon structure.60,61 From the XRD patterns, a reduced full width half maximum (FWHM) of the 002 peak was observed with the increasing mol fraction of Na2CO3, meaning a more crystalline carbon was obtained. There was also a clear shift to higher 2θ for the 002 peak with increasing Na2CO3, which was attributed to an increase in the crystallite size, Lc, associated with the stacking of carbon sheets (Fig. 2a). Fig. 2b highlights the trend between the mol fraction of Na2CO3 and d002 and Lc, wherein d002 decreases with increasing mol fraction of Na2CO3 and Lc increases. A larger Lc means more stacked sheets are present, likely because of reduced d-spacing via increased van der Waals interactions between sheets. This reduced d-spacing could limit mass transport of reactants during the ORR, limiting electrocatalytic performance. Additionally, the larger Lc corresponding to increased XNa means there are an increased number of edge sites exposed to the electrolyte, which have been linked to material degradation under alkaline environments.55 Thus, employing Na2CO3 as an activator could be detrimental to ORR performance and stability owing to the effects of this synthetic handle on the structure of the final activated carbon.
![]() | ||
Fig. 2 (Semi)crystallinity of the activated N-doped carbons. (a) XRD patterns, (b) d002 and Lc as a function of Na+ mol fraction. |
As aforementioned, the structure of the prepared N-doped carbons is delegated by the alkali cyanides that form the molten salt for pore generation. These alkali cyanides act as structure-guiding agents, retaining the local structure of their crystal forms, while having the physical properties of a liquid.62,63 Essentially, the molten phase of these templates has ordering capabilities while being a liquid consisting of salt ions at high temperatures,64 hence the term ‘ionothermal liquid,’ which can be used in place of ‘molten salt.’65 During synthesis, the cations and anions of the molten phase interact with the carbon precursor to generate a more ordered, porous N-doped carbon, as compared to porous carbons prepared via chemical etching, to give one example.36,46 The templating occurs via interaction of the cation and anion with the carbon precursor such that charge balance is achieved, wherein the generated pores are dependent on the ion pair and the formation of salt clusters.66 As such, the larger cation, K+ vs. Na+, will likely lead to more open structures being obtained. Herein, this appears to be the case—as the mol fraction of Na+ activator decreased (K+ activator, K2CO3, increased), larger d-spacings were observed along the 002 plane indicative of a more open structure.
Considering the cations involved in molten salt templating will affect the final carbon structure via their size but also through ionic interactions, the charge density of each activating mixture employed to prepare the carbons 0–100Na was also evaluated with respect to the fringe parameters discussed above (d002, Lc). As shown in Fig. 2, the d-spacing along the 002 plane decreased with increasing mol fraction of Na+, which corresponds to increasing charge density (Fig. S33). Then, Lc increased with increasing mol fraction of Na+, also corresponding to an increase in charge density (Fig. S34). Thus, increased charge density within the molten salt phase led to templated porous N-doped carbons with more restricted space between stacked carbon sheets and increased crystallinity. The smaller d002 with increasing mol fraction Na+ (increased charge density in activation mixture), could be disadvantageous towards ORR electrocatalysis via limiting mass transport of O2 and reaction intermediates.
The d101, corresponding to in-plane spacing, was similar for all XNa at ∼4.1 Å (Fig. S6), indicating that the interplanar spacings between the 101 planes were fairly consistent between the N-doped carbons prepared with differing activating mixtures. This observation likely means that the degree of N-doping was similar among 0Na–100Na along the 101 plane as differences in the prevalence of defects, such as dopants, would have likely led to larger deviations in d101 values. On the other hand, La varied from 29.5–33.5 Å, with larger La corresponding to the N-doped carbons prepared with an Na+ poor (K+ rich) mixture of the alkali carbonates (Fig. S6). This trend indicates that the average width of a graphitic carbon sheet increased with the increasing K+ mole fraction (decreasing Na+) in the activating mixture, which corresponds to a decrease in the charge density of the molten phase during templating. Essentially, decreased charge density during molten salt templating likely led to reduced interaction with the carbon matrix and reduced incurred defects with regard to individual carbon sheets, hence the observed increase in La indicative of larger, undisturbed graphitic-like sheets.
In summary, an increased charge density corresponding to an increased Na+ mol fraction in the activating mixture led to a decrease in the spacing between stacked carbon sheets (d002) and an increased number of stacked carbon sheets along the 002 plane. However, this increased charge density had little effect on the lateral spacing (d101) and was deleterious towards maintaining high order for a given carbon sheet, resulting in decreased sheet width as determined by a decrease in La. Thus, while an increase in Na2CO3 mol fraction led to increased crystallinity along the 002 plane, it led to smaller d002 and less continuous sheets of graphitic carbon. This result indicated that a more K2CO3 rich activating mixture may be preferable for the ORR instead, as the corresponding larger d002 could enable improved mass transport for the ORR via increased accessibility and the increased La corresponding to wider carbon sheets could imply less carbon oxidation under alkaline conditions via a reduction in edge sites exposed to the electrolyte.55
Beyond XRD, Raman spectroscopy was employed to determine the extent of in-plane defects with the changing mol fraction of Na+ activator, Na2CO3, in the activating mixture implemented for pore generation. Raman spectra were collected from 500 to 3500 cm−1 using a 514 nm laser as the excitation source. Each spectrum was fit with six bands at approximately 1250, 1350, 1510, 1600, 2700, and 2900 cm−1 corresponding to the D4, D, D3, G, 2D, and G + D bands (Fig. S7).37–41 The D band at 1350 cm−1 is observed because of defects within the carbon structure. The D4 band at ∼1250 cm−1 and D3 band at ∼1500 cm−1 are attributed to amorphous character and vibrations associated with the presence of heteroatoms and other defects.37,38 The G-band at ∼1580 cm−1 is associated with the stretching mode of sp2 bonded carbon. Then, the bands at higher wavenumber, 2D and G + D, are overtone bands corresponding to the major D (1350 cm−1) and G (∼1580 cm−1) bands, which represent higher vibrational energy transitions that are a multiple of the fundamentally observed frequencies (D and G bands).39–41 Following the fitting of the Raman spectra, the D and G bands were integrated to determine their intensities and to calculate the ID to IG ratio. As shown in Fig. S9, the ID/IG decreased with increasing Na+ mol fraction (decreasing K+ mol fraction) from 2.9 to 1.4, levelling off at 75% Na+. Given XPS results indicated similar N-content amongst the prepared carbon samples (Fig. 1 and S5), this difference in ID/IG ratio could only arise from differences in structure and porosity, leading to in-plane defects induced by varying the composition of the activating mixture. The observed decrease in ID/IG indicates that a Na+ poor (K+ rich) activating mixture results in a greater level of defects and, as a result, a decreased degree of graphitization. This observation aligns with XRD results, which indicated decreased crystallinity for carbons prepared with a K+-rich reaction mixture and increased crystallinity when the mixture was rich in Na+. In addition, La determined from Raman results decreased with increasing Na+ mol fraction (decreasing K+ mol fraction), once again, matching with results from XRD. An abundance of K+, as compared to Na+, in the activating mixture likely led to an increased number of defects but larger La owing to the differing charge densities of the in situ formed molten salt templates at elevated temperature. The K+ rich molten salts were less charge dense, likely leading to differing charge networks that interacted with the carbon at 1000 °C and upon cooling, such that more open structures but increased carbon sheet widths were realized as compared to those rich in Na+. Thus, from XRD and Raman spectroscopy, an activating mixture poor in Na+ (rich in K+), will result in a more defective and open structure with individual carbon sheets exhibiting increased order (larger carbon sheet widths).
Scanning electron microscopy (SEM) images showcased morphological changes of the activated carbons as the mol fraction of the activating mixture was varied. Specifically, larger defects were observed with decreasing Na+ mol fraction (increasing K+ mol fraction). These larger defects could positively influence electrocatalytic performance by improving mass transport to active sites within small pores. As shown in Fig. S10, all of the activated carbons consisted of micron-scale particles ranging from 10–40 μm in diameter. For the carbons activated with both K2CO3 and Na2CO3 (25Na, 50Na, and 75Na), additional particulates were formed on the surface of these micron-scale particles, ranging from a few microns in diameter to hundreds of nanometers. For those carbons activated with either pure K2CO3 or Na2CO3, only the 10–40 μm particles were observed. In addition, as shown in Fig. S10, at the micron scale, surface defects are observed for 0Na–75Na but not for 100Na at the maximum achievable magnifications for imaging. This observation indicates that increasing the mol fraction of Na+ led to reduced surface defects at the micron scale, meaning more K+ led to a more defective surface. This increased presence of defects with respect to K+ could be the result of its increased size and decreased charge density as compared to Na+ during the templating process. The defects observed from these SEM images likely represent macropores (>50 nm in diameter), which could enable improved transport of O2 to smaller pores for enhanced electrocatalytic ORR performance. STEM images were attempted of the 50Na sample to obtain more high-resolution images to observe these features, however, high-quality images could not be obtained at the nanometer scale due to the amorphous character of the carbon (Fig. S35). Nevertheless, the presence of macropores was confirmed via N2 porosimetry, which is discussed in more detail below (Fig. 3). Thus, the increased presence of defects that occur by employing an increased K+ activator in the activating mixture is likely beneficial towards the ORR via the generation of hierarchical porous networks that afford high accessibility of electroactive species to active sites within small pores.
![]() | ||
Fig. 3 Surface area and porous structure of activated N-doped carbons. (a) Cumulative pore volume, pore size distribution, and BET surface area, (b) relative pore size distribution. |
Because of the influence of surface area and pore size on electrocatalytic performance and, specifically, mass transport for the ORR, nitrogen isotherms were collected at 77 K for each N-doped carbon sample to determine their BET SA and extract pore size distribution information using kernels derived from NLDFT on carbon. The isotherms were mixed type I and IV isotherms with H4 hysteresis, indicating that the carbons consisted of both micro- and mesopores (Fig. S11). From the isotherms, it was clear that N2 uptake decreased with the increasing mol fraction of Na+ activator, meaning that a greater mol fraction of Na+ led to lower BET SAs. This observation/finding was confirmed by multi-point BET analysis, wherein point selection satisfied the Rouquerol criteria (positive C-value, the p/p0 value corresponding to monolayer coverage fell within the BET analysis range, etc.) (Fig. S12).67 As shown in Fig. 3a, the BET SA decreased with increasing mol fraction of Na+ from 1202 to 773 m2 g−1. Pore size distributions showed that absolute micropore volume decreased with increasing mol fraction of Na+, as observed with BET SA (Fig. 3a). The absolute mesopore volume ranged from 0.15–0.24 cm3 g−1, wherein the maximum mesopore volume was observed for the 50Na carbon prepared with an equimolar activating mixture of Na2CO3 and K2CO3. As such, the absolute mesopore volume increased with increasing charge density of the activating mixture until the activating mixture consisted of 75% Na2CO3. For 75Na and 100Na, the absolute mesopore volumes were essentially the same at ∼0.18 cm3 g−1. Mesopore formation may have been hindered because of pore collapse and deformation. This pore collapse and deformation likely arose via the loss of N as HCN, N2, and NCCN, as the precursor contained approximately 45 at% N while the activated carbons had 1.7 ± 0.5 at% N at the surface.46 However, similar N-loss attributable to pore collapse is expected between the samples because the carbons were prepared from the same precursor and had consistent final N-content. As such, this difference in mesoporosity likely resulted from molten salt templating effects. Specifically, the incorporation of different cations has been shown to alter pure molten salts, indicating that a mixed alkali-cation molten phase could give rise to drastically different charge networks for structure guidance and pore formation.68,69 Additionally, at increased temperature, the arrangement of the molten phase has been shown to vary, leading to the disruption or loss of charge networks,68 which could influence porogenesis upon cooling and, ultimately, O2 transport through porous carbon for electrocatalysis.
Nevertheless, given the inclusion of Na2CO3 in the activating mixture led to a limited volume of micropores, this sample set enables clarification of the impact of micropores on ORR electrocatalysis. When the micropore volume of the carbon is subdivided into two groups: 1–2 nm pore volume and <1 nm pore volume, only the pore volume for pores <1 nm decreased with increasing mol fraction of Na+ activator. This observation indicates that tunability via the mol fraction of Na+ arises from modulation of very small pores. This effect of Na+ mol fraction on the smallest pore sizes is supported by the relative pore size distribution shown in Fig. 3b, wherein the proportion of pore volume associated with <1 nm micropores decreased from 56 to 32% with increasing mol fraction of Na+ while that for 1–2 nm pore volumes was consistent between samples. Thus, the mol fraction of Na+ affects the volume of micropores <1 nm such that the BET SA decreases with increasing mol fraction of Na+. Such reduced micropore volume and BET SA could hinder mass transport of O2 for ORR electrocatalysis.
In addition to N2 porosimetry, CO2 porosimetry was performed to further characterize microporosity, as microporosity has been attributed to improved performance via trapping of O2 molecules and providing an increased number of accessible active sites for ORR electrocatalysis.23,52 The CO2 isotherms were collected at 0 °C and kernels derived from NLDFT on carbon were applied to determine pore size distributions. As shown in Fig. S14, CO2 adsorption was very similar amongst the activated N-doped carbons. The cumulative pore volume ranged from ∼0.03 to 0.05 cm3 g−1, and the pore size distributions were nearly identical between samples with pore widths of less than 0.4 nm, from 0.4 to 0.6 nm, and from 0.7 to 1.4 nm (Fig. S14). Deviations in pore volume were only observable for volumes associated with pores from 0.6 to 1.4 nm in diameter. Such a difference in pore volume associated with ∼1 nm diameter was also observed from N2 porosimetry measurements, which can probe pore widths as small as ∼0.7 nm in diameter (Fig. 3).67 Thus, while the mole fraction of Na+ does not appear to lead to extensive alterations to ultramicroporosity volume (<0.7 nm), there is a clear effect on pore volumes within the micropore regime (0.7 to 1.4 nm).
When the BET SA and N2 pore size distribution data were compared with structural data, strong correlations with fringe parameters determined from XRD patterns, especially the d-spacing of the 002 plane, were observed. Both the BET SA and small micropore volume (<1 nm) correlated positively with d002, while other variables (1–2 nm, 2–50 nm, > 50 nm, and total pore volume) reached their maximum at 50Na (Fig. S34). Given d002 depends on the mol fraction of Na+ activator as discussed above, the correlations with BET SA and pore volume < 1 nm indicate that the mol fraction of Na+ has a direct effect on the generated porous structure. Specifically, increased Na+ mol fraction leads to more ordered surfaces as indicated by the decrease in d002 and reduced FWMH of the d002 peak from XRD (Fig. 2). As such, the Na+ activator yields more ordered crystal growth for porogenesis via molten-salt-templating, which entails fewer defects, explaining the reduced micropore volume of the carbons prepared with Na+-rich activating mixtures. This trend also correlates with the charge density of the molten phase, which increased with increasing mol fraction of Na+ activator. Thus, a more charge-dense molten salt led to more ordered carbons at the expense of microporosity. Thus, the K+-rich activating mixtures (Na+-poor) with reduced charge density generated more open porous structures with increased microporosity, which could enable improved mass transport during the ORR to boost performance.
Following analysis of physicochemical properties, the activated N-doped carbons were evaluated as ORR electrocatalysts using a RRDE setup. Overall, the prepared carbons demonstrated similar values for Eonset (0.719–0.792 ± 0.004–0.015 V vs. RHE), which only varied significantly when comparing the carbons activated with the pure components of the activating mixture (Fig. 4). On the other hand, values of E1/2 did vary between the porous N-doped carbons, wherein E1/2 decreased with increasing Na+ mol fraction in the activating mixture from 0.691 to 0.557 V vs. RHE. However, there was not an obvious difference in selectivity, n, between the carbons with the exception of the 100Na sample, considering the uncertainty in n for 0–75Na (there was observed overlap of the standard deviation for n between these carbons). At 0.7 V vs. RHE, n = ∼3.9 for 0–75Na, while n = 3.5 for 100Na (Fig. 4d). Nevertheless, for all the carbons, n decreased with decreasing potential, stabilizing around 3.2–3.6 at potentials below 0.5 V vs. RHE (Fig. S26).
With regard to Eonset, the similarity in this value between the carbons (0.719–0.792 ± 0.004–0.015 V vs. RHE), indicated that Eonset was primarily tuned by the chemical composition, the N-content and N-motif distribution of the carbons. As aforementioned, N-sites function as inducers of active sites via a partial positive charge on C atoms.18,23 Given the carbons all have, statistically, the same N-content, similar Eonset is unsurprising. However, despite this predominant effect of N-content, a minute effect from BET SA and porosity is observed. As shown in Fig. 4b, c and f, 0Na and 100Na have different Eonset at 0.782 ± 0.004 and 0.719 ± 0.006 V, respectively. This difference arises from differences in their BET SA and porous structure, as their N-, O-, and C-content did not differ drastically from each other (Fig. S5). Of the carbons studied, 0Na had the largest d002, BET SA, and pore volume associated with pores < 1 nm in diameter, while 100Na had the smallest associated values (Fig. 2 and 3). The increased surface area and more open structure of 0Na, as indicated by its larger d002, BET SA, and small micropore volume, likely contributed to improved Eonset via increased access to N-sites for electrocatalysis. Interestingly, 75Na had a more similar Eonset to 0Na than 100Na (Fig. 4f). Given the standard deviation of Eonset for 75Na falls within the spread of Eonset for other carbons prepared with K+ in addition to Na+, these results imply that there may be threshold pore volumes and BET SA that correlate with increased access to N-sites such that Eonset becomes increasingly positive. For the studied sample set, these threshold values appear to be a pore volume associated with <1 nm pores of greater than 0.15 cm3 g−1 and a BET SA of greater than 770 m2 g−1 (Fig. 3). Thus, the structure and porosity of N-doped carbons can influence kinetic parameters such as Eonset by modulating the number of exposed active sites to the electrolyte. Additionally, the positive correlation between Eonset and small micropore volume, as shown in Fig. 5a, indicates that these small pores are beneficial towards ORR electrocatalysis and likely contain the catalytic centers for electrocatalysis.
![]() | ||
Fig. 5 Correlations of d002, micropore volume (<1 nm) determined from N2 porosimetry at 77 K with (a) Eonset, (b) E1/2, and (c) n at 0.7 V vs. RHE. |
Increased variability was observed with respect to differences in E1/2 between the N-doped carbons. Overall, the value of E1/2 increased with decreasing mol fraction of Na+ (increasing mol fraction of K+) employed for pore generation, with no overlap in the standard deviation of E1/2 values further supporting this observed trend, with the exception of 50Na and 75Na, which had nearly identical relative pore size distribution (Fig. 4c and e). Given E1/2 resides in the mixed control region of the LSV, it is influenced by both kinetics and mass transport limitations of the carbon materials. The similarity in Eonset indicated that kinetics are quite similar between the carbons because of similarities in chemical composition, implying that changes in E1/2 are primarily tuned by differences in mass transport that arise from differences in textural properties. Specifically, E1/2 was most positive at 0.691 ± 0.004 V vs. RHE for 0Na with the largest BET SA, micropore volume, largest d002, and ID/IG ratio (Fig. 2, 3, and S9). The E1/2 decreased with increased Na+ mol fraction, which corresponded to a decrease in BET SA, micropore volume, d002, and ID/IG ratio: an increase in Na+ mol fraction led to reduced BET SA, micropore volume, and ‘openness’ between carbon sheets such that mass transport of O2 is increasingly limited. As discussed above, this decrease in hierarchical porous structure originates from an increase in charge density of the molten salt phase during templating at 1000 °C. Thus, using K2CO3 alone for molten salt templating appears most beneficial for imparting the optimal porous, (semi)crystalline, and defective structure for this sample set with respect to ORR electrocatalysis.
While the activator(s) employed and the fractions of them used for synthesis had a clear effect on E1/2, there was no clear influence on selectivity, n, beyond there being a clear difference between 100Na and the 0–75Na samples (Fig. 4). At 0.7 V vs. RHE, for 0–75Na, n was ∼3.8–3.9, indicative of high selectivity for the 4e− pathway of the ORR. On the other hand, 100Na had an n of 3.5 ± 0.1, implying that the 2e− pathway is occurring alongside the 4e− pathway to a greater extent than for the carbons activated with a proportion of the K+ activator. Given the similar chemical composition between the 0–75Na carbons and 100Na (Fig. S5), this difference arises from their differences in structural and textural properties. With regard to structural properties, d002, indicative of the spacing between stacked carbon sheets, was reduced for 100Na as compared to 0–75Na (Fig. 2). The larger d002 of the 0–75Na samples, from ∼3.95 to 3.80 Å, respectively, indicates a more open structure for O2 transport through the carbon material, as compared to 100Na with a d002 of 3.69 Å. Despite the mere ∼0.10–0.25 Å difference in d002, this difference in n likely occurs owing to the kinetic diameter of O2 at ∼3.5 Å, as O2 would experience more limited transport to active sites through the more confined space between carbon sheets for 100Na as compared to 0–75Na. In addition, as seen from the Raman spectra in Fig. S7, the extent of defects dramatically decreased from 0 to 100Na, wherein the peak intensity of the G-band surpassed that of the D-band. This result indicates that, when no K+ is present for activation, a less defective doped carbon is produced, leading to a more restricted structure, preventing transport to active sites. This concept is supported by the reduced BET SA and pore volume observed for 100Na as compared to the 0–75Na samples (Fig. 3). The smaller BET SA and pore volume of 100Na likely led to restricted access of O2 to smaller pores, as probed by CO2 at 0 °C (Fig. S14). Despite 100Na having nearly identical CO2 accessible pore volume to the 0–75Na carbons at 0 °C, this volume was less accessible owing to its reduced volume associated with larger pores, as probed by N2 at 77 K (Fig. 3). Thus, the reduced selectivity, n, at 0.7 V vs. RHE for 100Na vs. 0–75Na is likely a result of a portion of active sites being essentially blocked from participation in the ORR and a lack of an interconnected pore network for the facilitation of O2 transport.
However, as the applied potential decreased, the selectivity, n, was no longer different between 100Na and 0–75Na (Fig. S26). This result implies that with decreasing potential, there is increased driving force for the ORR to proceed via the 4e− pathway. Further analysis of the LSVs indicates that this apparent favorability for the 4e− pathway for 100Na is actually the result of indirect 4e− transfer rather than the direct 4e− pathway. Specifically, as shown in Fig. 4b, for 100Na, a pseudo-limiting current density is obtained at ∼0.4 V vs. RHE, only for the current density to decrease further and reach a new limiting current density. This “two-wave” behavior is indicative of a 2-step 2-electron reduction reaction. Specifically, O2 is reduced to HO2− via 2-electrons followed by 2-electron reduction of HO2− to hydroxide ions.33,70 This 2-step reduction process can occur as a result of tortuous networks within the N-doped carbon. Essentially, due to these tortuous networks, HO2− molecules have a longer lifetime at the working electrode, leading to further reduction to hydroxide ions via 2-electrons rather than being transported to and detected at the Pt ring of the RRDE. Thus, this ‘two-wave’ behavior can lead to an observed selectivity, n, nearing 4 (Fig. 4b and S26). As discussed, this 4-electron transfer consists of two discrete steps and not the direct 4-electron pathway of the ORR. Thus, when this behavior is observed, the extent to which the direct 4-electron pathway occurs may be inflated, especially at more negative potentials.
Upon further observation of the LSV shapes of the activated carbons, this behavior is also clearly visible for 75Na and 50Na, although to a lesser extent. Thus, for 75Na and 50Na it is likely that the indirect 4e− pathway is being catalyzed in addition to direct 4e− transfer, which is supported by the decrease in n from ∼3.8 at 0.7 V (vs. RHE) to less than 3.8 with decreasing potential. From Fig. S26, this observation appears likely as there is a more dramatic change in n with decreasing potential for 50–100Na than 0Na and 25 Na. This difference in selectivity between the N-doped, porous carbons trends with the relative micropore volume for pores < 1 nm, as determined from N2 porosimetry at 77 K. Specifically, 0Na and 25Na consisted of 56 and 43% of pores < 1 nm in diameter, while 50Na, 75Na, and 100Na were at 36, 35, and 32%, respectively. Thus, the tuning of micropores via modulation of the activating mixture resulted in less porous carbons, such that tortuous porous networks contributed to indirect 4e− transfer via a 2-step 2-electron reduction rather than direct 4e− transfer. Specifically, a decreased mol fraction of Na+ (increased K+) led to increased microporosity such that the direct 4e− pathway was selectively achieved. This hypothesis is further supported by the double layer capacitance (Cdl), which acts as a proxy for electrochemically active surface area. As shown in Fig. S29 and S30, Cdl is largest for 0Na and 25Na, matching with N2 porosimetry results that indicated these samples had the largest and most accessible surface for the ORR.
In addition to selectivity, variability was also observed between samples concerning limiting current (Fig. 4b). An increased limiting current is often associated with increased electrochemically active surface area or surface area in general, as limiting current is reached when the reaction rate is entirely controlled by mass transfer and not kinetics. However, this correlation is not always observed, as for the carbons studied herein. This trend between limiting current and surface area is likely not observed for multiple reasons: (1) variability in surface roughness and electrode geometry leading to differing thicknesses in the diffusion layer at the disk of the RRDE that impacts the current response in the mass-transfer controlled region,71 (2) those carbons (50–100Na) undergoing an additional 2-electron reduction step, as compared to 0 and 25Na, indicative of additional kinetic effects and not pure mass transfer control, which makes direct comparison of the current response past 0.4 V vs. RHE challenging, (3) limiting current densities falling within a standard deviation of one another as result of similarities in pore size distributions, such as for 0 and 25Na, and the limitations of drop-casting for electrode preparation.72 As such, it is difficult to compare samples in terms of limiting current at reduced potentials.
Nevertheless, given the most significant differences in electrochemical performance and selectivity were observed for the two extremes of the sample set, 0Na and 100Na, these two carbons were evaluated for long-term stability. Chronoamperometry was performed for 24 h at an applied potential of 0.4 V vs. RHE. In addition, SEM images were taken before and after chronoamperometry to gauge changes to morphology. As shown in Fig. S31, both 0Na and 100Na exhibited a steep decay in signal over the first 3–4 hours of long-term testing. Then, a gradual decay in current density was observed from 4–24 hours. Of the two samples, 100Na demonstrated increased decay, as compared to 0Na, with current density decaying by 25% compared to 20% for 0Na at 3.4 h. Additionally, the current density for 100Na also exhibited increased decay from 3.4 h to 24 h at 33% compared to 0Na at 24%. While these decays in current density are large, the reduced decay in current density of 0Na as compared to 100Na indicates that increased in-plane order and a reduced number of edge sites contribute to improved stability. As discussed above, 0Na with an activation mixture consisting entirely of K2CO3 and, as a result, K+ ions during molten salt templating, had a larger La as compared to 100Na prepared with 100% Na2CO3 for pore generation. This increased La for 0Na vs. 100Na was observed with both XRD and Raman spectroscopy (Fig. S6) and corresponded to increased in-plane order and width of graphitic carbon sheets. This increased lateral order led to a reduced number of edge sites, which have been considered as ‘hot spots’ for carbon oxidation under alkaline conditions.55 Thus, 0Na demonstrated increased stability compared to 100Na.
While a clear decay in current density was observed for 0Na and 100Na, the comparison of SEM images taken before and after long-term testing showed no obvious changes in morphology on the micron scale (Fig. S32). Thus, current density decay is likely the result of changes to the nanostructure of the N-doped, porous carbons or changes in chemical composition, as has been discussed in past literature.24,48 Given the similar chemical composition of 0Na and 100Na (Fig. 1), current decay related to surface group transformation was likely similar, as reflected by the ∼5% difference in current decay after 3.4 h of long-term testing. However, as discussed above, 100Na possibly underwent further oxidation as compared to 0Na with time because of decreased lateral order prior to ORR electrocatalytic testing. Essentially, the increased number of edge sites of 100Na likely led to increased oxidation and a further disordering of the microstructure, which would lead to even further degradation of the carbon material.
To further gauge material transformation during stability testing, Raman spectroscopy was performed before and after long-term testing (24 h, chronoamperometry) of the 0Na sample on a rotating ring disk electrode (RRDE) (Fig. S33). Following this long-term test, the ID/IG ratio decreased from 2.16 to 1.57, indicating a loss in disorder to the carbon electrocatalysts. Given the D peak at 1350 cm−1 arises because of defects in the carbon material, this decrease implies that those defects are somehow being lost. A reduction in the level of defects is likely the result of either the loss of N-dopants or the agglomeration of nanometer diameter pores. Thus, although N-doped carbon electrocatalysts are operating under reducing potentials, there is still material degradation occurring.
XPS spectra, Raman spectra, fringe parameters, SEM images, N2 isotherm data, BET plots, cumulative pore volume and pore size distribution data, CO2 adsorption data, electrochemical data, and STEM images. See DOI: https://doi.org/10.1039/d5ta04494d.
This journal is © The Royal Society of Chemistry 2025 |