Mohammad Yeganeh Ghotbi*a,
Behzad Felia,
Marziyeh Azadfalaha and
Masoumeh Javaherib
aMaterials Engineering Department, Faculty of Engineering, Malayer University, Malayer, Iran. E-mail: m.yeganeh@malayeru.ac.ir; yeganehghotbi@gmail.com
bMaterials and Energy Research Center, P. O. Box 14155-4777, Tehran, Iran
First published on 15th October 2015
The extensive research on the synthesis of nitrogen-doped carbon materials (NCMs) as non-precious metal catalysts (NPMCs) has shown a promising future for applying the NPMCs to catalyze the slow oxygen reduction reaction (ORR) in polymer electrolyte fuel cells (PEFCs) and, therefore, the widespread use of devices based on PEFCs. However, the main reasons for the delay in starting the practical use of NCMs in PEFCs are the use of very specific organic chemicals as well as multiple stages of the catalyst synthesis, lack of stability, activity and/or selectivity limitations. Here we show that the NCMs can be produced by a simple route on a very large scale using any organic anions for the first time. The synthesized carbon catalysts showed highly porous structures, tunable nitrogen content and high electrochemical performance. As the best performing catalyst it had an open circuit potential (OCP) of as high as 1.04 V, and a Tafel slope of as low as 38.0 mV per decade with an exchange current density of as high as 1.34 × 10−4 A cm−2, which is a much higher performance compared to other NPMCs.
In general, NCMs as NPMCs are mostly produced by coating macrocycles, chelates, polymers (also monomers and oligomers) or other organic compounds containing nitrogen with or without some transition metals, in particular Ni, Co and Fe, onto carbon nanoparticles and then heat-treating the obtained composites in an inert or nitrogen-rich atmosphere (N2, NH3, or HCN).5–7 The carbonic layers consisting of M–N sites onto carbon nanoparticles are formed by the coating process. It is known that due to a radical enhancement, these active electrocatalyst centers facilitate the ORR and improve the performance and durability of the cell.5,8,9 Accordingly, NCMs with high electrocatalytic activity and durability have been produced by various methods such as the heat-treatment process, arc discharge, chemical vapor deposition, plasma treatment and ball milling.5 In almost all cases, the type of the organic materials, the heat-treatment temperature, synthetic conditions, etc. play important roles in determining the efficiency of the NCMs as catalysts for the ORR.9–16 Consequently, the significant factors to improve the electrocatalytic activity are a high nitrogen content within the carbon structure (as active sites), high porosity to increase the surface area, and the presence of functional groups such as oxide, hydroxide, carbonyl and carboxide in addition to some transition metals on the carbon surface.5,15,17,18 To this end, there are still serious problems due to the production of NCMs in a large scale and cost-effective way. For instance, multiple stages are needed to obtain the product in addition to a low nitrogen content in the produced carbons.5,6,16 Meanwhile, most macrocycles, chelates and other organic chemicals used for producing NCMs are poisonous and expensive.
Herein, we present an easy method for a large scale production of highly efficient nitrogen-doped carbon nanosheet catalysts for the ORR. In this method, an α-phase layered zinc hydroxide (αLH) was used as a layered nanoreactor to prepare the carbon nanosheets. αLH, typically zinc hydroxide nitrate (Zn5(OH)8(NO3)2·2H2O) with card no. 24-1460, is a synthetic mineral with positively charged brucite-type layers constructed solely with one type of cation (Zn2+).19–21 These positive charges can be neutralized by anions, which together with water molecules are located inside the interlayer spaces of the inorganic layers. The α-phase zinc hydroxide nitrate (ZHN) was chosen as the inorganic layered host and succinate/salicylate anions (without and with benzene rings) as the organic guests, which were encapsulated into the intergallery of the host to produce the organic–inorganic layered nanohybrids by a simple ion-exchange process. Heat-treatment of the nanohybrids resulted in the production of the nitrogen-doped carbon catalysts.
XRD patterns for the as-synthesized ZHN and also its resultant nanohybrids with succinate and salicylate anions with different ion-exchange times (ZHSu and ZHSa) are shown in Fig. 2. As observed, all the samples show the α-phase brucite-like structures with the basal spacing peaks around 9.7 Å.19–21 The inorganic–organic nanohybrids with succinate (ZHSu) and salicylate (ZHSa) anions have a new sharp peak at 11.8 and 15.6 Å, respectively. This shows that the organic anions could be intercalated into the interlayer of the parent material via the ion-exchange process. Due to the bigger size of the organic anions compared with that of nitrate, the interlayer space has expanded to accommodate succinate/salicylate, resulting in the expansion of the basal spacing from 9.7 Å for the host material to 11.8 and 15.6 Å for the ZHSu and ZHSa nanohybrids, respectively. The basal spacing peak at 9.7 Å due to the parent material (ZHN) can also be seen owing to the incompleteness of the ion-exchange process for all the nanohybrids. However, with the increase of the ion-exchange times, the peak intensities of the basal spacing for the nanohybrids increased at the expense of that of the as-synthesized ZHN. It indicates that both the anions of succinate (or salicylate) and nitrate with different mass ratios are simultaneously present inside the interlayer space of the nanohybrids. The nitrogen-doped carbon catalysts with various nitrogen percentages could be produced using the heat-treatment of the nanohybrids via a chemical reaction between the nitrate and organic groups which are encapsulated between the inorganic layered nanoreactors.
![]() | ||
Fig. 2 XRD patterns for the as-synthesized zinc hydroxide nitrate (ZHN), its resulting nanohybrids with succinate (ZHSu) and salicylate (ZHSa) anions and the resultant CSu3 and CSa6 carbon catalysts. |
Fig. 2 also shows XRD patterns for the CSu3 and CSa6 carbon catalysts. A very broad peak around 26° can be seen in both patterns. This peak reflects the 002 plane of the disordered graphite structure within the c-planes. The carbon patterns also reveal that the catalysts consist of mainly amorphous carbon material without any remaining zinc oxide after acid etching.20
Fig. 3 shows the FTIR spectra for the as-synthesized ZHN, succinic and salicylic acids, the ZHSu3 and ZHSa6 nanohybrids and the CSu3/CSa6 carbon catalysts. In the ZHN sample, lattice vibrations of the metal–oxygen bonds Zn–O and Zn–OH are observed at 436 and 468 cm−1, respectively.20–22 Two weak bands at 837 and 1018 cm−1 and an intense band at 1384 cm−1 can be ascribed to ν2, ν1 and ν3 of the intercalated anion, nitrate.21 The water-bending vibration (δH2O) of the adsorbed/intercalated water molecules is observed at 1623 cm−1.20,21 The FTIR spectrum of succinic acid shows two bands at 918 and 1417 cm−1, which are due to the C–O–H out-of-plane and in-plane bending vibrations, respectively.23 The band at 1202 cm−1 is attributed to the C–O stretching vibration of the acidic group.23 Two bands at 1687 cm−1 (asymmetric) and at 1314 cm−1 (symmetric) are due to CO stretching vibrations.23 The FTIR spectrum of ZHSu3, the zinc hydroxide succinate nanohybrid, is composed of spectral band features of both salicylic acid and ZHN.23,24 The bands at 440 and 500 cm−1 are due to metal oxide and hydroxide vibrations. The bands at 1035 and 1060 cm−1 are attributed to C–O–H vibrations in the nanohybrid. The band at 1230 cm−1 indicates the presence of a C–O group. Two bands at 1350 and 1550 cm−1 are assigned to the symmetric and asymmetric vibrations of the carboxylic group. These red-shifted bands reveal that hydrogen bonding in the intercalated succinate with hydroxide and oxide groups within the brucite-like layers is much stronger than that for free succinic acid.20,22 The intense peak at 1384 cm−1 is due to the nitrate anion. It means that both nitrate and succinate anions are simultaneously between the nanohybrid layers. Fig. 3 also shows the FTIR spectrum for the CSu3 catalyst obtained by heat-treating the ZHSu3 nanohybrid followed by acid etching. There are two bands around 1334 and 1605 cm−1 due to C
N/C–C and C
C/C
O stretching vibrations, respectively.7,9,20 As can be seen, the spectrum doesn’t contain any bands in the range of 400–600 cm−1 due to M–O vibrations, confirming that the N-doped carbon catalyst is free of the metal oxide after the acid washing process.20 Similar results can be observed for the ZHSa6 nanohybrid and its resultant CSa6 carbon catalyst in Fig. 3.
Linear sweep voltammograms (LSVs) show the electrocatalysis activity on different carbon catalysts and also a Pt/C reference electrode (10 wt% E-TEK, 200 μgPt cm−2) in Fig. 4. Also, the related data are listed in Table 1. According to RDE measurements (Fig. 4a and b), all the samples are catalytically active toward oxygen reduction as observed in the ORR polarization plots. Also, the Tafel plots on different carbon catalysts are presented to compare their activity in Fig. 4c and f, according to their kinetic current densities by correcting the mass transport.25
Catalyst | ORR onset potential (V) | Tafel slope (mV dec−1) | i0a (A cm−2) | D1/2c (mol cm−2 s−1/2) | C (%) | N (%) | O (%) |
---|---|---|---|---|---|---|---|
a The i0 has been calculated based on the geometrical areas of the electrodes. | |||||||
CSu1 | 0.93 | 40.5 | 8.96 × 10−5 | 1.27 × 10−7 | 70.16 | 15.99 | 13.85 |
CSu2 | 1.02 | 38.3 | 9.61 × 10−5 | 7.53 × 10−8 | 72.62 | 13.86 | 13.52 |
CSu3 | 1.04 | 38.0 | 1.34 × 10−4 | 1.36 × 10−7 | 75.94 | 11.57 | 12.49 |
CSu4 | 0.97 | 42.2 | 1.12 × 10−4 | 1.69 × 10−7 | 80.50 | 9.20 | 10.30 |
CSa2 | 0.87 | 43.9 | 7.12 × 10−5 | 8.36 × 10−8 | 76.46 | 14.57 | 8.97 |
CSa4 | 0.97 | 50.6 | 6.67 × 10−5 | 5.19 × 10−8 | 80.85 | 9.64 | 9.51 |
CSa6 | 1.02 | 40.1 | 1.62 × 10−4 | 1.68 × 10−7 | 84.17 | 7.47 | 8.36 |
CSa8 | 0.97 | 40.6 | 9.85 × 10−5 | 1.21 × 10−7 | 87.43 | 4.84 | 7.73 |
Pt/C | 0.96 | 44.2 | 4.73 × 10−4 | 6.74 × 10−6 | 90 | — | — |
The LSV tests were also performed using carbon paper as the working electrode (Fig. 4d and e). As presented in Table 1, the ORR onset is as high as 0.87 V for the CSa2 catalyst (versus a reversible hydrogen electrode, RHE), prepared in a 2 h ion-exchange process. By changing the ion-exchange time (changing the C/N ratio), more improvements in the performance of the catalysts are gradually observed as evidenced by the lower ORR overpotentials and higher current densities. Specifically, the onset potentials show a positive shift from 0.93 to 1.04 V for the N-doped catalysts of CSu1 to CSu3 and from 0.87 to 1.02 V for the CSa2 to CSa6 catalysts.
The Tafel slopes in the low current density region can clearly identify the activity difference of the samples as a function of ion-exchange time as listed in Table 1. The table summarizes the onset potentials, the Tafel slopes and the exchange current densities (obtained from the Tafel equation,26 η = blog(i/i0), where η is the overpotential, b is the Tafel slope and i0 is the exchange current density) of the ORR on different catalysts. According to both the ORR tests (RDE and carbon paper), it is evident that the activity of the catalysts increases in the order:
CSu3 > CSu2 > CSu4 > CSu1 |
CSa6 > CSa8 > CSa4 > CSa2 |
The Tafel slopes of 38.0 and 40.1 mV per decade were measured for CSu3 and CSa6, as the best performing catalysts derived from succinate and salicylate anions, respectively. The Tafel slope values are lower than those for other NPMCs.1,5,6,8,27 The small values of the Tafel slopes are ascribed to the very fast kinetics of the transfer of electrons on the catalyst surfaces. Regarding the large thickness of the carbon catalyst layers that cause high electrical resistance and also mass transfer resistance, the observed high ORR performance and the observed current densities for the carbon catalysts are valued.25
Although the exchange current densities were changed slightly with varying the ion-exchange time, the i0 values are nearly similar for all the catalysts. However, in comparison with other NPMCs, the i0 values are at least three orders of magnitude higher.7,26,28
Oxygen permeability on different electrodes was measured by chronoamperometry according to the modified Cottrell equation.26 Permeability is the product of Db1/2cb, where Db is the diffusion coefficient and cb is the concentration of oxygen. Fig. 5 shows I vs. t−1/2 plots for oxygen reduction on different electrodes, indicating the existence of the linear relationship for all the catalysts. As observed in Table 1, the oxygen permeability values show a small increase with an increase in the ion-exchange time. This reflects the porous nature of each catalyst and the abundant presence of active sites on the catalyst surface, resulting in the faster permeation and, therefore, reduction of oxygen on the catalyst surface. In accordance with the electrochemical tests and the elemental analysis data presented in Table 1, one can deduce that higher activity is not directly related to higher nitrogen content. An optimum amount of nitrogen provides the highest activity toward the ORR.
As stated earlier, the XRD patterns of the CSu3 and CSa6 catalysts (Fig. 2) reflect that the carbon is mainly in an amorphous form with a disordered graphite structure, so that it may enhance the active site numbers via facilitating the incorporation of nitrogen within the carbon structure.5,7 The presence of C–N, C–C and C–O bonds in the carbon catalysts was confirmed via FTIR spectroscopy (Fig. 3). These carbon–nitrogen and carbon–oxygen bonds are active sites and/or facilitate the ORR activity.5,15,17 As we stated earlier, the doping of carbon with nitrogen and oxygen (or sulfur, phosphorous, etc.) causes an enhancement in the ORR activity. For example, the authors7,29 showed that pure carbon cannot show a high ORR activity. They proved that these carbon–nitrogen and carbon–oxygen bonds are active sites for facilitating the ORR. Nitrogen (and other doping agents) is an n-type carbon dopant which donates electrons to the carbon and facilitates the ORR.7,30 By acid washing and thus from the presence of C–O groups in a nitrogen-doped carbon catalyst, Nallathambi et al.29 showed higher ORR activity of the obtained carbon material.
Fig. 6 shows the cyclic voltammograms in Ar-saturated 0.5 M H2SO4 for all the carbon catalysts with two separate oxidation and reduction peaks which can be attributed to the characteristic changes of oxidation state of the carbon catalysts (quinine–hydroquinone).25,28,31
![]() | ||
Fig. 6 The cyclic voltammograms of different carbon catalysts in Ar-saturated 0.5 M H2SO4 at a 50 mV s−1 scan rate. |
These high performances for our carbon catalysts are due to the preparation method which results in the production of the highest percentage of pyridinic nitrogen (ca. 398.5 eV) among other C–N bonds within the nitrogen-doped carbon structure, as observed in Fig. 7e. It is well known that the pyridinic nitrogen is positioned on the edge of the graphite plane and the lone pair of electrons from the nitrogen is due to the ORR performance.13,32 The higher exposure of the smaller ion size of succinate (compared with the salicylate anion) to nitrate anions between the layers of ZHN is expected to render more pyridinic nitrogen, which enhances the ORR activity.
The morphologies of the catalysts were studied by FESEM as illustrated in Fig. 7. The images show highly porous nanosheets which result in high surface areas and greater access to active sites and, therefore, the catalysts obtain higher ORR performances.10,12,13,15 The microstructure shown in Fig. 7f affords a velvet-like texture with micro/mesopores of 1–3 nm and with a narrow size distribution which are excellent sites to capture reactants for the ORR. Large pores in addition to pore tortuosity increase the material’s electrical resistance, resulting in lowering of the conductivity of the material and its ability to capture the reactants.33 Also, the images show a higher surface area for CSu3 in comparison with that for CSa6. It is probably due to the lower heat-treatment temperature for the CSu catalysts (650 °C) in comparison with that for the CSa catalysts (850 °C). Moreover, as mentioned earlier, CSu3 has a greater percentage of pyridinic nitrogen than that for CSa6. Accordingly, it is expected that the ORR performance of CSu3 is higher than that of CSa6.
The catalyst inks were prepared by ultrasonic dispersion (45 min) of the carbon catalysts in an alcohol solution containing Nafion solution (5 wt% solution), distilled water (2 ml) and isopropyl alcohol (1 ml). The ink was then put into an oven at 90 °C for about 20 min to obtain a wet gel. The gel was painted onto the carbon paper (TGPH-0120T, Toray) by a brush, and the electrode was dried in an oven at 90 °C for 30 min to obtain a catalyst loading of 4 mg cm−2 (with a mass ratio of Nafion/carbon = 30/70), after which the electrode was also put in an oven at 120 °C for 30 min to ensure the removal of the solvents. Finally, the electrode was sintered at 200 °C for 1 h.26,35 For RDE measurements, the ink was deposited onto the glassy carbon disk electrode and dried in an oven at 50 °C for 1 h. The catalyst loading, the counter and reference electrodes and also the electrolyte and the potential scan rates were the same as those used in the LSV tests using the carbon paper. The RDE tests were measured at a 900 rpm rotating speed.
This journal is © The Royal Society of Chemistry 2015 |