Abstract
Understanding the correlation between exposed surfaces and performances of controlled nanocatalysts can aid effective strategies to enhance electrocatalysis, but this is as yet unexplored for the nitrogen reduction reaction (NRR). Here, we first report controlled synthesis of well-defined Pt3Fe nanocrystals with tunable morphologies (nanocube, nanorod and nanowire) as ideal model electrocatalysts for investigating the NRR on different exposed facets. The detailed electrocatalytic studies reveal that the Pt3Fe nanocrystals exhibit shape-dependent NRR electrocatalysis. The optimized Pt3Fe nanowires bounded with high-index facets exhibit excellent selectivity (no N2H4 is detected), high activity with NH3 yield of 18.3 μg h−1 mg−1cat (0.52 μg h−1 cm−2ECSA; ECSA: electrochemical active surface area) and Faraday efficiency of 7.3% at −0.05 V versus reversible hydrogen electrode, outperforming the {200} facet-enclosed Pt3Fe nanocubes and {111} facet-enclosed Pt3Fe nanorods. They also show good stability with negligible activity change after five cycles. Density functional theory calculations reveal that, with high-indexed facet engineering, the Fe-3d band is an efficient d-d coupling correlation center for boosting the Pt 5d-electronic exchange and transfer activities towards the NRR.
Keywords: Pt3Fe, nanowire, high-index, facet-controlled, N2 reduction
We have reported facile synthesis of Pt3Fe nanocrystals with tunable morphologies (nanocubes, nanorods and nanowires) to evaluate the performance of N2 electroreduction on different exposed surface structures.
INTRODUCTION
Ammonia (NH3), an essential composition for fertilizer feedstock, industrial and chemical precursors [1], not only plays a vital role in the development of ammonia fuel cells but is also a promising candidate for the hydrogen economy because of its high energy density and large hydrogen capacity [2]. However, from a thermodynamic point of view, the high bond energy of the triple bond in nitrogen (N2) makes it an intricate multi-step reaction to convert N2 to NH3 [3]. Currently, the Haber-Bosch method is most widely applied because efficient synthesis of NH3 from N2 and H2 has reached industrial-scale yields [4]. However, this method requires conditions of high temperature of 400–500 °C and pressure of 200–250 bar [5]. It also consumes 1–2% of the global annual energy supply and is responsible for >1% of global CO2 emissions [6]. Therefore, a more effective process for NH3 synthesis under mild conditions is highly desirable.
Recently, tremendous efforts for converting N2 to NH3 have been made in pursuit of efficient and sustainable catalysis with use of biocatalytic, photocatalytic and electrocatalytic methods [7–14]. One possible strategy for NH3 synthesis is electroreduction of N2 to NH3 in which the N2 reduction reaction (NRR) process can be operated by renewable electricity energy, and operated at mild temperature and pressure [15–18]. There have been several investigations on NRR with noble-metal catalysts (Ru [19], Au [20], Pd [21] and Rh [22]), as summarized in Table S1. More recently, several efforts have been devoted to optimizing electrocatalysts to enhance NRR, including size regulation [23], crystal engineering [24], ion incorporation [25], introduction of defect sites [26,27] and component regulation [28]. However, to the best of our knowledge, surface structure regulation, as one of the most effective strategies to precisely tune catalytic properties, has not yet been demonstrated for NRR.
Previous research revealed that the NRR performance is gravely limited by linear scaling of the two vital intermediate energetics between *N2H and *NH2 (* indicates the adsorption site). Further theoretical analyses disclose that Fe and Pt atoms can effectively address the energetics of *N2H and *NH2, respectively, synergistically providing efficient active sites to enhance NRR activity [29,30]. Herein, we report a facile method for selectively growing Pt3Fe nanocubes (NCs), Pt3Fe nanorods (NRs) and Pt3Fe nanowires (NWs) (Supplementary Fig. 1), which give rise to active and stable electrocatalysts for NRR. Optimized Pt3Fe NWs bounded with high-index facets exhibit much enhanced NH3 yield (18.3 μg h−1 mg−1cat, 0.52 μg h−1 cm−2ECSA; ECSA: electrochemical active surface area), Faradaic efficiency (7.3%) and selectivity (no N2H4 is detected) at −0.05 V versus reversible hydrogen electrode (RHE), and are much better than those of {200} facet-enclosed Pt3Fe NCs and {111} facet-enclosed Pt3Fe NRs. The Pt3Fe NWs also show durable stability with negligible activity decay for five cycles. Density functional theory (DFT) calculation reveals that, on the high-indexed surface engineering, strong orbital interaction between Pt and neighboring Fe sites induces an obvious correlation effect for boosting up Pt-5d electronic activities for efficient NRR.
RESULTS AND DISCUSSION
A simple wet-chemical method is adopted to control the synthesis of Pt3Fe nanocrystals. Taking Pt3Fe NWs synthesis as an example, potassium tetrachloroplatinate (II) (K2PtCl4) and iron nonacarbonyl (Fe2(CO)9) are chosen as metal precursors, ribose is used as the reducing agent, cetyltrimethyl ammonium chloride (CTAC) and oleylamine (OAm) are applied as the surfactant and solvent, respectively. Uniform Pt3Fe NWs with average diameter of 15 nm and zigzag border along the whole length are obtained (Fig. 1a and Supplementary Fig. 2), characterized by high-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM) and TEM. The X-ray diffraction (XRD) pattern (Fig. 1b) of Pt3Fe NWs shows distinct diffraction peaks at 40.32, 46.92, 68.42, 82.52 and 87.2o, which are readily indexed to (111), (200), (220), (311) and (222) reflections of face-centered cubic (fcc) Pt3Fe (JCPDS number 29–0717), being consistent with those of Pt3Fe NCs and Pt3Fe NRs (Supplementary Fig. 3) [31]. The crystal structure model of the Pt3Fe NWs represents a primitive cubic structure. It is composed of a periodic square matrix of Fe and Pt, which are located at the corner and face center of each unit cell, respectively. The HAADF-STEM image and elemental mappings show the elemental distributions of Pt and Fe (Fig. 1c). The Pt (green), Fe (red) and mixed images indicate that all elements are evenly distributed on the Pt3Fe NWs, confirming the alloyed structure (Supplementary Fig. 4). The selected area electron diffraction (SAED) image (Fig. 1d) further indicates that the Pt3Fe NWs have good crystallinity and fcc structure. Spherical aberration correction HRTEM images reveal the distinct lattice fringes. The measured value of lattice spacing is 0.195 nm. In addition, a high-index facet of {311} can be readily observed (Fig. 1e and Supplementary Fig. 5), which is reported to exhibit much higher catalytic performance compared to most common facets ({200} and {111}) because of the high density of atomic steps and ledges [32–34]. Notably, when ribose is replaced by maltose, while maintaining the other synthetic parameters unchanged, Pt3Fe NCs with average diameter of 5 nm are obtained (Fig. 1f and Supplementary Fig. 6). Meanwhile, Pt3Fe NRs with average diameter of 3 nm are realized by merely changing the precursor of Pt (Fig. 1g; Supplementary Figs 7 and 8). Detailed X-ray photoelectron spectroscopy (XPS) is carried out to determine the electronic properties of surface Pt atoms. Compared with commercial Pt/C, the Pt0 4f7/2 binding energy of Pt3Fe nanocrystals has a negative shift from 71.2 to 70.6 eV, indicating that the charge is transferred from Fe to Pt, because of the lower electronegativity of Fe (Supplementary Fig. 9a) [35]. We can see that the Pt 4f spectra show two distinct peaks, which are assigned to the Pt 4f7/2 and Pt 4f5/2 orbit levels. Each peak can be further split into two doublets, which are assigned to Pt0 and Pt2+ chemical valence states. It is clearly shown that the majority of the Pt of Pt3Fe is in mainly metallic state. The Pt(II)/Pt(0) ratio is summarized in Table S2. The XPS spectra of Fe 2p reveal two distinct peaks at 710.3 and 724.1 eV, which are assigned to the characteristic peaks of Fe 2p3/2 and Fe 2p1/2 orbit levels, respectively. The results indicate that the majority of the Fe is mainly in an oxidized state (Supplementary Fig. 9b) [36]. The lattice spacings of Pt3Fe NCs and Pt3Fe NRs are measured to be 0.195 and 0224 nm, corresponding to the {200} and {111} facets, respectively (Supplementary Figs 10 and 11). Hence, well-defined Pt3Fe nanocrystals bounded with distinct facets have been successfully created and can be adopted as ideal model electrocatalysts for fundamental understanding of the relationship between surface structure and catalysis.
Figure 1.

(a) TEM image and schematic illustration, (b) XRD pattern and crystal structure, (c) elemental mappings and HAADF-STEM image, (d) SAED image (white circles indicate the diffraction spots) and (e) spherical aberration correction HRTEM image of Pt3Fe NWs. HAADF-STEM images and schematic illustrations of (f) Pt3Fe NCs and (g) Pt3Fe NRs.
Pt NCs, Fe3O4 NPs (Supplementary Fig. 12) and Pt3Fe nanocrystals were first loaded onto carbon (C, Vulcan XC-72) to investigate the NRR properties. The resulting electrocatalysts were called Pt NCs/C, Fe3O4 NPs/C, Pt3Fe NCs/C, Pt3Fe NRs/C and Pt3Fe NWs/C. A schematic for the electrochemical NRR is shown in Supplementary Fig. 13. During each NRR measurement, pure N2 flowed into the cathode electrolyte at a flow rate of 30 standard mL/min, wherein N2 combined with electrons to form N2 reduction product. Using Pt3Fe nanocrystals as the cathodic catalysts, only NH3 without the by-product of N2H4 could be detected, highlighting the good selectivity for NH3 (Supplementary Fig. 14). The standard calibration curves are given in Supplementary Figs 15 and 16. Saturated calomel reference electrode is calibrated on reversible hydrogen electrode (Supplementary Fig. 17). During the NRR progress, the Pt3Fe NWs/C exhibited negligible decay in current density under different applied potential, indicating the good corrosion resistance ability (Fig. 2a). The Pt NCs/C (Supplementary Fig. 18) had very poor NRR activity compared with Pt3Fe NCs/C, indicating that introduction of Fe is essential for the NRR. In addition, the NRR activity of Pt3Fe NRs/C exhibited much improvement compared with Pt3Fe NCs/C. The average NH3 yield and Faraday efficiency (FE) of the Pt3Fe NWs/C under different applied potentials is given in Fig. 2b and c. The NRR was performed at 0 V, the measured yield of NH3 was 10.2 μg h−1 mg−1cat (0.29 μg h−1 cm−2ECSA) with the highest FE of 12.3%. The NH3 yield increased to a maximum value of 18.3 μg h−1 mg−1cat (0.52 μg h−1 cm−2ECSA) until −0.05 V, indicating that it consumed less energy to effectively convert N2 into NH3. The NRR performance of Pt3Fe NWs/C was much better than that of the Fe3O4 catalysts (Supplementary Fig. 19). The maximum yield of NH3 was further quantitatively determined by an indophenol blue method [37]. As observed, the calculation value of NH3 yield was close to the quantitative result from the Nessler reagent method (Supplementary Figs 20 and 21). The FE linearly decreased when it reached a more negative potential, because of the existence of a competition reaction between NRR and hydrogen evolution reaction (HER) [38]. We compared the NRR activity with different reaction temperatures at −0.05 V to assess the apparent activation energy and investigate the different exposed facets effect of Pt3Fe nanocrystals. As observed, the NH3 yield enhanced with increasing reaction temperature due to faster mass transfer rate of reactants (Supplementary Fig. 22). The estimated apparent activation energies were 11.0, 16.9 and 27.3 kJ mol−1 for Pt3Fe NWs/C, Pt3Fe NRs/C and Pt3Fe NCs/C, respectively (Fig. 2d). That is, the Pt3Fe NWs/C bounded with high-index facets could significantly decrease the apparent activation energy and hence enhance the NRR activity. Significantly, the NH3 yield and FE of Pt3Fe NWs/C were higher than those of Pt3Fe NCs/C and Pt3Fe NRs/C (Fig. 2e), indicating that different surface structures of Pt3Fe nanocrystals indeed have essential influence on the NRR activity and the high-index facets of Pt3Fe nanocrystals play a vital role in the improvement of NRR activity.
Figure 2.

(a) I-t curves, (b) NH3 yield and (c) FE of Pt3Fe NWs/C at different applied potentials. (d) Apparent activation energy for NRR with different Pt3Fe electrocatalysts. (e) Histograms of the NH3 yield and FE of Pt3Fe NCs/C, Pt3Fe NRs/C and Pt3Fe NWs/C. The error bars in (b) indicate the standard deviations of three independent tests in the same conditions.
We then carefully examined the N source of the produced NH3. There were no distinguishable peaks in the N 1s region, indicating that no nitrogenous species existed on the surface of synthetic catalysts (Supplementary Fig. 23). Replacing N2 by Ar, while maintaining other experimental parameters unchanged, no NH3 was detected (Supplementary Fig. 24). The same result was observed when Pt3Fe NWs/C was replaced by carbon powder loaded on carbon paper (Supplementary Fig. 25). The amount of NH3 increased linearly to the electrocatalysis time in N2-saturated electrolyte, which indicated that the produced NH3 came from the NRR process (Supplementary Fig. 26). Considering the small amounts of NH3 and NOx in atmosphere and feeding gas, we kept high-purity Ar and N2 continuously flowing into the KOH electrolyte without applied potential. The UV-vis results (Supplementary Figs 27 and 29) revealed that no NH3 and NOx were detected. Finally, 15N (99% 15N atom) isotope labeling experiments were carried out to further confirm the NH3 source. After continuous electrolysis at −0.05 V using 15N2 as the supplying gas, the 1H nuclear magnetic resonance spectra show a double coupling peak of 15NH4+ without the triple coupling peak of 14NH4+. Hence, the result confirmed that the NH3 was derived from the electroreduction of N2 in the presence of Pt3Fe NWs/C (Supplementary Fig. 30). In addition, by varying the N2 flow rate, the current density exhibited small change. The tiny fluctuation of FE and NH3 yield implied that N2 diffusion was not the rate-determining step (Supplementary Fig. 31) [39].
The intrinsic reason for higher catalytic activity of Pt3Fe NWs/C was investigated. The Tafel slope is a vital parameter to evaluate the HER mechanism [40,41]. The higher value of 163.8 mV dec−1 gained from Pt3Fe NWs/C suggested its sluggish HER kinetics, which may in turn enhance the NRR performance (Supplementary Fig. 32), further confirmed by theoretical calculation (Supplementary Fig. 33). After that, surface valance spectra was created to investigate the relationship between electronic effect and the binding strength of adsorbates. As observed, the d-band center shifts upwards from Pt NCs/C (−3.69) to Pt3Fe NCs/C (−3.31) when introducing Fe to Pt NCs/C. For transition metals, the lower binding energy of d-band center would cause weaker bonding between the adsorbates and metal surface [42–44]. Accordingly, the Pt3Fe NWs/C bounded with high-index facets could enable strong bonding with N2 and may boost NRR activity (Supplementary Fig. 34).
We used DFT calculations to further interpret the NRR activity differences of Pt3Fe (311), (111) and (200). The bonding and antibonding orbitals near the Fermi level (EF) exhibit an electron-rich distribution on the Pt3Fe (311) surface (Fig. 3a), while (111) and (200) surfaces present less electron-localizing on the surface (Fig. 3b and c). The preeminent d-electron exchange and transfer activities (d-EXTA) on the high indexed surface have been demonstrated through projected partial density of states (PDOSs) analysis (Fig. 3d–f). The edge of dominant peak of Pt-5d band reflects a direct determination of 5d-EXTA for efficient N2 fixation. Taking this trend, on the (311) surface, the Fe-3d orbital clearly merges at the EF without an evident gap between the eg and t2g components. The energetic interval between Pt-5d and Fe-3d is nearly 0.95 eV, which cost less energetic barrier for transferring d-electrons above EF to N2 2p orbital (Fig. 3d). The (111) and (200) present nearly two-times higher to freely cross EF (Fig. 3e and f). The comparison of energetic trends reveals that the NRR pathway on the (311) is the most energetically beneficial with an energetic barrier of merely 0.16 eV and overall gain of −4.86 eV. However, the (111) and (200) denote a higher energy gain of about −2.0 eV and the energy barriers are 0.75 and 0.34 eV, respectively (Fig. 3g). In the distal reaction pathway, the N atom furthest away from the catalyst surface preferentially undergoes hydrogenation, to form the first NH3 molecule. The NH3 molecule is released after the N≇N triple bond breaks. The remaining N atom continues under hydrogenation to generate another NH3 molecule. According to the above results, the reaction path of our work could be a distal reaction pathway. We further reason that the better activity of NRR on the (311) is attributed to the excellent energetic preference of N2 fixation, while the underperformance of (200) for NRR is ascribed to overbinding of adsorbing H for efficient N-hydrogenation (Fig. 3h). The energetic trend is consistent with analysis of electronic activities from both experimental and theoretical perspectives.
Figure 3.
(a–c) The real spatial contour plots of bonding and anti-bonding orbitals near EF on Pt3Fe (311), (111) and (200) surfaces, respectively. (d–f) PDOSs of d-bands for surface Pt-5d and Fe-3d sites on the (311), (111) and (200) are given, respectively. (g) The NRR energetic pathway on the surfaces of (311), (111) and (200), respectively. (h) Direct adsorption comparison for N2+6H and N2 on the surfaces of (311), (111) and (200), respectively.
A chronoamperometric test was first conducted at −0.05 V in N2-saturated KOH electrolyte to evaluate the durability of Pt3Fe NWs/C. The current density (Fig. 4a) exhibited negligible decay after 30 h electrolysis. The stability of Pt3Fe NWs/C was also evaluated by successive cycle electrolysis at −0.05 V. After five successive cycles, the total current density exhibited no evident fluctuation (Supplementary Fig. 35). The NH3 yield and FE of Pt3Fe NWs/C were measured after each cycle, with no obvious changes (Fig. 4b). After stability testing, TEM image, elemental mapping (Supplementary Fig. 36) and SEM-EDS (Supplementary Fig. 37) of the Pt3Fe NWs/C confirmed that its structure and composition were largely maintained, demonstrating that the Pt3Fe NWs/C was stable enough for NRR.
Figure 4.

(a) I-t curve of the Pt3Fe NWs/C in N2-saturated KOH electrolyte at −0.05 V and (b) NH3 yield and FE calculated after each cycle at −0.05 V.
CONCLUSION
In summary, we have demonstrated facile synthesis of Pt3Fe nanocrystals with tunable morphologies (NC, NR and NW) to evaluate the NRR performance on different exposed facets. The detailed studies show that the Pt3Fe nanocrystals exhibit shape-dependent electrocatalytic activity towards NRR. Notably, the Pt3Fe NWs bounded with high-index facets exhibit much improvement in NH3 yield (18.3 μg h−1 mg−1cat, 0.52 μg h−1 cm−2ECSA), FE (7.3%) as well as selectivity (no N2H4 is detected) at −0.05 V under ambient conditions, which are much better than that those of {200} facet-enclosed Pt3Fe NCs and {111} facet-enclosed Pt3Fe NRs. The Pt3Fe NWs also show durable electrochemical stability with no obvious activity decay after five successive electrolysis cycles. DFT calculation reveals that strong d-d coupling between Pt and Fe sites bridges the electron transfer for prominent NRR. This work provides the first example of the fundamental correlation between exposed surfaces and NRR performances of distinct nanocrystals.
METHODS
Preparation of Pt3Fe NWs
In preparation of monodisperse Pt3Fe NWs, K2PtCl4 (10.4 mg), Fe2(CO)9 (4.6 mg), ribose (45 mg), CTAC (32 mg) and OAm (5 mL) were added to a reaction bottle (volume: 35 mL), which was capped and ultrasonicated for 1 h. The reaction bottle was heated from room temperature to 180°C within 0.5 h and maintained at 180°C for 5 h in an oil bath. After cooling to room temperature, the obtained products were collected by centrifugation and washed three times with a cyclohexane/ethanol mixture.
Preparation of Pt3Fe NRs
The synthesis of monodisperse Pt3Fe NRs was similar to that of Pt3Fe NWs, except Pt(acac)2 (9.8 mg) was used as precursor and the amount of ribose was 90 mg at the beginning.
Preparation of Pt3Fe NCs
The synthesis of monodisperse Pt3Fe NCs was similar to that of Pt3Fe NWs, except maltose (108 mg) was used as reductant at the beginning.
Preparation of Pt NCs
In preparation of monodisperse Pt NCs, Pt(acac)2 (10 mg), PVP (200 mg), formaldehyde solution (40%, 2.5 mL) and benzyl alcohol (10 mL) were added to a reaction bottle. After sonication for 0.5 h, the homogeneous solution was transferred to a 20 mL Teflon-lined stainless autoclave and then heated at 150°C for 10 h. The product was collected via centrifugation and further washed with an ethanol-acetone mixture.
Preparation of Fe3O4 NPs
In preparation of monodisperse Fe3O4 NPs, Fe2(CO)9 (9.2 mg), ribose (45 mg), OAm (4.8 mL) and OAC (0.2 mL) were added to a reaction bottle, which was capped and then ultrasonicated for 1 h. The reaction bottle was heated from room temperature to 200°C within 0.5 h and maintained at 200°C for 5 h in an oil bath. After cooling to room temperature, the obtained products were collected by centrifugation and washed three times with a cyclohexane/ethanol mixture.
Catalytic measurements
Firstly, different morphology of Pt3Fe nanocrystals, Fe3O4 NPs, commercial carbon, cyclohexane (1 mL) and ethanol (8 mL) were added to a reaction bottle. The Pt NCs, commercial carbon, ethanol (1 mL) and acetone (8 mL) were also added to a reaction bottle. After sonication for 1 h, the homogeneous solution was centrifuged and dried naturally to obtain a powder. Then, the powder was annealed at 150°C for 1 h under ambient atmosphere. To prepare catalysts, the above powder (5 mg), Nafion solution (6 μL, 5 wt%) and absolute isopropyl alcohol (500 μL) were mixed and sonicated for 0.5 h to form homogeneous ink, which was then dropped (10 μL) evenly on carbon paper with geometric area of 1 × 1 cm2. The carbon paper was dried under ambient conditions.
The electrochemical tests were performed in a gas-tight two-chamber electrolytic cell separated by Nafion 115 membrane. Before electrochemical NRR measurements, the Nafion 115 membrane was first pretreated in 80°C H2O2 (5%) aqueous solution for 1 h and then washed in 80°C ultrapure water for another 1 h. The electrochemical experiments were conducted on a CHI660E electrochemical analyzer (CHI Instruments) using a three-electrode configuration (working electrode of synthetic catalysts, reference electrode of saturated calomel electrode and counter electrode of carbon rod). The mentioned potentials have been converted to RHE. The cyclic voltammetry (CV) tests were carried out in 0.1 M HClO4 solution with a scan rate of 50 mV s−1 under ambient conditions. The ECSA was obtained by integrating the hydrogen adsorption charge (QH) between −0.25 V and 0.1 V on the CV curves. The value of adsorbed single-layer hydrogen (qH) on the Pt surface is 210 μC cm−2 and the formula is ECSA = QH/(qH × m) (Supplementary Fig. 38).
Potentiostatic tests were carried out in electrochemical NRR. Before the measurement, highly pure N2 gas was flowed continuously into the cathode electrolytic cell with a proper position for 0.5 h.
Calculation of NH3 yield, Faraday efficiency and apparent energy
In NRR tests, FE is defined as the amount of charge for the synthesis of NH3 divided by the total charge through the electrodes during the electrolysis process. The NH3 yield was determined by a colorimetric method using Nessler's reagent. The formation of each NH3 molecule requires three electrons, hence the FE of NH3 can be determined by the below formula:
![]() |
The NH3 yield was determined by the below formula:
![]() |
where
: quantity of electric charge;
Faraday constant, 96485 C mol−1;
the KOH electrolyte volume;
the calculated NH3 concentration; t: the reduction time;
: the ECSA of the catalysts.
According to the Arrhenius equation, the apparent energy calculations process is below:
![]() |
where K: NH3 yield rate at temperature T; R: Molar gas constant (J mol−1 K−1); Ea: Apparent energy (kJ mol−1), T: Absolute reaction temperature (K). We plot lnK against 1000/T, and the slope is k, so Ea = -kR.
Determination of ammonia yield
The yield of NH3 was determined via a colorimetric method using Nessler's reagent. The calibration curve was obtained as follows: first, a series of reference solutions was prepared, by pipetting known NH4Cl solutions and 0.1 M KOH working electrolyte into colorimetric tubes. These were made up to the mark (10 mL) with 0.1 M KOH solution. Next, 1 mL of 0.2 M potassium sodium tartrate (KNaC4H4O6, chelating soluble metal ion) ultrapure water solution was pipetted into each of the tubes and these were mixed thoroughly, followed by pipetting of 1 mL Nessler's reagent into each of the tubes and further mixing. The mixed solutions were left for 0.5 h. Using a blank solution for background correction, the absorbance of the solutions was measured at 425 nm in a 10 mm glass cuvette. The calibration curve (y = 0.162x-0.005, R2 = 0.998) exhibited a good linear relationship between absorbance and NH4+ concentration according to three independent calibration tests.
Determination of hydrazine hydrate
The hydrazine hydrate was determined by a colorimetric method using Watt-Chrisp reagent. Para (dimethylamino) benzaldehyde (5.99 g), concentrated HCl (30 mL) and absolute ethanol (300 mL) were mixed as color reagents. A calibration curve was obtained as follows: first, different reference solutions were prepared, by pipetting hydrazine hydrate-nitrogen 0.1 M HCl solution into colorimetric tubes. These were made up to 5 mL with diluted hydrochloric acid electrolyte (pH = 1), then 5 mL was pipetted above the color reagent and this was left to stand for 0.5 h for color development under ambient conditions. The absorbance of the color solution was measured at 455 nm with a 10 mm glass cuvette, and the yields of hydrazine were determined by the standard calibration curve using a mixture of 5 mL remaining solution and 5 mL color reagent. The calibration curve (y = 1.208x-0.088, R2 = 0.977) was obtained with hydrazine monohydrate solutions of different known concentrations, and exhibited a good linear relationship between absorbance and N2H4·H2O concentration according to three independent calibration tests.
Determination of NOx
The concentration of NOx was measured with a colorimetric method using N-(-1-naphthyl) ethylenediamine dihydrochloride as color reagent. A mixture of 0.5 g sulfanilic acid, 5 mg n-(1-naphthyl)-ethylenediamine dihydrochloride, 90 mL H2O and 5 mL acetic acid was stirred to form a homogeneous solution. This was transferred to a 100 mL volumetric flask to obtain the color agent. A mixture of 1 mL electrolyte and 4 mL color agent was left for 0.5 h in the dark. The absorbance of the solutions was measured at 540 nm from the UV-vis absorption spectrum. The calibration curve was obtained by using different known concentrations of potassium nitrite solution in 0.1 M KOH.
15. 15N isotope labeling experiment
An 15N isotopic labeling experiment was carried out to verify the source of produced ammonia. After 15N2 electroreduction at −0.05 V in KOH electrolyte for 10 h, the obtained product was qualitatively determined by 1H nuclear magnetic resonance (NMR, Agilent 600 MHz). In detail, 30 mL of the electrolytic solution was moved out and then acidized to pH ∼ 3. The solution was concentrated to 2 mL at 80°C. Afterwards, 0.9 mL of concentrated solution and 0.1 mL D2O containing 100 ppm dimethyl sulphoxide (99.99%) as an internal standard were mixed for 1H NMR test.
DFT calculations
The DFT+U calculations were carried out with CASTEP code [45]. In this framework, we use rotationally invariant (Anisimov type) functional DFT+U [46]. The Hubbard U parameter is self-consistently determined for the pseudized C-2p, Fe-3d and Pt-5d orbital by the new linear response way [47–53]. The geometry optimization used the Broyden-Fletcher-Goldfarb-Shannon (BFGS) algorithm in all DFT+U calculations. The PBE functional was chosen for PBE+U calculations with a kinetic cutoff energy of 750 eV, with the valence electron states expressed in a plane-wave basis set. The ensemble DFT (EDFT) method of Marzari et al. [54] was used for convergence on the transition metal contained compounds.
The 2 × 2 × 1 supercell of Pt3Fe (311) surface model was chosen with 112 atoms (i.e. Pt84Fe28) and seven layers thick. The 2 × 2 × 1 supercell of Pt3Fe (111) surface model was chosen with 128 atoms (i.e. Pt96Fe32) and eight layers thick. The Pt3Fe (200) surface model was chosen with 144 atoms (i.e. Pt108Fe36) and eight layers thick. The vacuum thickness was set to be 10 Å. The reciprocal space integration was performed using the mesh of 2 × 2 × 1 [55] with Gamma-center-off, which was self-consistently selected for total energy minimization. With these special k-points, the total energy converged to less than 5.0 × 10−7 eV per atom. The Hellmann-Feynman forces on the atom converged to less than 0.001 eV/Å.
The Fe and Pt norm-conserving pseudopotentials were generated using the OPIUM code in the Kleinman-Bylander projector form [56], the non-linear partial core correction [57] and a scalar relativistic averaging scheme [58] are used to treat the spin-orbital coupling effect. For this treatment, we actually similarly chose a non-linear core correction technique for correcting the valence-core charge density overlapping in such heavy fermions elements, the details of this method are given in a previous work [47,48]. In particular, we treated the (3d, 4s, 4p) and (5d, 6s, 6p) states as the valence states of both Fe and Pt atoms. The RRKJ method was chosen for optimization of the pseudopotentials [59].
For all of the electronic state calculations, we used self-consistent determination for the U correction on the localized d orbitals to correct the on-site Coulomb energy of the electron spurious self-energy. By that method, the Hubbard U parameters on the half-filled shell of 3d6 orbitals of Fe are self-consistently determined to be Ud = 6.03 eV, and Ud = 5.26 eV for Pt-5d9 within the Pt3Fe system.
Supplementary Material
Acknowledgements
The characterizations were made at the Testing & Analysis Center in Soochow University.
Contributor Information
Wu Tong, College of Chemistry, Chemical Engineering and Materials Science, Soochow University, Suzhou 215123, China.
Bolong Huang, Department of Applied Biology and Chemical Technology, Hong Kong Polytechnic University, Hong Kong, China.
Pengtang Wang, College of Chemistry, Chemical Engineering and Materials Science, Soochow University, Suzhou 215123, China.
Qi Shao, College of Chemistry, Chemical Engineering and Materials Science, Soochow University, Suzhou 215123, China.
Xiaoqing Huang, College of Chemistry, Chemical Engineering and Materials Science, Soochow University, Suzhou 215123, China.
FUNDING
This work was supported by the Ministry of Science and Technology (2016YFA0204100 and 2017YFA0208200), the National Natural Science Foundation of China (21571135), Young Thousand Talented Program, Jiangsu Province Natural Science Fund for Distinguished Young Scholars (BK20170003), the Priority Academic Program Development of Jiangsu Higher Education Institutions, and the start-up funding from Soochow University.
AUTHOR CONTRIBUTIONS
H.-X.Q. and S.Q. proposed and supervised the project. T.W. and W.-P.T. conceived and designed the experiments. T.W. carried out the material synthesis, characterizations and electrochemical tests. H.-B.L. carried out DFT calculations. T.W., W.-P.T., H.-B.L., S.Q. and H.-X.Q. co-wrote the manuscript. All authors participated in discussing and analyzing the experimental results.
Conflict of interest statement. None declared.
REFERENCES
- 1. Chen JG, Crooks RM, Schrock RRet al. Beyond fossil fuel-driven nitrogen transformations. Science 2018; 360: eaar6611. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2. Jia H, Du A, Zhang CYet al. Site-selective growth of crystalline ceria with oxygen vacancies on gold nanocrystals for near-infrared nitrogen photofixation. J Am Chem Soc 2019; 141: 5083–6. [DOI] [PubMed] [Google Scholar]
- 3. Liu S, Wang M, Yan CLet al. Facilitating nitrogen accessibility to boron-rich covalent organic frameworks via electrochemical excitation for efficient nitrogen fixation. Nat Commun 2019; 10: 3898. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4. Li L, Tang C, Qiao SZet al. Two-dimensional mosaic bismuth nanosheets for highly selective ambient electrocatalytic nitrogen reduction. ACS Catal 2019; 9: 2902–8. [Google Scholar]
- 5. Yu X, Han P, Zheng GFet al. Boron-doped graphene for electrocatalytic N2 reduction. Joule 2018; 2: 1610–22. [Google Scholar]
- 6. Guo C, Ran J, Qiao SZet al. Rational design of electrocatalysts and photo(electro)catalysts for nitrogen reduction to ammonia (NH3) under ambient conditions. Energy Environ Sci 2018; 11: 45–56. [Google Scholar]
- 7. Hu C, Chen X, Xiong YJet al. Surface plasmon enabling nitrogen fixation in pure water through a dissociative mechanism under mild conditions. J Am Chem Soc 2019; 141: 7807–14. [DOI] [PubMed] [Google Scholar]
- 8. Qiu W, Xie X, Sun XPet al. High-performance artificial nitrogen fixation at ambient conditions using a metal-free electrocatalyst. Nat Commun 2018; 9: 3485. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9. Zhang N, Jalil A, Xiong YJet al. Refining defect states in W18O49 by Mo doping: a strategy for tuning N2 activation towards solar-driven nitrogen fixation. J Am Chem Soc 2018; 140: 9434–43. [DOI] [PubMed] [Google Scholar]
- 10. Yao Y, Zhu S, Shao MHet al. A spectroscopic study on the nitrogen electrochemical reduction reaction on gold and platinum surfaces. J Am Chem Soc 2018; 140: 1496–501. [DOI] [PubMed] [Google Scholar]
- 11. Tang C, Qiao SZ. How to explore ambient electrocatalytic nitrogen reduction reliably and insightfully. Chem Soc Rev 2019; 48: 3166–80. [DOI] [PubMed] [Google Scholar]
- 12. Cheng H, Ding LX, Chen GFet al. Molybdenum carbide nanodots enable efficient electrocatalytic nitrogen fixation under ambient conditions. Adv Mater 2018; 30: 1803694. [DOI] [PubMed] [Google Scholar]
- 13. Cheng H, Cui P, Wang HHet al. High efficiency electrochemical nitrogen fixation achieved with a lower pressure reaction system by changing the chemical equilibrium. Angew Chem Int Ed 2019; 58: 15541–7. [DOI] [PubMed] [Google Scholar]
- 14. Luo Y, Chen G-F, Ding Let al. Efficient electrocatalytic N2 fixation with MXene under ambient conditions. Joule 2019; 3: 279–89. [Google Scholar]
- 15. Hui L, Xue Y, Li YLet al. Highly efficient and selective generation of ammonia and hydrogen on a graphdiyne-based catalyst. J Am Chem Soc 2019; 141: 10677–83. [DOI] [PubMed] [Google Scholar]
- 16. Geng Z, Liu Y, Zeng Jet al. Achieving a record high yield rate of 120.9 μg h−1 mg−1cat for N2 electrochemical reduction over Ru single-atom catalysts. Adv Mater 2018; 30: 1803498. [DOI] [PubMed] [Google Scholar]
- 17. Hong W, Lu W, Qiang Wet al. Ambient electrosynthesis of ammonia: electrode porosity and composition engineering. Angew Chem Int Ed 2018; 57: 12360–4. [DOI] [PubMed] [Google Scholar]
- 18. Zhang L, Ji X, Sun XPet al. Electrochemical ammonia synthesis via nitrogen reduction reaction on a MoS2 catalyst: theoretical and experimental studies. Adv Mater 2018; 30: 1800191. [DOI] [PubMed] [Google Scholar]
- 19. Tao H, Choi C, Sun Zet al. Nitrogen fixation by Ru single-atom electrocatalytic reduction. Chem 2018; 5: 204–14. [Google Scholar]
- 20. Wang M, Liu S, Yan CLet al. Over 56.55% Faradaic efficiency of ambient ammonia synthesis enabled by positively shifting the reaction potential. Nat Commun 2019; 10: 341. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21. Shi M, Bao D, Jiang Qet al. Anchoring PdCu amorphous nanocluster on graphene for electrochemical reduction of N2 to NH3 under ambient conditions in aqueous solution. Adv Energy Mater 2018; 8: 1800124. [Google Scholar]
- 22. Wang H, Li Y, Wang Let al. One-pot synthesis of bi-metallic PdRu tripods as an efficient catalyst for electrocatalytic nitrogen reduction to ammonia. J Mater Chem A 2019; 7: 801–5. [Google Scholar]
- 23. Shi M, Bao D, Jiang Qet al. Au sub-nanoclusters on TiO2 toward highly efficient and selective electrocatalyst for N2 conversion to NH3 at ambient conditions. Adv Mater 2017; 29: 1606550. [DOI] [PubMed] [Google Scholar]
- 24. Bao D, Zhang Q, Zhang XBet al. Electrochemical reduction of N2 under ambient conditions for artificial N2 fixation and renewable energy storage using N2/NH3 cycle. Adv Mater 2017; 29: 1604799. [DOI] [PubMed] [Google Scholar]
- 25. Chen G, Cao X, Wang Het al. Ammonia electrosynthesis with high selectivity under ambient conditions via a Li+ incorporation strategy. J Am Chem Soc 2017; 139: 9771–4. [DOI] [PubMed] [Google Scholar]
- 26. Lv C, Qian Y, Yu GHet al. Defect engineering metal-free polymeric carbon nitride electrocatalyst for effective nitrogen fixation under ambient conditions. Angew Chem Int Ed 2018; 57: 10246–50. [DOI] [PubMed] [Google Scholar]
- 27. Lv C, Yan C, Yu GHet al. An amorphous noble-metal-free electrocatalyst enables N2 fixation under ambient conditions. Angew Chem Int Ed 2018; 130: 6181–4. [DOI] [PubMed] [Google Scholar]
- 28. Yao Y, Wang H, Shao MHet al. Electrochemical nitrogen reduction reaction on ruthenium. ACS Energy Lett. 2019; 4: 1336–41. [Google Scholar]
- 29. Skúlason E, Bligaard T, Nørskov JKet al. A theoretical evaluation of possible transition metal electro-catalysts for N2 reduction. Phys Chem Phys 2012; 14: 1235–45. [DOI] [PubMed] [Google Scholar]
- 30. Liu X, Jiao Y, Qiao SZet al. Building up a picture of the electrocatalytic nitrogen reduction activity of transition metal single-atom catalysts. J Am Chem Soc 2019; 141: 9664–72. [DOI] [PubMed] [Google Scholar]
- 31. Bai S, Bu L, Huang XQet al. Multicomponent Pt-based zigzag nanowires as selectivity controllers for selective hydrogenation reactions. J Am Chem Soc 2018; 140: 8384–7. [DOI] [PubMed] [Google Scholar]
- 32. Luo M, Sun Y, Guo SJet al. Stable high-index faceted Pt skin on zigzag-like PtFe nanowires enhances oxygen reduction catalysis. Adv Mater 2018; 30: 1705515. [DOI] [PubMed] [Google Scholar]
- 33. Tian N, Zhou ZY, Wang ZLet al. Synthesis of tetrahexahedral platinum nanocrystals with high-index facets and high electro-oxidation activity. Science 2007; 316: 732–5. [DOI] [PubMed] [Google Scholar]
- 34. Zhang L, Zhang JW, Zheng LSet al. Cu2+-assisted synthesis of hexoctahedral Au-Pd alloy nanocrystals with high-index facets. J Am Chem Soc 2011; 133: 17114–7. [DOI] [PubMed] [Google Scholar]
- 35. Bu LZ, Shao Q, Huang XQet al. PtPb/PtNi intermetallic core/atomic layer shell octahedra for efficient oxygen reduction electrocatalysis. J Am Chem Soc 2017; 139: 9576–82. [DOI] [PubMed] [Google Scholar]
- 36. Wang Y, Cui X, Zheng Wet al. Rational design of Fe-N/C hybrid for enhanced nitrogen reduction electrocatalysis under ambient conditions in aqueous solution. ACS Catal 2018; 9: 336–44. [Google Scholar]
- 37. Wang J, Yu L, Feng Xet al. Ambient ammonia synthesis via palladium-catalyzed electro-hydrogenation of dinitrogen at low overpotential. Nat Commun 2018; 9: 1795. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38. Zhang L, Ding LX, Wang Het al. Ammonia synthesis under ambient conditions: selective electroreduction of dinitrogen to ammonia on black phosphorus nanosheets. Angew Chem Int Ed 2019; 131: 2638–42. [DOI] [PubMed] [Google Scholar]
- 39. Liu Y, Su Y, Zhao Jet al. Facile ammonia synthesis from electrocatalytic N2 reduction under ambient conditions on N-doped porous carbon. ACS Catal 2018; 8: 1186–91. [Google Scholar]
- 40. Zhang X, Luo Z, Zhang Het al. Lithiation-induced amorphization of Pd3P2S8 for highly efficient hydrogen evolution. Nat Catal 2018; 1: 460–8. [Google Scholar]
- 41. Lai J, Huang B, Guo SJet al. Strongly coupled nickel–cobalt nitrides/carbon hybrid nanocages with Pt-like activity for hydrogen evolution catalysis. Adv Mater 2019; 31: 1805541. [DOI] [PubMed] [Google Scholar]
- 42. Kim D, Resasco J, Yang PDet al. Synergistic geometric and electronic effects for electrochemical reduction of carbon dioxide using gold-copper bimetallic nanoparticles. Nat Commun 2014; 5: 4948. [DOI] [PubMed] [Google Scholar]
- 43. Duchesne PN, Li ZY, Zhang Pet al. Golden single-atomic-site platinum electrocatalysts. Nat Mater 2018; 17: 1033–9. [DOI] [PubMed] [Google Scholar]
- 44. Huang H, Jia H, Zeng Jet al. Understanding of strain effects in the electrochemical reduction of CO2: using Pd nanostructures as an ideal platform. Angew Chem Int Ed 2017; 129: 3648–52. [DOI] [PubMed] [Google Scholar]
- 45. Clark SJ, Segall MD, Payne MCet al. First principles methods using CASTEP. Zeitschrift Fur Kristallographie 2005; 220: 567. [Google Scholar]
- 46. Vladimir IA, Lichtenstein AI.. First-principles calculations of the electronic structure and spectra of strongly correlated systems: the LDA+ U method. J Phys Condens Matter 1997; 9: 767–807. [Google Scholar]
- 47. Huang B, Gillen R, Robertson J. Study of CeO2 and its native defects by density functional theory with repulsive potential. J Phys Chem C 2014; 118: 24248–56. [Google Scholar]
- 48. Huang B. Superiority of DFT+U with non-linear core correction for open-shell binary rare-earth metal oxides: a case study of native point defects in cerium oxides. Philo Mag 2014; 94: 3052–71. [Google Scholar]
- 49. Huang B. 4f fine-structure levels as the dominant error in the electronic structures of binary lanthanide oxides. J Comput Chem 2016; 37: 825–35. [DOI] [PubMed] [Google Scholar]
- 50. Huang B. Intrinsic deep hole trap levels in Cu2O with self-consistent repulsive Coulomb energy. Solid State Commun 2016; 230: 49–53. [Google Scholar]
- 51. Huang B. Unraveling energy conversion modeling in the intrinsic persistent upconverted luminescence of solids: a study of native point defects in antiferromagnetic Er2O3. Phys Chem Phys 2016; 18: 13564–82. [DOI] [PubMed] [Google Scholar]
- 52. Huang B. The screened pseudo-charge repulsive potential in perturbed orbitals for band calculations by DFT+U. Phys Chem Phys 2017; 19: 8008–25. [DOI] [PubMed] [Google Scholar]
- 53. Hu J, Huang B, Yang Set al. Engineering stepped edge surface structures of MoS2 sheet stacks to accelerate the hydrogen evolution reaction. Energy Environ Sci 2017; 10: 593–603. [Google Scholar]
- 54. Marzari N, Vanderbilt D, Payne MC. Ensemble density-functional theory for Ab initio molecular dynamics of metals and finite-temperature insulators. Phys Rev Lett 1997; 79: 1337–41. [Google Scholar]
- 55. Probert MIJ, Payne MC. Improving the convergence of defect calculations in supercells: an ab initio study of the neutral silicon vacancy. Phys Rev B 2003; 67: 075204–15. [Google Scholar]
- 56. Kleinman L, Bylander DM. Efficacious form for model pseudopotentials. Phys Rev Lett 1982; 48: 1425–8. [Google Scholar]
- 57. Louie SG, Cohen ML. Nonlinear ionic pseudopotentials in spin-density-functional calculations. Phys Rev B 1982; 26: 1738–42. [Google Scholar]
- 58. Grinberg I, Rappe AM. Transferable relativistic Dirac-Slater pseudopotentials. Phys Rev B 2000; 62: 2311–5. [Google Scholar]
- 59. Rappe AM, Joannopoulos JD. Optimized pseudopotentials. Phys Rev B 1990; 41: 1227–31. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.




