Efficient co-production of ammonia and formic acid from nitrate and polyester via paired electrolysis

Mengmeng Du a, Tao Sun b, Xuyun Guo c, Mingzhu Han a, Yu Zhang *ad, Wenxuan Chen a, Mengxiang Han a, Jizhe Ma a, Wenfang Yuan a, Chunyu Zhou b, Valeria Nicolosi c, Jian Shang *e, Ning Zhang *fg and Bocheng Qiu *a
aJiangsu Key Laboratory of Pesticide Sciences, Department of Chemistry, College of Sciences, Nanjing Agricultural University, Nanjing, 210095, China
bDepartment of Radiology, Xinqiao Hospital, Army Medical University, Chongqing, 400037, China
cSchool of Chemistry, Centre for Research on Adaptive Nanostructures and Nanodevices (CRANN) and Advanced Materials Bio-Engineering Research Centre (AMBER), Trinity College Dublin, Dublin, D02PN40, Ireland
dCollege of Energy and Power Engineering, Nanjing Institute of Technology, Nanjing, 211167, China. E-mail: yu_zhang1225@163.com
eLow-Dimensional Energy Materials Research Center, Shenzhen Institute of Advanced Technology, Chinese Academy of Sciences, Shenzhen, 518055, China. E-mail: jian.shang@siat.ac.cn
fKey Laboratory of Precision and Intelligent Chemistry, Department of Environmental Science and Engineering, University of Science and Technology of China, Hefei, 230026, China. E-mail: zhangning18@ustc.edu.cn
gSustainable Energy and Environmental Materials Innovation Center, Suzhou Institute for Advanced Research, University of Science and Technology of China, Suzhou, 215123, China

Received 22nd January 2025 , Accepted 6th March 2025

First published on 7th March 2025


Abstract

Paired electrolysis, which integrates a productive cathodic reaction, such as the nitrate reduction reaction (NO3RR) with selective oxidation at the anode, offers an intriguing way to maximize both atomic and energy efficiency. However, in a conventional design, the NO3RR is often coupled with the anodic oxygen evolution reaction, leading to substantial energy consumption while yielding low-value oxygen. Here, we report a hybrid electrolysis system that combines cathodic reduction of nitrate to ammonia and anodic oxidation of polyethylene-terephthalate-derived ethylene glycol (EG) to formic acid (FA), utilizing oxygen-vacancy-rich (OV) Co3O4 and Cu doped Ni(OH)2 as the cathode and anode, respectively. Remarkably, this paired electrolysis system demonstrates a faradaic efficiency (FE) of 92% for cathodic ammonia production and a FE of 99% for anodic FA production, while reducing the cell voltage by 0.54 V compared to the conventional NO3RR system at the same current density of 100 mA cm−2. Experimental investigations combined with theoretical calculations reveal that the OV introduction effectively addresses the insufficient NO3 adsorption and hydrogenation on bare Co3O4. Additionally, Cu incorporation increases the Ni–O covalency, resulting in an improved EG adsorption ability. This work presents a promising way for waste management via paired electrolysis.



New concepts

Electrochemical ammonia (NH3) production from nitrate-containing wastewater streams presents a sustainable route to concurrent chemical synthesis and waste management, which is thus recognized as an alternative to the conventional, energy-intensive Haber–Bosch process. However, the nitrate (NO3) reduction reaction (NO3RR) typically pairs with the oxygen evolution reaction (OER) at the anode that suffers from sluggish kinetics and high energy input while yielding low-value oxygen. To tackle this issue, we implement paired electrolysis that couples the cathodic NO3RR with anodic oxidation of plastic-derived ethylene glycol. This integrated system enables simultaneous valorization of nitrate wastewater and upcycling of plastic waste within a single electrolyzer, thereby maximizing the utilization of waste resource. This study offers an environmentally benign and energy-efficient pathway for transforming waste streams into valuable chemicals.

Introduction

Ammonia (NH3) plays a crucial role in the production of fertilizers and is recognized as a critical precursor for the synthesis of a plethora of nitrogen-containing chemicals, including urea and amino acids.1–3 The industrial-scale NH3 synthesis is currently being implemented via the Haber–Bosch process, which, however, features energy-intensive nature and high carbon emission.4–6 The electrochemical nitrate (NO3) reduction reaction (NO3RR) driven by renewable electricity offers a cost-effective and low-carbon way for sustainable NH3 production.7–10 The NO3RR for NH3 production involves a sluggish eight-electron transfer kinetics, which necessitates catalysts endowed with abundant active sites.11–14 Moreover, the byproducts (e.g., NO2, N2, and NH2OH) evolved in the NO3RR and the competitive hydrogen evolution reaction (HER) significantly impede the selectivity and current efficiency of the NO3RR.15–17 Thus far, developing high-performance electrocatalysts for the NO3RR remained a huge challenge, primarily due to the limited understanding of the catalytic mechanism and the structure–activity relationship.

Spinel oxides, characterized by their tetrahedral and octahedral coordination centers, such as Co3O4, have garnered substantial interest in the NO3RR, potentially owing to their tunable electronic structure and rich redox chemistry.18–20 Furthermore, the octahedral Co3+ sites with empty d orbitals exhibit a higher affinity for atomic hydrogen adsorption (*H), which thereby serve as hydrogenation centers in converting NO2 to NH3.21,22 In spite of these aforementioned advantages, bare Co3O4 encounters several limitations in the NO3RR, particularly poor NO3 adsorption capability, which arises from the elevated energy of the lowest unoccupied molecular π* orbital (LUMO π*) of NO3.23,24 Defect engineering has been deemed as a powerful tool for optimizing the reaction intermediate adsorption on spinel structures.25–27 Notably, the incorporation of an oxygen vacancy (OV) into Co3O4 leads to an upshift of the d-band center of Co 3d orbitals. This contributes to charge injection into the π* orbitals of NO3, thus facilitating promoting NO3 adsorption and activation.

In this context, creation of an OV in Co3O4 can be expected to play a pivotal role in transforming NO3 into NO2.

However, further enhancement of the overall performance of the NO3RR still remains essential for its widespread application. Notably, a limiting factor increasing the energy input of the conventional system is the anodic oxygen evolution reaction (OER), which requires a high oxidative potential yet produces low-cost oxygen (O2).28–30 Electrochemical oxidation of plastic-derived feedstocks, benefitting from the low operation potential and the production of value-added chemicals, represents a promising alternative to the OER.31–34 Consequently, integrating the NO3RR and plastic upcycling into one electrolyzer is anticipated to yield valuable chemicals at minimal investment. Nevertheless, the operational stability and economic feasibility of this paired system still face challenges.

Herein, we designed a paired electrochemical system that couples the cathodic NO3RR for NH3 production with anodic polyethylene-terephthalate-derived (PET-derived) ethylene glycol (EG) oxidation for formic acid (FA) evolution. Through constructing an oxygen vacancy Co3O4 (OV–Co3O4) cathode and a Cu doped Ni(OH)2 (Cu–Ni(OH)2) anode, we achieved a FE of NH3 (92%) at the cathode and a FE of FA (99%) at the anode at 100 mA cm−2. Based on experimental observations and theoretical insights into the improved ability for NO3 adsorption and activation on OV–Co3O4, as well as the enhanced EG binding on Cu–Ni(OH)2, we propose a continuous co-production of NH4Cl and formate from liter-scale nitrate wastewater and hundred-gram-scale PET bottles. Techno-economic analysis (TEA) suggests that this paired electrolysis system remains profitable, thus attracting industrial investment.

Results and discussion

Cathode catalyst characterization studies

The preparation procedure of the OV–Co3O4 cathode initiates with the growth of Co(OH)F nanorods on Ni foam (Co(OH)F/NF) through a hydrothermal method (Fig. 1a). Scanning electron microscopy (SEM) and X-ray diffraction (XRD) patterns clearly show the uniform growth of Co(OH)F nanorods throughout the entire NF substrate (Fig. S1 and S2, ESI). The following calcination treatment in air successfully achieves the conversion of Co(OH)F/NF into Co3O4/NF, as evidenced by the SEM and transmission electron microscopy (TEM) observations (Fig. S3 and S4, ESI). The OV is subsequently introduced into Co3O4via reduction treatment in a H2 atmosphere for 1 h. Unless otherwise specified, the OV–Co3O4/NF catalyst refers to the sample subjected to H2 reduction for 1 h. The crystal structure of OV–Co3O4 is elucidated by using XRD analysis, which reveals all the peaks observed corresponding to the standard spinel structure (Fig. 1b). Furthermore, SEM and TEM observations collectively exhibit the retention of the nanorod morphology in OV–Co3O4 (Fig. 1c and d). Selected area electron diffraction (SAED) patterns (Fig. 1d, inset) exhibit the exposure of the characteristic (220), (311), (400), and (511) planes of Co3O4. High-resolution TEM reveals a set of lattice fringes with a spacing of 0.24 nm (Fig. 1e), which is assigned to the (311) plane of Co3O4.35 High-angle annular dark field scanning TEM (HAADF-STEM) imaging and electron energy loss spectroscopy (EELS) demonstrate the uniform distribution of Co and O elements throughout the OV–Co3O4 nanorod (Fig. 1f). The OV was identified by electron spin resonance (ESR) spectra (Fig. 1g), in which a weak ESR signal at g = 2.002 is observed for Co3O4. In contrast, OV–Co3O4 exhibits a significantly enhanced OV signal, suggesting the successful introduction of OV. The OV introduction inevitably leads to a change in the valence state of Co atoms adjacent with OV. As illustrated in Co 2p X-ray photoelectron spectroscopy (XPS) (Fig. 1h), the proportion of Co2+ is substantially improved from 39% to 87% after OV introduction, which is attributed to the transformation of octahedral Co3+ to Co2+ upon OV incorporation.36 Furthermore, the local electrons at the OV sites have the potential to reduce the adsorbed water molecule to hydrogen and hydroxyl.37,38 This phenomenon accounts for the higher content of surface hydroxyls on OV–Co3O4/NF (79%) compared with bare Co3O4/NF (22%), as observed in the O 1s XPS spectra (Fig. 1i).
image file: d5mh00130g-f1.tif
Fig. 1 Morphology and structure characterization of the cathode catalyst. (a) Schematic illustration of the OV–Co3O4/NF preparation process. (b) XRD patterns of Co3O4 and OV–Co3O4 powders scraped from the Ni substrate. (c) SEM image, (d) TEM image (the inset shows the corresponding SAED patterns), (e) HRTEM image, (f) HAADF-STEM image and its corresponding EELS maps of OV–Co3O4. (g) ESR spectra of OV–Co3O4 and Co3O4 powders scraped from the NF substrate. (h) Co 2p and (i) O 1s XPS spectra of Co3O4/NF and OV–Co3O4/NF.

NO3RR activity and mechanism analysis

The electrochemical NO3RR activity of OV–Co3O4/NF was measured in a three-electrode H-type cell under ambient conditions. Colorimetric methods, utilizing a calibration curve (Fig. S5, ESI), were employed to quantitatively determine ammonia concentrations. All potentials are referenced to the reversible hydrogen electrode (RHE), unless otherwise specified. A slight enhancement in the HER activity can be observed for OV–Co3O4/NF relative to Co3O4/NF in the absence of NO3 (Fig. 2a). Upon NO3 addition, OV–Co3O4/NF enables a dramatically increased current density compared with Co3O4/NF, highlighting the key role of OV in the NO3RR. To evaluate the NO3RR performance more fairly, specific activities were calculated by normalizing linear sweep voltammetry (LSV) curves with the electrochemical active surface area (ECSA) (Fig. S6, ESI). The results suggest that the OV–Co3O4/NF still allows better NO3RR activity than Co3O4/NF. Additionally, OV–Co3O4/NF exhibits more favorable reaction kinetics and enhanced charge transfer than Co3O4/NF, as corroborated by the reduced Tafel slope and charge transfer resistance (Fig. S7 and S8, ESI).
image file: d5mh00130g-f2.tif
Fig. 2 NO3RR activity and enhanced mechanism. (a) Polarization curves of Co3O4/NF and OV–Co3O4/NF in 1 M KOH solution with (solid line) and without (dotted line) 0.1 M NO3. (b) The NO3RR activity comparison between Co3O4/NF and OV–Co3O4/NF. (c) 1H NMR spectra of post-reaction electrolytes using 15NO3 and 14NO3 feedstocks. (d) Cycling experiments of OV–Co3O4/NF for the NO3RR conducted at −0.3 V. (e) PDOS of the (311) planes of Co3O4 and OV–Co3O4. (f) The band center calculations of O 2p and Co 3d in Co3O4 (311) and OV–Co3O4 (311). (g) The Gibbs free energy for H adsorption over Co3O4 (311) and OV–Co3O4 (311). (h) In situ DEMS analysis of nitrogen-containing intermediates during the NO3RR over OV–Co3O4/NF. (i) In situ FTIR spectra of the NO3RR over OV–Co3O4/NF recorded under varying potentials. (j) Free energy diagram of each intermediate state within the NO3RR on Co3O4 (311) and OV–Co3O4 (311).

The FE and productivity of NH3 over OV–Co3O4/NF and Co3O4/NF were measured to reveal the role of OV. As presented in Fig. 2b, OV–Co3O4/NF exhibits a higher FE of NH3 compared with Co3O4/NF across a wide potential range from −0.1 to −0.45 V. Impressively, OV–Co3O4/NF achieves a maximum FE of 95.2% at −0.3 V, which outperforms most of recently reported NO3RR catalysts (Table S1, ESI). Additionally, OV–Co3O4/NF displays an NH3 productivity of 0.31 mmol cm−2 h−1 at −0.3 V, notably higher than that of bare Co3O4/NF (0.17 mmol cm−2 h−1). This enhanced performance is attributed to the promotion of NO3 adsorption and activation by OV, thus accelerating the NO3-to-NO2 conversion step and improving the NH3 yield rate.

To identify the impact of the OV concentration on the NO3RR, OV–Co3O4/NF with varying OV contents were synthesized by adjusting the reductive treatment time in a H2 atmosphere (Fig. S9 and S10, ESI). The results indicate that the OV–Co3O4/NF catalyst treated with H2 reduction for 1 h exhibits the highest FE and productivity of NH3. To exclude the environmental contaminants in NH3 detection, isotope-labeling tests were carried out (Fig. 2c). The results confirm that the produced NH3 originates from the NO3RR rather than the environmental contaminants. The electrochemical NO3RR durability of OV–Co3O4/NF was assessed in ten consecutive cycles (Fig. 2d). The productivity and FE of NH3 exhibited slight fluctuations in each cycle but remained stable. Strikingly, no obvious changes in morphology, crystal structure, and chemical state can be observed for the recovered OV–Co3O4/NF, as confirmed by a combination of SEM, TEM, XRD, and XPS characterization studies (Fig. S11–S13, ESI).

The role of OV in the NO3RR was further investigated via density functional theory (DFT) calculations (Fig. S14, ESI). Note that the distance between the Co 3d and O 2p band centers (denoted as DM) is recognized as a descriptor of the degree of Co–O covalency, with a smaller DM indicating an increased Co–O covalency.39,40 The partial density of state (PDOS) combined with the band center calcination reveals that OV–Co3O4 enables a reduced DM value (0.17 eV) compared with bare Co3O4 (0.19 eV), suggesting a higher overlap between Co 2p and O 1s in OV–Co3O4 (Fig. 2e and f). Such improved Co–O covalency contributes to the NO3 adsorption and activation.41 Moreover, the OV sites promote the generation of active adsorbed H (H*), as indicated by the free energy for H* evolution being closer to zero on OV–Co3O4 compared to Co3O4 (Fig. 2g). The enhanced activity for H* generation on OV–Co3O4 contributes to the hydrogenation of NO2 to NH3.42

The reaction pathway was identified using in situ differential electrochemical mass spectrometry (DEMS). There exist two potential pathways for the conversion of NO3 to NH3, as illustrated in Fig. S15 (ESI). The appearance of *NO signal, coupled with the absence of *NH2OH signal, implies that NH3 is most likely formed via the *NOH intermediate pathway (Fig. 2h). In situ Fourier transform infrared (FTIR) spectroscopy was employed to differentiate the reaction intermediates (Fig. 2i). The peaks at 1338 and 1665 cm−1 are ascribed to *NO2 and *NO intermediate species, respectively.43 Note that the absence of the characteristic NH2OH band in FTIR is consistent with the results of in situ DEMS. Based on the reaction pathway confirmed by in situ spectroscopy, we calculated the Gibbs free energies of reaction intermediates on OV–Co3O4 (311) and Co3O4 (311). As displayed in Fig. 2j, the energy barrier for NO3 adsorption on bare Co3O4 reaches up to 1.28 eV, which is recognized as the rate-determining step (RDS). The introduction of OV results in a significantly reduced energy barrier for NO3 adsorption, revealing the pivotal role of OV in the NO3-to-NO2 step. Furthermore, OV–Co3O4 enables improved hydrogenation ability in the NO2-to-NH3 step compared with Co3O4, as uncovered by the reduced energy barrier for hydrogenation of *NO to *NOH. This observation agrees with the results of the calculated energy for *H adsorption (Fig. 2g).

Anode catalyst characterization studies

The Cu doped Ni(OH)2 (Cu–Ni(OH)2) anode catalyst was synthesized by using a two-step approach (Fig. 3a), involving hydrothermal growth of Cu–Ni(OH)2·xH2O nanosheets on NF (Fig. S16, ESI), followed by annealing to obtain Cu–Ni(OH)2/NF. The crystal structure of the as-prepared Cu–Ni(OH)2 is indexed into the β-Ni(OH)2 phase (PDF#73-1520), as elucidated in XRD patterns (Fig. 3b). The molar ratio of Cu to Ni was determined to be 1[thin space (1/6-em)]:[thin space (1/6-em)]3 using inductively coupled plasma mass spectrometry (ICP-MS). Additionally, SEM and TEM images of Cu–Ni(OH)2/NF indicate that Cu–Ni(OH)2 exists as a nanosheet structure on the NF substrate (Fig. 3c and d), with no significant difference compared with bare Ni(OH)2/NF (Fig. S17, ESI). The lattice fringe of a spacing of 0.27 nm, observed in the HRTEM image (Fig. 3e), is attributed to the (100) plane of Ni(OH)2.44 The elemental mapping demonstrates that Cu, Ni, and O elements are uniformly dispersed on the nanosheet (Fig. 3f). The chemical states of Cu–Ni(OH)2/NF were investigated via XPS. The Cu 2p XPS spectrum reveals that the Cu dopant is present in Cu2+ (Fig. S18, ESI). In Ni 2p XPS spectra, a positive shift of 0.3 eV in the binding energy of Ni 2p can be observed for Cu–Ni(OH)2/NF relative to Ni(OH)2/NF (Fig. 3g), indicative of decreased charge density on the Ni site after Cu incorporation. This change in charge density results in a reduced DM value in Cu–Ni(OH)2, suggesting a greater Ni–O covalency and an enhanced ability for oxygen-containing adsorption (Fig. 3h and i).
image file: d5mh00130g-f3.tif
Fig. 3 Morphology and structure characterization of the anode catalyst. (a) Schematic illustration of Cu–Ni(OH)2/NF preparation. (b) XRD patterns of Ni(OH)2 and Cu–Ni(OH)2 powders scraped from the NF substrate. (c) SEM image, (d) TEM image (the inset shows the corresponding SAED patterns), (e) HRTEM image, (f) HAADF-STEM image and its corresponding EELS maps of Cu–Ni(OH)2. (g) Ni 2p XPS spectra of Ni(OH)2/NF and Cu–Ni(OH)2/NF. (h) PDOS of Ni(OH)2 (001) and Cu–Ni(OH)2/NF (001). (i) Band centers of Ni(OH)2 (001) and Cu–Ni(OH)2/NF (001).

EGOR activity and mechanism analysis

The EGOR performances of Cu–Ni(OH)2/NF and Ni(OH)2/NF were assessed in 1 M KOH containing 0.3 M EG at room temperature under atmospheric pressure. The reason for fixing the EG concentration at 0.3 M is elucidated in Fig. S19 (ESI). The LSV curves show that Cu–Ni(OH)2/NF requires a potential of 1.39 V to achieve a current density of 100 mA cm−2, which is 60 mV lower than that of bare Ni(OH)2/NF, indicating the enhanced EGOR activity due to Cu doping (Fig. 4a). Additionally, Cu–Ni(OH)2/NF exhibits a higher specific EGOR activity (normalized by ECSA) than Ni(OH)2/NF (Fig. S20, ESI). Furthermore, the current density observed in LSV curves dramatically increases for EGOR on both Ni(OH)2/NF and Cu–Ni(OH)2/NF compared with the OER, uncovering more thermodynamically favorable conditions for EGOR over the OER. To identify the impact of Cu doping concentration on EGOR activity, the Cu–Ni(OH)2/NF catalysts with different Cu doping levels were synthesized (Fig. S21, ESI). The results demonstrate that the Cu–Ni(OH)2/NF catalyst with the Cu/Ni molar ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]3 delivers optimal EGOR activity (Fig. S22, ESI). Unless otherwise noted, the Cu–Ni(OH)2/NF catalyst in this work refers to the sample with a Cu/Ni molar ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]3. In addition to the high EGOR activity observed from LSV curves, Cu–Ni(OH)2/NF allows faster reaction kinetics and reduced charge transfer resistance with respect to bare Ni(OH)2/NF, as evidenced by the Tafel slopes (Fig. 4b) and EIS spectra (Fig. S23, ESI).
image file: d5mh00130g-f4.tif
Fig. 4 EGOR activity and enhanced mechanism. (a) Polarization curves of Ni(OH)2/NF and Cu–Ni(OH)2/NF measured in 1 M KOH with and without 0.3 M EG. (b) Tafel slopes of Ni(OH)2/NF and Cu–Ni(OH)2/NF. (c) The FE and productivity of FA on Ni(OH)2/NF and Cu–Ni(OH)2/NF under various potentials. (d) Concentrations of FA, GA, EG, and FE of FA as a function of consumed charges. (e) Stability tests using chronopotentiometry measurements at the current density of 100 mA cm−2, and the electrolyte was refreshed every 5 h. (f) and (g) The computed pCOHP diagrams of Ni–Oads bond on Cu–Ni(OH)2 (f) and Ni(OH)2 (g). (h) The calculated energy for EG adsorption on Cu–Ni(OH)2 (001) and Ni(OH)2 (001). (i) and (j) Bode phase plots of the operando EIS on Cu–Ni(OH)2/NF in 1 M KOH (i) and 1 M KOH + 0.3 EG (j). (k) and (l) In situ Raman tests on Cu–Ni(OH)2 under varying potentials in 1 M KOH (k) and 1 M KOH + 0.3 EG (l).

The EGOR products during 2-h potentiostatic tests on Cu–Ni(OH)2/NF and Ni(OH)2/NF were identified using HPLC (Fig. S24, ESI). Compared with Ni(OH)2/NF, Cu–Ni(OH)2/NF exhibits a higher FE of FA (FEFA) across a wide potential range of 1.4 to 1.6 V (Fig. 4c). When the potential is fixed at 1.5 V, Cu–Ni(OH)2/NF achieves a maximum FEFA of 98%, along with a FA productivity of 5.92 mmol cm−2 h−1, surpassing the performance of most recently reported EGOR catalysts (Table S2, ESI). Intriguingly, when a total electric charge of 2000 C is transferred through the entire cell, Cu–Ni(OH)2/NF achieves an EG conversion of 97% and a FA yield of 89%, with the production of a small amount of glycolic acid (GA) byproduct (Fig. 4d and Fig. S25, ESI), indicating the highly selective conversion of EG to FA. The long-term stability of Cu–Ni(OH)2/NF for EGOR was performed at a current density of 100 mA cm−2 (Fig. 4e). Impressively, the FEFA value is consistently maintained at around 90% throughout the 100-h operation period. The recovered Cu–Ni(OH)2/NF catalyst was characterized by using various techniques, including SEM, TEM, XRD, and XPS (Fig. S26–S28, ESI). These results show that the used catalyst retains the original morphology and structure, indicating the stability of Cu–Ni(OH)2/NF.

To understand the underlying mechanism responsible for the enhanced EGOR performance on Cu–Ni(OH)2/NF, the projected crystal orbital Hamilton populations (pCOHP) were performed (Fig. S29, ESI), which is used to elucidate the Ni–Oads bonding state (Oads is from the adsorbed EG species).45 It is well-accepted that a negative value of integrated pCOHP (IpCOHP) indicates a stable Ni–Oads bonding state.41 As depicted in Fig. 4f and g, the IpCOHP value relative to the Fermi level is calculated to be −0.69 for Ni–Oads on the Cu–Ni(OH)2 (001) surface, lower than that (−0.60) for Ni–Oads on the Ni(OH)2 (001) surface, suggesting an enhanced bonding interaction between Ni and EG upon Cu doping. Furthermore, the EG adsorption energy calculations confirm a more thermodynamically favorable EG adsorption on Cu–Ni(OH)2 (Fig. 4h). To experimentally verify the enhanced EG adsorption on Cu–Ni(OH)2, open-circuit potential (OCP) measurements were performed. As shown in Fig. S30 (ESI), the OCP change over Cu–Ni(OH)2 is determined to be 0.32 V, significantly higher than bare Ni(OH)2 (0.06 V), indicating the favorable EG adsorption on Cu–Ni(OH)2.46 The interfacial behaviors were investigated using Operando EIS at various potentials. When the EG feedstock is added, the Nyquist semicircle in the low-frequency region at 1.4 V becomes apparent (Fig. 4i and j), indicating the appearance of EGOR.47 This observation agrees with the oxidation potential observed from the LSV curves (Fig. 4a). Upon further increasing the potentials, the EGOR peak gradually shifts to lower phase degrees. These results suggest that EGOR is thermodynamically more favorable than OER in the presence of EG.48 The EGOR mechanism over Cu–Ni(OH)2 was investigated using in situ Raman measurements. Under OER conditions, a pair of characteristic bands at 470 and 557 cm−1, corresponding to NiOOH,49 are observed at 1.4 V (Fig. 4k). The band intensities of NiOOH increase with increasing potential, indicating surface reconstruction on Cu–Ni(OH)2. Upon EG addition, the NiOOH signals completely disappeared and the catalyst reverts to its initial Cu–Ni(OH)2 (Fig. 4l), indicating that EGOR proceeds via an indirect oxidation way on Cu–Ni(OH)2.50

Co-production of NH3 and FA

Encouraged by the high activities for half-reaction NO3RR and EGOR, we attempted to integrate the NO3RR and EGOR in a two-electrode membrane–electrode assembly (MEA) system separated by a proton exchange membrane (PEM) (Fig. 5a). The as-prepared Cu–Ni(OH)2/NF and OV–Co3O4/NF served as the anode and cathode, respectively. The LSV curves show that, to obtain a current density of 100 mA cm−2, the paired NO3RR//EGOR system only requires a cell voltage of 1.66 V, markedly lower than that (2.20 V) for the conventional NO3RR//OER system (Fig. 5b). This disparity suggests that the replacement of the OER by EGOR contributes to reducing the overall energy input. To assess the coupled performance in the two-electrode system, the FEs of FA and NH3 were measured under different current densities during 2-h electrolysis. It can be observed that Cu–Ni(OH)2/NF enables stable and high FE for FA production (>95%) across a wide current density range from 25 to 100 mA cm−2. In contrast, NO3RR activity exhibits a volcano-type trend, with a current density of 100 mA cm−2 being recognized as optimal for NH3 production. At higher current densities (>100 mA cm−2), the observed reduction in the FE of NH3 is attributed to the competitive hydrogen evolution reaction.
image file: d5mh00130g-f5.tif
Fig. 5 Integration of the NO3RR and EGOR. (a) Schematic diagram showing the two-electrode MEA reactor for NO3RR coupled with EGOR. (b) Comparison of the LSV curves for NO3RR//OER and NO3RR//EGOR system. (c) The FEs of NH3 and FA measured under different current densities. (d) Schematic illustration of formate and NH4Cl production from PET bottle and nitrate wastewater. (e) and (f) XRD patterns of synthesized formate (e) and NH4Cl (f). (g) TEA of NO3RR//OER and NO3RR//EGOR system.

The practical application of this paired system was subsequently carried out using real-world PET bottles and simulated nitrate wastewater as the initial feedstock (Fig. 5d and Fig. S31, ESI). Prior to co-electrolysis, 100 g of post-consumer plastic bottles were collected and hydrolyzed into their constituent monomers using KOH (Fig. S32, ESI). The resulting hydrolysate was directly fed into the anodic chamber. As for the cathodic feedstock, a simulated high-concentration nitrate solution was prepared via electrodialysis, following established protocols from previous literature.15 Of note, KOH was added as electrolyte prior to electrochemical valorization. After 210-h electrolysis (Fig. S33, ESI), FA was identified as the primary product in the anolyte, evidenced by 1H nuclear magnetic resonance (NMR) (Fig. S34, ESI). Following separation and purification, 76.67 g of formate were obtained (FA exists as formate under alkaline conditions), with an overall yield of 83%. In parallel, NH3 was generated at the cathode with an 80% yield, resulting in 13.76 g of NH4Cl crystals after acidification and evaporation. The high purity of formate and NH4Cl was confirmed via XRD patterns, FTIR spectra, and 1H NMR (Fig. 5e and f and Fig. S35 and S36, ESI). To assess the economic viability of our constructed NO3RR//EGOR system, a TEA was conducted (Fig. 5g and Note S1, ESI). The TEA results indicate that our coupled system achieves a higher gross profit (79665.9 $ d−1), corresponding to a 15-fold increase than the conventional NO3RR//OER system (5173.4 $ d−1), which indicates the potential market competitiveness of our built electrosynthesis system.

Conclusions

In this work, we constructed a paired electrolysis system with low energy input comprising cathodic reduction of NO3 to NH3 and anodic oxidation of PET-derived EG to FA, which is made possible by the discovery of the OV–Co3O4 cathode and Cu–Ni(OH)2 anode. The OV incorporation compensates for the deficiencies of bare Co3O4 in terms of NO3 adsorption and activation. Cu doping enlarges the Ni–O covalency, thus resulting in an enhanced capability for EG adsorption. Under optimal conditions, this paired electrochemical system achieves 13.76 g of high-purity NH4Cl crystals with an 80% yield from simulated waster and 76.67 g of formate with an 83% yield from post-consumer PET bottles. Overall, this work provides a cost-effective and low-carbon route to mitigate wastes while producing valuable chemicals.

Author contributions

B. Qiu and Y. Zhang supervised the project and conceived the idea. B. Qiu, N. Zhang, M. Du, and M. Han designed all experiments and co-wrote the manuscript. M. Du, T. Sun, and X. Guo performed the materials synthesis and characterization studies. M. Du, W. Chen, M. Han, and J. Ma carried out electrocatalytic measurements. J. Shang performed the DFT simulations. V. Nicolosi and C. Zhou provided suggestions for the characterization studies. All authors provided comments on the manuscript.

Data availability

The data supporting this article have been included as part of the ESI.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

This work was supported by the National Natural Science Foundation of China (22302095, 22378208, and 22305243), the Fundamental Research Funds for the Central Universities (KJYQ2025012), the Shenzhen Science and Technology Planning Project (JCYJ20220531100613030), and the Natural Science Foundation of Jiangsu Province (BK20210382 and BK20230272). V. N. and X. G. wish to thank the support of the SFI-funded AMBER Research Centre and the SFI Frontiers for the Future award (Grant No. 12/RC/2278_P2 and 20/FFP-A/8950 respectively). Furthermore, V. N. and X. G. wish to thank the Advanced Microscopy Laboratory in CRANN for the provision of their facilities.

References

  1. G.-F. Han, F. Li, Z.-W. Chen, C. Coppex, S.-J. Kim, H.-J. Noh, Z. Fu, Y. Lu, C. V. Singh, S. Siahrostami, Q. Jiang and J.-B. Baek, Nat. Nanotechnol., 2021, 16, 325–330 CrossRef CAS PubMed.
  2. S. Li, Y. Zhou, X. Fu, J. B. Pedersen, M. Saccoccio, S. Z. Andersen, K. Enemark-Rasmussen, P. J. Kempen, C. D. Damsgaard, A. Xu, R. Sažinas, J. B. V. Mygind, N. H. Deissler, J. Kibsgaard, P. C. K. Vesborg, J. K. Nørskov and I. Chorkendorff, Nature, 2024, 629, 92–97 CrossRef CAS PubMed.
  3. S. Li, X. Fu, J. K. Nørskov and I. Chorkendorff, Nat. Energy, 2024, 9, 1344–1349 CrossRef CAS.
  4. M. Wang, M. A. Khan, I. Mohsin, J. Wicks, A. H. Ip, K. Z. Sumon, C.-T. Dinh, E. H. Sargent, I. D. Gates and M. G. Kibria, Energy Environ. Sci., 2021, 14, 2535–2548 RSC.
  5. D. Ye and S. C. E. Tsang, Nat. Synth., 2023, 2, 612–623 CrossRef CAS.
  6. K. Chu, Y. Luo, P. Shen, X. Li, Q. Li and Y. Guo, Adv. Energy Mater., 2022, 12, 2103022 CrossRef CAS.
  7. S. Han, H. Li, T. Li, F. Chen, R. Yang, Y. Yu and B. Zhang, Nat. Catal., 2023, 6, 402–414 CrossRef CAS.
  8. J. Hao, T. Wang, R. Yu, J. Cai, G. Gao, Z. Zhuang, Q. Kang, S. Lu, Z. Liu and J. Wu, Nat. Commun., 2024, 15, 9020 CrossRef CAS PubMed.
  9. K. Kim, A. Zagalskaya, J. L. Ng, J. Hong, V. Alexandrov, T. A. Pham and X. Su, Nat. Commun., 2023, 14, 823 CrossRef CAS PubMed.
  10. M. Du and B. Qiu, Nat. Water, 2024, 2, 1053–1054 CrossRef.
  11. P. H. van Langevelde, I. Katsounaros and M. T. Koper, Joule, 2021, 5, 290–294 CrossRef.
  12. M. Duca and M. T. Koper, Energy Environ. Sci., 2012, 5, 9726–9742 RSC.
  13. G. Qing, R. Ghazfar, S. T. Jackowski, F. Habibzadeh, M. M. Ashtiani, C.-P. Chen, M. R. Smith, III and T. W. Hamann, Chem. Rev., 2020, 120, 5437–5516 CrossRef CAS PubMed.
  14. M. Li, X. Liu, L. Wang, F. Hou, S. X. Dou and J. Liang, EcoMat, 2021, 3, e12096 CrossRef CAS.
  15. F.-Y. Chen, A. Elgazzar, S. Pecaut, C. Qiu, Y. Feng, S. Ashokkumar, Z. Yu, C. Sellers, S. Hao, P. Zhu and H. Wang, Nat. Catal., 2024, 7, 1032–1043 CrossRef CAS.
  16. Y. Liu, Z. Zhuang, Y. Liu, N. Liu, Y. Li, Y. Cheng, J. Yu, R. Yu, D. Wang and H. Li, Angew. Chem., Int. Ed., 2024, 63, e202411396 CAS.
  17. C. Li, H. Li, B. Zhang, H. Li, Y. Wang, X. Wang, P. Das, Y. Li, X. Wu, Y. Li, Y. Cui, J. Xiao and Z.-S. Wu, Angew. Chem., Int. Ed., 2024, 63, e202411542 CrossRef CAS PubMed.
  18. Q. Hu, S. Qi, Q. Huo, Y. Zhao, J. Sun, X. Chen, M. Lv, W. Zhou, C. Feng, X. Chai, H. Yang and C. He, J. Am. Chem. Soc., 2024, 146, 2967–2976 CrossRef CAS PubMed.
  19. S. Qi, Z. Lei, Q. Huo, J. Zhao, T. Huang, N. Meng, J. Liao, J. Yi, C. Shang, X. Zhang, H. Yang, Q. Hu and C. He, Adv. Mater., 2024, 36, 2403958 CrossRef CAS PubMed.
  20. J. Yan, P. Liu, J. Li, H. Huang and W. Song, Small, 2024, 20, 2308617 CrossRef CAS PubMed.
  21. Z. Fan, C. Cao, X. Yang, W. Yuan, F. Qin, Y. Hu, X. Sun, G. Liu, Y. Tian and L. Xu, Angew. Chem., Int. Ed., 2024, 63, e202410356 CrossRef CAS PubMed.
  22. J.-Y. Fang, Q.-Z. Zheng, Y.-Y. Lou, K.-M. Zhao, S.-N. Hu, G. Li, O. Akdim, X.-Y. Huang and S.-G. Sun, Nat. Commun., 2022, 13, 7899 CrossRef CAS PubMed.
  23. K. Zhang, P. Sun, Y. Huang, M. Tang, X. Zou, Z. Pan, X. Huo, J. Wu, C. Lin, Z. Sun, Y. Wan, X. Zhang and L. An, Adv. Funct. Mater., 2024, 34, 2405179 CrossRef CAS.
  24. M. Zhang, Z. Ma, S. Zhou, C. Han, V. Kundi, P. V. Kumar, L. Thomsen, B. Johannessen, L. Peng, Y. Shan, C. Tsounis, Y. Yang, J. Pan and R. Amal, ACS Catal., 2024, 14, 11231–11242 CrossRef CAS.
  25. S. Kang, X. Guo, D. Xing, W. Yuan, J. Shang, V. Nicolosi, N. Zhang and B. Qiu, Small, 2024, 20, 2406068 CrossRef CAS PubMed.
  26. K. Chu, W. Zong, G. Xue, H. Guo, J. Qin, H. Zhu, N. Zhang, Z. Tian, H. Dong, Y.-E. Miao, M. B. J. Roeffaers, J. Hofkens, F. Lai and T. Liu, J. Am. Chem. Soc., 2023, 145, 21387–21396 CrossRef CAS PubMed.
  27. Y. Li, J. Ma, Z. Wu and Z. Wang, Environ. Sci. Technol., 2022, 56, 8673–8681 CrossRef CAS PubMed.
  28. J. Li, H. Li, K. Fan, J. Y. Lee, W. Xie and M. Shao, Chem Catal., 2023, 3, 100638 CrossRef CAS.
  29. Y. Yang, J. Liu, M. Ahmad, X.-Q. Sun, C. Liu, S. Chen, J. Zhang, J.-L. Luo and X.-Z. Fu, Chem. Eng. J., 2025, 505, 159653 CrossRef CAS.
  30. S.-Q. Liu, S. Xie, S. Wu, Y. Yang, P.-X. Lei, S. Luo, R. Feng, X.-Z. Fu and J.-L. Luo, Adv. Funct. Mater., 2024, 34, 2404105 CrossRef CAS.
  31. M. Du, R. Xue, W. Yuan, Y. Cheng, Z. Cui, W. Dong and B. Qiu, Nano Lett., 2024, 24, 9768–9775 CrossRef CAS PubMed.
  32. T. Ren, Z. Duan, H. Wang, H. Yu, K. Deng, Z. Wang, H. Wang, L. Wang and Y. Xu, ACS Catal., 2023, 13, 10394–10404 Search PubMed.
  33. M. Du, M. Xing, S. Kang, Y. Ma, B. Qiu and Y. Chai, EcoMat, 2022, 4, e12259 CrossRef CAS.
  34. Y. Ma, X. Guo, M. Du, S. Kang, W. Dong, V. Nicolosi, Z. Cui, Y. Zhang and B. Qiu, Green Chem., 2024, 26, 3995–4004 RSC.
  35. Y. Ma, Y. Zhang, M. Xing, S. Kang, M. Du, B. Qiu and Y. Chai, Chem. Commun., 2022, 58, 6642–6645 RSC.
  36. Z. Xiao, Y.-C. Huang, C.-L. Dong, C. Xie, Z. Liu, S. Du, W. Chen, D. Yan, L. Tao and Z. Shu, J. Am. Chem. Soc., 2020, 142, 12087–12095 CrossRef CAS PubMed.
  37. H. Idriss, Surf. Sci., 2021, 712, 121894 CrossRef CAS.
  38. T. J. Frankcombe and Y. Liu, Chem. Mater., 2023, 35, 5468–5474 CrossRef CAS.
  39. Y. Ma, H. Tao, X. Guo, P. Yang, D. Xing, V. Nicolosi, Y. Zhang, C. Lian and B. Qiu, J. Mater. Chem. A, 2024, 12, 27220–27228 RSC.
  40. Y. Sun, H. Liao, J. Wang, B. Chen, S. Sun, S. J. H. Ong, S. Xi, C. Diao, Y. Du and J.-O. Wang, Nat. Catal., 2020, 3, 554–563 CrossRef CAS.
  41. N. Zhang, J. Shang, X. Deng, L. Cai, R. Long, Y. Xiong and Y. Chai, ACS Nano, 2022, 16, 4795–4804 CrossRef CAS PubMed.
  42. C. Zhou, Y. Zhang, C. Xie, J. Bai, J. Li, H. Zhang, H. Zhu, M. Long, B. Zhou and G. Zheng, Environ. Sci. Technol., 2024, 58, 14940–14948 CrossRef CAS PubMed.
  43. J. Ni, J. Yan, F. Li, H. Qi, Q. Xu, C. Su, L. Sun, H. Sun, J. Ding and B. Liu, Adv. Energy Mater., 2024, 14, 2400065 CrossRef CAS.
  44. F. Ma, Y. Wen, P. Fu, J. Zhang, Q. Tang, T. Chen, W. Luo, Y. Zhou and J. Wang, Small, 2024, 20, 2305767 CrossRef CAS PubMed.
  45. Y.-C. Hao, Y. Guo, L.-W. Chen, M. Shu, X.-Y. Wang, T.-A. Bu, W.-Y. Gao, N. Zhang, X. Su and X. Feng, Nat. Catal., 2019, 2, 448–456 CrossRef CAS.
  46. M. Du, Y. Zhang, S. Kang, C. Xu, Y. Ma, L. Cai, Y. Zhu, Y. Chai and B. Qiu, Small, 2023, 19, 2303693 CrossRef CAS PubMed.
  47. M. Song, Y. Wu, Z. Zhao, M. Zheng, C. Wang and J. Lu, Adv. Mater., 2024, 36, 2403234 CrossRef CAS PubMed.
  48. F. Ma, Z. Li, R. Hu, Z. Wang, J. Wang, J. Li, Y. Nie, Z. Zheng and X. Jiang, ACS Catal., 2023, 13, 14163–14172 CrossRef CAS.
  49. H. Zhong, Q. Zhang, J. Yu, X. Zhang, C. Wu, Y. Ma, H. An, H. Wang, J. Zhang and X. Wang, Adv. Energy Mater., 2023, 13, 2301391 CrossRef CAS.
  50. W. Luo, H. Tian, Q. Li, G. Meng, Z. Chang, C. Chen, R. Shen, X. Yu, L. Zhu and F. Kong, Adv. Funct. Mater., 2024, 34, 2306995 CrossRef CAS.

Footnotes

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5mh00130g
These authors contributed equally to this work.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.