Main

The surface coordination passivation on Cu was first created by hydrothermally treating Cu foils in an aqueous solution of sodium formate at 200 °C. As shown in Fig. 1a, the surface of untreated Cu turned fully dark after being kept in 0.1 M NaOH at 25 °C for 8 h. By contrast, the Cu foil after formate treatment (denoted as Cu-FA) retained its metallic lustre under the same conditions, and outperformed even widely used Cu-alloys (that is, brass and bronze), graphene-coated Cu and benzotriazole (BTA)-treated Cu (Extended Data Fig. 1a, b). As revealed by optical, scanning electron microscopy (SEM), Raman and X-ray diffraction (XRD) analyses (Fig. 1b, c, Extended Data Fig. 1a–d), whereas the surfaces of brass, bronze and graphene-coated Cu foils were heavily oxidized, with the formation of black-coloured CuO after air exposure in NaOH, no obvious oxidation species were detected on Cu-FA.

Fig. 1: Anti-corrosion properties of Cu foils after the formate treatment.
figure 1

a, Photographs of formate-modified Cu foil (Cu-FA), Cu foil, brass and graphene-coated Cu foils (Cu-G) before (top) and after (bottom) exposure to 0.1 M NaOH for 8 h. b, Optical microscope photographs and corresponding Raman spectra of Cu-FA (left) and Cu foils (right) taken after NaOH exposure for 8 h. c, Representative SEM images of Cu-FA and Cu foils before (top) and after (bottom) exposure to NaOH for 8 h.

Source data

The hydrothermal temperature of the formate treatment is critical for building up an anti-corrosive surface on Cu. Lowering the temperature from 200 °C to 160 °C failed to create effective passivation (Extended Data Fig. 1e). By switching the treatment to solvothermal conditions with the presence of sodium formate and oleylamine in a mixed solvent of dimethylformamide (DMF) and H2O, the passivation on Cu (Extended Data Fig. 1f–j) was readily developed at 160 °C. Quantitative electrochemical measurements (see Extended Data Fig. 1h, Methods) demonstrated that the oxidative corrosion rate of Cu-FA in 0.1 M NaOH was reduced to 3.89 μm yr−1, corresponding to a 20-fold enhancement in anti-corrosion performance compared to bare Cu (78.2 μm yr−1)14,15,16. The anti-corrosion performance achieved by the formate treatment was much better than that reported for small organic molecules4,10,16. The corrosion rate of Cu-FA was one order of magnitude lower than that of Cu foil treated by 1-dodecathiol (Cu-DT; 50.4 μm yr−1) and BTA (Cu-BTA; 59.7 μm yr−1) (Extended Data Fig. 2a). In addition, the passivation layer was effective at preventing the oxidation of Cu under other harsh conditions (Extended Data Fig. 2b–d). After electrochemical tests (1 h in NaOH), whereas the untreated Cu foil was heavily corroded, with the formation of a rough surface containing oxidized products, the surface of Cu-FA remained almost unchanged. Electrochemical measurements revealed a much enhanced corrosion resistance of Cu in 1 M NaOH after the formate treatment. The surface of Cu-FA remained intact after exposure to H2O2 (30%) for 15 min. Oxidation started to take place after 60 min. In comparison, the surface of untreated Cu was severely oxidized after 15 min. After 1 h of heating at 160 °C in air, the surface passivation layer was still highly effective at preventing oxidation of Cu in alkaline conditions. Moreover, as revealed by a home-built integrated scanning reference electrode technique and scanning tunnelling microscopy (SRET/STM) system, even when Cu-FA was scratched, oxidative corrosion was restricted locally to the breached areas and was not accelerated (Extended Data Fig. 2e), indicating that the method developed here is superior to techniques using pinhole-free cathodic coating8.

Based on the observation that the anti-corrosion performance of Cu-FA depends on the formate treatment time (Extended Data Fig. 3a, b), we investigated the underlying mechanism. Although switching from the formate–H2O system to the formate–oleylamine system in mixed DMF–H2O decreased the temperature from 200 °C to 160 °C, a long reaction time was still required to achieve effective passivation. Surprisingly, regardless of which treatment system was used and which original crystalline orientation of the polycrystalline Cu foil was selected, the relative intensity of the Cu(110) diffraction peak increased with treatment time, indicating that deep lattice reconstruction of the Cu surface took place during the formate treatment. Cross-sectional samples of the Cu foils before and after the formate treatment were prepared by the focused-ion-beam technique for the direct imaging of lattice changes using transmission electron microscopy (TEM) (Extended Data Fig. 3c, d). Whereas the surface of the untreated Cu foil was mainly Cu(111), the surface of the formate-treated Cu foil was reconstructed to Cu(110), which was not limited to surface atomic layers, but extended to a thickness of over 200 nm.

To identify the atomic structure of the surface passivation layer on Cu, we carried out combined scanning tunnelling microscopy (STM) and atomic-force microscopy (AFM) experiments using a qPlus sensor17,18. Contaminants on the Cu-FA foils were removed by annealing in an ultrahigh vacuum (UHV) chamber (Extended Data Fig. 4a). Although annealing at 150 °C already gave STM images showing step planes and sharp step edges, it was impossible to obtain high-resolution STM images. Increasing the annealing temperature to 300 °C produced un-contaminated surfaces for high-resolution imaging. Although microcrystalline Cu foils were used for the formate treatment, large-scale single-crystalline domains (Fig. 2a) were observed on Cu-FA. In comparison, untreated Cu foils showed poorly defined rough surfaces after the same annealing. The steps on Cu-FA were uniform with a height of about 2.55 Å, corresponding to the diatomic step height of Cu(110), indicating that the microcrystalline Cu surface was mostly reconstructed to Cu(110) after the formate treatment. The same step heights were observed for the samples annealed at 150 °C and 300 °C, confirming that the reconstruction was not due to the annealing at 300 °C, but induced by the formate treatment, as suggested by XRD and TEM.

Fig. 2: STM and AFM imaging of formate-treated Cu.
figure 2

a, STM topography of the formate-treated Cu foil. Inset, line profile across an atomic step. Before imaging, the sample was annealed under UHV at 300 °C. b, Zoom-in STM image showing a c(6 × 2) superlattice. The unit cell is highlighted by a white rectangle. Set points of STM images: 1 V, 20 pA (a); −1 V, 50 pA (b). c, High-resolution constant-height AFM image (Δf, frequency shift). The \([1\,\bar{1}\,0]\) row of Cu(110) is highlighted by a red arrow. Inset, zoom-in AFM image of the paired triangular lobes (top) and corresponding structural model (bottom). The AFM image was recorded with a tip height of −0.3 nm, referenced to the STM set point (1 V, 5 pA). The oscillation amplitude is 0.1 nm. d, Simulated AFM image based on the DFT structure shown in e, using a neutral-tip model (effective lateral stiffness k = 0.25 N m−1, oscillation amplitude A = 0.1 nm). e, Dehydrated structure model of Cu(110) passivated by dinuclear Cu(ii)–formate motifs. The structure was built by removing H2O from the structure model of a Cu(110)-c(6 × 2) surface fully passivated by [Cu(μ-HCOO)(OH)2]2 and O2− (Extended Data Fig. 4c). Blue spheres, Cu on the surface; green spheres, Cu on the subsurface; light-grey spheres, Cu in the bulk; white spheres, H; red spheres, O; grey spheres, C.

From the high-resolution STM images of Cu-FA, a perfect Cu(110)-c(6 × 2) superlattice (Fig. 2b) was identified. It should be noted that the c(6 × 2) superstructure (Extended Data Fig. 4b) was also observed on single-crystal Cu(110) treated with sodium formate solution, followed by annealing at 150 °C; this superstructure was comparable to that created by the formate treatment in providing effective passivation, implying that it did not result from annealing at 300 °C. In the constant-height AFM image, each round protrusion in the STM image was further resolved to be paired non-spherical lobes (Fig. 2c). In addition, the \([1\,\bar{1}\,0]\) rows of the underlying Cu(110) are visible, as indicated by the red arrow in Fig. 2c. In the zoom-in AFM image, the paired lobes have a distinct triangular shape with a separation of 3.73 ± 0.03 Å (standard deviation of four measurements) (inset of Fig. 2c), which is larger than the distance between neighbouring Cu atoms (2.5 Å) on the \([1\,\bar{1}\,0]\) row, precluding that the bright protrusions originate from bidentate formates on the \([1\,\bar{1}\,0]\) row19.

Cu easily forms paddle-wheel dinuclear Cu(ii) carboxylate coordination compounds20. When the ligands are formates, the distance of C atoms on each pair of perpendicularly bridging formates is typically in the range 3.2–4.0 Å (Extended Data Fig. 4c). Moreover, the presence of bridging formates on Cu-FA annealed under similar conditions as those used for STM imaging was confirmed by Fourier transform infrared (FTIR), Raman spectroscopy and temperature-programmed desorption–mass spectrometry (TPD-MS) studies (Extended Data Fig. 4d). With the presence of formates, the Cu-FA foils annealed in UHV still exhibited the same anti-corrosion performance as before annealing (Extended Data Fig. 4e). On the basis of these results, a (6 × 2) supercell of Cu(110) with a five-layer-thickness slab was built for periodic density functional theory (DFT) calculations using the Perdew–Burke–Ernzerhof (PBE) functional as implemented in the Vienna ab initio simulation package21. The slab was coated by a layer consisting of dinuclear [Cu(μ-HCOO)(OH)2]2 units and O2− arranged in a c(6 × 2) configuration (Extended Data Fig. 4c). Although different Cu(110) structures modified with formate have been extensively investigated22,23,24,25, a structure modified with dinuclear Cu–formate motifs has not yet been reported in the literature. This structure can be viewed as a gradual material with the formal valence of Cu decreased from +2 on the surface dinuclear Cu motifs to +1 in the subsurface and to 0 in the inner Cu atoms, as revealed by Bader charge analysis and confirmed by linear sweep voltammetry and by Auger and X-ray photoelectron (XPS) spectroscopies (Extended Data Fig. 4f-g).

XPS measurements also suggested that a dehydration process occurred on Cu-FA upon annealing under conditions (UHV, 300 °C) similar to those of the pretreatment for the high-resolution AFM/STM imaging. Whereas the presence of OH was observed for the unannealed sample, the annealed Cu-FA displayed XPS signals of O from the formate and O2− on Cu, but no obvious OH signal. According to the observed dehydration process, the dehydrated structure model (Fig. 2e) was optimized by DFT calculations. The simulated AFM image of the optimized structure (Fig. 2d) shows not only the carbon–carbon distance of 3.66 Å between the two coordinated formates, but also the fine triangular structure of the formate moieties, consistent with the experimental observations (Fig. 2c). The dehydrated c(6 × 2) structure should therefore be the structure observed by STM after annealing at 300 °C. However, once the dehydrated structure was exposed to the atmosphere, the occurrence of the hydration process was revealed by STM (Extended Data Fig. 5a), with the appearance of defect-like dark depressions, which are attributed to interstitial O2− species on the hydrated c(6 × 2) structure (Extended Data Fig. 4c). Therefore, the hydrated c(6 × 2) structure must be the real passivating structure, although the dehydrated structure is easily obtained upon annealing (Extended Data Fig. 5b).

The passivation layer reveals the two important features. (1) Whereas the valence of the outmost surface Cu is +2, that of the subsurface Cu atoms is +1. All exposed Cu sites are already in the oxidized state. (2) The subsurface Cu atoms are fully bound by O2− and OH groups. In alkaline conditions, the binding of OH on oxidized Cu species is so strong that, for example, Cu(OH)2 has a very small solubility product constant of 2 × 10−22. These features make it difficult for O2 or other adsorbates to interact with Cu underneath the passivation layer. The perfect symmetry match between Cu(110) and the four OH binding sites on [Cu(μ-HCOO)(OH)2]2 is crucial to the passivation. It is therefore critical to have the Cu surface reconstructed into Cu(110) before a robust passivation layer can be created, which explains why a long treatment time was required to achieve effective anti-corrosion. As expected, when a single-crystal Cu(110) surface was treated with formate solution at 100 °C (Fig. 3a–c, Extended Data Fig. 5c, d), an anti-corrosion effect was readily achieved, even when the treatment time was reduced to 1 h. In comparison, no anti-corrosion protection was identified on Cu with (111) and (100) surfaces, even by extending the treatment to 10 h under the same conditions. However, when the Cu(111) surface was scratched to introduce defect sites, the reconstruction of the single-crystal surface was readily accelerated to give anti-corrosion protection (Extended Data Fig. 5e), suggesting that the reconstruction must be initiated at the domain boundaries. These results confirmed the importance of rearranging the Cu surface into Cu(110) before the formate passivation becomes effective.

Fig. 3: Importance of Cu(110) for effective passivation.
figure 3

ac, SEM images of formate-treated single-crystalline surfaces after exposure to 0.1 M sodium hydroxide for 12 h: Cu(110)-FA (a), Cu(100)-FA (b) and Cu(111)-FA (c). d, Comparison of adsorption free energies of O2 and Cl at 298 K on a clean Cu(110) surface and on Cu(110)-FA with surface co-passivation by [Cu(μ-HCOO)(OH)2]2 and O2−. e, Comparison of the mass of dissolved Cu(ii) from Cu-FA/DT, Cu-FA and Cu after the salt spray test in 5% NaCl at 47 °C for 96 h. The data are averages of three independent measurements. Error bars represent the standard errors. f, Optical photographs of Cu-FA/DT (1), Cu-FA (2) and Cu (3) foils after a 10-min exposure to Na2S (50 mM).

Source data

As demonstrated by DFT calculations (Fig. 3d, Extended Data Fig. 5f), whereas O2 binds strongly with clean Cu(110), the Cu(110)-c(6 × 2) surface passivated by [Cu(μ-HCOO)2(OH)2]2 and O2− leaves no accessible sites for O2 to be adsorbed and activated so that the oxidation of Cu can be switched off. DFT calculations also suggested that the anti-corrosion properties of Cu-FA surfaces should be enhanced if very strong alkanethiol ligands are used to coordinate with steps or defect sites that are not protected by [Cu(μ-HCOO)(OH)2]2 units (Extended Data Fig. 6a). 1-dodecathiol (DT) was therefore chosen for the demonstration. The Cu-FA foil that was further treated with DT (denoted as Cu-FA/DT) exhibited a substantially enhanced anti-corrosion performance (Fig. 3e, f, Extended Data Fig. 6b–g) with a corrosion rate of 0.755 μm yr−1 in 0.1 M NaOH, two orders of magnitude lower than that of bare Cu (78.2 μm yr−1). The surface of Cu-FA/DT remained almost intact after exposure to salt spray for 24 h. Moreover, compared with Cu-FA, the Cu-FA/DT foil increased its Na2S-resistant concentration by two orders of magnitude. No obvious formation of dark copper sulfide was observed on Cu-FA/DT after 5 h of immersion in 10 mM Na2S. Cu-FA/DT also exhibited enhanced oxidation resistance in 720-h seawater inundation and 30% H2O2.

Understanding of this mechanism provided an alternative strategy to achieve effective passivation by introducing Cu(HCOO)2 as the metal precursor for the reductive growth of a Cu(110) surface on Cu foils. This strategy allows us to create well passivated a Cu(110) surface at 120 °C, making the technique applicable to Cu materials of different forms and sizes (Extended Data Fig. 7). For example, the electric conductivity of treated Cu wires was well maintained even after 60 h of ageing in 0.1 M NaOH at 60 °C. No obvious surface change was observed for treated Cu wires subjected to 24 h of air oxidation at 160 °C. The Cu-FA/DT meshes were highly corrosion-resistant to salt spray and Na2S. Moreover, no obvious colour change was observed on Cu-FA/DT tubes and foils after a 96-h wet mechanical test in which salty water (3.5% NaCl, 1% Na2CO3, 1% Na2SO4 and 0.1% NaOH) was running through the samples intermittently, suggesting good wet mechanical properties for the passivation layer.

The developed anti-corrosion protocol also works well for Cu nanomaterials (Extended Data Fig. 8). Cu nanowires (NWs) with an average diameter of ~35 nm and length of ~220 μm were synthesized for the assessment26. The unmodified Cu NWs in solution were unstable in air and were easily oxidized to form a surface oxide layer, thus dramatically reducing the electronic conductivity27. After the formate treatment, the Cu NWs (denoted as Cu NWs-FA) showed high stability in air for 90 days owing to their fully passivated Cu(110) surface. No formation of Cu oxide species was detected from Cu NWs-FA after 48 h of thermal treatment at 80 °C in air (Fig. 4a–d). However, because of their high surface area, the Cu NWs-FA did not survive under harsh alkaline corrosion conditions of 1 M NaOH in air. When further treated with DT, the resulting Cu NWs-FA/DT exhibited an enhanced anti-corrosion performance under harsh conditions (that is, 1 M NaOH in air, salt spray at 47 °C), which makes them a good candidate for fabricating transparent conductive electrodes. Using Cu NWs-FA/DT, we obtained a transparent conductive electrode with sheet resistance of 25.8 Ω per square (light transmission of 87% at 550 nm) and with good air and mechanical stabilities (Extended Data Table 1). Highly corrosion-resistant Cu nanoparticles were also prepared using the developed protocol. Moreover, the technique was successfully applied to kilogram-scale production of Cu pastes under mild conditions to replace Ag pastes for radiofrequency identification applications (Extended Data Fig. 9a–d).

Fig. 4: Anti-corrosion strategy for stabilizing Cu NWs and room-temperature electrochemical anti-corrosion technique.
figure 4

ac, SEM images of formate-treated Cu NWs (Cu NWs-FA) (a) and untreated Cu NWs (b) after 48-h air oxidation at 80 °C, and their corresponding XRD patterns (c). Numbers in parentheses show the powder diffraction file card numbers of the samples (JCPDS, joint committee on powder diffraction standards). d, Resistance change profile of Cu NWs and Cu NWs-FA with heating time at 80 °C. e, XRD pattern of Cu foil before and after (Cu-FA(EC)) electrochemical treatment in the presence of sodium formate. f, Optical photographs of a Cu foil obtained by roll-to-roll electrochemical treatment followed by thiol treatment (Cu-FA(EC)/DT) and untreated Cu foil after ageing in 50 mM Na2S for 10 min.

Source data

The importance of the Cu anti-corrosion technique reported here is supported by the development of a facile, quick and efficient electrochemical process to fabricate anti-corrosive Cu surface at ambient conditions (Extended Data Fig. 9e–i). Lattice reconstruction for Cu surfaces under electrochemical redox conditions has been well documented28,29,30. In our electrochemical process, an anodic sweep from −0.80 V to 0.20 V (versus a saturated calomel electrode, SCE) in an electrolyte containing 1.0% HCOONa (pH 8–10) at a scan rate of 10 mV s−1 was applied to generate oxidized Cu species, followed by a 5-min electrochemical reduction at −0.7 V (versus SCE), during which the cathodic reduction of Cu(i) into Cu(0) occurred. As illustrated in Fig. 4e, the diffraction peak of Cu(220) was obviously enhanced after the electrochemical treatment, suggesting that the electrochemical treatment also induced reconstruction similar to that achieved by the solvothermal method. The as-prepared Cu-FA(EC) foil also exhibited identical anti-corrosion properties to the Cu-FA foils made by the solvothermal method. Similarly, further DT treatment can enhance the performance to make Cu highly corrosion-tolerant to Na2S. The electrochemical method developed in this work makes it possible to continuously produce anti-corrosive Cu materials using roll-to-roll processes for practical applications (Fig. 4f). However, we should point out that in the application of the formate technique, acidic conditions with pH < 3 must be avoided, because the formate passivation layer cannot survive under such conditions.

Methods

Materials

The Cu foil (99.8%, metal basis, thickness 25 μm, No. 046986) and polyvinylpyrrolidone (PVP; molecular weight 55,000) were purchased from Alfa Aesar Chemical Reagent Co. Ltd (Tianjin, China). Cu(111), Cu(110) and Cu(100) single crystals were obtained from Hefei Kejing Materials Technology Co., Ltd (Hefei, China). 1-DT (analytical reagent, AR), CuCl2·2H2O (AR), glucose (AR), oleylamine (80−90%) and oleic acid (90%) were purchased from Aladdin Reagent Co., Ltd (Shanghai, China). Sodium formate (HCOONa), sodium sulfide (Na2S), copper sulfate (CuSO4·5H2O), sodium hydroxide (NaOH), diethyldithiocarbamate (DDTC), sodium ethylenediamine tetraacetate (EDTA), ammonium citrate, ethylene glycol (EG), DMF, carbon tetrachloride (CCl4), acetone, ethanol and other reagents were purchased from Sinopharm Chemical Reagent Co., Ltd (Shanghai, China). Brass foil (H68), bronze foil, Cu wire (99.9%, diameter 2 mm) and Cu mesh (99.9%, thickness 50 μm) were obtained from Shenzhen Xinming Metal Material Co., Ltd (Shenzhen, China). All reagents were used as received without further purification. The water used in the studies was ultrapure water with a resistivity of ≥18 MΩ cm provided by a Millipore water purification system (Bedford, USA).

Hydrothermal and solvothermal formate treatments of copper (Cu-FA)

The following three methods were developed to treat Cu surfaces with formate for effective anti-oxidation passivation.

Method I

Hydrothermal treatment of Cu in the presence of formate. In a typical treatment, a 2 cm × 2 cm × 0.025 mm Cu foil or mesh or a 10-cm-long Cu wire is placed with 500 mg HCOONa in 10 ml of nitrogen-saturated H2O. The mixture is sealed in a 50-ml stainless-steel autoclave. The mixture is then heated from room temperature to 200 °C in 40 min and kept at 200 °C for 24 h to prepare Cu-FA. The Cu-FA products are collected and rinsed with deionized water to remove free HCOONa.

Method II

Solvothermal formate treatment of Cu in the presence of oleylamine (Cu-FA). In a typical treatment, a 2 cm × 2 cm × 0.025 mm Cu foil or mesh or a 10-cm-long Cu wire and 200 mg HCOONa in 1 ml H2O are dissolved in 12 ml DMF by sonication. Then, 1 ml oleylamine is added into the mixture and sonicated for another 3 min. The mixture is sealed in a 50-ml stainless-steel autoclave. The mixture is then heated from room temperature to 160 °C in 30 min and kept at 160 °C for 16 h to prepare Cu-FA. The Cu-FA products are collected and rinsed three times with deionized water and ethanol alternately to remove free HCOONa and oleylamine.

Method III

Solvothermal formate treatment of Cu in the presence of oleylamine and Cu(HCOO)2. In a typical treatment, a 2 cm × 2 cm × 0.025 mm Cu foil or mesh or a 10-cm-long Cu wire, 0.1 mg Cu(HCOO)2·4H2O and 200 mg HCOONa in 1 ml H2O are mixed in 12 ml DMF by sonication. Then, 1 ml oleylamine is added into the mixture and sonicated for another 3 min. The mixture is sealed in a 50-ml stainless-steel autoclave. The mixture is then heated from room temperature to 120 °C in 20 min and kept at 120 °C for 12 h to prepare Cu-FA. The Cu-FA products are collected and rinsed three times using deionized water and ethanol alternately to remove free HCOONa and oleylamine.

Formate treatments of Cu single crystals

In a typical treatment, two species of 1 cm × 1 cm × 0.5 cm Cu single crystals with exposure surfaces of Cu(110), Cu(100) or Cu(111) were placed in 10 ml H2O containing 200 mg HCOONa (N2-purged). The mixture was sealed in a 50-ml stainless-steel autoclave. The sample was then heated from room temperature to 100 °C in 20 min and kept at 100 °C for different times to prepare Cu(hkl)-FA. The single crystals were collected and rinsed with deionized water to remove free HCOONa.

Preparation of copper foils, meshes and wires co-stabilized by formate and 1-DT (Cu-FA/DT)

The as-prepared Cu-FA samples were dried in vacuum before use. In order to prepare Cu-FA/DT, the Cu-FA samples were immersed in 20 ml 1 mM DT/EtOH solution and kept still for 5 min, and were then washed with pure ethanol three times to produce Cu-FA/DT.

Preparation of graphene-coated Cu foils (Cu-G)

The Cu-G samples were grown by the chemical vapour deposition (CVD) method31. Cu foils (25 μm thick, 99.8%, Alfa Aesar) were placed on a flat oxide substrate and then loaded into a CVD system (Hefei Kejing Company, OTF1250). The system was heated to 1,000 °C in 1 h with Ar (500 standard cubic centimetres per minute, sccm), followed by annealing in additional H2 (10 sccm) for 40 min. Then, CH4 gas (0.1 sccm) was introduced as the carbon source for graphene growth. Finally, the system was cooled naturally with Ar (500 sccm) and H2 (10 sccm).

Preparation of Cu NWs

In a typical synthesis of Cu NWs32, CuCl2·2H2O (680 mg, 4 mmol) and glucose (792 mg, 4 mmol) were dissolved in 80 ml of deionized water by stirring in a beaker. In another beaker, 8 ml oleylamine and 80 μl oleic acid were dissolved in ethanol by stirring. Afterwards, these two solutions were put into a 500-ml beaker and diluted to 400 ml with water, followed by magnetic stirring for 16 h at 50 °C. After the colour of the mixture changed from blue to cyanish grey, the mixture was transferred to a stainless-steel autoclave and then heated from room temperature to 120 °C in 20 min. After heating at 120 °C for 6 h, the mixture was cooled to room temperature. The upper yellow suspension was discarded and the bottom reddish-brown Cu NWs were obtained. The Cu NWs were then centrifuged and washed with a 1:1 ethanol:water mixture (100 ml, 8,000 rpm, 2 min) and pure ethanol. Finally, the purified Cu NWs were dispersed into 100 ml ethanol to obtain a Cu–nanowire dispersion with a content of ~1 mg ml−1.

Preparation of Cu NWs-FA and Cu NWs-FA/DT

In a typical synthesis, 10 mg Cu NWs, 0.1 mg Cu(HCOO)2·4H2O and 200 mg HCOONa in 1 ml H2O were dissolved in 12 ml EG by sonication to form a homogeneous solution. Then, 1 ml oleylamine was added into the mixture and sonicated for another 3 min. The mixture was sealed in a 50-ml stainless-steel autoclave, and was heated from room temperature to 120 °C in 20 min and kept at 120 °C for 12 h to prepare Cu NWs-FA. Following the same protocol, when the EG was replaced by 12 ml DMF, Cu NWs-FA were also readily prepared by heating the mixture from room temperature to 160 °C in 20 min and then maintaining it at 160 °C for 16 h. The Cu NWs-FA products were collected and rinsed three times with ethanol and deionized water alternately to remove free HCOONa and oleylamine. The as-prepared Cu NWs-FA was dried in vacuum before use. To prepare Cu NWs-FA/DT, 10 mg of the as-prepared Cu NWs-FA was dispersed in 20 ml EtOH containing 1 mM DT by magnetic stirring for 10 min and was then centrifuged with a 1:1 ethanol:water mixture (100 ml, 8,000 rpm, 2 min) and pure ethanol.

Preparation of Cu nanoparticles (Cu NPs)

Cu NPs were synthesized using a previously reported method33. 11.1 g PVP (K30) and 4 g sodium hypophosphite were mixed into 40 ml EG inside a round-bottom flask while vigorously stirring at room temperature under ambient atmosphere. The mixture was heated to 90 °C at a rate of 5 °C min−1. Then, 10 ml of a 1 M solution of copper sulfate (CuSO4·5H2O) in EG at 90 °C was rapidly added into the PVP–sodium hypophosphite solution while stirring vigorously. As reduction occurred, the colour of the suspension turned from green to reddish brown within 2 min, indicating the formation of Cu NPs. The reaction was quenched, and the suspension was rapidly cooled by adding chilled deionized water. The Cu NPs were separated and washed with deionized water by centrifugation, and acetone was used as a non-solvent to remove excess PVP and side products. The resulting precipitates were dried under vacuum at 60 °C for 2–3 h before formate decoration.

Preparation of FA-treated Cu NPs (Cu NPs-FA)

In a typical synthesis, 10 mg Cu NPs, 0.1 mg Cu(HCOO)2·4H2O and 200 mg HCOONa in 1 ml H2O were dissolved in 10 ml EG by sonication to form a homogeneous solution. Then, 1 ml oleylamine was added into the mixture and sonicated for another 3 min. The mixture was sealed in a 50-ml stainless-steel autoclave, and was then heated from room temperature to 120 °C in 20 min and kept at 120 °C for 12 h to prepare Cu NPs-FA. The Cu NPs-FA products were collected and rinsed three times with ethanol and deionized water alternately to remove free HCOONa and oleylamine. The as-prepared Cu NPs-FA was dried in vacuum before use.

Preparation of formate-coated copper foil by the electrochemical method (Cu foil-FA-EC)

The Cu foil-FA-EC was investigated on a CHI 760E electrochemical workstation in a three-electrode configuration cell using polished Cu foil as the working electrode, a platinum plate (1 × 1 cm2) as the counter electrode and an SCE as the reference electrode in 1% HCOONa/NaOH aqueous electrolyte (pH 8–10), whereas the active area was 1 cm2. Cyclic voltammetry curves were collected in the potential window of −800 mV to 200 mV at a scan rate of 10 mV s−1. Electrochemical reconstruction was achieved by the chronoamperometry method, in which the applied potential is the redox potential of Cu(i) to Cu(0).

Electrochemical corrosion measurements

To quantitatively characterize the corrosion rates of Cu foils treated with different methods, electrochemical measurements were conducted with a standard three-electrode configuration in a 0.1 M NaOH or 1 M NaOH (Sigma Aldrich) solution at room temperature. The following Cu samples were compared: bare Cu foil, Cu-FA, Cu-FA/DT, Cu-BTA and Cu-DT. In all measurements, Cu foils with an area of 2 cm2 were used as the working electrode, a Pt wire was used as the counter electrode and Ag/AgCl was used as the reference electrode. All electrochemical measurements were carried out using a potentiostat with a voltage sweep rate of 20 mV s−1. The corrosion rates (CR) of the samples were determined by the kinetics of both anodic and cathodic reactions34 and calculated from the corrosion current density, Jcorr, as:

$${\rm{CR}}=\frac{{J}_{{\rm{corr}}}\times K\times {\rm{EW}}}{\rho },$$
(1)

where K = 3,272 mm A−1 cm−1 yr−1 is the corrosion rate constant, and EW and ρ are the equivalent weight and mass density of the corroding species, respectively.

Electrochemical imaging with scanning reference electrode

Potential variation images of the sample surface were recorded by an integrated SRET/STM system (Xiamen Le Gang Materials Technology Co. Ltd, China)35,36, which can simultaneously map the potential or current distribution and the topography on metal surfaces. The SRET/STM operates at SRET and STM scanning modes using a reference Pt–Ir probe and a local reference Pt–Ir probe, respectively. The sample is connected to the ground, the reference probe measures the average electrochemical potential (Eaverage) in bulk solution, whereas the local reference probe detects the local electrochemical potential (Elocal) near the sample surface, which is strongly influenced by local electrochemical reactions taking place on the sample surface. In SRET mode, the recorded potential signal is the potential difference ∆E between the reference and local reference probes, ∆E = Eaverage − Elocal. The two signals (Elocal and Eaverage) are fed into a differential electrometer to generate the resulting potential signal (∆E) to construct three-dimensional current images. As a result, local anodes have a high ∆E in the potential or current images.

Stability tests under different environments

Stability in alkaline solutions: for the alkaline-resistance test, the Cu materials were immersed in 0.1 M NaOH solutions at room temperature for different intervals. For the corrosion rate evaluation, the Cu materials were immersed in 0.1 M NaOH solutions under different voltages for electrochemical tests. Stability in salt spray conditions: the Cu materials were placed in a HD-E808-60 Salt Spray Test Chamber (Haida Instruments Co. Ltd, China) with 5% NaCl and relative humidity >100% at 47 °C. Stability in Na2S solution: the Cu materials were immersed in Na2S solutions of different concentrations (1 mM to 1 M) at room temperature for 0–120 min for the Na2S-resistance test. Stability in H2O2 solutions: the Cu materials were immersed in H2O2 of different concentrations for the anti-oxidation evaluation. Stability in air at 80 °C and 80% relative humidity: freshly made samples were placed in air at 80 °C and relative humidity of (80 ± 5)%. Stability under wet mechanical conditions: salty water (3.5% NaCl, 1% Na2CO3, 1% Na2SO4 and 0.1% NaOH) was continuously supplied through the tubing samples (inner diameter 1.6 cm) at a flow rate of 1,400 l h−1 or intermittently through foil samples at an overall flow rate of 400–800 l h−1.

Characterizations

Raman spectra were obtained on XploRA (Jobin Yvon-Horiba, France) confocal Raman microscopes with a dark-field function. The excitation wavelength was 532 nm from a He–Ne laser and the power on the sample was about 1 mW. We used a 50× magnification long-working-distance (8 mm) objective to focus the laser onto the sample and collect the backscattered light.

Optical microscope images of the surfaces of different Cu samples were recorded on an Olympus BH2-UMA optical microscope in reflectance mode with a Moticam 2000 2.0 M pixel camera, on a Nikon Eclipse Ti–U optical microscope in reflectance mode with an Ample Scientific 3.0 M pixel camera or on the confocal Raman microscopes.

SEM measurements were carried out on Zeiss SIGMA microscope at an accelerating voltage of 15 kV. The powder XRD experiments were conducted on a Rigaku Ultima IV XRD system using Cu Kα radiation. The operation voltage and current were 40 kV and 30 mA, respectively. The scanning speed was 10° min−1. In order to obtain the surface reconstruction information, a grazing-incidence XRD mode was used.

TEM studies were performed on a TECNAI F30 TEM operating at 300 kV. The samples were prepared by dropping an ethanol dispersion of the samples onto 300-mesh carbon-coated copper grids and immediately evaporating the solvent. High resolution-TEM and high resolution-STEM were performed on an atomic-resolution analytical microscope (JEM-ARM 200F) operating at 200 kV. The Cu NWs and Cu NWs-FA/DT were prepared by dropping an ethanol or cyclohexane dispersion of the samples onto 300-mesh carbon-coated copper grids and immediately evaporating the solvent. To further confirm the surface reconstruction, the focused-ion-beam technique was used to prepare the cross-sectional TEM samples of the Cu foil surfaces before and after the formate treatment for characterizing their surface layer structures using JEM-ARM 200F.

The infrared spectra of the solid samples were recorded from 4,000 cm−1 to 650 cm−1 on a Nicolet iS10 FTIR spectrometer (Thermo Scientific Corporation) in the attenuated total reflection–FTIR mode.

XPS and X-ray-induced Auger electron spectroscopy analyses were performed using a Thermo Fisher Scientific ESCALAB 250Xi spectrometer with focused monochromatic Al Kα radiation (1,486.6 eV; 150 W; 500 μm diameter of irradiated area). A hemispherical electrostatic analyser and a standard lens were used to maximize the signal. The basic pressure inside the analysis chamber was about 5 × 10−10 mbar. The takeoff angle between the direction of the analyser and the Cu foil was ±45°. The data were averaged over more than 10 scans during the measurement and were used without further smoothing. Peak fitting was processed using CasaXPS software. A Shirley-type background subtraction was performed before curve fitting, and the experimental curve for the Cu 2p3/2 peak fitting was set to a mixed Gaussian and Lorentzian function (defined in CasaXPS as GL(90))37.

TPD-MS was performed on a home-built TPD time-of-flight (TOF) analyser. 10.0 mg of Cu NWs or 200 mg of Cu wire, foil or mesh were pyrolyzed in a small tube heated by a heating coil. A K-type thermocouple was put inside the sample tube and insulated from the samples to measure the temperature. The heating coil was powered by a precise electric source and adjusted in intervals of 10 mV. The temperature of the sample tube was ramped smoothly from room temperature to 800 °C at a rate of 5 °C min−1, controlled by a computer. The desorbed species were ionized by an ultraviolet lamp positioned very close to the sample tube with a photon energy of 10.6 eV, and then transferred to the TOF analyser by an ion optical system. The TOF analyser had a resolution of over 5,000 and a sensitivity of the order of parts per billion. All these steps were performed in high vacuum (~3 × 10−5 Pa). The mass spectrum and sample temperature were recorded every second. Each spectrum is an accumulation of 10,000 spectra gathered at intervals of 100 μs.

The resistivity of the Cu samples was measured on a CHI 760E electrochemical workstation. The resistances of the copper foils were measured using a four-probe method (ST2263, Suzhou Jingge Electronic Co. Ltd). For the Cu wires, the electric conductivity (σ) is calculated according to the equation σL/(R × Sa), where L is the length of the Cu wire, Sa represents the cross-sectional area and R is the resistance. For the copper foil or films, the square resistance was acquired and σ = 1/(sheet resistance × thickness of the foils or films).

The thermal diffusivity (α; in m2 s−1) of the Cu foil was measured on an LFA Nanoflash 467 Light flash system. The in-plane thermal diffusivities were measured for the Cu foil and Cu foil-FA samples before and after a corrosion test with 0.1 M NaOH. The thermal conductivity (κ) is calculated according to the equation κ = α × ρ × Cp, with apparent density ρ = 8.96 g cm−3 and specific heat capacity Cp = 0.39 × 103 J kg−1 K−1.

The contact angles of different Cu samples were measured using a Ramé–Hart M500 digital goniometer equipped with a dispensing needle (VICI Precision Sampling, CA, USA). A 2-μl water droplet was generated using the automatic dispenser of the goniometer. The sessile droplet was formed by fixing the needle and approaching the substrate parallel to the needle direction with a gentle feed rate of a few micrometres per minute. All the tests were carried out in air at room temperature. The axisymmetric drop-shape analysis profile method was used to estimate the contact angle of the water droplet on the solid surface.

STM and AFM measurements

STM/AFM experiments were performed with a combined AFM/STM system (Createc, Germany) at 5 K with base pressure <5 × 10−9 Pa. Electrochemically etched Pt–Ir tips were sharpened by a focused ion beam and cleaned by alternate annealing and sputtering before the experiments, and further controlled by field-emission and voltage-pulse procedures during the scanning. Before the STM investigations, the samples were annealed in vacuum by heating up to 300 °C for an extended time (typically longer than 10 h) in a pre-chamber with base pressure <1.5 × 10−8 Pa. The Cu(110) single-crystal sample with the c(6 × 2) structure was obtained in UHV with the following procedure. 1) The Cu(110) single crystal was cleaned by five sputtering–annealing cycles. 2) The clean Cu(110) sample was quickly transferred from the UHV chamber to a load-lock chamber protected by N2 gas, where a droplet of ultrapure sodium formate solution was dripped onto the Cu(110) surface. 3) The Cu(110) sample covered by formate solution was transferred back to the UHV chamber, followed by low-temperature annealing (100 °C–300 °C) to obtain the c(6 × 2) structure. Bias voltage refers to the sample voltage with respect to the tip. All the STM topographic images were obtained in constant-current mode and the AFM images in constant-height mode.

Simulations of AFM images

The Δf images were simulated with a molecular mechanics model including the electrostatic force, according to the methods described in refs. 38,39. We used the following parameters in the flexible probe–particle tip model: effective lateral stiffness k = 0.25 N m−1 and effective atomic radius Rc = 1.748 Å. The parameters of the Lennard–Jones pairwise potential for all elements were taken from the force field developed by Zhao et al.40.

Quantification of Cu(ii) ions

The amount of Cu(ii) ions was quantified by the standard sodium diethydlthiocabamate spectrophotometric method (China Standard HJ 485-2009). In ammonia solution (pH 8-10), Cu(ii) reacts with DDTC to produce a yellow–brown complex. The Cu(ii)-DDTC complex can be extracted using CCl4. The absorbance is measured at a wavelength of 440 nm and the colour is stabilized for 1 h. Iron, manganese, nickel and cobalt react with DDTC to form coloured complexes, which can be masked by EDTA-ammonium citrate solution to eliminate interference.

Computational details

Spin-polarization calculations were carried out at the PBE level using the Vienna ab initio simulation package (VASP 5.3.5). The valence electrons were described by plane-wave basis sets with a cutoff energy of 400 eV, and the core electrons were replaced by projector-augmented wave pseudopotentials41,42,43,44,45.

It should be noted that different Cu(110) structures modified with formate species have been extensively investigated using both experiments and calculations23,24,25,46,47,48. Although the formate species that preferentially adsorb on Cu(110) have been well documented, the formate coordination inducing the reconstruction of Cu surfaces into Cu(110) has not been studied. In our DFT simulations, we adopted the c(6 × 2) structure with dinuclear [Cu(μ-HCOO)(OH)2]2 units on the Cu(110) surface. Accordingly, a (6 × 2) supercell with five-layer-thickness slabs was built. The vacuum regions between the slabs were set as 15 Å and k-point sampling was performed following the Monkhorst–Pack procedure with a 3 × 2 × 1 mesh49. During structural optimization, the bottom two layers of the slab were fixed at a bulk truncated position while the top three layers and the adsorbates were allowed to be fully relaxed. All internal structural parameters were allowed to relax until the Hellman–Feynman forces on each ion were lower than 0.02 eV Å−1.

The adsorption energies (ΔEads) were calculated using equation (2), where Ead/sub, Ead and Esub are the total energies of the optimized adsorbate/substrate system, the adsorbate and the clean substrate, respectively.

$$\Delta {E}_{{\rm{ads}}}=\Delta {E}_{{\rm{ad}}/{\rm{sub}}}-\Delta {E}_{{\rm{ad}}}-\Delta {E}_{{\rm{sub}}}.$$
(2)

Because the triplet O2 was poorly described by generalized gradient approximation functionals, we used gas-phase H2O and H2 as references to estimate the total energy of O2 using equation (3).

$$E({{\rm{O}}}_{2})=2E({{\rm{H}}}_{2}{\rm{O}})-2E({{\rm{H}}}_{2})-\Delta {E}_{{\rm{r}}}.$$
(3)

Here, ΔEr denotes the reaction heat at 0 K without a zero-point energy correction, which can be deduced from the experimental atomic energies of O2, H2 and H2O using equation (4).

$$\Delta {E}_{{\rm{r}}}=2{\rm{AE}}({{\rm{H}}}_{2}{\rm{O}})-2{\rm{AE}}({{\rm{H}}}_{2})-{\rm{AE}}({{\rm{O}}}_{2}),$$
(4)

where AE(H2O), AE(H2) and AE(O2) represent the atomic energies of H2O, H2 and O2, respectively. Our calculations showed that the corrected total energy of O2 was −9.41 eV.

Here, we neglected the zero-point energy and thermal correction, and used equation (5) to estimate the Gibbs free energy of O2 adsorption at 298 K (ΔGads). We assumed that adsorption would lose translational, rotational and vibrational entropies50 and adopted the experimental standard entropy (ΔS) of gaseous O2 (205.15 J K−1 mol−1).

$$\Delta {G}_{{\rm{ads}}}=\Delta {H}_{{\rm{ads}}}-T\Delta S\approx \Delta {E}_{{\rm{ads}}}+T\Delta S({{\rm{O}}}_{2},{\rm{gas}}).$$
(5)

Similarly, we estimate the adsorption free energy of Cl at 298 K as

$$\Delta {G}_{{\rm{a}}{\rm{d}}{\rm{s}}}^{{\prime} }\approx \Delta {E}_{{\rm{a}}{\rm{d}}{\rm{s}}}-{\rm{E}}{\rm{A}}-\varPhi +T\Delta S({{\rm{C}}{\rm{l}}}^{-},{\rm{g}}{\rm{a}}{\rm{s}})-\Delta {G}_{{\rm{s}}{\rm{o}}{\rm{l}}}({{\rm{C}}{\rm{l}}}^{-}),$$
(6)

where ΔEads(Cl) is the adsorption energy of the gaseous Cl atom, taken from DFT calculations; EA, Φ and ΔGsol(Cl) represent the electron affinity of Cl atom (−3.61 eV), the work function of Cu(110) (−4.48 eV) and the solvation free energy of Cl (−3.23 eV), respectively, taken from experimental values51. Considering that the gaseous Cl has only three-dimensional translational freedom, ΔS (Cl, gas) could be calculated using the Sackur–Tetrode formula50,52,

$$\frac{S}{\,{R}_{{\rm{g}}}}=\frac{5}{2}+\,\mathrm{ln}\,\left[\frac{V}{N}{\left(\frac{2{\rm{\pi }}m{k}_{{\rm{B}}}T}{{h}^{2}}\right)}^{3/2}\right],$$
(7)

where m is the molecular mass of Cl, V/N is the volume per molecule in the standard state, kB is the Boltzmann constant, Rg is the gas constant and h is the Planck constant. From equation (7), the translational entropy of Cl in the gas phase was calculated to be 153 J mol−1 K−1.

To identify the oxidation states of surface Cu, we carried out Bader charge analysis. In terms of molecular orbital theory, the wavefunction belongs to the whole system, and not to individual atoms. Bader’s theory of ‘atoms in molecules’53 provides a reliable scheme to partition the electronic charge density between the atoms according to zero-flux surfaces. It should be noted that the Bader charge is not the same as the oxidation state, but there are a few situations in which the two are correlated. To determine the oxidation state of an element (such as Cu), the Bader charges of a series of oxides of the element in different oxidation states (such as metallic Cu, Cu2O and CuO) are calculated first and used as references to calibrate the oxidation state of the element in an unknown system54,55.

Transparent conductive films based on Cu NWs

To make a transparent conductive thin film, a dilute suspension of Cu NWs in toluene was produced using sonication. A thin film was then fabricated by filtering down the nanowire suspensions onto a nitrocellulose porous membrane (pore size 220 nm) under vacuum. The nanowire network was transferred to a transparent substrate (glass or PET) by applying pressure to the back side of the membrane and forcing intimate contact with the substrate. The thin film was then annealed under forming gas (5% H2 and 95% Ar) at 160 °C for 60 min to improve junction contact before measurements.

Background substrate transmittances were subtracted from all the data. The characteristic transmittance and haze factors were acquired at a wavelength of 550 nm. The haze measurements were carried out using the D1003-13 standard. Four transmittance scans of a sample with different configurations were acquired for the haze calculations: T1, incident light; T2, total light transmitted by the specimen; T3, light scattered by the instrument; and T4, light scattered by the instrument and specimen. The haze factor of a specimen can be calculated by the equation haze = [(T4/T2) − (T3/T1)]%. A comparison of the performances of the transparent conductive film made of Cu NWs-FA/DT with the references is given in Extended Data Table 1.