In this study, a palladium/graphene modified stainless steel electrode was successfully prepared and applied in an electrochemical reduction device to remove Cr (VI) from the wastewater. Pd was modified onto the electrode mainly via interacting with the carboxyl group of graphene. The Cr (VI) removal efficiency was up to 99.70 ± 0.00% under the optimal condition (Pd content proportion of 3%, electrode potential of −0.9 V, pH = 2 and electrolyte concentration of 6 g/L). It was found that Cr (VI) was removed via the following processes: (1) direct electrochemical reduction by accepting electrons, (2) indirect electrochemical reduction by H2O2 that was generated from H2 in the presence of Pd, (3) adsorption through hydrogen bond, and (4) chemical reduction through alkoxy groups donating electrons. The indirect electrochemical reduction considerably promoted the Cr (VI) removal while a small amount of Cr (VI) was removed via adsorption and chemical reduction. The method could not only be used as a pretreatment technology to solve the problem of excessive Cr (VI) concentration of industrial wastewater, but also could provide reference for the electrochemical reduction of similar metal ions.

  • Pd/G-SS electrode was prepared for the electrochemical reduction of Cr (VI).

  • The Cr (VI) removal efficiency was up to 99.70% under the optimal condition.

  • Cr (VI) was removed via electrochemical reduction, adsorption and chemical reduction.

  • Cr (VI) removal was enhanced via indirect electrochemical reduction assisted by Pd.

Graphical Abstract

Graphical Abstract
Graphical Abstract

The discharge of wastewater is increasing with the accelerating process of industrialization. Industrial wastewater contains a variety of heavy metals, which cannot only pollute water bodies but also are harmful to human health. Therefore, it is necessary to give priority to treatment of industrial wastewater containing a large amount of heavy metals before discharged (Amanze et al. 2022). Chromium (Cr) is one of the important industrial raw materials, widely used in electricity plating, tanning, pigment manufacturing, metallurgy and mining, etc. (Wang et al. 2020), which results in the discharge of a large amount of Cr-containing wastewater. It was reported that 15% Cr(VI) in the metallurgical industry is difficult to be utilized, which enters into aquatic environment easily. In addition, a large amount of wastewater discharged from the tannery industry is also one of the main sources of Cr-containing wastewater. These wastewaters with heavy metal pollution characteristics currently lack a unified management program (Bhunia et al. 2022). The content of Cr ranks 24th in the earth's crust and can be represented from all valence states from −2 to +6 (Sinha et al. 2022). Among these forms, hexavalent chromium (Cr(VI)) and trivalent chromium (Cr(III)) can exist stably in water environment (Wang et al. 2020). Cr(VI) can cause respiratory damage, digestive damage and skin irritation in humans. Furthermore, Cr(VI) will affect the growth of plants and aquatic organisms (He et al. 2020). Therefore, 100 μg/L Cr(VI) was set as the concentration limit of industrial effluents discharged into groundwater by the United States Environmental Protection Agency (USEPA) (Monga et al. 2022). In contrast, Cr(III) is a trace element, which has a good effect on the prevention of diabetes and high blood pressure. And the toxicity of Cr(III) is approximately 100-fold less than Cr(VI) (Lyu et al. 2021). The most effective method for the removal of Cr(VI) contamination is reducing Cr(VI) to Cr(III).

A variety of sophisticated techniques, including biological, physical and chemical methods, have been formulated for the elimination of Cr (VI) from wastewater. The biological methods mainly rely on the characteristics of the organism itself to adsorb or reduce Cr (VI). At present, it has been found that many kinds bacteria (Sun et al. 2021b), fungi (Shi et al. 2018), plants (Saranya et al. 2022), and etc. have the ability to remove Cr(VI). However, biological methods take a long time to cultivate organisms and do not work efficiently. Physical removal is mainly based on adsorption (Kekes et al. 2021), membrane separation (Roy Choudhury et al. 2018) and ion exchange (Chen & Liu 2020). The removal of Cr(VI) by adsorption is relatively fast, but the addition of adsorbents can easily cause secondary pollution (Patel & Parikh 2020). And membranes and ion exchange resins are easily contaminated and the price of their replacement is expensive. Chemical methods mainly refer to chemical precipitation, electrocoagulation and electrochemical reduction methods. In the process of chemical precipitation, ferrous salt or sulfide is added to Cr(VI)-containing wastewater as a reducing agent. The advantage of this process is that Cr(VI) is rapidly precipitated and removed by filtration. However, the added agent may cause secondary pollution. Another chemical method is electrocoagulation (Pan et al. 2016), for instance, Al anodes are used to produce aluminum cations that have the same effect as the addition of Al based coagulants in conventional treatment systems (Jin et al. 2016b). The electrodes must be replaced regularly due to the corrosion (Sandi et al. 2020), which undoubtedly complicates the process. Electrochemical reduction has various advantages over other methods. It is robust, easy to operate and flexible in case of fluctuating wastewater streams (Muddemann et al. 2019). The electrochemical reduction of Cr(VI) is accomplished under an applied electric field, in which Cr(VI) is migrated to the cathode and reduced to Cr(III) (Chen et al. 2022) (Equation (1)). This method has been of particular interest as it does not require the addition of external reducing agents, thereby minimizing cost, and opportunities exist to increase activity and selectivity through electrode modifications (Stern et al. 2021).
(1)

The electrode material is crucial to the removal efficiency of an electrochemical process (Zhang et al. 2019). Conventional electrode materials are generally selected from carbon materials, stainless steel and other metals that are not very reactive (Stern et al. 2020; Villalobos-Neri et al. 2021). The metal electrodes have excellent electrical conductivity, which is beneficial for electron transfer in the reaction, but their small specific surface area limits their contact with Cr(VI). Carbon materials are often incorporated into the electrodes due to their large specific surface area, high conductivity and outstanding electrolyte accessibility (Wang et al. 2019). Graphite felt, carbon fiber felt and carbon aerogel have been used as work electrodes for electrochemical Cr(VI) reduction (Rana-Madaria et al. 2005; Huang et al. 2014; Chen et al. 2021). However, the strength of carbon materials is not high, which is not beneficial to the processing and repeated use of electrodes. At present, stainless steel/carbon composite electrodes are the development direction of new electrode materials.

Graphene is the name given to a single layer of carbon atoms densely packed into a benzene-ring structure. Individual graphene sheets have extraordinary electronic transport properties (Sun et al. 2022). With high aspect ratio (the ratio of lateral size to thickness), large surface, rich electronic states, and good mechanical properties, graphene is a suitable electrode candidate outperforming other carbon materials (Arvas et al. 2022). And graphene exhibits good adsorption (Ham et al. 2018) and catalytic (Wang et al. 2016) properties in wastewater treatment. For instance, compared with the stainless steel (SS) electrode, the graphene-modified stainless steel electrode had much lower charge transfer resistance and more oxygen-containing functional groups, which can promote the electrochemical reduction of Cr(VI) (Chen et al. 2022). Chemical vapour deposition (Zou et al. 2022) and chemical reduction are common methods for the preparation of graphene.

The electrochemical reduction of Cr(VI) is often accompanied by a hydrogen evolution reaction (HER) (Li et al. 2021). Pd, a kind of active metal, can store the H2 produced by HER on the electrode surface (Granja-DelRío et al. 2021) and catalyze the stored H2 to form H2O2 (Tian et al. 2020). H2O2 can reduce Cr(VI) in the aqueous solution under acidic conditions and in the presence of external power supply (Jialiang et al. 2021). However, agglomeration may occur if Pd nanoparticles are added to the aqueous solution alone (Hu et al. 2020), which make the catalytic sites of Pd cannot be fully exposed. Attaching Pd to graphene can effectively solve the above problems.

In this study, SS electrode was modified with palladium and graphene to fabricate a palladium/graphene-modified stainless steel (Pd/G-SS) electrode. With this electrode as the work electrode, the electrochemical Cr(VI) reduction under different conditions was investigated and the reaction mechanism for the electrochemical Cr(VI) reduction was explored via various trails.

Chemicals and materials

The preparation of the chemicals does not require additional purification. The natural graphite powder (C, ∼99.5%) was obtained from Jinrilai Graphite Company, Qingdao. The palladium chloride (PdCl2, >59% for Pd) powder, N-methyl pyrrolidone (C5H9NO, >99.5%), acetone (C3H6O, >99.5%), and ethanol (C2H6O, >95%), potassium dichromate (K2Cr2O7, >99.8%), sulfuric acid (H2SO4, 95% ∼ 98%), sodium sulfate (Na2SO4, >99.0%) and other chemicals (analytically pure) were purchased from Sinopharm Chemical Reagent Co., Ltd. Polyvinylidene fluoride ((C2F2)n, PVDF) was supplied by Shanghai Macklin Biochemical Co., Ltd. The 304 SS mesh (60 mesh) was provided by Wuxin Company, Changzhou.

Preparation of Pd/G-SS electrode

Pd powder was obtained by reducing divalent Pd (Dai et al. 2014). After graphene oxide was prepared by oxidizing natural graphite powder according to the improved Hummers method (Zomorodkia et al. 2021), reduced graphene oxide (termed as graphene in this text) was synthesized via hydrothermal reduction (Serrapede et al. 2020). The palladium/graphene (Pd/G) slurries with various Pd content proportions were prepared as follows: (1) Pd powder with various weight (0, 1 mg, 3 mg, 5 mg or 7 mg) and graphene with fixed weight (100 mg) were sequentially added into 5 mL N-methyl pyrrolidone, followed by dispersing evenly through magnetic stirring; (2) 100 mg PVDF was dispersed evenly into 5 mL N-methyl pyrrolidone; (3) the above two solutions were mixed to obtain five Pd/G slurries with different Pd content proportions (0, 1, 3, 5 and 7%). The Pd/G-SS electrodes were prepared by adhering Pd/G slurries to the SS meshes through adhesive PVDF. The preparation process is as follows: (1) the 304 SS mesh (60 mesh) was cut into 20 mm × 45 mm, then rinsed with acetone, ethanol and DI water in an ultra-sonication, and finally dried in the nitrogen gas; (2) SS was soaked in the Pd/G slurries for 30 s, then dried at 80 °C for 2 h; (3) by repeating the second step three times, electrodes with a fixed graphene loading (1.50 mg/cm2) and varied Pd content proportions (0, 1, 3, 5 and 7%) can be fabricated. Among these electrodes, the Pd/G-SS electrode with a Pd content proportion of 0 was termed as the graphene-modified stainless steel (G-SS) electrode in this text.

Test device

The test device was a cuboid reactor (Figure 1), made of plexiglass. A conventional potentiostat (CHI630E, CH Instruments Inc. USA) was used, with a Pd/G-SS electrode (20 mm × 30 mm), a graphite electrode (20 mm × 30 mm) and an Ag/AgCl electrode (0.197 V vs. standard hydrogen electrode) as the work electrode, counter electrode, reference electrode, respectively. The three electrodes were respectively connected to the three jacks of the potentiostat and the test device was placed on a magnetic stirrer with a constant speed of 800 r·min−1. All potentials were reported vs. Ag/AgCl electrode (unless otherwise specified).
Figure 1

Schematic of test device for electrochemical Cr(VI) reduction.

Figure 1

Schematic of test device for electrochemical Cr(VI) reduction.

Close modal

Experimental procedure

Synthetic Cr(VI)-containing wastewater was prepared by dissolving K2Cr2O7 in DI water. The pH was adjusted by adding H2SO4 and Na2SO4 solution was used as the electrolyte. Unless otherwise specified, the test device was run with the following parameters: Cr(VI) concentration of 10.0 mg/L, pH 2, Na2SO4 concentration of 10 g/L, electrode potential of −0.6 V and Pd/G-SS electrode with Pd content proportion of 3%. The system was in turn run at different Pd content proportions (0, 1, 3, 5 and 7%), different electrode potentials (−0.6 V, −0.7 V, −0.8 V, −0.9 V and −1.0 V), different pH (1, 1.5, 2, 2.5 and 3), and different electrolyte concentrations (2 g/L, 6 g/L, 10 g/L and 14 g/L). Under each condition, the test was repeated at least in triplicate and the Cr(VI) concentration was measured every 20 min.

Under the optimum working conditions, adsorption tests without external power supply and electrochemical Cr(VI) reduction tests were successively carried out to analyze the reaction mechanism. The Cr(VI) adsorption percent was examined by comparing the residual concentration after the reaction with the initial concentration. Fourier transform infrared radiation (FTIR) spectroscopy and X-ray photoelectron spectroscopy (XPS) were used to examine the change of functional groups and the valence states of Cr. The electrochemical performance of Pd/G-SS electrode was tested by cyclic voltammetry (CV) and electrochemical impedance spectroscopy (EIS) tests.

Analysis and calculations

Cr(VI) concentration was measured by the diphenylcarbazide spectrophotometry. H2O2 concentration was measured by the Titanium Potassium Oxalate Spectrophotometry. The functional groups were analyzed with a FTIR spectrometer (VERTEX 70, Bruker, Germany, 400–4,000 cm−1). The micromorphology of graphene and Pd/G slurry were observed by a transmission electron microscopy (TEM) (Tecnai G220, FEI Company, Netherlands). XPS (AXIS-ULTRADLD-600W, Kratos Company, UK) was performed with an Al X-ray monochromatic source. CV test (2.5 mV·s−1, negative scan first and then positive scan) and EIS test (frequency range: 0.1–10 kHz, amplitude: 5 mV) were performed using a potentiostat mentioned above.

Characterization of graphene and Pd/G slurry

In FTIR spectra of graphene and Pd/G slurry, four infrared peaks appeared at the wavenumbers of 1,058 cm−1, 1,400 cm−1, 1,640 cm−1and 3,440 cm−1 (Figure 2). After graphene was mixed with Pd, the infrared peaks at 1,400 cm−1, 1,640 cm−1 and 3,440 cm−1, respectively corresponding to C-OH stretching vibrations (Ren et al. 2018), C = O stretching vibration in –COOH (Yavari et al. 2022) and O-H stretching vibrations in -COOH (Tadjenant et al. 2020), decreased considerably. The infrared peaks at 1,058 cm−1, corresponding to C-O in alkoxy groups (Tadjenant et al. 2020), did not change obviously. The above results indicate that the Pd mainly interacted with the carboxyl group of graphene in the Pd/G slurry.
Figure 2

FTIR spectra of graphene and Pd/G slurry.

Figure 2

FTIR spectra of graphene and Pd/G slurry.

Close modal
As shown in Figure 3(a), graphene showed smooth but wrinkled and layered structure. Pd appeared agglomeration due to the nano-sized Pd particle owning to the high interfacial force for aggregating (Figure 3(b)). When the Pd content proportion of Pd/G slurry was set at 1 and 3%, Pd nanoparticles were discretely modified onto graphene (Figure 3(c) and 3(d)). However, when the Pd content proportion was further increased to 5 and 7%, Pd nanoparticles agglomerated onto graphene (Figure 3(e) and 3(f)). The agglomeration of Pd nanoparticles will reduce its specific surface area and catalytic sites, further affecting its catalytic performance (Hu et al. 2020).
Figure 3

TEM images for graphene (a), Pd (b), Pd/G slurry with Pd content proportions of 1% (c), 3% (d), 5% (e), 7% (f), ×68,000.

Figure 3

TEM images for graphene (a), Pd (b), Pd/G slurry with Pd content proportions of 1% (c), 3% (d), 5% (e), 7% (f), ×68,000.

Close modal

Electrochemical (VI) reduction under different conditions

The introduction of Pd will increase the catalyst costs but might reduce electricity costs and infrastructure costs via shortening reaction time. In current study, it is difficult to accurately evaluate the cost due to a lab-scale device being used. Therefore, removal efficiency and reaction time, rather than cost, should be the focus when analyzing the effects of various factors. Under most condition, the Cr(VI) removal efficiency showed a fast increase within 60 min, followed by a slow increase (Figure 4). Therefore, the Cr(VI) removal efficiency within 60 min was used to analyze the effect of each factor. The Cr(VI) removal efficiency within 60 min increased when the Pd content proportion was increased from 0 to 3%, but it decreased as the Pd content proportion was further increased to 5 and 7% (Figure 4(a)). The electrochemical reduction of Cr(VI) is often accompanied by HER. Pd, as a catalyst, can promote (Li et al. 2021) and store the H2 produced by the side reaction of HER on the electrode surface (Granja-DelRío et al. 2021), and also catalyze H2O2 formation from H2 in the presence of acid and external power supply (Zhang et al. 2022). This formed H2O2 can reduce Cr(VI) in the aqueous solution (Kim et al. 2015), enhancing the Cr(VI) removal efficiency at the Pd content proportions of 1 and 3%. The decrease of Cr(VI) removal efficiency at Pd content proportions of 5 and 7% can be attributed to two reasons: (1) too violent HER consumed a lot of electrons that should be used for electrochemical reduction of Cr(VI) (Liu et al. 2018). And the H2 generated by the HER could not be completely stored and converted into H2O2, so a large number of bubbles were formed on the electrode to hinder the contact between Cr(VI) and the electrode; (2) the catalytic site of Pd was decreased due to Pd agglomerating at high Pd content proportion (Hu et al. 2020), which can also be seen from the TEM images (Figure 3). The Cr(VI) removal efficiency within 60 min reached the highest level of 89.20 ± 0.20% at the Pd content proportion of 3%. Thus, Pd/G-SS electrode with a Pd content proportion of 3% was used in subsequent experiments.
Figure 4

Cr(VI) removal efficiency at different Pd content proportions (a), electrode potentials (b), initial pH (c), and Na2SO4 concentrations (d).

Figure 4

Cr(VI) removal efficiency at different Pd content proportions (a), electrode potentials (b), initial pH (c), and Na2SO4 concentrations (d).

Close modal

The Cr(VI) removal efficiency within 60 min increased when the electrode potential was decreased from −0.6 V to −0.9 V, but it decreased as the electrode potential was further decreased to −1.0 V (Figure 4(b)). The decrease of electrode potential from −0.6 V to −0.9 V promoted the HER and further promoted the generation of H2O2, resulting in the increase of Cr(VI) removal efficiency (Kim et al. 2015). The decrease of Cr(VI) removal efficiency when the electrode potential being dropped to −1.0 V can be attributed to the violent HER, which was similar to the case of high Pd content proportion (5 and 7%). Since the Cr(VI) removal efficiency within 60 min reached the highest level of 99.70 ± 0.30% at the electrode potential of −0.9 V, this electrode potential was used in subsequent experiments.

The Cr(VI) removal efficiency within 60 min increased obviously when the initial pH was decreased from 3.0 to 2.0, but it decreased as the initial pH was further decreased to 1.5 and 1.0 (Figure 4(c)). Low pH was beneficial to reduce the reaction overpotential of electrochemical Cr(VI) reduction, enhancing the removal of Cr(VI). Besides, both the formation of H2O2 and the reduction of Cr(VI) depended on strong acid conditions (Jin et al. 2016a; Zhang et al. 2022). Therefore, the Cr(VI) removal efficiency was extremely low at the initial pH of 3 (2.00 ± 0.30%) due to the electrochemical reduction being hindered. After the initial pH was decreased below 2.0, the reduction between Cr(VI) and H2O2 was limited by H2O2 formation due to limited catalytic sites of Pd. And the ability of Cr(VI) to obtain electrons directly from the electrode was weakened due to HER exacerbating electron consumption at low pH. As a result, the highest Cr(VI) removal efficiency of 99.70 ± 0.30% was obtained at the initial pH of 2.0, and this initial pH was used in subsequent experiments.

The Cr(VI) removal efficiency within 60 min increased when the electrolyte (Na2SO4) concentration was increased from 2 g/L to 6 g/L, but it decreased as the electrolyte concentration was further increased to 10 g/L and 14 g/L (Figure 4(d)). Increasing the electrolyte concentration can improve the electrochemical Cr(VI) reduction via reducing the solution resistance and promote the HER (Hou et al. 2012). Although the HER and the electrochemical Cr(VI) reduction competed for electrons, the electrochemical Cr(VI) reduction was improved due to the H2O2 formation in the presence of Pd on the electrode (Kim et al. 2015). However, more electrolyte cations would be attracted to the cathode with the increase of the electrolyte concentration, resulting in a difficulty for Cr(VI) species to reach the cathode (Chen et al. 2022). When the electrolyte concentration was set at 6 g/L, the Cr(VI) removal efficiency was up to 99.70 ± 0.00% and the residual Cr(VI) concentration was 0.03 ± 0.00 mg/L, and this electrolyte concentration was used in subsequent experiments.

Adsorption and chemical reduction performance of the Pd/G-SS electrode

The Cr(VI) adsorption percent increased fast to 2.80 ± 0.70% within 60 min and maintained stable later (Figure 5(a)). After absorption, the infrared peaks at 1,058 cm−1 (corresponding to C-O of alkoxy groups) and 3,440 cm−1 (corresponding to O-H of carboxyl groups) were decreased (Figure 5(b)). The consumption of alkoxy group might be attributed to it donating electrons to Cr(VI) (in the form of ) to form Cr(III) (Kwak et al. 2007). The change in O-H of carboxyl group was owing to two adsorption reactions: (1) the carboxyl group can combine with a H+ to form -COOH2+ under acidic conditions and thereby adsorb negatively charged Cr(VI) via hydrogen bond, (2) the carboxyl group can lose a H+ to form -COO and thereby adsorb the positively charged Cr(III) through electrostatic force (Liu et al. 2019). The O-H of the latter carboxyl group was destroyed after adsorbing Cr(III). As shown in Figure 5(c), two characteristic peaks at 576.6 eV and 579.1 eV (respectively representing Cr(III) and Cr(VI) (Hou et al. 2012)) were detected on the Cr 2p3/2 XPS spectra of the Pd/G-SS electrode after adsorption. This confirmed that Cr(VI) can not only be adsorbed onto the electrode but also chemically reduced to Cr(III). The relative area ratio of Cr(III) to [Cr(VI) + Cr(III)] was 21.20%, and it was estimated that approximately 21.20% of absorbed Cr(VI) was chemically reduced to Cr(III). Overall, a small amount of Cr(VI) can be removed via adsorption and chemical reduction.
Figure 5

Cr(VI) adsorption percent of the Pd/G-SS electrode during adsorption test (a), FTIR spectra (b), Cr 2p3/2 XPS spectra (c) of the Pd/G-SS electrode after adsorption.

Figure 5

Cr(VI) adsorption percent of the Pd/G-SS electrode during adsorption test (a), FTIR spectra (b), Cr 2p3/2 XPS spectra (c) of the Pd/G-SS electrode after adsorption.

Close modal

Electrochemical analysis of the Pd/G-SS electrode

There was a distinct reduction peak at −0.28 V (vs Ag/AgCl) in the CV curve of the Pd/G-SS electrode, while no obvious peak was observed for the SS electrode (Figure 6(a)). The appearance of the reduction peak represented the occurrence of electrochemical Cr(VI) reduction on the electrode. This result indicated that the Pd/G modification improved the electrochemical performance of the electrode. According to the EIS spectra (Figure 6(b)), various parameters were fitted. The Rct of the Pd/G-SS electrode was much lower than that of the G-SS electrode (113.6 Ω vs. 230.3 Ω). It was reported that the Rct of a SS electrode was higher than 500 Ω (Bansod et al. 2016). This implied that Pd/G modification considerably improved the conductivity of the electrode. Compared with the G-SS electrode, the Pd/G-SS electrode had lower CPE-P (0.3883 vs. 0.5524) and higher CPE-T (0.005875 vs. 0.002164). CPE-P represented the interface uniformity and quality (Sari et al. 2015) while CPE-T reflected the capacitance of the double layer between the electrode and the solutions (Sun et al. 2021a). This indicated that the interface between electrode and solution became rough after the Pd modification (Sun et al. 2021a). In the low frequency region, the slope of EIS spectra for the Pd/G-SS electrode was higher than that for the G-SS electrode. As a result, the fitted W1 of the Pd/G-SS electrode was higher than that of the G-SS electrode (20.4 Ω vs. 7.1 Ω). This indicated that the introduction of Pd would favor the ion diffusion. For the Pd/G-SS electrode, improvements in conductivity, interface roughness and ion diffusion would ultimately facilitate the electrochemical Cr(VI) reduction. EIS test further confirmed that the electrochemical performance of the electrode was enhanced after the Pd/G modification.
Figure 6

CV curves of the SS and Pd/G-SS electrodes (a), EIS spectra of the G-SS and Pd/G-SS electrodes and their equivalent circuit (where Rs is equivalent series resistance; Rct is charge transfer resistance; W1 is Warburg impedance and CPE is constant phase element) (b), FTIR spectra (c) and Cr 2p3/2 XPS spectra (d) of the Pd/G-SS electrode after the electrochemical reduction test.

Figure 6

CV curves of the SS and Pd/G-SS electrodes (a), EIS spectra of the G-SS and Pd/G-SS electrodes and their equivalent circuit (where Rs is equivalent series resistance; Rct is charge transfer resistance; W1 is Warburg impedance and CPE is constant phase element) (b), FTIR spectra (c) and Cr 2p3/2 XPS spectra (d) of the Pd/G-SS electrode after the electrochemical reduction test.

Close modal

After the electrochemical reduction, the infrared peak at 3,440 cm−1 (corresponding to O-H of carboxyl group (Tadjenant et al. 2020)) was further decreased as compared with the peak after adsorption (Figure 6(c)). Cr(VI) absorbed by -COOH2+ will be reduced to Cr(III) after the electrochemical reduction and -COOH2+ will lose 2 H+ to form -COO to absorb Cr(III) (Liu et al. 2019), resulting in the destruction of O-H of carboxyl groups. XPS spectra of the Pd/G-SS electrode showed that only a characteristic peak at 576.6 eV (representing Cr(III) (Hou et al. 2012)) was detected after the electrochemical reduction (Figure 6(d)). This implied that all absorbed Cr(VI) was electrochemically reduced. Besides, the Cr(VI) removal efficiency was up to 99.70 ± 0.0% in the electrochemical reduction test (section 3.2). These results implied that Cr(VI) was mainly removed via the electrochemical reduction.

Indirect electrochemical reduction performance of the Pd/G-SS electrode

As shown in Figure 7(a), H2O2 was detected when the Pd/G-SS electrode was used for the electrochemical Cr(VI) reduction test. The electrochemical Cr(VI) reduction was often accompanied by the HER (Li et al. 2021). The detection of H2O2 confirmed that the produced H2 was converted to H2O2 in the presence of Pd (Tian et al. 2020). The H2O2 concentration increased slowly to 0.70 ± 0.12 mg/L in first 60 min but increased fast to 1.40 ± 0.36 mg/L in the next 60 min. As stated in section 3.2, the Cr(VI) removal efficiency showed a fast increase within 60 min in current study. The variation in H2O2 concentration was attributed to the fact that the formed H2O2 was rapidly used for Cr(VI) reduction (Kim et al. 2015) within the first 60 min but stored in solution after most Cr(VI) was reduced. The electrochemical reduction of Cr(VI) with both G-SS and Pd/G-SS electrodes conformed to pseudo-first-order kinetics. The reaction constant for the Pd/G-SS electrode was 0.0479, 1.4-fold higher than that for the G-SS electrode (0.0342) (Figure 7(b)). As a result, the Cr(VI) removal efficiency with the Pd/G-SS electrode was always higher than that with the G-SS electrode (Figure 7(c)). These results indicated that the generated H2O2 on the Pd/G-SS electrode was used for the indirect electrochemical Cr(VI) reduction (Equation (2)) (Kim et al. 2015), thereby improving the overall removal efficiency of Cr(VI). The Cr(VI) removal efficiency with the Pd/G-SS electrode reached 99.70 ± 0.00% within 60 min while that for the G-SS electrode was 87.50 ± 1.50%. It can be estimated that approximate 12.30% of Cr(VI) was removed via the indirect electrochemical reduction.
(2)
Figure 7

H2O2 concentration (a), electrode reaction kinetics (b) and Cr(VI) removal efficiency (c) during the electrochemical reduction test.

Figure 7

H2O2 concentration (a), electrode reaction kinetics (b) and Cr(VI) removal efficiency (c) during the electrochemical reduction test.

Close modal

Electrochemical reduction method based on the Pd/G-SS electrode can effectively remove Cr(VI). Under the optimal condition with Pd content proportion of 3%, electrode potential of −0.90 V, initial pH = 2, and electrolyte concentration of 6 g/L, the Cr(VI) removal efficiency reached 99.70 ± 0.00% within 60 min and the residual Cr(VI) concentration was 0.03 ± 0.00 mg/L. The removal of Cr(VI) involved four pathways: (1) direct electrochemical reduction by accepting electrons, (2) indirect electrochemical reduction by H2O2 that was generated from H2 in the presence of Pd, (3) adsorption through hydrogen bond, and (4) chemical reduction through alkoxy groups donating electrons. The indirect electrochemical reduction considerably promoted the Cr(VI) removal while a small amount of Cr(VI) was removed via adsorption and chemical reduction. Electrochemical Cr(VI) reduction could be used as a pretreatment technology to solve the problem of excessive Cr(VI) concentration.

This work was supported by the National Natural Science Foundation of China under Grant number 21577108.

All relevant data are included in the paper or its Supplementary Information.

The authors declare there is no conflict.

Amanze
C.
,
Zheng
X.
,
Man
M.
,
Yu
Z.
,
Ai
C.
,
Wu
X.
,
Xiao
S.
,
Xia
M.
,
Yu
R.
,
Wu
X.
,
Shen
L.
,
Liu
Y.
,
Li
J.
,
Dolgor
E.
&
Zeng
W.
2022
Recovery of heavy metals from industrial wastewater using bioelectrochemical system inoculated with novel Castellaniella species
.
Environ. Res.
205
,
112467
.
Arvas
M. B.
,
Gürsu
H.
,
Gencten
M.
&
Sahin
Y.
2022
Supercapacitor applications of novel phosphorus doped graphene-based electrodes
.
J. Energy. Storage.
55
,
105766
.
Bansod
A. V.
,
Patil
A. P.
,
Moon
A. P.
&
Khobragade
N. N.
2016
Intergranular corrosion behavior of low-nickel and 304 austenitic stainless steels
.
J. Mater. Eng. Perform.
25
(
9
),
3615
3626
.
Bhunia
A.
,
Lahiri
D.
,
Nag
M.
,
Upadhye
V.
&
Pandit
S.
2022
Bacterial biofilm mediated bioremediation of hexavalent chromium: a review
.
Biocatal. Agric. Biotechnol.
43
,
102397
.
Chen
G.
&
Liu
H.
2020
Photochemical removal of hexavalent chromium and nitrate from ion-exchange brine waste using carbon-centered radicals
.
Chem. Eng. J.
396
,
125136
.
Chen
P.
,
Cheng
R.
,
Meng
G.
,
Ren
Z.
,
Xu
J.
,
Song
P.
,
Wang
H.
&
Zhang
L.
2021
Performance of the graphite felt flow-through electrode in hexavalent chromium reduction using a single-pass mode
.
J. Hazard. Mater.
416
,
125768
.
Dai
Z.
,
Tian
D.
,
Zhang
X.
,
Gong
S.
&
Yang
C.
2014
Influence of ethanol catalyzed oxidation on the activity of Pd/C with different ratio of Pd and carbon
.
Adv. Mat. Res.
860–863
,
835
838
.
Granja-DelRío
A.
,
Alducin
M.
,
Juaristi
J. I.
,
López
M. J.
&
Alonso
J. A.
2021
Absence of spillover of hydrogen adsorbed on small palladium clusters anchored to graphene vacancies
.
Appl. Surf. Sci.
559
,
149835
.
Ham
K.
,
Lee
J.
&
Electrochem Soc
I.
2018
Enhanced Capacitive Deionization of Graphene Nanoplatelet/Activated Carbon Composite Electrode
. In
Selected Proceedings Fron the 233ed Ecs Meeting
, Vol.
85
, pp.
1321
1327
.
Hou
Y.
,
Liu
H.
,
Zhao
X.
,
Qu
J.
&
Chen
J. P.
2012
Combination of electroreduction with biosorption for enhancement for removal of hexavalent chromium
.
J. Colloid. Interf. Sci.
385
,
147
153
.
Jialiang
L.
,
Xinmiao
H.
,
Jingwen
Y.
,
Yunyi
L.
,
Zhiwei
Z.
,
Yuanyuan
L.
,
Jiangyu
Y.
&
Yunmei
W.
2021
A review of the formation of Cr(VI) via Cr(III) oxidation in soils and groundwater
.
Sci. Total. Environ.
774
,
145762
.
Jin
W.
,
Du
H.
,
Yan
K.
,
Zheng
S.
&
Zhang
Y.
2016a
Improved electrochemical Cr(VI) detoxification by integrating the direct and indirect pathways
.
J. Electroanal. Chem.
775
,
325
328
.
Jin
W.
,
Du
H.
,
Zheng
S.
&
Zhang
Y.
2016b
Electrochemical processes for the environmental remediation of toxic Cr(VI): a review
.
Electrochim. Acta.
191
,
1044
1055
.
Kim
K.
,
Kim
J.
,
Bokare
A. D.
,
Choi
W.
,
Yoon
H.-I.
&
Kim
J.
2015
Enhanced removal of hexavalent chromium in the presence of H2O2 in frozen aqueous solutions
.
Environ. Sci. Technol.
49
(
18
),
10937
10944
.
Kwak
G.
,
Wang
H. Q.
,
Choi
K. H.
,
Song
K. H.
,
Kim
S. H.
,
Kim
H. S.
,
Lee
S. J.
,
Cho
H. Y.
,
Yu
E. J.
,
Lee
H. J.
,
Park
E. J.
&
Park
L. S.
2007
Poly(diphenylacetylene)s with electron-donating alkoxy and electron-withdrawing fluoroalkyl groups: effect of the substituent on fluorescence
.
Macromol. Rapid. Comm.
28
(
12
),
1317
1324
.
Li
X.
,
Wang
C.
,
Liu
F.
,
Zhang
Z.
,
Ali
J.
,
Long
Q.
&
Zhang
J.
2021
Electrocatalytic reduction of Cr(VI) over heterophase MoS2 film electrode
.
Chem. Eng. J.
404
,
126556
.
Liu
D.
,
Dai
Y.
,
Zou
J.
,
Wang
S.
,
Qi
L.
&
Ling
C.
2019
Sandwich-type ferrocene-functionalized magnetic nanoparticles: synthesis, characterization, and the adsorption of Cr(VI)
.
J. Materi. Sci-Mater. El.
30
(
15
),
13924
13932
.
Lyu
Y.
,
Yang
T.
,
Liu
H.
,
Qi
Z.
,
Li
P.
,
Shi
Z.
,
Xiang
Z.
,
Gong
D.
,
Li
N.
&
Zhang
Y.
2021
Enrichment and characterization of an effective hexavalent chromium-reducing microbial community YEM001
.
Environ. Sci. Pollut. Res. Int.
28
(
16
),
19866
19877
.
Monga
A.
,
Fulke
A. B.
&
Dasgupta
D.
2022
Recent developments in essentiality of trivalent chromium and toxicity of hexavalent chromium: implications on human health and remediation strategies
.
J. Haz. Mat. Adv.
7
,
100113
.
Muddemann
T.
,
Haupt
D.
,
Sievers
M.
&
Kunz
U.
2019
Electrochemical reactors for wastewater treatment
.
Chembioeng. Rev.
6
(
5
),
142
156
.
Pan
C.
,
Troyer
L. D.
,
Catalano
J. G.
&
Giammar
D. E.
2016
Dynamics of chromium(VI) removal from drinking water by iron electrocoagulation
.
Enviro. Sci. Technol.
50
(
24
),
13502
13510
.
Rana-Madaria
P.
,
Nagarajan
M.
,
Rajagopal
C.
&
Garg
B. S.
2005
Removal of chromium from aqueous solutions by treatment with carbon aerogel electrodes using response surface methodology
.
Ind. Eng. Chem. Res.
44
(
17
),
6549
6559
.
Roy Choudhury
P.
,
Majumdar
S.
,
Sahoo
G. C.
,
Saha
S.
&
Mondal
P.
2018
High pressure ultrafiltration CuO/hydroxyethyl cellulose composite ceramic membrane for separation of Cr (VI) and Pb (II) from contaminated water
.
Chem. Eng. J.
336
(
8
),
570
578
.
Sandi
,
Afriani
F.
&
Tiandho
Y.
2020
Application of electrocoagulation for textile wastewater treatment: A review
. In
2nd International Conference on Green Energy and Environment (Icogee 2020)
, Vol.
599
, p.
012069
.
Serrapede
M.
,
Fontana
M.
,
Gigot
A.
,
Armandi
M.
,
Biasotto
G.
,
Tresso
E.
&
Rivolo
P.
2020
A facile and green synthesis of a moo2-reduced graphene oxide aerogel for energy storage devices
.
Materials
13
(
3
),
594
.
Sinha
R.
,
Kumar
R.
,
Abhishek
K.
,
Shang
J.
,
Bhattacharya
S.
,
Sengupta
S.
,
Kumar
N.
,
Singh
R. K.
,
Mallick
J.
,
Kar
M.
&
Sharma
P.
2022
Single-step synthesis of activated magnetic biochar derived from rice husk for hexavalent chromium adsorption: equilibrium mechanism, kinetics, and thermodynamics analysis
.
Groundw. Sustain. Dev.
18
,
100796
.
Stern
C. M.
,
Hayes
D. W.
,
Kgoadi
L. O.
&
Elgrishi
N.
2020
Emerging investigator series: carbon electrodes are effective for the detection and reduction of hexavalent chromium in water
.
Environ. Sci-Wat. Res.
6
(
5
),
1256
1261
.
Stern
C. M.
,
Jegede
T. O.
,
Hulse
V. A.
&
Elgrishi
N.
2021
Electrochemical reduction of Cr(vi) in water: lessons learned from fundamental studies and applications
.
Chem. Soc. Rev.
50
(
3
),
1642
1667
.
Sun
D.
,
Frankel
G. S.
,
Brantley
W. A.
,
Heshmati
R. H.
&
Johnston
W. M.
2021a
Electrochemical impedance spectroscopy study of corrosion characteristics of palladium-silver dental alloys
.
J. Biomed. Mater. Res. B.
109
(
11
),
1777
1786
.
Sun
Y.
,
Jin
J.
,
Li
W.
,
Zhang
S.
&
Wang
F.
2021b
Hexavalent chromium removal by a resistant strain Bacillus cereus ZY-2009
.
Environ. Technol.
DOI:.
Tadjenant
Y.
,
Dokhan
N.
,
Barras
A.
,
Addad
A.
,
Jijie
R.
,
Szunerits
S.
&
Boukherroub
R.
2020
Graphene oxide chemically reduced and functionalized with KOH-PEI for efficient Cr(VI) adsorption and reduction in acidic medium
.
Chemosphere
258
,
127316
.
Tian
P.
,
Xuan
F.
,
Ding
D.
,
Sun
Y.
,
Xu
X.
,
Li
W.
,
Si
R.
,
Xu
J.
&
Han
Y.-F.
2020
Revealing the role of tellurium in palladium-tellurium catalysts for the direct synthesis of hydrogen peroxide
.
J. Catal.
385
,
21
29
.
Villalobos-Neri
E. E.
,
Paramo-Garcia
U.
,
Hernandez-Escotoz
H.
,
Mayen-Mondragon
R.
&
Gallardo-Rivas
N. V.
2021
Hexavalent chromium reduction at the polypyrrole-coated 304-stainless-steel-electrode in a filter-press-type reactor
.
Int. J. Electrochem Sci.
16
(
10
),
21103
.
Wang
Y.
,
Liu
S.
,
Li
R.
,
Huang
Y.
&
Chen
C.
2016
Electro-catalytic degradation of sulfisoxazole by using graphene anode
.
J. Environ. Sci.
43
,
54
60
.
Wang
X.
,
Liu
L.
&
Niu
Z.
2019
Carbon-based materials for lithium-ion capacitors
.
Mater. Chem. Front.
3
(
7
),
1265
1279
.
Wang
H.
,
Lei
W.
,
Li
X.
,
Huang
Z.
,
Jia
Q.
&
Zhang
H.
2020
Catalytic reductive degradation of Cr(VI)
.
Proc. Chem.
32
(
12
),
1990
2003
.
Zhang
M.
,
Shi
Q.
,
Song
X.
,
Wang
H.
&
Bian
Z.
2019
Recent electrochemical methods in electrochemical degradation of halogenated organics: a review
.
Environ. Sci. Pollut. R.
26
(
11
),
10457
10486
.
This is an Open Access article distributed under the terms of the Creative Commons Attribution Licence (CC BY 4.0), which permits copying, adaptation and redistribution, provided the original work is properly cited (http://creativecommons.org/licenses/by/4.0/).

Supplementary data