In recent years, carbon tetrachloride (CT) has been frequently detected in surface water and groundwater around the world; it is necessary to find an effective way to treat wastewater contaminated with it. In this study, Ni/Fe bimetallic nanoparticles were immobilized on reduced graphene oxide (NF@rGO), and used to dechlorinate CT in aqueous solution. Scanning electron microscopy (SEM) demonstrated that the two-dimensional structure of rGO could disperse nanoparticles commendably. The results of batch experiments showed that the 4N4F@rGO (Fe/GO = 4 wt./wt., and Ni/Fe = 4 wt.%) could reach a higher reduction capacity (143.2 mgCT/gcatalyst) compared with Ni/Fe bimetallic nanoparticles (91.7 mgCT/gcatalyst) and Fe0 nanoparticles (49.8 mgCT/gcatalyst) respectively. That benefited from the nickel metal as a co-catalyst, which could reduce the reaction activation energy of 6.59 kJ/mol, and rGO as an electrical conductivity supporting material could further reduce the reaction activation energy of 4.73 kJ/mol as presented in the conceptual model. More complete dechlorination products were generated with the use of 4N4F@rGO. Based on the above results, the reductive pathway of CT and the catalytic reaction mechanism have been discussed.

  • Ni/Fe nanoparticles were evenly immobilized on reduced graphene oxide (NF@rGO).

  • NF@rGO dramatically accelerated the hydrogenolysis pathway of carbon tetrachloride.

  • More hazardous intermediate products were reduced with the presence of NF@rGO.

  • Atomic hydrogen as a powerful reducing agent was involved in the dechlorination process.

  • The catalytic activities of two catalysts with different supporters were compared.

Graphical Abstract

Graphical Abstract
Graphical Abstract

Carbon tetrachloride (CT) was widely used in the past as a cleaning and degreasing solvent in industry due to its ability to dissolve a wide range of organic substances. The widespread use of CT led to its dispersion in the environment persistently (Doherty 2000). Due to the adverse health effects of CT on humans and animals, such as central nervous system depression, liver and kidney damage, CT has been established as a class of priority pollutants prohibited by the US EPA (Hua & Hoffmann 1996). However, chloroform (CF) as a reduction product of CT is more toxic (Semprini & McCarty 1992). Therefore, it is of great significance to develop an efficient method for transforming CT into non-hazardous products.

In general, degradation of CT via biological methods always requires quite a long time (Hashsham et al. 1995), and may have the risk of poisoning the organism in some situations (Santharam et al. 2014), which strongly restricts its application field. Boronina & Klabunde (1995) found that three zero-valent metals (magnesium, tin and zinc) can be used to reduce CT in aqueous solutions. Feng & Lim (2005) found that although Zn could reduce CT with a high reaction rate compared with Fe powder, it was hard to reduce CT to complete dechlorination products (CDCPs) during the reaction. Due to the low-cost and high efficiency, zero-valent iron (ZVI) has been widely investigated in the wastewater treatment of chlorinated organic compounds (COCs) (Gillham & Ohannesin 1994; Matheson & Tratnyek 1994; Balko & Tratnyek 1998; Cwiertny et al. 2006). Among these, nanoscale zero-valent iron (nZVI) has particularly attracted considerable attention due to its large specific surface area (Zhang 2003). However, this advantage could not be fully exploited while in practice (Liu & Lal 2012), such as nZVI easily formed a magnetic passive layer which blocked the reactive sites and resulted in aggregation of the nanoparticles (NPs) (Zhang et al. 2002). Therefore, it is highly required to develop a new composite catalyst to improve the catalytic activity of nZVI.

Recently, various support materials have been developed to prevent particle aggregation (e.g. MWCNTs, activated carbon, SiO2, carboxymethylcellulose, Fe3O4) (Kustov et al. 2014; Jin et al. 2018; Lv et al. 2018; Wu et al. 2018; Chen et al. 2019). Beyond these, graphene as a supporting material holds excellent mechanical, electrical and thermal characteristics, which is superior to other traditional materials (Dreyer et al. 2009). In order to obtain a similar structure to pristine graphene, reduced graphene oxide (rGO) was synthesized from graphite flake via a chemical method. It has some excellent characteristics: (1) the high specific surface area of the two-dimensional structure could allow nanoparticle catalyst to be efficiently loaded on it (Bai & Shen 2012), (2) the conjugated structure of graphene could enhance the adsorption capacity of the reactant molecules (Zhang et al. 2012), (3) the oxygen-containing groups could make the layers hydrophilic (Pei & Cheng 2012), (4) the characteristics of electrical conductivity were favorable during the catalytic reaction (Stankovich et al. 2007).

Deposition of a second metal (e.g. Ni, Pd, Cu, Co, Au and Pt) as a co-catalyst onto the surface of iron is a general process to enhance the dechlorination efficiency and reaction rate (Cwiertny et al. 2006). Therein, Pd/Fe bimetal greatly improved the dechlorination efficiency of COCs; however, nickel as a non-precious metal could replace palladium metal without loss of performance (Zhang et al. 1998). As a result of lower hydrogen overpotential, Ni as a co-catalyst accelerated the hydrogen production process compared with single Fe, and the produced H2 could reduce COCs indirectly with the existence of an effective catalyst (e.g. Pd, Ni, Cu and Co) (Matheson & Tratnyek 1994; Schreier & Reinhard 1995; Lowry & Reinhard 1999; Schrick et al. 2002). For this reason, we chose Ni as a cost-effective metal dopant on nZVI particles to improve the reactivity of the catalyst.

In this study, nanoscale nickeled zero-valent iron @ rGO (NF@rGO) composite was synthesized, characterized and used to degrade CT. Batch experiments evaluated the removal capacity and the catalytic activity of laboratory-synthesized catalysts; besides that, the reduction products of CT and its reaction kinetics were studied. Catalyst characterization was carried out by X-ray diffraction (XRD), X-ray photo-electron spectroscopy (XPS) and scanning electron microscope (SEM) with X-ray energy dispersive spectroscopy (EDS). The purpose of our work was to develop a highly active catalyst, and provide a theoretical reference for the catalytic hydrodechlorination of COCs.

Material and chemicals

All the chemicals used were of analytical reagent (AR) grade from commercial sources. Graphite flake (purity ≥ 99.8%, −325 mesh), NaNO3 (≥99.5%), H2SO4 (5 and 98%), KMnO4 (≥98.0%), H2O2 (30%) and HCl (5%) were used to synthesize graphene oxide. FeSO4·7H2O (≥99.5%). NiSO4·6H2O (≥98.5%) and NaBH4 (≥ 99.5%) were used to synthesize NF@rGO. All degradation experiments used a certain volume of freshly made stock solution of CT (15 mg/L). All deionized (DI) water was prepared with an Aquelix 5 (Millipore) pure water system and bubbled with nitrogen (99.2%) before use.

Synthesis of GO

The graphene oxide (GO) was prepared by Hummers method, and the whole process was carried out in an Erlenmeyer flask with water bath and magnetic stirring. Firstly, graphite flake (1 g) and NaNO3 (0.5 g) powder were mixed by using H2SO4 aqueous solution (23 mL, 98%). After 1 h stirring at 0 °C, 3 g of KMnO4 was divided into 6 parts and add to the dispersion within 2.5 h at 10–15 °C. Subsequently, the temperature of the water bath was adjusted to 35 °C and it was stirred for 24 h, and then the temperature was increased to 90 °C. Meanwhile, H2SO4 aqueous solution (100 mL, 5%) and deionized water (150 mL) was added to the mixture respectively for further dilution of the dispersion. After the H2O2 aqueous solution (5 mL, 30%) was added to the suspension, the resultant GO was immediately filtrated (using 20 μm qualitative filter paper) until the suspension stopped bubbling, and it was washed with HCl aqueous solution (5%) and DI water (until no SO42− was detected in the supernate fluid). Finally, the dried GO powder was obtained with a vacuum oven (60 °C for 48 h).

Synthesis of NF@rGO

The NF@rGO was synthesized in a nitrogen atmosphere through a two-step reductive deposition method, and the whole process was carried out in a four-necked flask with continuous stirring; various composition ratios (Fe/GO = 2:1, 3:1, 4:1 and 5:1 wt./wt.) were considered. First of all, GO was added to DI water to form a suspension buffer (0.75 g/L) with ultrasonic dispersion, and 100 mL of corresponding mass of FeSO4·7H2O aqueous solution was added dropwise into the solution with a peristaltic pump at the speed of 20 rpm. And then, NaBH4 aqueous solution (10.0 g/L) was added dropwise into the solution. At the end of dripping, ferrous iron and GO were reduced by NaBH4 and then nZVI particles were co-precipitated with rGO. The mechanism is NaBH4, as a strong reducing reagent, could be effective in reducing the carbonyl group to the alcohol group of GO (Periasamy & Thirumalaikumar 2000), and the reaction equation of ferrous iron reduced by NaBH4 is described as follows:
(1)
Secondly, the nZVI@rGO was collected by magnetic separation and washed with absolute ethyl alcohol repeatedly. After that, the semi-finished product was put into a four-necked flask using DI water. And then the NiSO4·6H2O aqueous solution was dropwise added into the solution to coat a thin layer of nickel on the surface of the nZVI particles. The method is according to the following equations for chemical reactions:
(2)

At last, the NF@rGO NPs were washed and freeze-dried for further use. Similarly, nZVI and Ni/Fe bimetallic NPs were also prepared with this procedure. The preparation method of 2N4F@MWCNTs (Fe/MWCNTs = 2 wt./wt., Ni/Fe = 4 wt.%) was according to our previous work (Chen et al. 2019).

Catalyst characterization

The surface morphology of nZVI, GO and NF@rGO was characterized by scanning electron microscope (SEM) (Hitachi, SU-8010, Japan), the Ni content of Ni/Fe bimetal was quantified by X-ray energy dispersive spectroscopy (EDS) (Horiba, Japan). The mineralogical properties of nZVI, fresh and used NF@rGO, were investigated by X-ray diffraction (XRD) (Bruker, AXS, D8 Advance). The chemical states of NF@rGO were examined by X-ray photo-electron spectroscopy (XPS) (Thermo Scientific, ESCALAB 250 XI).

Experimental system

A series of batch experiments were conducted to study the catalytic activity of laboratory-synthesized catalysts and investigate the intermediate products of CT degradation. All of the experiments were carried out in 100 mL serum bottles with a rotary shaker (220 rpm), each bottle was double sealed, Teflon-lined and had an aluminum cap, and each experiment was run in triplicate. Before the experiment, 0.01 g of laboratory-synthesized catalysts were added in the serum bottle respectively with 15 mg/L of CT stock solution until no headspace was left. At specific time intervals, 1 mL of samples were filtered with 0.22 μm filters and withdrawn to bottle headspace for analysis the concentration of CT and its degradation products by gas chromatography spectrometry (GC-2014, Shimadzu, Japan). In addition, the generated CH4 and CO in the bottle were gathered with an aluminum foil gas-collecting bag.

The sample from bottle headspace was determined by GC with electron capture detector (ECD). Samples were separated and evaluated with a Rtx-1 column (30.0 m × 0.25 mm ID × 0.25 μm thickness, Restek, USA), with nitrogen (99.99%) as the carrier gas. The oven temperature was programmed from 40 to 100 and finally to 200 °C at the rate of 8 and 6 °C/min, with a 5 min initial hold and 10 min final hold. The temperature of injector and detector were 220 and 320 °C, respectively.

Besides, the sample from the gas-collecting bag was determined by GC with a thermal conductivity detector (TCD). Samples were separated and evaluated with a TDX-01 column (2.0 m × 3.00 mm, China), with helium (99.99%) as the carrier gas. The oven temperature was 120 °C, held for 7 min, and the temperatures of the injector and detector were 100 and 150 °C, respectively. Blank experiments and repeated experiments were carried out and the results are described in the Supplementary Information.

Characterization of catalysts

XRD measurement

The XRD patterns in Figure 1 were determined to identify the crystal structures of nZVI, fresh and used 4N4F@rGO (Fe/GO = 4:1 wt./wt., Ni/Fe = 4 wt.%). The sharp and strong diffraction peaks at 2θ = 44.62° and 65.03° corresponded to the standard card of α-Fe (JCPDS, No. 06-0696). However, the intensity of the reflection peak at 2θ = 44.62° decreased significantly in fresh 4N4F@rGO, this may be caused by the small crystallites of Fe0 particles.

Figure 1

XRD patterns for nZVI, fresh and aged 4N4F@rGO.

Figure 1

XRD patterns for nZVI, fresh and aged 4N4F@rGO.

Close modal

Besides, it was noteworthy that the reflection peaks at 2θ = 27.55°, 36.29° and 46.91°, which belonged to aged 4N4F@rGO, corresponded to the standard card of lepidocrocite (γ-FeOOH) (JCPDS, No. 44-1415) and were assigned to the (210), (301) and (020) planes, respectively. The reflection peaks at 2θ = 30.07°, 35.87°, 56.91°, 62.50° corresponding to magnetite (Fe3O4) (JCPDS, No. 19-0629) were assigned to the (220), (311), (511) and (440) planes, and/or maghemite (γ-Fe2O3) (JCPDS, No. 25-1402), assigned to the (206) and (119) planes. The results exposited that the Fe0 was oxidized and deposited on the surface of 4N4F@rGO in the process of reaction.

However, no obvious characteristic diffraction peaks of Ni or graphite were detected either in fresh or recycled 4N4F@rGO, which was due to the diffraction peaks of nickel metal being very close to iron, and graphene sheets were prevented from forming a regular crystal structure as a result of the inserted NPs.

XPS characterization

In order to get more information on the chemical composition of NF@rGO, XPS analyses were performed to evaluate the chemical states including Ni, Fe, O and C (Figure 2(a)).

Figure 2

(a) Full survey XPS patterns for 4N4F@rGO, (b)–(e) high-resolution XPS patterns for 4N4F@rGO of Ni (2p), Fe (2p), O (1 s) and C (1 s).

Figure 2

(a) Full survey XPS patterns for 4N4F@rGO, (b)–(e) high-resolution XPS patterns for 4N4F@rGO of Ni (2p), Fe (2p), O (1 s) and C (1 s).

Close modal

According to Ni 2p spectrum (Figure 3(b)), the photoelectron peaks located at 852.4 and 869.0 eV demonstrated the existence of Ni (Ni2p3/2, 852.3 eV; Ni2p1/2, 869.7 eV) (Moulder et al. 1979), and the peaks of NiO (Ni2p3/2, 853.3 eV; Ni2p1/2, 871.7 eV) (Moulder et al. 1979) were also detected at 854.2 and 870.6 eV.

Figure 3

(a) SEM images of nZVI, (b) rGO, (c, d) 4N4F@rGO and (e) EDS spectra of the metal composition of 4N4F@rGO (corresponding to the box in part d).

Figure 3

(a) SEM images of nZVI, (b) rGO, (c, d) 4N4F@rGO and (e) EDS spectra of the metal composition of 4N4F@rGO (corresponding to the box in part d).

Close modal

Meanwhile, Figure 2(c) shows the high-resolution spectrum of Fe2p, two photoelectron peaks at 710.5 and 723.9 eV demonstrate the existence of oxidized iron (Fe(III) and Fe(II)), which are similar to magnetite (Fe3O4) (Fe2p3/2, 710.6 eV; Fe2p1/2, 724.1 eV) (Fiedor et al. 1998), hydrated ferric oxide (FeOOH) (Fe2p3/2, 711.2 eV; Fe2p1/2, 724.9 eV) (Echigo et al. 2012), hematite (α-Fe2O3) (Fe2p3/2, 711.0 eV; Fe2p1/2, 724.6 eV) (Yamashita & Hayes 2008) and maghemite (γ-Fe2O3) (Fe2p1/2, 710.9 eV) (Zhang et al. 2011). Moreover, Fe0 showed an obvious peak at 706.7 eV followed by its satellite peak at 719.5 eV (Fe2p3/2, 706.75; Fe2p1/2, 719.95 eV) (Moulder et al. 1979).

The O (1 s) spectrum (Figure 2(d)) showed three types of oxygen species, which corresponded to the anionic oxygen in FeXOY (530.7 eV), the oxygen-containing functional groups in rGO (531.9 eV) (Fan et al. 2011), and the adsorbed H2O (536.0 eV) (Ertl et al. 1997).

The C (1 s) spectra (Figure 2(e)) presented four types of carbon bonds (Moulder et al. 1979; Ertl et al. 1997; Hayes 2001): (1) non-oxygenated carbon (including C-C, C = C, and C-H) (284.4 eV), (2) the carbon in C-O (C-O-C and/or C-OH groups) (285.4 eV), (3) the carbonyl carbon in C = O (287.8 eV), (4) carboxylate carbon in O-C = O (288.9 eV). The peak area of C-C was greater than other carbon-based oxygen-containing functional groups, which proved that GO was successfully reduced to rGO by NaBH4 during the preparation of the catalyst.

SEM characterization

The SEM image of nZVI (a), rGO (b), 4N4F@rGO (c, d) is presented in Figure 3; the partition ratio of the alloy element (corresponding to the box in part d) was measured and is presented in Figure 3(e). It can be seen that nZVI particles interconnected to form a necklace-like structure due to surface tension and magnetic interactions between each particle. Figure 3(b) exhibits the sheet-like structures of GO. Moreover, it can be clearly seen 4N4F@rGO from Figure 3(c) and 3(d) that the Ni/Fe bimetal particles were evenly distributed on the surface of rGO. The EDS analysis indicated that the NPs were Ni/Fe bimetal and reached a desirable ratio.

Batch experiments

Loading ratio of rGO

Various mass of Ni/Fe bimetallic NPs (Ni/Fe = 4 wt.%) were loaded onto the surface of rGO. As shown in Figure 4, the reduction of CT followed pseudo-first-order kinetics, 4N4F@rGO (Fe/GO = 4 wt./wt., Ni/Fe = 4 wt.%) reached a higher efficiency in CT degradation.

Figure 4

Different loading ratio of Ni/Fe bimetal on rGO.

Figure 4

Different loading ratio of Ni/Fe bimetal on rGO.

Close modal

Furthermore, the properties of four laboratory-synthesized catalysts were investigated by calculating the removal capability (mg CT/g catalyst). The results of the calculation are listed in Table 1; 4N4F@rGO could degrade CT with a high removal rate in 35 min, the observed rate constant (kobs) indicated that the reaction rate of 4N4F@rGO was 10 times that of nZVI. By comparison with the removal capability and reaction rate, 4N4F@rGO achieved better results than 2N4F@MWCNTs (Fe/MWCNTs = 2 wt./wt., Ni/Fe = 4 wt.%) in CT degrading, which means rGO as a supporter could be loaded with more metal to improve its catalytic performance.

Table 1

Results of laboratory-synthesized catalysts on degradation of CTa

CatalystRemoval rate (%)Removal capability (mg/g)kobs (min−1)
nZVI 37.2 55.8 0.0102 
N461.2 91.7 0.0267 
2N4F@MWCNTs 90.8 136.2 0.1073 
4N4F@rGO 95.5 143.2 0.1106 
CatalystRemoval rate (%)Removal capability (mg/g)kobs (min−1)
nZVI 37.2 55.8 0.0102 
N461.2 91.7 0.0267 
2N4F@MWCNTs 90.8 136.2 0.1073 
4N4F@rGO 95.5 143.2 0.1106 

aReaction conditions: initial concentration of CT = 15 mg/L, catalyst dosage = 0.1 g/L, initial pH = 7.0, reaction time = 35 min and the temperature was 30 °C.

Reaction activation energy

The degradation of CT by nZVI, N4F and 4N4F@rGO was fitted to the pseudo first-order kinetic model, the experimental results are summarized in Table 2 and Figure S3 (in the Supplementary Information). The kinetic model was used to simulate the relationships between the reaction rate and temperature (Figure 5) for obtaining the activation energy via the Arrhenius equation (Dahm 1994):
(3)
where K stands for the pseudo-first-order rate constant, A for the frequency factor, R for the universal gas constant (8.314 J·mol−1 K−1), T for the temperature (K), and then the value of activation energy (Ea) can be obtained by calculating the slope of the plot line via the following equation:
(4)
Table 2

Results of laboratory-synthesized catalysts on degradation of CT under different temperatureb

T (K)nZVI
N4F
4N4F@rGO
Removal rate (%)Ka (min−1)Removal rate (%)Ka (min−1)Removal rate (%)Ka (min−1)
283.15 21.7 0.0030 27.7 0.0082 69.8 0.0340 
293.15 29.1 0.0053 41.8 0.0140 83.3 0.0607 
303.15 37.2 0.0102 61.2 0.0267 95.5 0.0926 
308.15 46.2 0.0163 68.9 0.0345 98.3 0.1268 
313.15 54.0 0.0219 81.0 0.0452 99.0 0.1616 
T (K)nZVI
N4F
4N4F@rGO
Removal rate (%)Ka (min−1)Removal rate (%)Ka (min−1)Removal rate (%)Ka (min−1)
283.15 21.7 0.0030 27.7 0.0082 69.8 0.0340 
293.15 29.1 0.0053 41.8 0.0140 83.3 0.0607 
303.15 37.2 0.0102 61.2 0.0267 95.5 0.0926 
308.15 46.2 0.0163 68.9 0.0345 98.3 0.1268 
313.15 54.0 0.0219 81.0 0.0452 99.0 0.1616 

bReaction conditions: initial concentration of CT = 15 mg/L, catalyst dosage = 0.1 g/L, initial pH = 7.0 and reaction time = 35 min.

Figure 5

Kinetic analysis of (a) nZVI, (b) N4F, (c) 4N4F@rGO under different temperatures, and fitting curve of ln(K)/T (d).

Figure 5

Kinetic analysis of (a) nZVI, (b) N4F, (c) 4N4F@rGO under different temperatures, and fitting curve of ln(K)/T (d).

Close modal

Figure 5(d) clearly demonstrates that nickel as a co-catalyst decreased the activation energy of 6.59 kJ/mol in N4F compared with nZVI, and rGO as a conductive material could further reduce the activation energy of 4.73 kJ/mol by accelerating electron transfer during the redox reaction.

Reduction products and pathway of CT

In order to explore the reduction pathway of CT, various products were monitored by GC in the process of reaction, including chloroform (CF), dichloromethane (DCM), carbon monoxide (CO) and methane (CH4), trace amounts (around the detection limit of GC) of perchloroethylene (PCE) and trichloroethylene (TCE) were also detected after a certain time of reaction. Among these, CO and CH4 may be considered as complete dechlorination products (CDCPs).

As shown in Figure 6, for CF as a daughter product from CT, the amount decreased significantly after 10 min in 2N4F@MWCNTs and 4N4F@rGO catalytic system; however, DCM was hardly further reduced by these two catalysts within 2 h through independent experiments, so the generated CDCPs were attributed to the reduction products of CF. It is noteworthy that CDCPs were not detected in the nZVI catalytic system during the experimental period, which means Ni as a co-catalyst could promote the generation of CDCPs. Besides, benefiting from the high reactivity of 2N4F@MWCNTs and 4N4F@rGO, considerable amounts of CDCPs were produced in 2 h.

Figure 6

The changes of substrate concentration in various catalytic reduction systems: (a) nZVI, (b) N4F, (c) 2N4F@MWCNTs and (d) 4N4F@rGO.

Figure 6

The changes of substrate concentration in various catalytic reduction systems: (a) nZVI, (b) N4F, (c) 2N4F@MWCNTs and (d) 4N4F@rGO.

Close modal
Figure 7

Reduction process of CT.

Figure 7

Reduction process of CT.

Close modal
Based on the kinds of degradation products and the concentration variation, it was inferred that the reduction pathway of CT was initiated by dissociative electron transfer based on carbon-centered radical (Balko & Tratnyek 1998):
(5)
The resulting trichloromethyl radicals (or dichloromethyl radicals) can be further electronically transferred to produce CF (or DCM):
(6)
It was inferred that the generated CDCPs were the reduction products of CF, CF was degraded by losing a proton first and then a chloride ion to form dichlorocarbene (Satterfield 1991):
(7)
The dichlorocarbene was then hydrolyzed to form CO, and CO could be further reduced to methane via Fischer–Tropsch process (Smith & March 2001):
(8)
(9)
The detected perchloroethylene (PCE) may be formed by dimerization of dichlorocarbene radicals, and trichloroethylene (TCE) should be the reduction product of PCE.
(10)

Kinetics of CT reduction

According to the reaction equations (Equations (5)–(9)), the major reduction pathway of CT could be described by Figure 7. Based on the pseudo first-order kinetic model, the rate equations of CT reduction can be expressed as follows:
(11)
(12)
(13)
(14)
where C is the concentration of CT and its reduction products, the values of k1-k3 can be obtained by fitting the experimental data with the curve of the following equation:
(15)
(16)
(17)
(18)
The fitting curve of the rate equations is shown in Figure 6. The variance was used as a criterion to select the best set of rate constants (all of the R2 were greater than 0.92), and the k values are listed in Table 3.
Table 3

Simulated rate constants of CT reduction processes with various catalysts

Catalystsk1 (h−1)k2 (h−1)k3 (h−1)
nZVI 0.57 0.16 0.00 
N41.60 0.15 0.01 
2N4F@MWCNTs 13.87 0.24 0.70 
4N4F@rGO 19.81 0.15 1.20 
Catalystsk1 (h−1)k2 (h−1)k3 (h−1)
nZVI 0.57 0.16 0.00 
N41.60 0.15 0.01 
2N4F@MWCNTs 13.87 0.24 0.70 
4N4F@rGO 19.81 0.15 1.20 

Table 3 reveals that all CT were reduced through the hydrogenolysis pathway to CF and DCM by these four catalysts. However, the presence of nickel not only accelerated the reduction of CT to CF, but also promoted the complete reduction pathway of CF. Moreover, N4F obtained more reactive sites due to immobilization on rGO, greatly increasing the reaction rate of CT reduction, including reduction of CT to CF (k1) and reduction of CF (k2+k3). As a result, rGO could support more N4F NPs, 4N4F@rGO has the capability of reducing more CF to CDCPs (k3).

Mechanism of CT reduction by 4N4F@rGO

As shown in Figure 8, the reduction of CT by 4N4F@rGO in aqueous solution involved four reactions. In the Fe-H2O system, Fe0 serving as electron donor could directly react with COCs:
(19)
Figure 8

Mechanism of CT reduction by 4N4F@rGO.

Figure 8

Mechanism of CT reduction by 4N4F@rGO.

Close modal
Meanwhile, Fe2+ as the oxidation product of Fe0 could still be involved in the reduction of COCs (Equation (20)). However, this reaction occurred slowly due to the lower electron-donating ability of Fe2+ (E0 (Fe2+/Fe0) = −0.440 V, E0 (Fe3+/Fe2+) = +0.770 V). This process can be demonstrated by the formation of a passivation layer, which contained trivalent iron on the surface of the 4N4F@rGO after repeated use.
(20)
A benefit from nickel metal is that it has lower hydrogen overpotential compared with iron. Ni/Fe bimetal could accelerate the indirect reduction of COCs by the hydrogen produced (Equation (21)). Nickel metal is indispensable in the process of hydrogen reduction, H2 could not reduce COCs directly in the absence of an effective catalyst (e.g. Pd, Ni, Cu and Co). The generated hydrogen could be embedded in the crystal lattice of Ni to form reactive atomic hydrogen (·H) (Equations (23)–(25)) (Cwiertny et al. 2006), and ·H as a very powerful reducing agent played an important role in the complete dechlorination of CT (Wang et al. 2009).
(21)
(22)
(23)
(24)
(25)
As a carbon material, rGO can form primary batteries with Fe0 in aqueous solution. In the process of redox reaction, the electrons transferred from anode (Fe0) to cathode (rGO), COCs as the electron acceptor could be reduced at the surface of rGO without contact with Fe0:
(26)
(27)

In this research, 4N4F@rGO has been successfully synthesized by a two-step method and its removal capacity and CT reduction rate was assessed and compared with three other laboratory-synthesized catalysts. The results suggested that 4N4F@rGO can degrade CT with a high reaction rate and selectivity towards non-chlorinated products. Furthermore, nickel metal reduced the active energy and accelerated the complete dechlorination reaction by the H-radical produced, and the Ni/Fe bimetal gained more reactive sites by loading on rGO to further accelerate the reduction of CT to CDCPs.

Based on the kinds of degradation products of CT and its concentration variation in reduction experiments, it was inferred that the reduction process of CT was initiated by dissociative electron transfer based on carbon-centered radicals. Meanwhile, the mechanism of CT reduction by 4N4F@rGO was also discussed, it was concluded that rGO as an excellent electroconductive supporter could not only disperse the bimetal but also participated in the galvanic interaction with Fe0 to reduce CT indirectly.

The catalytic activities of two catalysts (4N4F@rGO and 2N4F@MWCNTs) with different supporters were compared. It was found that rGO could load more Ni/Fe and provide a better fluidity for substrate, which was conducive to reduce more intermediate products to CDCPs.

This work was supported by the National Natural Science Foundation of China (50578151), the National Science and Technology Major Project of China (2009ZX07207-008, 2009ZX07419-002, 2009ZX07207-001), the Beijing Municipal Education Commission School-Enterprise Cooperation Projects (51900265005), portable, in car, on-line monitoring instrument development and demonstration for focusing on prevention and control of heavy metals like mercury, chromium, lead, cadmium, arsenic (2012YQ060115), the Fundamental Research Funds for the Central Universities (2652013101, 2652013086, 2652013087), and the Key Project of Air Pollution Causes and Control Technology Research (2016YFC0209205).

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

Bai
S.
Shen
X. P.
2012
Graphene-inorganic nanocomposites
.
Rsc Advances
2
(
1
),
64
98
.
Balko
B. A.
Tratnyek
P. G.
1998
Photoeffects on the reduction of carbon tetrachloride by zero-valent iron
.
Journal of Physical Chemistry B
102
(
8
),
1459
1465
.
Cwiertny
D. M.
Bransfield
S. J.
Livi
K. J. T.
Fairbrother
D. H.
Roberts
A. L.
2006
Exploring the influence of granular iron additives on 1,1,1-trichloroethane reduction
.
Environmental Science & Technology
40
(
21
),
6837
6843
.
Dahm
C. N.
1994
Chemical kinetics and process dynamics in aquatic systems by Patrick L. Brezonik
.
Journal of the North American Benthological Society
14
(
2
),
354
.
Dreyer
D. R.
Park
S.
Bielawski
C. W.
Ruoff
R. S.
2009
The chemistry of graphene oxide
.
Chemical Society Reviews
39
(
1
),
228
240
.
Ertl
G.
Kniizinger
H.
Weitkamp
J.
Professor
S.
Suslick
K. S.
1997
Handbook of heterogeneous catalysis, Vol. III
.
Zeitschrift Für Physikalische Chemie
208
(
Part_1_2
),
274
278
.
Fan
Z. J.
Kai
W.
Yan
J.
Wei
T.
Zhi
L. J.
Feng
J.
Ren
Y. M.
Song
L. P.
Wei
F.
2011
Facile synthesis of graphene nanosheets via Fe reduction of exfoliated graphite oxide
.
ACS Nano
5
(
1
),
191
198
.
Fiedor
J. N.
Bostick
W. D.
Jarabek
R. J.
Farrell
J.
1998
Understanding the mechanism of uranium removal from groundwater by zero-valent iron using X-ray photoelectron spectroscopy
.
Environmental Science & Technology
32
(
10
),
1466
1473
.
Gillham
R. W.
Ohannesin
S. F.
1994
Enhanced degradation of halogenated aliphatics by zero-valent iron
.
Ground Water
32
(
6
),
958
967
.
Hashsham
S. A.
Scholze
R.
Feedman
D. L.
1995
Cobalamin-enhanced anaerobic biotransformation of carbon tetrachloride
.
Environmental Science & Technology
29
(
11
),
2856
2863
.
Hayes
M.
2001
Review of: Rase, H. F. Commercial heterogeneous catalysis. Handbook of commercial catalysts – heterogeneous catalysts
.
Platinum Metals Review
45
(
2
),
83
.
Hua
I.
Hoffmann
M. R.
1996
Kinetics and mechanism of the sonolytic degradation of CCl4: intermediates and byproducts
.
Environmental Science & Technology
30
(
3
),
864
871
.
Jin
X.
Li
Q.
Yang
Q.
2018
The reactivity of Fe/Ni colloid stabilized by carboxymethylcellulose (CMC-Fe/Ni) toward chloroform
.
Environmental Science & Pollution Research
25
(
21
),
21049
21057
.
Kustov
L. M.
Al-Abed
S. R.
Virkutyte
J.
Kirichenko
O. A.
Shuvalova
E. V.
Kapustin
G. I.
Mishin
I. V.
Nissenbaum
V. D.
Tkachenko
O. P.
Finashina
E. D.
2014
Novel Fe-Pd/SiO2 catalytic materials for degradation of chlorinated organic compounds in water
.
Pure and Applied Chemistry
86
(
7
),
1141
1158
.
Liu
R.
Lal
R.
2012
Nanoenhanced materials for reclamation of mine lands and other degraded soils: a review
.
Journal of Nanotechnology
2012
,
1
18
.
Lowry
G. V.
Reinhard
M.
1999
Hydrodehalogenation of 1-to 3-carbon halogenated organic compounds in water using a palladium catalyst and hydrogen gas
.
Environmental Science & Technology
33
(
11
),
1905
1910
.
Matheson
L. J.
Tratnyek
P. G.
1994
Reductive dehalogenation of chlorinated methanes by iron metal
.
Environmental Science & Technology
28
(
12
),
2045
2053
.
Moulder
J. F.
Chastain
J.
King
R. C.
Jr
1979
Handbook of X-ray photoelectron spectroscopy : a reference book of standard spectra for identification and interpretation of XPS data
.
Chemical Physics Letters
220
(
1
),
7
10
.
Pei
S.
Cheng
H. M.
2012
The reduction of graphene oxide
.
Carbon
50
(
9
),
3210
3228
.
Periasamy
M.
Thirumalaikumar
P.
2000
Methods of enhancement of reactivity and selectivity of sodium borohydride for applications in organic synthesis
.
Journal of Organometallic Chemistry
609
(
1–2
),
137
151
.
Santharam
S.
Davis
L. C.
Erickson
L. E.
2014
Biodegradation of carbon tetrachloride in simulated groundwater flow channels
.
Environmental Progress & Sustainable Energy
33
(
2
),
444
453
.
Satterfield
C. N.
1991
Heterogeneous Catalysis in Industrial Practice
, 2nd edn.
McGraw-Hill Inc.
,
New York, NY
.
Schrick
B.
Blough
J. L.
Jones
A. D.
Mallouk
T. E.
2002
Hydrodechlorination of trichloroethylene to hydrocarbons using bimetallic nickel-iron nanoparticles
.
Chemistry of Materials
14
(
12
),
5140
5147
.
Smith
M. B.
March
J.
2001
Advanced Organic Chemistry: Reaction, Mechanisms, and Structure
, 5th edn.
Wiley
,
New York, NY
, pp.
465
.
Stankovich
S.
Dikin
D. A.
Piner
R. D.
Kohlhaas
K. A.
Kleinhammes
A.
Jia
Y.
Wu
Y.
Nguyen
S. T.
Ruoff
R. S.
2007
Synthesis of graphene-based nanosheets via chemical reduction of exfoliated graphite oxide
.
Carbon
45
(
7
),
1558
1565
.
Wang
X.
Chen
C.
Chang
Y.
Liu
H.
2009
Dechlorination of chlorinated methanes by Pd/Fe bimetallic nanoparticles
.
Journal of Hazardous Materials
161
(
2
),
815
823
.
Yamashita
T.
Hayes
P.
2008
Analysis of XPS spectra of Fe2+ and Fe3+ ions in oxide materials
.
Applied Surface Science
254
(
8
),
2441
2449
.
Zhang
W. X.
2003
Nanoscale iron particles for environmental remediation: an overview
.
Journal of Nanoparticle Research
5
(
3–4
),
323
332
.
Zhang
W. X.
Wang
C. B.
Lien
H. L.
1998
Treatment of chlorinated organic contaminants with nanoscale bimetallic particles
.
Catalysis Today
40
(
4
),
387
395
.
Zhang
Y.
Chen
L. X.
Lei
Y. Q.
Wang
Q. D.
2002
The reduction of cycling capacity degradation of Mg-Ni-based electrode alloys by Fe substitution
.
International Journal of Hydrogen Energy
27
(
5
),
501
506
.
Zhang
N.
Zhang
Y. H.
Xu
Y. J.
2012
Recent progress on graphene-based photocatalysts: current status and future perspectives
.
Nanoscale
4
(
19
),
5792
5813
.

Supplementary data