Hydrogen production performance of the non-platinum-based MoS2/CuS cathode in microbial electrolytic cells

Pingping HAO Fangfang LI Yawen WANG Houfen LI Xiao ZHANG Rui LI Lei WANG Jianxin LIU

Citation:  Pingping HAO, Fangfang LI, Yawen WANG, Houfen LI, Xiao ZHANG, Rui LI, Lei WANG, Jianxin LIU. Hydrogen production performance of the non-platinum-based MoS2/CuS cathode in microbial electrolytic cells[J]. Chinese Journal of Inorganic Chemistry, 2024, 40(9): 1811-1824. doi: 10.11862/CJIC.20240054 shu

微生物电解槽中非铂基材料MoS2/CuS阴极产氢性能

    通讯作者: 王磊, 56218605@qq.com
    刘建新, liujx0519@163.com
  • 基金项目:

    国家自然科学基金 22008167

摘要: 以二水合钼酸钠、硫脲、草酸以及三水合硝酸铜为原料,使用一步水热法成功合成了MoS2/CuS复合催化剂。考察了不同Mo、Cu前驱体物质的量之比(nMonCu)制备的MoS2/CuS阴极催化剂在双室微生物电解槽(MEC)体系中的产氢性能。采用X射线衍射(XRD)、X射线光电子能谱(XPS)、扫描电子显微镜(SEM)、透射电子显微镜(TEM)、线性扫描伏安法(LSV)、电化学阻抗谱(EIS)和循环伏安法(CV)对合成的催化剂进行表征和产氢性能分析。结果表明MoS2/CuS-20%(nMonCu=5∶1)的析氢性能优于铂(Pt)网,此外,在0.8 V的外加电压下,MoS2/CuS-20%作为阴极在MEC中72 h的产氢速率为(0.203 1±0.023 7) mH23·m-3·d-1,略高于Pt网的(0.188 6±0.013 4) mH23·m-3·d-1。一定量CuS的加入不仅调节了MoS2的电子转移能力,而且增加了活性位点的密度。

English

  • Hydrogen is a clean energy source and feedstock used in numerous industries[1-2]. Its calorific is as high as 22 MJ·kg-1, nearly 2.75 times higher than hydrocarbon fuels[3], which is considered the most important energy carrier and fuel of the future. However, the commonly used methods of hydrogen production, such as hydrogen production from fossil energy sources, photocatalytic hydrogen production[4-5], and hydrogen production from water electrolysis[6], inevitably suffer from high energy consumption, low efficiency, and environmental pollution. Microbial electrolysis cell (MEC) is a new technology developed in recent years, which can simultaneously treat organic wastewater and produce hydrogen to convert biomass energy into hydrogen energy. The concept of MEC originated in 2005 to explore the waste-to-energy generation route based on sustainability and renewability. MEC is based upon bio-catalyzed reactions and is also named a bio-catalyzed electrolysis cell (BEC) or bio-electrochemically assisted microbial reactor (BEAMR)[7]. MECs are capable of chemical conversion of organic matter to hydrogen in the complete absence of oxygen with a small voltage supply. In MEC, microorganisms at the anode decompose organic matter to produce electrons and protons, the electrons produced at the anode reach the cathode through an external circuit, and the protons reach the cathode through the solution, where they combine to produce hydrogen at an applied auxiliary voltage[8].

    The hydrogen evolution reaction (HER) takes place at the cathode. In an MEC system, the cathode is the main electron donor[9]. The cathode base materials and the catalyst used on it have a major impact on the performance of the MECs[10]. To date, platinum (Pt)-based materials have exhibited the most effective catalytic performance toward hydrogen production in MECs. However, its practical application is limited by its high cost, rapid poisoning by cations, and short lifetime[11]. Therefore, it is critical to find catalysts that can replace Pt for the commercialization of MEC systems.

    Recently, many catalysts have been explored. These novel catalysts include non-precious metal bifunctional electrocatalysts—two-dimensional cobalt-doped MnPSe3 nanosheets (CMPS) for alkaline seawater splitting HER and oxygen evolution reaction (OER) bifunctional catalysts[12]; advanced electrocatalysts for energy-efficient hydrogen production systems such as methanol oxidization-assisted hydrogen production (MOAHP), HzOR-assisted hydrogen production (HOAHP), and urea oxidization-assisted hydrogen production (UOAHP)[13]; and nanoporous gold supported multi-functional layered double hydroxides for oxygen/hydrogen evolution reactions, hydrazine oxidation reaction, and overall hydrazine splitting[14]. Molybdenum disulfide (MoS2), a transition metal sulfide with a layered inorganic compound structure, has been explored for use as a replaceable catalyst for platinum-based materials due to its large specific surface area, low price, earth abundance, and good stability[15-16]. To the best of our knowledge, there are several examples of the use of molybdenum disulfide-based catalysts as cathode catalysts for hydrogen production in MEC (Table S1, Supporting information): Hwang et al.[17] developed a novel MoS2 nano-carbon (NC) coated cathode for hydrogen production in a MEC. The MoS2-NC200 cathode, electrodeposited at -200 μA·cm-2, showed the maximum hydrogen production rate (HPR) of (0.152±0.002) m3·m-2·d-1 at an applied voltage of 0.9 V which is comparable to the previously reported Pt electrodes. However, the vertical stacking of multiple S-Mo-S layers of MoS2 results in poor conductivity and insufficient exposed active sites, leading to its poor HER[8].

    Copper sulfide (CuS) is also a transition metal sulfide with low cost, good stability, low cytotoxicity[18], excellent optical[19], electronic[20], catalytic, and other physicochemical properties. Due to these excellent properties, CuS has been widely studied for a variety of applications, such as biosensors[21], electro-catalytic hydrogen evolution[22], and batteries[23]. It might be an effective approach if CuS and MoS2 could be combined to form a composite catalyst. The excellent ability of CuS to regulate electrons and to separate reactive species[21, 24] could just compensate for the inadequacy of MoS2 alone as a catalyst for electrochemical reactions.

    Inspired by this, we used a one-step hydrothermal method to complex two transition metal sulfides, MoS2 and CuS. After analyzing the electrochemical hydrogen evolution reaction of MoS2/CuS, it was applied to the hydrogen production system of a two-chamber MEC and further evaluated in terms of HPR, Coulombic efficiency (CE), cathodic hydrogen recovery rate (CHR), and energy recovery. It is expected that the addition of CuS can improve the conductivity of MoS2 as well as increase its edge active sites, thus enhancing the hydrogen production performance.

    All reagents were analytical grade and used without further purification. Sodium molybdate dehydrate (Na2MoO4·2H2O) and ammonium chloride (NH4Cl) were bought from Tianjin Guangfu Technology Development Co., Ltd. Thiourea (NH2CSNH2) and sodium acetate (CH3COONa) were bought from Tianjin Tianli Chemical Reagent Co., Ltd. Oxalic acid (C2H2O4) and anhydrous calcium chloride (CaCl2) were bought from Tianjin Komeo Chemical Reagent Co., Ltd. Copper nitrate trihydrate (Cu(NO3)2·3H2O) was bought from Tianjin Beichen Fangzheng Reagent Factory. Anhydrous disodium hydrogen phosphate (Na2HPO4) and anhydrous sodium dihydrogen phosphate (NaH2PO4) were bought from Shanghai Ball Chemical Reagent Co., Ltd. Potassium bicarbonate (KHCO3), magnesium sulfate heptahydrate (MgSO4·7H2O), and potassium ferricyanide (K3[Fe(CN)6]) were bought from Shanghai McLean Biochemical Technology Co., Ltd. Deionized water was used for the preparation of all solutions and rinsing.

    The catalyst MoS2/CuS was prepared by a step hydrothermal method. The precursors Na2MoO4·2H2O and NH2CSNH2 in a molar ratio of 1∶2 (5 mmol Na2MoO4·2H2O and 10 mmol NH2CSNH2) were added to 60 mL of distilled water. This mixture was mildly stirred for 30 min to get a homogeneous solution and labeled as solution-1. Then, 0.6 g C2H2O4 and 1 mmol Cu(NO3)2·3H2O were dissolved in 20 mL of distilled water and added into solution-1 dropwise at constant stirring for 1 h. Further, the final solution was transferred into the Teflon-lined stainless-steel autoclave and treated thermally at 180 ℃ for 24 h. The naturally cooled resulting colloidal solution was centrifuged with distilled water and anhydrous ethanol and the product was vacuum-dried overnight. The synthesized catalyst was designated as MoS2/CuS-20%. Then, we adjusted the amount of added Cu(NO3)2·3H2O substances to 0.25, 0.5, and 1.5 mmol to prepare MoS2/CuS-5%, MoS2/CuS-10%, and MoS2/CuS-30%, respectively.

    The carbon cloth electrodes were modified with MoS2/CuS-20% composites. An unmodified MoS2 electrode and a Pt mesh (Gaoss Union PTX, Wuhan Gaoshi Ruilian Technology Co. Ltd., Wuhan, China) were used as controls. Each electrode had a projected surface area of 4 cm2 (2 cm×2 cm). For modification, 40 mg MoS2/CuS-20% composite was mixed with 1 mL 0.5% Nafion solution and dropped on a carbon cloth. The mixture was sonicated for 0.5 h to obtain a homogeneous suspension. The suspension was applied to the carbon cloth uniformly using a pipette (both sides were coated). Finally, the modified carbon cloth was dried in the air. The final loading quantity of the modifying material was 10 mg·cm-2. The effect of catalyst loading on catalytic performance is shown in Fig.S1 (Supporting information). Other catalyst cathodes were also prepared by the same procedure.

    The constructed MEC is shown in Fig.S2. The MEC consisted of a dual chamber (150 mL, Zhonghua Instrument No.1 Store, Beijing, China) separated by a proton exchange membrane (N-117, Dupont). The liquid volume of the anode chamber consisted of 90 mL anolyte and 20 mL bacterial solution. The liquid volume of the cathode chamber consisted of 110 mL of phosphate buffer of K3[Fe(CN)6]. The anolyte contained NH4Cl (0.1 g·L-1), KHCO3 (0.5 g·L-1), MgSO4·7H2O (0.2 g·L-1), CaCl2 (0.075 5 g·L-1), NaH2PO4 (4.536 0 g·L-1), Na2HPO4 (8.8 g·L-1), acid trace element (1 mL·L-1, Table S2), and basic trace element (1 mL·L-1, Table S3). The catholyte contained: K3[Fe(CN)6] (6.6 g·L-1), NaH2PO4 (4.536 0 g·L-1), and Na2HPO4 (8.8 g·L-1). The anode was a carbon felt (Carbon Graphite Products Factory Co., Ltd., Tianjin, China) that had been cultivated under an MEC mode for three months. An external voltage of 0.8 V was applied to the MEC by connecting the cathode to the negative pole of a power supply (RYI3005-D, Jinqunhao Instrument Co., Ltd., Shenzhen, China), and the anode to the positive pole. The current was monitored with a multimeter (UNI-T 803, Unido Smart Test Store, Shanghai, China).

    It is necessary to evaluate various chemical properties of composite catalyst-coated electrodes, which are directly related to their performance. To confirm the chemical composition and the crystalline nature of the nanocomposites, the prepared catalysts were characterized using an X-ray diffractometer (XRD, Bruker D8 Advance Powder Diffractometer, operating voltage of 40 kV, tube current of 15 mA, λ=0.154 06 nm, in Cu radiation mode at a 2θ range from 10° to 80°; scan rate: 5 (°)·min-1). To further analyze the chemical valence of each element of MoS2/CuS, X-ray photoelectron spectroscopy (XPS, Thermo Fischer, ESCALAB 250Xi, UK) was used. The morphologies and chemical composition of MoS2/CuS were investigated with a field emission scanning electron microscope (FE-SEM, JSM-7900F, acceleration voltage of 10 kV) equipped with an energy-dispersive X-ray spectroscopy (EDX) instrument. The high-resolution transmission electron microscopy (TEM, JEM-F200, acceleration voltage of 200 kV) was employed to determine the crystal plane orientation of the catalysts.

    The hydrogen production was determined by gas chromatography (GC, Changzhou PANNA Instrument Co.) using a thermal conductivity detector (TCD) at 160 ℃ under argon as a gas carrier. The content of sodium acetate as a substrate was determined by UV-visible spectrophotometry (Flash Version: 3.00). When light passed through the sodium acetate solution, the sodium acetate solution selectively absorbed the light, and the absorbance of the sodium acetate solution varied with the wavelength of the light. The sodium acetate solution was scanned using a UV spectrophotometer in the wavelength range of 190 to 250 nm. 100 mg·L-1 sodium acetate had the maximum absorption peak at 191 nm, and 191 nm was selected as the measured wavelength[25].

    The electrochemical tests were carried out using an electrochemical station (Admiral, Beijing Xinlifang Technology Development Co., Ltd., Beijing, China) in a three-electrode system with anolyte (pH=7.0) as electrolyte. For these tests, prepared cathodes were applied as the working electrode equipped with an Ag/AgCl electrode (GB/T 59977—2010) as the reference electrode and platinum wire (1 mm×37 mm) as the counter electrode. Linear sweep voltammetry (LSV) tests were carried out to investigate the activity of hydrogen evolution on different cathodes with a range of-0.9 to 0 V (vs RHE) (the same as below) at a rate of 5 mV·s-1. The onset overpotential (ηonset) at a current density of 1 mA·cm-2 was used to compare the rates of hydrogen evolution at different cathodes. Tafel plots were calculated to evaluate the electrode kinetics, where the linear portions of the Tafel plots were fitted to the Tafel equation (η=blg|j|+a, where j is the current density, a is the cathodic intercept related to the exchange current density, and b is the Tafel slope). To investigate the charge transfer resistance of cathode materials, electrochemical impedance spectroscopy (EIS) was performed in the frequency range of 100 kHz to 10 MHz. Cyclic voltammetry (CV) tests performed at different scan rates (10, 20, 30, 40, 50, 60, 70, 80, 90, and 100 mV·s-1) were used to calculate the electrochemically active surface area (ECSA). Chronoamperometry (current-time) curves were established to evaluate the stability of cathode materials at a fixed current density of-1 mA·cm-2.

    To identify the microbial community of anode biofilm, total community genomic DNA extraction was performed using a Quant-iT PicoGreen dsDNA Assay Kit assay (Life, USA) following the manufacturer's instructions. The prokaryotic V3 and V4 regions of 16S rRNA genes were amplified using polymerase chain reaction (PCR; 2×Hieff®Robust PCR Master Mix). Two general PCR primers for bacterial gene amplicon were used: forward primer 341FCCTACGGGNGGCWGCAG and reverse primer 805R-GACTACHVGGGTATCTAATCC. Amplification products were purified using an Illumina MiSeq system (Illumina MiSeq, USA) and subjected to base-pair sequencing. The sequencing results were further processed and the operational taxonomic units were assigned following the manufacturer's instructions[26].

    1.8.1   Calculation of hydrogen production yield (HPY, mmol)

    According to the standard curve of H2 (Fig.S3), the total moles of H2 in the headspace volume (63 mL, volume of water quantity) were calculated based on the peak area and the following Eq.1:

    $ y=152\;845.79 x-9\;986.45 \quad\left(R^2=0.995\;91\right) $

    (1)

    where y is the peak area of gas chromatography and x (mmol) is the HPY of the reaction system.

    1.8.2   Calculation of HPR

    $ \mathrm{HPR}=\frac{10^{-3} V_{\mathrm{m}} \mathrm{HPY}}{V T} $

    (2)

    where Vm is the molar volume of gas (22.4 L·mol-1), V is the cathode volume (L), and T is the reaction time (d).

    1.8.3   Calculation of sodium acetate content (ns, g·L-1) in the substrate

    The content of sodium acetate in the reaction system was calculated using the standard curve of sodium acetate (Fig.S4), as follows:

    $ y=5.828\;8 x+0.111\;73 \quad\left(R^2=0.998\;18\right) $

    (3)

    where y is the absorbance value measured using a UV-visible spectrophotometer and x represents the ns.

    1.8.4   Calculation of CE (%) and CHR (%)

    $ \mathrm{CE}=\frac{M \int_0^t I \mathrm{d} t}{F b V n_{\mathrm{s}}} $

    (4)

    $ \mathrm{CHR}=\frac{n_{\mathrm{H}_2}}{n_{\mathrm{CE}}}=\frac{n_{\mathrm{H}_2}}{\int_0^t I \mathrm{d} t} $

    (5)

    where M is the molar mass of sodium acetate (82 g·mol-1), I is the electric current generated in MECs (A), F is Faraday constant (96 485 C·mol-1), t is the reaction time, and b is the number of electrons transferred (b=8), nH2 is the amount of substance of hydrogen that could be theoretically produced, nCE is the amount of substance of hydrogen that could be produced from the measured current.

    1.8.5   Calculation of energy recovery

    The energy efficiencies based on input electricity (ηE, %), energy efficiency related to the consumed substrate (ηS, %), and overall energy efficiency (ηE+S, %) were calculated with Eq.6-8:

    $ \eta=\frac{W_{\mathrm{H}_2}}{W_{\mathrm{E}}} $

    (6)

    $ \eta=\frac{W_{\mathrm{H}_2}}{W_{\mathrm{s}}} $

    (7)

    $ \eta_{\mathrm{E}+\mathrm{S}}=\frac{W_{\mathrm{H}_2}}{W_{\mathrm{E}}+W_{\mathrm{S}}} $

    (8)

    where WH2 is the heat of combustion of the hydrogen produced, calculated with WH2 = ΔHH2nH2HH2=285.83 kJ·mol-1, is the hydrogen heat value); WE is the energy added to the circuit by the power source, computed with $ W_{\mathrm{E}}=E_{\text {ap }} \int_0^t I \mathrm{d} t$ (Eap is the applied voltage to MECs); WS is the heat of combustion of the substrate (acetate), calculated as WSHSNSHS=870.28 kJ·mol-1, is the substrate heat value). In this study, the volumetric current density (A·m-2) was calculated by normalizing the current with the volume of the cathode chamber.

    The XRD patterns of the catalyst MoS2/CuS are clearly shown in Fig. 1. The main characteristic peaks of 14.378°, 29.026°, 35.870°, and 44.151° in 2θ correspond to (002), (004), (102) and (006) crystal panes of MoS2, respectively. The peaks of MoS2 (PDF No.37-1492) and CuS (PDF No.06-0464) can also be found in the XRD patterns of MoS2/CuS, which indicates the successful preparation. Moreover, it can be seen from the figure that the diffraction peaks of MoS2/CuS were sharp and intense compared with pure MoS2, indicating its highly crystalline nature.

    Figure 1

    Figure 1.  XRD patterns of MoS2 and MoS2/CuS

    XPS analysis was employed to investigate both elemental compositions of the electrode surface and their corresponding oxidation states ascribing the electronic configuration of respective atoms. Fig. 2a shows the XPS survey spectra of MoS2 and MoS2/CuS-20%, where element C (284.40 eV) was the background element used for measurement. It is easy to find that Mo, S, and Cu elements were distributed on the surface of MoS2/CuS-20%. To further confirm the specific chemical states of Mo, S, and Cu, the fine spectra of these three elements were also analyzed. From Fig. 2b, the two peaks at 228.87 and 231.92 eV correspond to Mo4+3d5/2 and Mo4+3d3/2 respectively, and also to the 2H phase of MoS2. The peaks at 229.97 and 233.32 eV correspond to Mo4+3d5/2 and Mo4+3d3/2 respectively, and also to the 1T phase of MoS2. The peaks at 232.74 and 235.80 eV correspond to Mo6+3d5/2 and Mo6+3d3/2, which are bound to form Mo—O bonds due to the presence of oxygen in the air[27]. The peak at 226.20 eV corresponds to the S2s[28-29]. Cu was not detected in MoS2, and the Cu2p spectrum of MoS2/CuS-20% is shown in Fig. 2c. The peaks at 932.44 and 952.10 eV correspond to Cu+2p3/2 and Cu+2p1/2, respectively, and the presence of Cu+ suggests that a small amount of Cu2S is generated. The peaks at 933.93 and 953.12 eV correspond to Cu2+2p3/2 and Cu2+2p1/2, but it can be seen that the intensity of the Cu2+ peak was higher than that of the Cu+ peak. This indicates that Cu2+ was dominant and most of the Cu was bound to form CuS. The residual binding energy of 956.54 eV corresponds to the satellite peak of Cu2+2p1/2 and the peak at 945.80 eV as well as 949.96 eV corresponds to the satellite peaks of Cu+2p1/2. Fig. 2d shows the fine XPS spectra of S2p in MoS2 and MoS2/CuS-20%. It can be seen that the binding energies of S2-2p3/2 and S2-2p1/2 are 161.70 and 162.90 eV in the 1T phase of MoS2, 163.18 and 164.44 eV in the 2H phase of S2-2p3/2 and S2-2p1/2. This indicates that MoS2 in both MoS2 and MoS2/CuS-20% existed as a mixture of 1T and 2H phases, and the addition of CuS did not change the existence morphology of MoS2.

    Figure 2

    Figure 2.  XPS spectra of MoS2 and MoS2/CuS-20%: (a) survey; (b) Mo3d; (c) Cu2p; (d) S2p

    Fig. 3a-3c shows the SEM images of MoS2 and MoS2/CuS-20%, it can be seen that pure MoS2 was a nanospheres composed of nanosheets. After the composite of 20% CuS, the overall morphology of the nanospheres did not change significantly, but it was observed that the nanospheres made of nanosheets had more folds, which suggests that MoS2/CuS-20% had more edge active sites, that is, the density of active sites increased. It is well known that the HER active site of MoS2 is located at the edge, and its basal surface is catalytically inert with poor HER performance[30].

    Figure 3

    Figure 3.  SEM image of (a) MoS2 and (b, c) MoS2/CuS-20%; TEM images of (d, e) MoS2 and (f, g) MoS2/CuS-20%; (h) High-resolution TEM image of MoS2/CuS-20%; (i) EDS mappings of Mo, S, and Cu in MoS2/CuS-20%; Lattice structure image of (j) CuS (108) plane and (k) MoS2 (102) plane

    Fig. 3d and 3e are the TEM images of the MoS2 nanosheet. Fig. 3e illustrates the TEM image of the rectangular region from Fig. 3d. Fig. 3f-3h all show TEM images of MoS2/CuS-20%, where Fig. 3g shows a localized enlargement of the rectangular region in Fig. 3f. By observing and comparing the TEM images of the two catalyst nanosheets, it can be seen that the crystal features of MoS2/CuS-20% were more obvious, which is consistent with the results of XRD. Compared with pure MoS2, the surface of MoS2/CuS-20% nanosheets was rougher and more disordered, which weakened the adhesion of bubbles at the catalyst interface and promoted the timely release of H2 bubbles[31]. In addition, clear lattice fringes can be observed in the high-resolution TEM of Fig. 3h, in which 0.251 and 0.174 nm correspond to the (102) crystall plane of MoS2 and the (108) crystal plane of CuS, respectively, which matches with the XRD results and also demonstrates the successful composite of MoS2 and CuS. The lattice structure images of the CuS (108) crystal plane and MoS2 (102) crystal plane are shown in Fig. 3j and 3k, respectively, and more detailed images of the lattice fringes are shown in Fig.S5. In addition, the EDS elemental distribution of MoS2/CuS-20% was shown in Fig. 3i, and the uniform distribution of Mo, Cu, and S elements can be seen, which further indicates that we successfully synthesized MoS2/CuS-20%.

    The LSV polarization curves of all catalysts measured are shown in Fig. 4a. Among them, the MoS2/CuS-20% sample showed remarkable HER performance with an overpotential of 135 mV when the current density was increased to 1 mA·cm-2 (Fig. 4b), which was much smaller than the other catalysts. The Tafel slopes of these samples were obtained by fitting the polarization curves, as shown in Fig. 4c. The Tafel slope of MoS2/CuS-20% (100.34 mV·dec-1) was lower than that of Pt mesh (110.89 mV·dec-1), MoS2 (206.43 mV·dec-1), MoS2/CuS-5% (186.22 mV·dec-1), MoS2/CuS-10% (157.25 mV·dec-1), and MoS2/CuS-30% (129.86 mV·dec-1). By comparing these values, the specific path of HER can be confirmed as the Volmer-Heyrovsky mechanism[32], and the smaller the Tafel value, the better the HER of the corresponding catalyst.

    Figure 4

    Figure 4.  (a) LSV, (b) overpotential, (c) Tafel slope curves, and (d) EIS of Pt mesh, MoS2, and MoS2/CuS

    Inset: the fitted equivalent circuit diagrams.

    The EIS can reflect the charge transfer kinetics in HER (Fig. 4d). Comparing the semicircular diameter of different electrodes, it can be found that MoS2/CuS-20% had the smallest radius in EIS, which had the smallest charge-transfer resistance, fast electron transfer, and high catalytic activity[33]. The measured impedance data were fitted using ZView based on the circuit diagram shown in Fig. 4d. Detailed impedance fitting data is shown in Table S4. The fitting equivalent circuits include charge transfer resistance (Rct), solution resistance (Rs), constant phase element (CPE), and solid phase diffusion Warburg (ZW) resistance. The calculated Rct of MoS2/CuS-20% was minimized to 34.40 Ω, which was smaller than the charge transfer resistance of Pt mesh of 41.42 Ω, indicating that MoS2/CuS-20% had a stronger charge transfer capability compared to Pt mesh. The double-layer capacitance (Cdl) values were used to evaluate the ECSA of the catalysts, and ECSA=Cdl/Cs, where Cs is the specific capacitance of the correspondent surface-smoothed sample under the same conditions, and the Cdl values were proportional to the ECSA. As shown in Fig. 5a-5c, the capacitive currents of Pt, MoS2, and MoS2/CuS-20% were measured in the potential range of non-Faraday intervals and the obtained capacitance currents were plotted as a function of different scanning speeds[34]. The capacitance currents obtained were plotted as a function of different scan rates with a slope of Cdl[35], as shown in Fig. 5d. Compared with MoS2, Pt, and MoS2/CuS-20% produced the highest Cdl (1.110 0 mF·cm-2) with the largest ECSA, which implied that the addition of a certain amount of CuS gives MoS2 more active reactive sites and a larger electrochemically active surface area to promote the electrocatalytic process effectively.

    Figure 5

    Figure 5.  (a-c) CV curves and (d) the fitted Cdl of Pt mesh, MoS2, and MoS2/CuS-20%

    Δj was the current density difference at 0.687 7 V (vs RHE) in CV.

    The MoS2, MoS2/CuS-5%, MoS2/CuS-10%, MoS2/CuS-20%, MoS2/CuS-30%, and Pt mesh cathodes were assembled in MECs to investigate their hydrogen evolution performance at an applied voltage of 0.8 V, 37 ℃ and 1.5 g·L-1 sodium acetate were studied in MEC. The HPR was calculated based on the amount of hydrogen produced. The average HPR of the MECs with MoS2/CuS-20% cathode was 0.203 1 mH23·m-3·d-1, which was even slightly higher than that of the MEC with Pt mesh cathode (0.188 6 mH23·m-3·d-1) (Fig. 6a). From the overall current trend of MEC, the prepared cathode catalysts with MoS2/CuS-20% produced a higher current density (Fig. 6b). The MEC with MoS2/CuS-20% as the cathode produced a maximum current density of 8.25 A·m-2, which was also higher than the maximum current density produced with the Pt mesh as the cathode (5.50 A·m-2). The decrease in current density during the experiment may be due to the gradual consumption of sodium acetate substrate and the decrease of available carbon source of microorganisms, which led to the slowdown of their metabolism and the decrease of HPR[33].

    Figure 6

    Figure 6.  (a) HPY and HPR of Pt mesh, MoS2, and MoS2/CuS and (b) variation of current density with reaction time during the reaction of Pt mesh, MoS2, and MoS2/CuS as cathodic catalysts

    To investigate the stability of the catalyst, we performed 12 h-chronoamperometry tests (Fig. 7a), and a MEC hydrogen production cycling experiment was carried out using MoS2/CuS-20% as the cathode (Fig. 7b). It is clear from Fig. 7a that the potential change was negligible after 12 h of chronoamperometry test, which indicates that MoS2/CuS-20% had high stability. The results of the cycling experiments also showed good cycling ability of MoS2/CuS-20% in MEC (Fig. 7b). After each cycle, anolyte and catholyte were replaced and 1.5 g·L-1 sodium acetate was added. There was no significant decrease in current density over the three cycles, indicating that the MoS2/CuS-20% cathode had high stability and reproducibility in MEC.

    Figure 7

    Figure 7.  (a) Current density of MoS2/CuS-20% as the cathode for cyclic reaction three times in the MEC; (b) Chronopotentiometry curve of MoS2/CuS-20% at-1 mA·cm-2

    The performance of the MEC was evaluated by CE, CHR, and energy recovery. The CEs of MoS2/CuS-20% and Pt mesh were calculated to be 16.07% and 10.33%, respectively, based on the current measured over time and the sodium acetate content, as shown in Fig. 8a. The recorded current data indicate that the MoS2/CuS-20% catalyst had higher CE values, while the CHR was lower than that of the Pt mesh. The ηE of the MECs were all higher than 100%. The ηE of Pt as a cathode reached 587.63%, which was higher than that of MoS2/CuS-20% (422.8%). As shown in Fig. 8b, the ηs and ηE+S of MoS2/CuS-20% were higher than that of Pt mesh. The above data can also be used to illustrate that the HPR and hydrogen production performance of MoS2/CuS-20% can be compared with that of Pt mesh.

    Figure 8

    Figure 8.  (a) CE and CHR of the MEC hydrogen production system of catalysts and (b) corresponding ηE, ηs, and ηE+S

    Our microorganisms were taken from coking wastewater, and wastewater usually contains a large number of microorganisms, which are very suitable for biological treatment[36-38]. It is well known that the electrochemically active microorganisms attached to the anode surface play a major role in the MEC hydrogen production system. Therefore, the microbial community of anode biofilm was analyzed by macrogenetic sequencing. The length distribution of high-quality sequences contained in all samples was statistically analyzed using R language scripts (Fig.S6a), laying the groundwork for subsequent diversity analyses of species composition. To evaluate the diversity of the reactor anode biofilm community, the Shannon diversity evaluation index of the microbial community in each group was evaluated (Fig.S6b). The gentleness of the curve indicates that the sequencing results can adequately reflect the diversity contained in the current samples[39-40].

    Based on the statistical map of the number of taxa of microorganisms at each level (Fig. 9a), we analyzed the distribution at the phylum (Fig. 9b) and genus (Fig. 9c) levels to investigate the microbial community of the anode biofilm. The phylum-level shows that Proteobacteria (34.5%), Actinobacteria (23.6%), and Chloroflexi (12.7%) were the dominant bacterium in anode biofilm. In the coking wastewater, 5.5% of Bacteroidotas and 1% of Firmicutes were also detected. Proteobacteria are well known for playing important roles in extracellular electron transfer[41], and Firmicutes were commonly reported in dark fermentative hydrogen production systems[42-44]. At the genus level, JG30-KF-CM45 accounts for 4%, followed by 3.6% Methylotenera, 2.7% Nitrospira, and 2.4% Blastococcus. At present, we have not studied these species deeply enough and there are few reports in the literature. The analysis of microbial phylum and genus in coking wastewater shows that the anode biofilm is composed of mixed microbial communities, and there are some unknown species, so the molecular mechanism of microorganisms in electron transport has not been studied in this paper.

    Figure 9

    Figure 9.  Microbial community of the anode biofilm: (a) statistical map of the number of microbial taxa at each level; Percentage of the average relative abundance of phylum (b) and genus (c) levels

    Dominant biofilm genera (relative abundance >2%) are shown individually, while the rest are grouped as others in b and c.

    In summary, the MoS2/CuS-20% composite catalyst was prepared by a one-step hydrothermal method. The addition of a certain amount of CuS made the structure of MoS2 more dense and the surface more rough and disordered, which not only increased the edge active site of MoS2 but also facilitated the release of hydrogen on the catalyst surface. Electrochemical tests showed that MoS2/CuS-20% had a Tafel slope of 100.34 mV·dec-1, and its conductivity and electroactive area are better than pure MoS2 and Pt mesh. The HPR of MoS2/CuS-20% was (0.203 1±0.023 7) mH23·m-3·d-1, which was slightly higher than that of the Pt mesh in a 72 h-MEC hydrogen production system. The feasibility of MoS2/CuS-20% as an alternative catalyst for Pt mesh in MEC has been demonstrated, and it is expected to be an excellent low-cost catalyst for hydrogen production in MEC in the future.

    Supporting information is available at http://www.wjhxxb.cn


    1. [1]

      Li Y F, Ren N Q, Yang C P, Wang A J, Zadsar M, Li J Z, Hu L J. Molecular characterization and hydrogen production of a new species of anaerobe[J]. J. Environ. Sci. Health, 2005, 40:  1929-1938. doi: 10.1080/10934520500184483

    2. [2]

      Ren N, Li J Z, Li B K, Wang Y, Liu S R. Biohydrogen production from molasses by anaerobic fermentation with a pilot-scale bioreactor system[J]. Int. J. Hydrog. Energy, 2006, 31:  2147-2157. doi: 10.1016/j.ijhydene.2006.02.011

    3. [3]

      Kapdan I K, Kargi F. Bio-hydrogen production from waste materials[J]. Enzyme Microb. Technol., 2006, 38:  569-582. doi: 10.1016/j.enzmictec.2005.09.015

    4. [4]

      Wang J P, Song Y N, Hu J, Li Y, Wang Z Y, Yang P, Wang G, Ma Q, Che Q D, Dai Y, Huang B B. Photocatalytic hydrogen evolution on P-type tetragonal zircon BiVO4[J]. Appl. Catal. B-Environ., 2019, 251:  94-101. doi: 10.1016/j.apcatb.2019.03.049

    5. [5]

      Zhang B B, Dong G J, Wang L, Zhang Y J, Ding Y, Bi Y P. Efficient hydrogen production from MIL-53(Fe) catalyst-modified Mo: BiVO4 photoelectrodes[J]. Catal. Sci. Technol., 2017, 7:  4971-4976. doi: 10.1039/C7CY01765K

    6. [6]

      Guan X, Zhou Z, Luo P, Wu F S, Dong S J. Effects of preparation method on the hydrolytic hydrogen production performance of Al-rich alloys[J]. J. Alloy. Compd., 2019, 796:  210-220. doi: 10.1016/j.jallcom.2019.05.053

    7. [7]

      Logan B E, Call D, Cheng S, Hamelers H V M, Sleutels T H J A, Jeremiasse A W, Rozendal R A. Microbial electrolysis cells for high yield hydrogen gas production from organic matter[J]. Environ. Sci. Technol., 2008, 42(23):  8630-8640. doi: 10.1021/es801553z

    8. [8]

      代红艳, 杨慧敏, 刘宪, 简选, 郭敏敏, 曹乐乐, 杨镇海. MoS2/石墨烯复合阴极材料的制备及微生物电解池催化产氢性能[J]. 高等学校化学学报, 2018,39,(2): 351-358. DAI H Y, YANG H M, LIU X, JIAN X, GUO M M, CAO L L, YANG Z H. Preparation of MoS2/graphene composite cathode materials and hydrogen production performance catalyzed by microbial electrolytic cell[J]. Chem. J. Chinese Universities, 2018, 39(2):  351-358.

    9. [9]

      Geelhoed J S, Hamelers H V M, Stams A J M. Electricity-mediated biological hydrogen production[J]. Curr. Opin. Microbiol., 2010, 13:  307-315. doi: 10.1016/j.mib.2010.02.002

    10. [10]

      Savla N, Guin M, Pandit S, Malik H, Khilari S, Mathuriya A S, Gupta P K, Thapa B S, Bobba R, Jung S P. Recent advancements in the cathodic catalyst for the hydrogen evolution reaction in microbial electrolytic cells[J]. Int. J. Hydrog. Energy, 2022, 47:  15333-15356. doi: 10.1016/j.ijhydene.2022.03.058

    11. [11]

      Ivanov I, Ahn Y T, Poirson T, Hickner M A, Logan B E. Comparison of cathode catalyst binders for the hydrogen evolution reaction in microbial electrolysis cells[J]. Int. J. Hydrog. Energy, 2017, 42:  15739-15744.

    12. [12]

      Zhang H, Meng G, Wei T R, Ding J Y, Liu Q, Luo J, Liu X J. Co doping promotes the alkaline overall seawater electrolysis performance over MnPSe3 nanosheets[J]. Chem. Commun., 2023, 59:  12144-12147.

    13. [13]

      Liu W X, Niu X X, Tang J W, Liu Q, Luo J, Liu X J, Zhou Y T. Energy-efficient anodic reactions for sustainable hydrogen production via water electrolysis[J]. Chem. Synth., 2023, 3:  44.

    14. [14]

      Qin Y J, Cao H J, Liu Q, Yang S Q, Feng X C, Wang H, Lian M L, Zhang D X, Wang H, Luo J, Liu X J. Multi-functional layered double hydroxides supported by nanoporous gold toward overall hydrazine splitting[J]. Front. Chem. Sci. Eng., 2023, 18(1):  6.

    15. [15]

      Liu B, Zhao W J, Ding Z J, Verzhbitskiy I, Li L J, Lu J P, Chen J Y, Eda G, Loh K P. Engineering bandgaps of monolayer MoS2 and WS2 on fluoropolymer substrates by electrostatically tuned many-body effects[J]. Adv. Mater., 2016, 28(30):  6457-6464.

    16. [16]

      Naylor C H, Kybert N J, Schneier C, Xi J, Romero G, Saven J G, Liu R Y, Johnson A T C. Scalable production of molybdenum disulfide based biosensors[J]. ACS Nano, 2016, 10(6):  6173-6179.

    17. [17]

      Hwang J H, Fahad S, Ryu H, Rodriguez K L, Domingo J S, Kushima A, Lee W H. Recycling urine for bioelectrochemical hydrogen production using a MoS2 nano carbon coated electrode in a microbial electrolysis cell[J]. J. Power Sources, 2022, 527:  231209.

    18. [18]

      Zhou M, Zhang R, Huang M A, Lu W, Song S L, Melancon M P, Tian M, Liang D, Li C. A chelator-free multifunctional[64Cu] CuS nanoparticle platform for simultaneous micro-PET/CT imaging and photothermal ablation therapy[J]. J. Am. Chem. Soc., 2010, 132:  15351-15358.

    19. [19]

      Basu M, Sinha A K, Pradhan M, Sarkar S, Negishi Y, Govind , Pal T. Evolution of hierarchical hexagonal stacked plates of CuS from liquid-liquid interface and its photocatalytic application for oxidative degradation of different dyes under indoor lighting[J]. Environ. Sci. Technol., 2010, 44:  6313-6318.

    20. [20]

      Chunga J S, Sohnb H J. Electrochemical behaviors of CuS as a cathode material for lithium secondary batteries[J]. J. Power Sources, 2002, 108:  226-231.

    21. [21]

      Kim W B, Lee S H, Cho M, Lee Y. Facile and cost-effective CuS dendrite electrode for non-enzymatic glucose sensor[J]. Sensors Actuat. B: Chem., 2017, 249:  161-167.

    22. [22]

      Wu X Q, Huang D D, Wu Y P, Zhao J, Liu X, Dong W W, Li S, Li D S, Li J R. In situ synthesis of nano CuS-embedded MOF hierarchical structures and application in dye adsorption and hydrogen evolution reaction[J]. ACS Appl. Energy Mater., 2019, 2:  5698-5706.

    23. [23]

      Wang Y C, Chao D L, Wang Z Z, Ni J F, Li L. An energetic CuS-Cu battery system based on CuS nanosheet arrays[J]. ACS Nano, 2021, 15:  5420-5427.

    24. [24]

      Liu S Q, Wen H R, Ying G, Zhu Y W, Fu X Z, Sun R, Wong C P. Amorphous Ni(OH)2 encounter with crystalline CuS in hollow spheres: A mesoporous nano-shelled heterostructure for hydrogen evolution electrocatalysis[J]. Nano Energy, 2018, 44:  7-14.

    25. [25]

      石倩萍, 程玉周, 黄韬, 孙强, 夏彬勇. 分光光度法测定污水处理用乙酸钠含量的研究[J]. 能源与环境, 2002,5,78-80. SHI Q P, CHENG Y Z, HUANG T, SUN Q, XIA B Y. Determination of sodium acetate for sewage treatment by spectrophotometric method[J]. Energy and Environment, 2002, 5:  78-80.

    26. [26]

      Dong Z S, Zhao Y, Fan L, Wang Y X, Wang J W, Zhang K. Simultaneous sulfide removal and hydrogen production in a microbial electrolysis cell[J]. Int. J. Electrochem. Sci., 2017, 12:  10553-10566.

    27. [27]

      Kim M, Anjum M A R, Choi M, Jeong H Y, Choi S H, Park N, Lee J S. Covalent 0D-2D heterostructuring of Co9S8-MoS2 for enhanced hydrogen evolution in all pH electrolytes[J]. Adv. Funct. Mater., 2020, 30:  2002536.

    28. [28]

      Kim M S, Tran D T, Nguyen T H, Dinh V A, Kim N H, Lee J H. Ni single atoms and Ni phosphate clusters synergistically triggered surface-functionalized MoS2 nanosheets for high-performance freshwater and seawater electrolysis[J]. Energy Environ. Mater., 2022, 5:  1340-1349.

    29. [29]

      Zhou Q S, Feng J R, Peng X W, Zhong L X, Sun R C. Porous carbon coupled with an interlaced MoP-MoS2 heterojunction hybrid for efficient hydrogen evolution reaction[J]. J. Energy Chem., 2020, 45:  45-51.

    30. [30]

      Zhang X, Jia F F, Song S X. Recent advances in structural engineering of molybdenum disulfide for electrocatalytic hydrogen evolution reaction[J]. Chem. Eng. J., 2021, 405:  127013.

    31. [31]

      Lu Z Y, Li Y J, Lei X D, Liu J F, Sun X M. Nanoarray based "superaerophobic" surfaces for gas evolution reaction electrodes[J]. Mater. Horiz., 2015, 2:  294-298.

    32. [32]

      An L, Huang L, Zhou P P, Yin J, Liu H Y, Xi P X. A self-standing high-performance hydrogen evolution electrode with nanostructured NiCo2O4/CuS heterostructures[J]. Adv. Funct. Mater., 2015, 25:  6814-6822.

    33. [33]

      Li F J, Liu W F, Sun Y, Ding W J, Cheng S A. Enhancing hydrogen production with Ni-P coated nickel foam as cathode catalyst in single chamber microbial electrolysis cells[J]. Int. J. Hydrog. Energy, 2017, 42:  3641-3646.

    34. [34]

      Liu L, Liu X L, Jiao S L. CuS@defect-rich MoS2 core-shell structure for enhanced hydrogen evolution[J]. J. Colloid Interface Sci., 2020, 564:  77-87.

    35. [35]

      Ren X, Wu D, Ge R X, Sun X, Ma H M, Yan T, Zhang Y, Du B, Wei Q, Chen L. Self-supported CoMoS4 nanosheet array as an efficient catalyst for hydrogen evolution reaction at neutral pH[J]. Nano Res., 2018, 11:  2024-2033.

    36. [36]

      Liu W Z, Wang A J, Ren N Q, Zhao X Y, Liu L H, Yu Z G, Le D J. Electrochemically assisted biohydrogen production from acetate[J]. Energy Fuel, 2008, 22:  159-163.

    37. [37]

      Kiely P D, Cusick R, Call D F, Selembo P A, Regan J M, Logan B E. Anode microbial communities produced by changing from microbial fuel cell to microbial electrolysis cell operation using two different wastewaters[J]. Bioresour. Technol., 2011, 102:  388-394.

    38. [38]

      Liu W Z, Wang A J, Sun D, Ren N Q, Zhang Y Q, Zhou J Z. Characterization of microbial communities during anode biofilm reformation in a two-chambered microbial electrolysis cell (MEC)[J]. J. Biotechnol., 2012, 157:  628-632.

    39. [39]

      Shannon B C E. A mathematical theory of communication[J]. Bell System Technical Journal., 1948, 27:  379-423.

    40. [40]

      Shannon B C E. A mathematical theory of communication[J]. Bell System Technical Journal., 1948, 27:  623-656.

    41. [41]

      Fykse E M, Aarskaug T, Madslien E H, Dybwad M. Microbial community structure in a full-scale anaerobic treatment plant during start-up and first year of operation revealed by high-throughput 16S rRNA gene amplicon sequencing[J]. Bioresour. Technol., 2016, 222:  380-387.

    42. [42]

      Cho S K, Jeong M W, Choi Y K, Shin J, Shin S G. Effects of low-strength ultrasonication on dark fermentative hydrogen production: Start-up performance and microbial community analysis[J]. Appl. Energy, 2018, 219:  34-41.

    43. [43]

      Liu Z D, Zhang C, Wang L J, He J W, Li B M, Zhang Y H, Xing X H. Effects of furan derivatives on biohydrogen fermentation from wet steam-exploded cornstalk and its microbial community[J]. Bioresour. Technol., 2015, 175:  152-159.

    44. [44]

      Vesga-Baron A, Etchebehere C, Schiappacasse M C, Chamy R, Tapia-Venegas E. Controlled oxidation-reduction potential on dark fermentative hydrogen production from glycerol: Impacts on metabolic pathways and microbial diversity of an acidogenic sludge[J]. Int. J. Hydrog. Energy, 2021, 46:  5074-5084.

  • Figure 1  XRD patterns of MoS2 and MoS2/CuS

    Figure 2  XPS spectra of MoS2 and MoS2/CuS-20%: (a) survey; (b) Mo3d; (c) Cu2p; (d) S2p

    Figure 3  SEM image of (a) MoS2 and (b, c) MoS2/CuS-20%; TEM images of (d, e) MoS2 and (f, g) MoS2/CuS-20%; (h) High-resolution TEM image of MoS2/CuS-20%; (i) EDS mappings of Mo, S, and Cu in MoS2/CuS-20%; Lattice structure image of (j) CuS (108) plane and (k) MoS2 (102) plane

    Figure 4  (a) LSV, (b) overpotential, (c) Tafel slope curves, and (d) EIS of Pt mesh, MoS2, and MoS2/CuS

    Inset: the fitted equivalent circuit diagrams.

    Figure 5  (a-c) CV curves and (d) the fitted Cdl of Pt mesh, MoS2, and MoS2/CuS-20%

    Δj was the current density difference at 0.687 7 V (vs RHE) in CV.

    Figure 6  (a) HPY and HPR of Pt mesh, MoS2, and MoS2/CuS and (b) variation of current density with reaction time during the reaction of Pt mesh, MoS2, and MoS2/CuS as cathodic catalysts

    Figure 7  (a) Current density of MoS2/CuS-20% as the cathode for cyclic reaction three times in the MEC; (b) Chronopotentiometry curve of MoS2/CuS-20% at-1 mA·cm-2

    Figure 8  (a) CE and CHR of the MEC hydrogen production system of catalysts and (b) corresponding ηE, ηs, and ηE+S

    Figure 9  Microbial community of the anode biofilm: (a) statistical map of the number of microbial taxa at each level; Percentage of the average relative abundance of phylum (b) and genus (c) levels

    Dominant biofilm genera (relative abundance >2%) are shown individually, while the rest are grouped as others in b and c.

  • 加载中
计量
  • PDF下载量:  0
  • 文章访问数:  112
  • HTML全文浏览量:  22
文章相关
  • 发布日期:  2024-09-10
  • 收稿日期:  2024-02-18
  • 修回日期:  2024-07-11
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

/

返回文章