Next Article in Journal
Evolutions on Microstructure and Impact Toughness of G115 Steel after Long-Term Aging at 700 °C
Previous Article in Journal
Effect of Temperature and Strain on Bonding of Similar AA3105 Aluminum Alloys by the Roll Bonding Process
Previous Article in Special Issue
Thermodynamic Simulation Model of Copper Side-Blown Smelting Process
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

An Efficient and Stable MXene-Immobilized, Cobalt-Based Catalyst for Hydrogen Evolution Reaction

School of Chemistry and Chemical Engineering, Jiangxi University of Science and Technology, Ganzhou 341000, China
*
Author to whom correspondence should be addressed.
Metals 2024, 14(8), 922; https://doi.org/10.3390/met14080922
Submission received: 3 July 2024 / Revised: 6 August 2024 / Accepted: 12 August 2024 / Published: 14 August 2024

Abstract

:
Hydrogen (H2) is considered to be the best carbon-free energy carrier that can replace fossil fuels because of its high energy density and the advantages of not producing greenhouse gases and air pollutants. As a green and sustainable method for hydrogen production, the electrochemical hydrogen evolution reaction (HER) has received widespread attention. Currently, it is a great challenge to prepare economically stable electrocatalysts for the HER using non-precious metals. In this study, a Co/Co3O4/Ti3C2Tx catalyst was synthesized by supporting Co/Co3O4 with Ti3C2Tx. The results show that Co/Co3O4/Ti3C2Tx has excellent HER activity and durability in 1 mol L−1 KOH, and the overpotential and Tafel slope at 10 mA·cm−2 were 87 mV and 61.90 mV dec−1, respectively. The excellent HER activity and stability of Co/Co3O4/Ti3C2Tx can be explained as follows: Ti3C2Tx provides a stable skeleton and a large number of attachment sites for Co/Co3O4, thus exposing more active sites; the unique two-dimensional structure of Ti3C2Tx provides an efficient conductive network for rapid electron transfer between the electrolyte and the catalyst during electrocatalysis; Co3O4 makes the Co/Co3O4/Ti3C2Tx catalyst more hydrophilic, which can accelerate the release rate of bubbles; Co/Co3O4 can accelerate the adsorption and deionization of H2O to synthesize H2. This study provides a new approach for the design and preparation of low-cost and high-performance HER catalysts.

1. Introduction

In order to deal with the energy crisis and environmental protection issues, countries around the world have been increasing their attention to new clean energy. Among many kinds of new clean energy sources available on the globe, hydrogen (H2) is considered to be the best carbon-free energy carrier that can replace fossil fuels. In the current hydrogen production technology, the electrochemical decomposition of water to produce hydrogen is considered to be one of the best and the most efficient and greenest hydrogen production methods [1,2,3]. An electrochemical water decomposition reaction is composed of the oxygen evolution reaction (OER) and hydrogen evolution reaction (HER), in which the HER is the key step. A catalyst that can reduce the reaction barrier and energy consumption, increase the reaction efficiency and improve the reaction rate plays a very important role in HER process. Therefore, the study of efficient electrocatalysts has become the most important part of HER technology exploration.
Numerous studies have found that precious metals such as platinum (Pt), ruthenium (Ru), palladium (Pd), gold (Au) and iridium (Ir) have more suitable ΔGH* values than other metallic elements. This may be due to the fact that precious metals have a surface defect, which makes the active substance easily adsorbed on their surface and facilitates the formation of active intermediates [4]. Pt-based material is currently recognized as the best HER electrocatalyst, and its ΔGH* value is almost zero. However, the limited reserves and high price of Pt limit its large-scale application, and therefore, it is not suitable for commercial hydrogen production [5]. Therefore, it is urgent to develop high-efficiency and low-cost HER electrocatalysts.
At present, a large number of studies have found that transition metals (such as Fe, Co, Ni, etc.) and transition metal oxides show high electrocatalytic activity for HER in alkaline solutions and are attractive candidates for HER electrocatalyst. Compared with the small global reserves and high market prices of precious metals, transition metals have abundant global reserves and cost advantages. The ΔGH* value of Co is moderate, and Co-based materials such as cobalt phosphide, cobalt selenide, cobalt sulfide and cobalt layered double hydroxides (Co-LDHs) have diversified characteristics in terms of their composition, crystal structure and reversible morphology. Results show that Co-based materials have great potential for the HER. However, despite the advantages of Co-based materials in terms of HER catalytic activity, there are still some shortcomings in their industrial applications. For example, cobalt oxide can dissolve in acidic solutions, resulting in the loss of active sites; agglomeration often occurs during the preparation of cobalt selenide; low conductivity can reduce the HER activity of cobalt sulfide, cobalt phosphide and Co-LDHs [6]. In order to realize the practical application of cobalt-based materials in the HER, some researchers have tried various strategies, such as constructing nanostructures, carbon material loading, morphological control and doping. For example, Wang et al. synthesized Co@CoO nanowires and found that they are excellent catalysts for hydrogen production at low overpotential [7]. He et al. improved the HER efficiency of Co-Ni phosphide (Co-Ni-P) in an alkaline solution by anchoring hollow Co-Ni-P nanospheres on reduced graphene oxide nanosheets (RGO) [8]. Gao et al. synthesized a nuclear/shell structure Co3O4@CoS2 nanoarray with excellent HER activity in acidic solutions [9]. Zhang et al. synthesized a highly efficient HER electrocatalyst, Ni2P/CoP@C-NSG, suitable for acidic solutions by loading highly dispersed Ni2P and CoP onto carbon nanospheres and growing Ni2P/CoP on a N and S co-doped graphene substrate [10]. In general, transition metal oxides are good for the dissociation of water but bad for the recombination of adsorbed hydrogen atoms (Had) to form H2. However, transition metals can effectively adsorb hydrogen atoms and promote the conversion of Had to H2. Therefore, researchers have set out to construct transition metal/metal oxide catalysts suitable for the HER. For example, Wang et al. reported a cobalt-based electrocatalyst, Co/CoO/CoMoO3, with a highly dispersed metal/metal oxide active component and a tight series structure that together promoted the efficient conduct of the HER under a high current density in a medium/alkaline electrolyte. Through experiments and DFT calculations, it was found that the strong OH adsorption at CoO site promotes the dissociation of water, and the appropriate H adsorption at Co sites promotes the precipitation of hydrogen, thus achieving the high current density of the HER in alkaline and neutral media [11].
In recent years, a large number of studies have found that the introduction of suitable substrates to disperse and anchor Co-based catalytic species is a feasible way to improve its utilization efficiency, because it can expose as many active points on the catalyst surface as possible [12,13,14,15]. Therefore, the HER performance of Co-based materials can be improved by selecting a suitable carrier [16]. MXene (Ti3C2Tx) is the general term for a series of novel two-dimensional (2D) transition metal carbide and carbon nitride materials with graphene-like structures. Compared to traditional three-dimensional (3D) catalysts, Ti3C2Tx has a larger surface area and can provide a large number of attachment points for active substances. At the same time, the good electrical conductivity of Ti3C2Tx can accelerate the transfer of electrons from the electrode to the catalyst surface. Therefore, it may be a suitable carrier for loading cobalt-based active components. However, due to the influence of van der Waals forces, Ti3C2Tx nanosheets are easy to accumulate or stack, which easily leads to the catalytic active site being covered up. It is worth noting that the layer spacing of MXene obtained via HF etching of the MAX phase can be expanded. The large layer spacing facilitates the entry of active substances into the interior of Ti3C2Tx, and the loaded active components can play a role in supporting the structure to prevent the stacking of Ti3C2Tx. Therefore, this study attempted to use Ti3C2Tx loaded with Co-based active components. Through the above attempts, the HER activity of the Co-based active component can be improved by dispersing and anchoring Ti3C2Tx. On the other hand, the stacking of Ti3C2Tx can be prevented by Co-based active components. Therefore, it is expected that the electrocatalyst with high HER activity can be prepared through the synergistic action of the carrier and cobalt-based catalyst.
Metal-supported electrocatalysts are generally prepared via the high-temperature H2 reduction of metal. However, due to the hightemperature, the H2 reduction method has the characteristics of a strong heating effect, and it may accelerate the growth of particles on the catalyst surface and the agglomeration of metal active centers, resulting in the reduction of metal dispersion and specific surface area of the catalyst, which has a negative impact on the catalytic performance. In order to reduce the agglomeration of metal active centers during the reduction process, using NaBH4 as a reducing agent is an effective strategy for preparing metal-supported electrocatalysts. Compared with the H2 reduction method, the NaBH4 reduction process is carried out in a solution at room temperature so that the solvent can disperse the reactants and quickly absorb the heat generated by the exothermic reaction, thus slowing down the agglomeration of metal active centers in the reduction process.
However, the synthesis of catalysts suitable for the HER by supporting cobalt-based active components with Ti3C2Tx is rarely reported. Therefore, in this study, Ti3C2Tx was prepared using ultrasonic-assisted HF etching of Ti3AlC2 and then used as the carrier to support the Co-based active component (Co/Co3O4). Subsequently, the Co/Co3O4/Ti3C2Tx catalyst was prepared through the liquid-phase reduction method using NaBH4 as the reducing agent. The structure, morphology, crystal phase and elemental valence of the obtained catalyst were analyzed using various characterization methods. The activity and stability of Co/Co3O4/Ti3C2Tx during electrocatalysis of the HER were studied in electrolytes of 1 mol L1 KOH and 0.5 mol L1 H2SO4.

2. Materials and Methods

2.1. Reagents

The reagents used are as follows: Co(NO3)2·6H2O (analytical grade, Sinopharm Chemical Reagent Co., Ltd., Shanghai, China), NaBH4 (analytical grade, Sinopharm Chemical Reagent Co., Ltd., Shanghai, China), Ti3AlC2 (analytical grade, Zhejiang Yamei Nano Technology Co., Ltd., Jiaxing, China) and HF (analytical grade, Shanghai Aladdin Technology Co., Ltd., Shanghai, China).

2.2. Synthesis of Co/Co3O4/Ti3C2Tx Catalyst

Ti3C2Tx preparation was performed as follows: First, 20 mL of 48% aqueous HF solution was transferred to a Teflon beaker with a volume of 200 mL. Subsequently, 1.00 g of Ti3AlC2 powder was slowly added to the above-mentioned beaker within 30 min and electromagnetically stirred for 24 h at 150 rpm and 55 °C. The resulting solution was centrifuged at 6400 rpm for 10 min to yield the sediment and recycle the supernatant (including the HF solution and etched Al layers), respectively. The obtained sediment was cleaned with deionized water several times until the pH of the dispersion was close to 6. Finally, the resulting powder (Ti3C2Tx) was dried in a vacuum oven at 60 °C for 24 h. The dried powder was obtained and identified as Ti3C2Tx (Tx stands for O, F and OH functional groups).
The Co/Co3O4/Ti3C2Tx catalyst preparation occurred as follows: First, 0.30 g Ti3C2Tx was added to a beaker containing 25 mL of deionized water and treated with an ultrasonic device in an ice water bath for 2 to 3 h. Subsequently, the Ti3C2Tx solution was obtained after the above-mentioned solution was clearly stratified; second, 0.8 g of Co(NO3)2·6H2O was added to the Ti3C2Tx solution and electromagnetically stirred for 30 min at 800 rpm and 25 °C. After standing for 3 h, the solution became clear and stratified to obtain a Co/Co3O4/Ti3C2Tx solution; third, some NaBH4 was added until the solution was clear and layered (the upper layer was colorless and transparent), magnetically stirred at 800 rpm for 30 min and left for 12 h. The white liquid in the beaker was poured away, and the remaining sediment was dried in the freeze dryer for 48 h to obtain Co/Co3O4/Ti3C2Tx. The preparation process of Co/Co3O4/Ti3C2Tx is shown in Figure 1.

2.3. Catalyst Characterization

The functional groups of Ti3C2Tx were identified using Fourier infrared spectroscopy (FT-IR) (Nicolet 5700, Madison, WI, USA). The crystalline phases of the obtained catalyst were analyzed using X-ray diffraction (XRD) (PANalytical, Empyean, Almelo, The Netherlands) with Cu Kα radiation (λ = 0.15418 nm) at 40 kV and 40 mA over a 2θ range from 5° and 85°, and the XRD results of the samples were compared with existing standard cards to determine their phase structure. The surface morphologies and chemical compositions of the catalysts were investigated using a scanning electron microscope (SEM) (MLA650F, FEI Company, Hillsboro, OR, USA) equipped with an energy-dispersive X-ray spectroscope (EDS, Quantax 400, Bruker, Ettlingen, Germany) at an accelerating voltage of 20 kV. The size, shape and distribution of the obtained catalyst were analyzed using a transmission electron microscope (TEM) (TecnaiG2-20, FEI Company, Hillsboro, OR, USA). The chemical valence states of the elements in the catalyst were measured using X-ray photoelectron spectroscopy (XPS) (Escalab Xi+, Thermo Fisher Scientific Inc., Oxford, UK) under the vacuum condition of 8 × 1010 Pa. A monochromatic Al Kα X-ray source (hv = 1486.6 eV) was used with a working voltage of 12 kV. The specific surface area and pore size distribution of the catalyst were measured using a N2 physical adsorption instrument (ASAP 2020, Micromeritics Instrument Ltd., Norcross, GA, USA) at 77 K. The contact angle of the catalyst was measured using a contact angle measuring instrument (JY-Pha, Chengde Yote Instrument Manufacturing Co. Ltd., Chengde, China) to determine the surface wettability.

2.4. Electrocatalytic Hydrogen Evolution Test of Catalyst

For the preparation of test electrode, first, 1 mg of the catalyst was weighed and ground to powder; second, it was dispersed in a solution composed of 0.25 mL of isopropyl alcohol, 0.75 mL of deionized water and 20 μL of 5 wt% Nafion solution; third, the suspension was dispersed via ultrasonic treatment under the condition of an ice-water bath until 1 mL of uniformly dispersed suspension was formed. Finally, the solution containing the catalyst was coated on the surface of the carbon paper with an area of 1 cm2, and the working electrode supported by the catalyst was obtained after completely drying.
The HER performance of the catalyst was tested on a CHI660E electrochemical workstation at 25 °C. A standard three-electrode system was adopted, with a spectral pure graphite rod as the auxiliary electrode and a saturated calomel electrode (SCE) as the reference electrode. The potentials are all relative to the reversible hydrogen electrode (RHE). The conversion formula is shown in Equation (1):
V ( R H E ) = V ( S C E ) + 0.0591 × p H + 0.2415
where V(RHE) is the potential value (V) of the RHE, V(SCE) is the potential value (V) of the SCE, and pH is the pH value of the solution.
In this study, the non-Faraday, dual-layer capacitance currents corresponding to the cyclic voltammetry curve (CV) of the catalyst under different scanning rates (10, 20, 30, 40, 50, 60, 70, 80, 90 and 100 mV s1) were tested. After linear fitting, the cyclic voltammetry curves were developed. The capacitance of the electrochemical double-layer capacitance (Cdl) was calculated using Equation (2):
I = V × C d l
where V is the scanning rate (mV s1), and I is the current density at different scanning rates (mA cm2).
In order to evaluate the true electrocatalytic activity of the catalyst, the electrochemically active surface area (ECSA, cm2) involved in the electrochemical catalytic reaction should be calculated. The ECSA value can be calculated from the Cdl value using the Randles–Sevcik equation, as shown in Equation (3):
E C S A = C d l C s
where the Cs value is the specific capacitance (mF) of the corresponding surface smooth sample under the same conditions. The Cs value is usually in the range of 20–60 mF [17], and the Cs value was set as 40 mF in this study.
In general, the electrocatalytic activities of different catalysts can be compared by their overpotential (η) values at a current density of 10 mA cm2. η values can be calculated using the linear sweep voltammetry (LSV) curve derived from the linear sweep voltammetry (LSV) method, which is usually applied to the electrode with a continuously linear voltage (−0.6 V vs. RHE), and the resulting current is monitored. Since this type of current depends on the rate of oxidation or reduction occurring at the electrode, the size of the current in the LSV curve can reflect the degree of HER. In this study, the scanning rate was set at 5 mV s−l for LSV testing. Then, a line at y = −10 was made, and the abscissa of the intersection point of the LSV curve was the η value after taking the absolute value.
The Tafel slope (b) is the rate at which the current increases with respect to η, which is mainly determined by the transfer coefficient. The smaller the b value is, the faster the current density increases and the smaller the η change, indicating that the rate-determining step is at the end of the multi-electron transfer reaction, which usually marks an electrocatalyst with excellent performance. The current density and potential data of the LSV diagram are converted into a graph of log(current density i) and η values; that is, the Tafel diagram is obtained. Then, the linear fitting was carried out through Origin 8 software, and the b value was calculated using Equation (4):
η = b log I + a
where η is the overpotential (mV), I is the current density (mA cm2), a is the constant, and b is the Tafel slope.
The larger the surface area of the catalyst, the more active sites are exposed to the electrocatalyst surface, and the mass activity (MA) value is more indicative of the true activity of the catalyst. In general, a smaller-sized catalyst has higher MA value because a smaller-sized particle means that the ratio of surface atoms to total atoms is larger, exposing more electrocatalytic active sites. In order to more intuitively compare the activities of different catalysts, the MA values can be calculated according to Equation (5):
M A = I / M
where MA (mA mg1) is the mass activity of the catalyst, I is the current density (mA cm2), and M is the catalyst load (mg cm2).
The stability of the catalyst was tested using a chronocurrent method (i − t), which is usually performed at a constant potential and monitors changes in the electrode current. In addition, the stability of the electrocatalyst was tested using the CV method with more than 1000 cycles, which was achieved by comparing the LSV data before and after the stability test. All CV, LSV, i − t and other tests were performed in electrolytes of 1 mol L1 KOH or 0.5 mol L1 H2SO4, and the obtained data were not corrected by IR compensation.

3. Results

3.1. Characterization of Catalyst

The FT-IR result of the Ti3C2Tx is shown in Figure 2. There are three characteristic peaks of Ti3C2Tx in Figure 2, corresponding to C=O stretching vibrations, O-H bending vibrations and O-H stretching vibrations at 1633, 1393 and 3446 cm−1, respectively.
The surface morphologies of Ti3C2Tx and Co/Co3O4/Ti3C2Tx were characterized using an SEM, and the results are shown in Figure 3. It can be observed from Figure 3a that the Ti3C2Tx obtained via HF etching Ti3AlC2 has an obvious “organ-like” morphology, which is conducive to the active substance to penetrate the internal structure of Ti3AlC2. The morphology is formed due to the destruction of Ti-Al bonds between the etched layers, which is consistent with other researchers’ reports [18]. It can be clearly observed from Figure 3b that the surface of the obtained Co/Co3O4/Ti3C2Tx has particle clusters with different sizes and shapes. The formation of these particles can be attributed as follows: during the reduction reaction, a large amount of Co2+ is reduced by NaBH4 to zero-valent Co, which accumulates and deposits on the surface of Ti3C2Tx [19].
In order to understand the types and distribution of elements in the catalyst, the Co/Co3O4/Ti3C2Tx catalyst shown in Figure 4a was analyzed using element surface scanning, and the results are shown in Figure 4b–f.
Many bright and evenly distributed areas can be seen in Figure 4b, indicating that the elements Co and O are evenly distributed on the surface of Ti3C2Tx, which helps expose more active sites on the catalyst surface. Then, six square regions were randomly selected on the surface of Co/Co3O4/Ti3C2Tx, and the element information of each square region was detected. The results are shown in Figure 4g. It can be seen from Figure 4g that the atomic proportion of the same element at each detection point is close, which further indicated that the distribution of the elements Co and O is relatively uniform. The average atomic contents of C, O, Ti and Co are 22.16%, 58.10%, 4.95% and 14.80%, respectively. The high O content in Co/Co3O4/Ti3C2Tx indicates the presence of cobalt oxide.
In order to further explore the structural information of catalysts, the morphologies of Ti3C2Tx and Co/Co3O4/Ti3C2Tx were observed using a TEM, and the results are shown in Figure 5. It can be seen from Figure 5a that Ti3C2Tx presents a unique two-dimensional sheet structure with many active edge sites, which can provide a large number of attachment sites for the loaded active substances. As shown in Figure 5b, a large number of clusters are attached to the surface of the layered Co/Co3O4/Ti3C2Tx. This may be because NaBH4 reduced a large amount of Co2+ to Co, and then, Co particles gathered and deposited on the surface of Ti3C2Tx [20], which was consistent with the results of the SEM analysis.
The crystal structures and compositions of Ti3AlC2, Ti3C2Tx and Co/Co3O4/Ti3C2Tx were obtained using XRD. It can be seen from the XRD patterns of Ti3AlC2 and Ti3C2Tx shown in Figure 6a that after HF etching of Ti3AlC2, obvious characteristic diffraction peaks appeared at 2θ of 9.5°, 19.3° and 39.0°. By comparing the standard card of Ti3AlC2 (JCPDS NO.52-0875), it can be inferred that the above characteristic diffraction peaks correspond to the (002), (004) and (104) crystal faces, respectively. The characteristic diffraction peak (104) of Al in the MAX structure almost completely disappeared, indicating that the Al layer had been completely etched. In addition, Ti3AlC2 shows a new wide characteristic diffraction peak at 2θ of 6.1°, indicating that the (002) characteristic diffraction peak corresponding to MXene moved from 9.5° to a lower angle of 6.1°. According to the Bragg equation, the layer spacing of MXene obtained by etching the MAX phase was enlarged. The larger layer spacing facilitates the active substances to enter the interior of Ti3C2Tx, and the active components can support the structure and prevent the stacking of Ti3C2Tx [20,21]. The XRD pattern of Co/Co3O4/Ti3C2Tx is shown in Figure 6b. Four characteristic diffraction peaks appeared at 2θ of 41.6°, 44.6°, 47.4° and 75.5°, respectively. By comparing the standard card of Co (JCPDS NO.05-0727), it can be inferred that the above characteristic diffraction peaks correspond to the (100), (002), (101) and (110) crystal faces, respectively. At the same time, six characteristic diffraction peaks appeared at 2θ of 34.1°, 39.6°, 57.3°, 68.4, 71.9 and 85.4°, respectively. By comparing the standard card of CoO (JCPDS NO. 42-1300), it can be inferred that the above characteristic diffraction peaks correspond to the (111), (200), (220), (311), (222) and (400) crystal faces. In addition, the XRD pattern of Co/Co3O4/Ti3C2Tx shows that there are seven characteristic diffraction peaks at 2θ of 19.0°, 31.3°, 36.7°, 38.5°, 44.8°, 59.3° and 65.3°. Compared with the standard card of Co3O4 (JCPDS NO.43-1003), it can be inferred that the above characteristic peaks correspond to the crystal faces of (111), (220), (311), (222), (400), (511) and (440), respectively. This also proves that the Co element in Co/Co3O4/Ti3C2Tx exists in metallic and oxidized states (Co, CoO and Co3O4). In addition, Co/Co3O4/Ti3C2Tx showed eleven characteristic diffraction peaks corresponding to MXenes at 2θ of 9.5°, 19.0°, 33.9°, 36.6°, 39.1°, 41.6°, 48.3°, 56.7°, 60.0°, 65.3°, 70.2° and 73.9°, respectively. In summary, the Co/Co3O4/Ti3C2Tx catalyst was successfully prepared.
The elemental composition of the Co/Co3O4/Ti3C2Tx catalyst was analyzed using XPS. Figure 7a shows the full XPS spectrum of Co/Co3O4/Ti3C2Tx, and binding energy peaks corresponding to Co, O, Ti, C and other elements can be clearly seen in the full spectrum. Furthermore, the fine spectrum of C1s was obtained via narrow-area scanning (Figure 7b), which can be divided into three peaks with binding energies of 284.8 eV, 286.4 eV and 289.0 eV, corresponding to C-C, O-C and O-C=O, respectively [22,23]. Figure 7c shows the fine spectrum of Co2p, which can be divided into three peaks with binding energies of 780.1 eV, 781.1 eV and 782.7 eV, respectively, attributed to Co0, Co3+ and Co2+ [24], indicating that Co exists in the form of metallic Co and in the oxidation state as Co3O4, which is consistent with the XRD analysis results. It is worth noting that most of the Co2p peaks are Co2+. Figure 7d shows the fine spectrum of O1s, which can be divided into three characteristic peaks with binding energies of 530.7 eV, 531.2 eV and 532.2 eV, respectively, attributed to Co-O, −OH and O in the adsorbed H2O [25], which means that O in the catalyst exists in two forms: lattice oxygen and surface hydroxyl oxygen. The surface of the catalyst contains both cobalt oxide (Co-O) and cobalt hydroxide (Co-OH), which can effectively accelerate the dissociation of water [26,27]. In addition, the appearance of the binding energy peak corresponding to oxygen in the adsorbed H2O also indicates that the catalyst has a certain hydrophilicity.
The hydrophilicity of the catalyst has a great influence on the HER. Hence, the hydrophilicity of the catalyst was detected by measuring the contact angles of Ti3C2Tx, Co and Co/Co3O4/Ti3C2Tx, and the results are shown in Figure 8a–c. As can be seen from Figure 8a–c that the contact angles of Ti3C2Tx, Co and Co/Co3O4/Ti3C2Tx are 141.9°, 122.1° and 11.7°, respectively. The contact angle of Co/Co3O4/Ti3C2Tx is much smaller than those of Ti3C2Tx and Co, which indicates that the hydrophilicity of Co/Co3O4/Ti3C2Tx is much stronger than those of Ti3C2Tx and Co. The possible reasons are as follows: a portion of Co on the surface of Co/Co3O4/Ti3C2Tx is oxidized to Co3O4, which endows the catalyst to have a higher degree of hydrophilicity. When the HER occurs, the stronger the hydrophobicity of the catalyst surface, the shorter the retention time of the generated bubbles on the catalyst surface and the smaller the bubble size when leaving the surface, which means that the reaction speed of the HER is faster [28].
The specific surface area and aperture of Ti3C2Tx, Co and Co/Co3O4/Ti3C2Tx were measured using the specific surface area tester. Figure 9a shows the N2 adsorption–desorption curve, in which the N2 adsorption–desorption isothermal curve of Co/Co3O4/Ti3C2Tx is type IV, indicating that there was a H3 hysteresis loop at a relative pressure p/p0 of 0.45–0.9 caused by the capillary condensation of adsorbate. Therefore, the Co/Co3O4/Ti3C2Tx had a mesoporous structure, and the pore structure was irregular. The pore diameters and specific surface areas of the catalysts are listed in Table 1. The specific surface area, pore volume and pore size of Ti3C2Tx were 2.94 m2 g1, 0.0076 cm3 g1 and 10.71 nm. For comparison, the specific surface area of Co was 64.02 m2 g1, the pore volume was 0.30 cm3 g1, and the pore size was 186.79 nm. When Ti3C2Tx was loaded with Co/Co3O4, the specific surface area of Co/Co3O4/Ti3C2Tx became 17.64 m2 g1, and the pore volume was 0.04 cm3 g1; both of them were increased. The pore size of Co/Co3O4/Ti3C2Tx was 10.07 nm, which was similar to the pore size of Ti3C2Tx. It is worth noting that compared with Ti3C2Tx, the specific surface area, pore volume and pore diameter of Co/Co3O4/Ti3C2Tx are greatly increased, which may be due to Co/Co3O4 infiltrating the interior of Ti3C2Tx and effectively inhibiting the accumulation of Ti3C2Tx flakes. This helps widen the layer spacing of the catalyst, thereby increasing the contact area between the catalyst and electrolyte, providing more active sites for the HER. In addition, the pore size distributions of Ti3C2Tx, Co and Co/Co3O4/Ti3C2Tx are shown in Figure 9b. Compared with Ti3C2Tx and Co/Co3O4/Ti3C2Tx, the pore size distribution of Co is wider, and the mesoporous distributions of Co and Co/Co3O4/Ti3C2Tx were mainly in the range of 35–40 nm and 5–20 nm, respectively.

3.2. Electrocatalytic HER Performance Test in Acidic Electrolyte

The HER electrocatalytic activities of Ti3C2Tx, Co and Co/Co3O4/Ti3C2Tx were tested in a 0.5 mol L−1 H2SO4 solution, and the LSV results are shown in Figure 10a. It can be seen from Figure 9a that the electrocatalytic activities of Ti3C2Tx and Co are very low. However, when Ti3C2Tx and Co formed Co/Co3O4/Ti3C2Tx, the electrocatalytic performance was significantly improved. Furthermore, the η values corresponding to Ti3C2Tx, Co and Co/Co3O4/Ti3C2Tx at a current density of 10 mA cm−2 were obtained, and they were 599 mV, 517 mV and 86 mV. However, after only four LSV tests, the η value of Co/Co3O4/Ti3C2Tx was changed from an initial 86 mV to 225 mV (see Figure 10b). The results show that Co/Co3O4/Ti3C2Tx has poor durability in a 0.5 mol L−1 H2SO4 solution and is not suitable for long-term use in an acidic medium. The reasons for the sharp decline in the catalytic activity of Co/Co3O4/Ti3C2Tx in an acidic electrolyte may be as follows: Co and Co3O4 were easily dissolved in strong acids. The stability test shows that the Co/Co3O4/Ti3C2Tx catalyst is not suitable for long-term use in an acidic medium.

3.3. Electrocatalytic HER Performance Test in Alkaline Electrolyte

The HER performances of Ti3C2Tx, Co and Co/Co3O4/Ti3C2Tx were tested in a 1 mol L−1 KOH solution, and the obtained LSV results are shown in Figure 11a. The η values corresponding to the above catalysts at a current density of 10 mA cm−2 were obtained: Co/Co3O4/Ti3C2Tx has a smaller η value (87 mV) than Co (443 mV) and Ti3C2Tx (541 mV). Thus, the trend of catalytic activity was Co/Co3O4/Ti3C2Tx > Co > Ti3C2Tx. In 1 mol L−1 KOH, both Co and Ti3C2Tx require large η values to perform the HER, which indicates that both Co and Ti3C2Tx have poor HER activity. The CV test results of Ti3C2Tx, Co and Co/Co3O4/Ti3C2Tx under different scanning rates (10, 20, 30, 40, 50, 60, 70, 80, 90 and 100 mV s−1) are shown in Figure 11b–d. Based on the CV test results of each catalyst, their Cdl values were obtained according to Equation (2). The Cdl values of Ti3C2Tx, Co and Co/Co3O4/Ti3C2Tx were 9.86 mF cm−2, 7.59 mF cm−2 and 28.34 mF cm−2, respectively (see Figure 11e). According to Equation (3), the Cdl value can be converted to the ECSA value, and the ECSA values of Ti3C2Tx, Co and Co/Co3O4/Ti3C2Tx were 246.50 cm2 189.75 cm2 and 708.5 cm2, respectively. It can be seen that Co/Co3O4/Ti3C2Tx has the largest electrochemically active area. The Tafel slope values of Ti3C2Tx, Co and Co/Co3O4/Ti3C2Tx in a 1 mol L−1 KOH electrolyte were calculated according to Equation (4). The Tafel slope value of Co/Co3O4/Ti3C2Tx was 61.90 mV dec−1, which is better than the 263.09 mV dec−1 of Co and 324.46 mV dec−1 of Ti3C2Tx (seen Figure 11f). The smaller Tafel slope value indicates that the Co/Co3O4/Ti3C2Tx catalyst has a fast kinetic process in a 1 mol L−1 KOH solution. Meanwhile, the Tafel slope values indicate that the Co/Co3O4/Ti3C2Tx catalyst follows the Volmer–Heyrovsky mechanism. Co/Co3O4 can accelerate the adsorption and deionization of H2O to synthesize H2. When water molecules are adsorbed on the Co/Co3O4 interface, the Co3O4 phase helps to split the HO-H bond and produce adsorbed hydroxide ions (OH) and hydrogen atoms (Hads), while the Co phase can integrate Hads to produce H2 [29].
The long-term durability of the Co/Co3O4/Ti3C2Tx catalyst was tested in a 1 mol L−1 KOH electrolyte, and the LSV results are shown in Figure 11g. Compared with the initial curve, the LSV curve did not change significantly after 1000 CV cycles, and the η value was only 8 mV different, indicating that Co/Co3O4/Ti3C2Tx had good cyclic stability in a 1 mol L−1 KOH electrolyte. In addition, in order to better reflect the activity changes of Co/Co3O4/Ti3C2Tx before and after 1000 CV tests, the MA value of Co/Co3O4/Ti3C2Tx before and after 1000 CV cycles was calculated from Equation (5). Before the stability test, the MA value of Co/Co3O4/Ti3C2Tx was 67.57 mA (mgCo)−1. After the stability test, the MA value of Co/Co3O4/Ti3C2Tx was 60.14 mA (mgCo)−1. The MA value of Co/Ti3C2Tx did not change much after 1000 CV cycles, and the activity of Co/Co3O4/Ti3C2Tx could still be maintained at 89% after 1000 CV cycles. The results show that Co/Co3O4/Ti3C2Tx has good catalytic stability in a 1 mol L1 KOH solution.
The electrocatalytic performance for the HER of the synthesized Co/Co3O4/Ti3C2Tx was compared with that of MXene-based catalysts reported in the literature, and the results are shown in Table 2. It could be seen from the table that the PtCo/Ti3C2Tx catalyst prepared in this study had better performance in terms of its overpotential and Tafel slope, making it a promising electrocatalyst for the HER.

4. Conclusions

In this study, the Co/Co3O4/Ti3C2Tx catalyst was synthesized by supporting Co/Co3O4 with Ti3C2Tx. In a 1 mol L1 KOH solution, the Co/Co3O4/Ti3C2Tx catalyst showed excellent HER performance, with a low overpotential of 87 mV at 10 mA cm−2 and a Tafel slope of 105.17 mV dec1. The smaller Tafel slope value indicates that the Co/Co3O4/Ti3C2Tx catalyst has a fast kinetic process and follows the Volmer–Heyrovsky mechanism in a 1 mol L1 KOH solution. The results also show that Co/Co3O4/Ti3C2Tx has good catalytic stability in a 1 mol L1 KOH solution, and the activity of Co/Co3O4/Ti3C2Tx could still be maintained at 89% after 1000 CV cycles. It can be expected that these results provide a possible new way to design and prepare a low-cost, high-performance HER catalyst.

Author Contributions

Conceptualization, B.W. and Q.S.; methodology, B.W. and W.G.; validation, W.G., B.W. and Q.S.; formal analysis, Q.S.; investigation, B.W.; resources, Q.S.; data curation, B.W. and W.G.; writing—original draft preparation, B.W. and W.G.; writing—review and editing, Q.S.; supervision, Q.S.; project administration, Q.S.; funding acquisition, Q.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The original contributions presented in the study are included in the article, further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Song, H.; Luo, S.; Huang, H.; Deng, B.; Ye, J. Solar-driven hydrogen production: Recent advances, challenges, and future perspectives. ACS. Energy. Lett. 2022, 7, 1043–1065. [Google Scholar] [CrossRef]
  2. Cheng, R.; Min, Y.; Li, H.; Fu, C. Electronic structure regulation in the design of low-cost efficient electrocatalysts: From theory to applications. Nano Energy 2023, 115, 108718. [Google Scholar] [CrossRef]
  3. Yusuf, B.A.; Yaseen, W.; Xie, M.; Zayyan, R.S.; Muhammad, A.I.; Nankya, R.; Xie, J.; Xu, Y. Recent advances in understanding and design of efficient hydrogen evolution electrocatalysts for water splitting: A comprehensive review. Adv. Colloid Int. Sci. 2023, 311, 102811. [Google Scholar] [CrossRef] [PubMed]
  4. Zhao, L.; Wu, R.; Wang, J.; Li, Z.; Wei, X.; Chen, J.S.; Chen, Y. Synthesis of noble metal-based intermetallic electrocatalysts by space-confined pyrolysis: Recent progress and future perspective. J. Energy Chem. 2021, 60, 61–74. [Google Scholar] [CrossRef]
  5. Wu, H.; Alkhatami, A.G.; Farhan, Z.A.; AbdalSalam, A.G.; Hamadan, R.; Aldarrji, M.Q.; Izzat, S.E.; Yosif, A.A.; Hadrawi, S.K.; Riyahi, Y.; et al. Recent developments in the production of hydrogen: Efficiency comparison of different techniques, economic dimensions, challenges and environmental impacts. Fuel Process. Technol. 2023, 248, 107819. [Google Scholar] [CrossRef]
  6. Wang, Y.; Yan, D.; Hankari, S.E.; Zou, Y.; Wang, S. Recent progress on layered double hydroxides and their derivatives for electrocatalytic water splitting. Adv. Sci. 2018, 5, 1800064. [Google Scholar] [CrossRef] [PubMed]
  7. Wang, C.; Li, Y.; Gu, C.; Zhang, L.; Wang, X.; Tu, J. Active Co@CoO core/shell nanowire arrays as efficient electrocatalysts for hydrogen evolution reaction. Chem. Eng. J. 2022, 429, 132226. [Google Scholar] [CrossRef]
  8. Liu, H.; He, P.; Wang, S.; Gao, J.; Zhou, L.; Li, C.; Zhang, Y.; Yang, D.; He, M.; Jia, L.; et al. Facile one-step fabrication of bimetallic Co–Ni–P hollow nanospheres anchored on reduced graphene oxide as highly efficient electrocatalyst for hydrogen evolution reaction. Int. J. Hydrogen. Energy 2019, 44, 24140–24150. [Google Scholar] [CrossRef]
  9. Li, T.; Niu, K.; Yang, M.; Shrestha, N.K.; Gao, Z.; Song, Y. Ultrathin CoS2 shells anchored on Co3O4 nanoneedles for efficient hydrogen evolution electrocatalysis. J. Power Sources 2017, 356, 89–96. [Google Scholar] [CrossRef]
  10. Zhang, X.; Bi, Y.; Wang, Y.; Liu, Q.; Zhang, Y.; Wang, J.; Chen, L. S co-doped graphene coupled with Ni2P/CoP-loaded carbon nanospheres: A novel catalyst for hydrogen evolution reaction. Mater. Today Sustain. 2024, 25, 100677. [Google Scholar] [CrossRef]
  11. Lin, B.; Chen, J.; Yang, R.; Mao, S.; Qin, M.; Wang, Y. Multi-hierarchical cobalt-based electrocatalyst towards high rate H2 production. Appl. Catal. B Environ. 2022, 316, 121666. [Google Scholar] [CrossRef]
  12. Hanan, A.; Lakhan, M.N.; Bibi, F.; Khan, A.; Soomro, I.A.; Hussain, A.; Aftab, U. MOFs coupled transition metals, graphene, and MXenes: Emerging electrocatalysts for hydrogen evolution reaction. Chem. Eng. J. 2024, 482, 148776. [Google Scholar] [CrossRef]
  13. Ali, S.A.; Sadiq, I.; Ahmad, T. Superlative porous organic polymers for photochemical and electrochemical CO2 reduction applications: From synthesis to functionality. Langmuir 2024, 40, 10414–10432. [Google Scholar] [CrossRef] [PubMed]
  14. Ali, S.A.; Ahmad, T. MBenes for energy conversion: Advances, bottlenecks, and prospects. Langmuir 2024, 40, 10835–10846. [Google Scholar] [CrossRef]
  15. Karthikeyan, S.C.; Sidra, S.; Ramakrishnan, S.; Kim, D.H.; Sagayaraj, P.J.; Sekar, K.; Yoo, D.J. Heterostructured NiO/IrO2 synergistic pair as durable trifunctional electrocatalysts towards water splitting and rechargeable zinc-air batteries: An experimental and theoretical study. Appl. Catal. B Environ. Energy 2024, 355, 124196. [Google Scholar] [CrossRef]
  16. Shaikh, Z.A.; Laghari, A.A.; Litvishko, O.; Litvishko, V.; Kalmykova, T.; Meynkhard, A. Liquid-phase deposition synthesis of ZIF-67-derived synthesis of Co3O4@TiO2 composite for efficient electrochemical water splitting. Metals 2021, 11, 420. [Google Scholar] [CrossRef]
  17. Sheng, H.; Ye, K.; Gao, Y.; Zhu, K.; Yan, J.; Wang, G.; Cao, D. Simultaneously boosting hydrogen production and ethanol upgrading using a highly-efficient hollow needle-like copper cobalt sulfide as a bifunctional electrocatalyst. J. Colloid Interface Sci. 2021, 602, 325–333. [Google Scholar] [CrossRef] [PubMed]
  18. Yang, Z.; Liu, A.; Wang, C.; Liu, F.; He, J.; Li, S.; Wang, J.; You, R.; Yan, X.; Sun, P.; et al. Improvement of gas and humidity sensing properties of organ-like MXene by alkaline treatment. ACS Sens. 2019, 4, 1261–1269. [Google Scholar] [CrossRef] [PubMed]
  19. Liu, S.; Deng, T.; Hu, X.; Shi, X.; Wang, H.; Qin, T.; Zhang, X.; Qi, J.; Zhang, W.; Zheng, W. Increasing surface active Co2+ sites of MOF-derived Co3O4 for enhanced supercapacitive performance via NaBH4 reduction. Electrochim. Acta 2018, 289, 319–323. [Google Scholar] [CrossRef]
  20. Cao, S.; Shen, B.; Tong, T.; Fu, J. 2D/2D heterojunction of ultrathin MXene/Bi2WO6 nanosheets for improved photocatalytic CO2 reduction. Adv. Funct. Mater. 2018, 28, 1800136. [Google Scholar] [CrossRef]
  21. Feng, A.; Yu, Y.; Wang, F.; Jiang, Y.; Yu, L.; Mi, L.; Song, L. Two-dimensional MXene Ti3C2 produced by exfoliation of Ti3AlC2. Mater. Design. 2017, 114, 161–166. [Google Scholar] [CrossRef]
  22. Ma, Y.Y.; Lang, Z.L.; Yan, L.K.; Wang, Y.H.; Tan, H.Q.; Feng, K.; Xia, Y.J.; Zhong, J.; Liu, Y.; Kang, Z.H.; et al. Highly efficient hydrogen evolution triggered by a multi-interfacial Ni/WC hybrid electrocatalyst. Energy Environ. Sci. 2018, 11, 2114–2123. [Google Scholar] [CrossRef]
  23. Eid, K.; Lu, Q.; Abdel-Azeim, S.; Soliman, A.; Abdullah, A.M.; Abdelgwad, A.M.; Forbes, R.P.; Ozoemena, K.I.; Varma, R.S.; Shibl, M.F. Highly exfoliated Ti3C2Tx MXene nanosheets atomically doped with Cu for efficient electrochemical CO2 reduction: An experimental and theoretical study. J. Mater. Chem. A 2022, 10, 1965–1975. [Google Scholar] [CrossRef]
  24. Islam, M.M.; Faisal, S.N.; Akhter, T.; Roy, A.K.; Minett, A.I.; Konstantinov, K.; Shi, X.D. Liquid-crystal-mediated 3D macrostructured composite of Co/Co3O4 embedded in graphene: Free-standing electrode for efficient water splitting. Part. Part. Syst. Character. 2017, 34, 1600386. [Google Scholar] [CrossRef]
  25. Ding, D.; Shen, K.; Chen, X.; Chen, H.; Chen, J.; Fan, T.; Wu, R.; Li, Y. Multi-level architecture optimization of MOF-templated Co-based nanoparticles embedded in hollow N-doped carbon polyhedra for efficient OER and ORR. ACS Catal. 2018, 8, 7879–7888. [Google Scholar] [CrossRef]
  26. McCrum, I.T.; Koper, M.T.M. The role of adsorbed hydroxide in hydrogen evolution reaction kinetics on modified platinum. Nat. Energy 2020, 5, 891–899. [Google Scholar] [CrossRef]
  27. Wu, J.; Fan, J.; Zhao, Y.; Wang, D.; Wang, H.; Liu, L.; Gu, Q.; Zhang, L.; Zheng, X.; Cui, D.J.; et al. Atomically dispersed MoOx on rhodium metallene boosts electrocatalyzed alkaline hydrogen evolution. Angew. Chem. Int. Edit. 2022, 61, 202207512. [Google Scholar]
  28. Do, H.H.; Tekalgne, M.A.; Tran, V.A.; Le, Q.V.; Cho, J.H.; Ahn, S.H.; Kim, S.Y. MOF-derived Co/Co3O4/C hollow structural composite as an efficient electrocatalyst for hydrogen evolution reaction. Fuel 2022, 329, 125468. [Google Scholar] [CrossRef]
  29. Iqbal, M.F.; Gao, W.; Mao, Z.; Hu, E.; Gao, X.; Zhang, J.; Chen, Z. Strategies to enhance the electrocatalytic behavior of metal selenides for hydrogen evolution reaction: A review. Int. J. Hydrogen. Energy 2023, 48, 36722–36749. [Google Scholar] [CrossRef]
  30. Ma, C.; He, H.; Qin, J.; Luo, L.; Lan, Y.; Zhang, J.; Yang, L.; Jiang, Q.; Huang, H. The marriage of hydrazone-linked covalent organic frameworks and MXene enables efficient electrocatalytic hydrogen evolution. Small Struct. 2024, 5, 2300279. [Google Scholar] [CrossRef]
  31. Ma, C.; He, H.; Qin, J.; Hao, L.; Jia, L.; Yang, L.; Huang, H. Combining MXene nanosheets with iron-based metal-organic frameworks for enhanced electrocatalytic hydrogen evolution reaction. Mater. Today Chem. 2023, 30, 101531. [Google Scholar] [CrossRef]
  32. Seh, Z.W.; Fredrickson, K.D.; Anasori, B.; Kibsgaard, J.; Strickler, A.L.; Lukatskaya, M.R.; Gogotsi, Y.; Jaramillo, T.F.; Vojvodic, A. Two-dimensional molybdenum carbide (MXene) as an efficient electrocatalyst for hydrogen evolution. ACS Energy Lett. 2016, 1, 589–594. [Google Scholar] [CrossRef]
  33. Wu, X.; Wang, Z.; Yu, M.; Xiu, L.; Qiu, J. Stabilizing the MXenes by carbon nanoplating for developing hierarchical nanohybrids with efficient lithium storage and hydrogen evolution capability. Adv. Mater. 2017, 29, 1607017. [Google Scholar] [CrossRef] [PubMed]
  34. Liang, J.; Ding, C.; Liu, J.; Chen, T.; Peng, W.; Li, Y.; Zhang, F.; Fan, X. Heterostructure engineering of Co-doped MoS2 coupled with Mo2CTx MXene for enhanced hydrogen evolution in alkaline media. Nanoscale 2019, 11, 10992–11000. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic diagram of Co/Co3O4/Ti3C2Tx preparation.
Figure 1. Schematic diagram of Co/Co3O4/Ti3C2Tx preparation.
Metals 14 00922 g001
Figure 2. FT-IR spectrum of Ti3C2.
Figure 2. FT-IR spectrum of Ti3C2.
Metals 14 00922 g002
Figure 3. SEM images of Ti3C2Tx and Co/Co3O4/Ti3C2Tx. (a) Ti3C2Tx; (b) Co/Co3O4/Ti3C2Tx.
Figure 3. SEM images of Ti3C2Tx and Co/Co3O4/Ti3C2Tx. (a) Ti3C2Tx; (b) Co/Co3O4/Ti3C2Tx.
Metals 14 00922 g003
Figure 4. Elemental analysis diagram of Co/Co3O4/Ti3C2Tx. (a) SEM diagram of Co/Co3O4/Ti3C2Tx; (b) EDS mapping of Co/Co3O4/Ti3C2Tx and (c) Co element, (d) O element, (e) C element and (f) Ti element; (g) atomic ratios of Co, O, C and Ti elements.
Figure 4. Elemental analysis diagram of Co/Co3O4/Ti3C2Tx. (a) SEM diagram of Co/Co3O4/Ti3C2Tx; (b) EDS mapping of Co/Co3O4/Ti3C2Tx and (c) Co element, (d) O element, (e) C element and (f) Ti element; (g) atomic ratios of Co, O, C and Ti elements.
Metals 14 00922 g004
Figure 5. TEM images of (a) Ti3C2Tx and (b) Co/Co3O4/Ti3C2Tx.
Figure 5. TEM images of (a) Ti3C2Tx and (b) Co/Co3O4/Ti3C2Tx.
Metals 14 00922 g005
Figure 6. XRD patterns of Ti3AlC2, Ti3C2Tx and Co/Co3O4/Ti3C2Tx. (a) Ti3AlC2 and Ti3C2Tx; (b) Co/Co3O4/Ti3C2Tx.
Figure 6. XRD patterns of Ti3AlC2, Ti3C2Tx and Co/Co3O4/Ti3C2Tx. (a) Ti3AlC2 and Ti3C2Tx; (b) Co/Co3O4/Ti3C2Tx.
Metals 14 00922 g006
Figure 7. XPS diagram of Co/Co3O4/Ti3C2Tx. (a) XPS full spectrum, (b) C1s spectrum, (c) Co2p spectrum and (d) O1s spectrum.
Figure 7. XPS diagram of Co/Co3O4/Ti3C2Tx. (a) XPS full spectrum, (b) C1s spectrum, (c) Co2p spectrum and (d) O1s spectrum.
Metals 14 00922 g007aMetals 14 00922 g007b
Figure 8. Contact angles of Ti3C2Tx, Co and Co/Co3O4/Ti3C2Tx. (a) Ti3C2Tx, (b) Co and (c) Co/Co3O4/Ti3C2Tx.
Figure 8. Contact angles of Ti3C2Tx, Co and Co/Co3O4/Ti3C2Tx. (a) Ti3C2Tx, (b) Co and (c) Co/Co3O4/Ti3C2Tx.
Metals 14 00922 g008
Figure 9. N2 adsorption–desorption curves and pore size distributions of Ti3C2Tx, Co and Co/Co3O4/Ti3C2Tx. (a) N2 adsorption–desorption curve; (b) pore size distribution (0–600 nm).
Figure 9. N2 adsorption–desorption curves and pore size distributions of Ti3C2Tx, Co and Co/Co3O4/Ti3C2Tx. (a) N2 adsorption–desorption curve; (b) pore size distribution (0–600 nm).
Metals 14 00922 g009
Figure 10. LSV diagram of (a) Ti3C2Tx, Co and Co/Co3O4/Ti3C2Tx; (b) Co/Co3O4/Ti3C2Tx (stability test) in 0.5 mol L−1 H2SO4 electrolyte.
Figure 10. LSV diagram of (a) Ti3C2Tx, Co and Co/Co3O4/Ti3C2Tx; (b) Co/Co3O4/Ti3C2Tx (stability test) in 0.5 mol L−1 H2SO4 electrolyte.
Metals 14 00922 g010
Figure 11. (a): LSV diagram of Ti3C2Tx, Co and Co/Co3O4/Ti3C2Tx in 1 mol L−1 KOH electrolyte; CV results of (b) Ti3C2Tx, (c) Co and (d) Co/Co3O4/Ti3C2Tx under different scanning rates (10, 20, 30, 40, 50, 60, 70, 80, 90 and 100 mV s−1); (e) Cdl value and (f) Tafel diagram of Ti3C2Tx, Co and Co/Co3O4/Ti3C2Tx; (g) LSV diagram of Co/Co3O4/Ti3C2Tx (stability test) in 1 mol L−1 KOH electrolyte.
Figure 11. (a): LSV diagram of Ti3C2Tx, Co and Co/Co3O4/Ti3C2Tx in 1 mol L−1 KOH electrolyte; CV results of (b) Ti3C2Tx, (c) Co and (d) Co/Co3O4/Ti3C2Tx under different scanning rates (10, 20, 30, 40, 50, 60, 70, 80, 90 and 100 mV s−1); (e) Cdl value and (f) Tafel diagram of Ti3C2Tx, Co and Co/Co3O4/Ti3C2Tx; (g) LSV diagram of Co/Co3O4/Ti3C2Tx (stability test) in 1 mol L−1 KOH electrolyte.
Metals 14 00922 g011
Table 1. Specific surface areas (BET), pore volumes (single-point method) and average pore diameters of Ti3C2Tx, Co and Co/Co3O4/Ti3C2Tx.
Table 1. Specific surface areas (BET), pore volumes (single-point method) and average pore diameters of Ti3C2Tx, Co and Co/Co3O4/Ti3C2Tx.
CatalystSpecific Surface Area (m2/g)Pore Volume
(cm3/g)
Average Pore Diameter (nm)
Ti3C2Tx2.940.007610.71
Co64.020.30186.79
Co/Co3O4/Ti3C2Tx17.640.04410.07
Table 2. Comparison of electrocatalytic performances for HERs of representative MXene-based catalysts.
Table 2. Comparison of electrocatalytic performances for HERs of representative MXene-based catalysts.
Type of ElectrocatalystElectrolyteElectrochemical PerformanceRefs.
COF/Ti3C2Tx0.5 M H2SO4A very low onset potential of 19 mV and a small Tafel slope of 50 mV dec−1[30]
MIL/Ti3C2Tx0.5 M H2SO4An operating overpotential of 107 mV at 10 mA/cm2[31]
Mo2CTx0.5 M H2SO4A low overpotential of 283 mV at 10 mA cm−2 and an average TOF of −0.02 H2 s−1 at 200 mV[32]
MoS2/Ti3C2/C0.5 M H2SO4A low overpotential of 135 mV at 10 mA cm−2 and a Tafel slope of 45 mV dec−1[33]
Co-MoS2/Mo2CTx1 M KOHA low overpotential of 112 mV at 10 mA cm−2 and a Tafel slope of 82 mV dec−1[34]
Co/Co3O4/Ti3C2Tx1 M KOHA low overpotential of 87 mV at 10 mA cm−2 and a Tafel slope of 105.17 mV dec−1This work
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Guo, W.; Wang, B.; Shu, Q. An Efficient and Stable MXene-Immobilized, Cobalt-Based Catalyst for Hydrogen Evolution Reaction. Metals 2024, 14, 922. https://doi.org/10.3390/met14080922

AMA Style

Guo W, Wang B, Shu Q. An Efficient and Stable MXene-Immobilized, Cobalt-Based Catalyst for Hydrogen Evolution Reaction. Metals. 2024; 14(8):922. https://doi.org/10.3390/met14080922

Chicago/Turabian Style

Guo, Wei, Buxiang Wang, and Qing Shu. 2024. "An Efficient and Stable MXene-Immobilized, Cobalt-Based Catalyst for Hydrogen Evolution Reaction" Metals 14, no. 8: 922. https://doi.org/10.3390/met14080922

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop